Esteban
Zuñiga-Puelles
*a,
Ayberk
Özden
b,
Raul
Cardoso-Gil
c,
Christoph
Hennig
de,
Cameliu
Himcinschi
b,
Jens
Kortus
b and
Roman
Gumeniuk
*a
aInstitut für Experimentelle Physik, TU Bergakademie Freiberg, Leipziger Straße 23, 09596 Freiberg, Germany. E-mail: estebanzp1@outlook.com; gumeniuk@physik.tu-freiberg.de
bInstitut für Theoretische Physik, TU Bergakademie Freiberg, Leipziger Straße 23, 09596 Freiberg, Germany
cMax-Planck-Institut für Chemische Physik Fester Stoffe, Nöthnitzer Straße 40, 01187 Dresden, Germany
dHelmholtz-Zentrum Dresden-Rossendorf, Institute of Resource Ecology, Bautzner Landstraße 400, 01328 Dresden, Germany
eThe Rossendorf Beamline at ESRF, CS 40220, 38043 Grenoble Cedex 9, France
First published on 3rd March 2025
The crystal structures of low- (α) and high-temperature (β) modifications of In2Te3 are refined for the first time. They crystallize with unique noncentrosymmetric face-centered cubic and half-Heusler arrangements, respectively. Both models are related with each other via a group–subgroup formalism. The small changes in unit-cell volume and entropy at the α → β structural phase transition indicate that it is of 2nd order. The strong structural disorder in α-In2Te3 results in a temperature-dependent behavior of the electrical resistivity similar to that of doped semiconductors (with a well-defined maximum) and extremely low thermal conductivity [κ(T) ≤ 0.7 W m−1 K−1]. Additional reasons for the poor electrical and thermal transport are low charge-carrier concentration and mobility, as well as enhanced phonon scattering at point defects, together with the presence of four-phonon processes and a ‘rattling’ effect. α-In2Te3 is found to be an n-type indirect semiconductor with an energy gap of 1.03 eV. All these physical characteristics are intrinsic properties of a high-quality stoichiometric crystal.
Noteworthily, despite being in focus for such important technological uses, the basic knowledge on In2Te3 seems to still remain questionable, requiring novel insights. In particular, the structural arrangement and nature of the phase transition are less explored. The available structural models for the low-temperature (LT) α-In2Te3 (within space groups F3m, Imm2 and I4mm)48,49 and high-temperature (HT) β-In2Te3 (fcc ZnS prototype)50 were assumed based on the indexing of powder X-ray diffraction patterns, as well as applying group–subgroup crystallographic relations. No structural refinements are known up to now.51,52 Importantly, quantum mechanical calculations performed assuming all these LT-models resulted in metallic properties for α-In2Te3 (material ID: 622511, 1223866 and 1105025),53 whereas those done for β-In2Te3 indicated it to be a semiconductor with a band gap of 1.52 eV.54 These results are contradictory and suggest that the existing structural models are likely to be incorrect.
The band-gap values obtained for α-In2Te3 from different optical spectroscopic measurements vary in the Eoptg ≈ 0.99–1.22 eV range.55–61 The nature of the transition (i.e., direct/indirect) remains unclear due to the absence of trustworthy structural data. Studies of the temperature dependencies of the Hall55,62 and Seebeck63 coefficients indicated hole- or electron-mediated conduction mechanisms in different samples. These findings agree well with the small homogeneity range reported for α-In2Te3.64–66 However, a relation between the composition and type of conductivity is still less understood. The charge-carrier concentrations at room temperature, of ≈1010 cm−3 for p- and ≈1012 cm−3 for n-type samples as deduced from Hall-effect measurements, are in line with the observed optical band gaps.67 Polycrystalline α-In2Te3 reveals a low thermal conductivity of 1–1.5 W m−1 K−1 in a broad temperature range of 300−700 K.34 However, the underlying physical mechanisms for such an effect have not been elucidated.
In this work, we revisit the structural models of LT and HT polymorphs of In2Te3 by using temperature-dependent high-resolution synchrotron X-ray diffraction. To elucidate the nature of the phase transition, determination of the intrinsic physical and thermodynamic properties, and magnetic, electrical and thermal transport measurements combined with optical and spectroscopic analyses were performed. All these characterization studies were carried out on a well-established high-quality stoichiometric In2Te3 crystal.
A representative sample's surface was analyzed via scanning electron microscopy (SEM) and spectroscopic ellipsometry. Local chemical composition and microstructure analyses were carried out using a SEM-JEOL JSM 7800F microscope equipped with a Bruker Quantax 400, XFlash 6‖30 (silicon drift detector) EDXS spectrometer. The latter confirmed the sample to be single-phase with a chemical composition of In1.9(1)Te3.1(1) determined from 10 randomly selected points (Fig. S1†).
Differential thermal analysis (DTA) with thermogravimetry (TG) was performed on a bulk piece of ≈20 mg, inspected with an optical microscope. The measurement in the temperature range of 300–1000 K with a heating/cooling rate of 10 K min−1 was done using a Netzsch STA 449F3 device.
Non-polarized Raman spectra were recorded with a Horiba LABRAM System-HR-800, with a CCD camera, a 600 grooves per mm grating, a HeNe laser (633 nm), and a 50× objective with a N.A. of 0.5 (calibrated with a Si[111]-standard by use of its 520.6 cm−1 peak). Neon calibration-lamp spectra were recorded after each measurement to track and correct any shift during the Raman measurements. Temperature-dependent Raman spectra and Raman thermal conductivities were obtained with a Linkam THMS-600 cooling–heating stage. The latter was continuously cooled using a liquid-nitrogen vapor-flow and the samples were fixed with silver paste to the stage.
Low-temperature (LT, T ≤ 300 K) magnetic susceptibility [χ(T)], specific heat capacity [cp(T)], charge-carrier concentration [n(T)] and mobility [μ(T)], electrical resistivity [ρ(T)], Seebeck coefficient [S(T)] and total thermal conductivity [κ(T)] measurements were performed with the VSM, HC, VdP-Hall, and TTO modules of a DynaCool-12 from Quantum Design, respectively. The χ(T) and cp(T) measurements were performed on sample pieces of ≈19.5 mg and ≈10 mg, respectively.
The Hall-effect measurements using the Van der Pauw four-probe method were performed on a square-shaped plate of ∼5 × 5 × 0.4 mm3, in contact with platinum wires (25 μm) in a magnetic field of 10 T. Further, n(T) and μ(T) were calculated from the measurements. High-temperature (HT, 300 K ≤ T ≤ 600 K) ρ(T) and S(T) were measured with an ULVAC ZEM-3 device. Both LT and HT ρ(T), S(T) and κ(T) properties were obtained from a bar-shaped sample with dimensions of ∼1.5 × 1.4 × 6.1 mm3. The HT specific heat was measured on a sample piece of ∼36 mg placed inside an Al2O3 crucible using a differential scanning calorimeter, DSC 8500, from PerkinElmer. The heating rate during the measurement was 20 °C min−1.
![]() | ||
Fig. 1 Selected 2θ-region of synchrotron PXRD patterns for 840 K < T < 900 K together with (hkl) indices for LT α- (red) and HT β- (blue) In2Te3. |
Therefore, in the next step we tried to deduce the atomic positions in the studied telluride by applying direct methods. The attempt resulted in the same Te2–Te5 positions as reported earlier (cf. Table S2† and data from the ICSD) as well as Te1 at 4c and In at 16e, 24g and 48h, which differed from those found in the literature.49,71–73 Refinement of such a model converged with RI = 0.089, RP = 0.134, the wrong In2.4Te3 composition and the residual electron density (RED) −2.2/+3.7e Å−3. Performed differential Fourier syntheses indicated additional RED at the 4b and 24g Wyckoff positions. The latter one was already known to be occupied by In-atoms in the model derived from the ZnS type. Having now too much In-concentration in the structure, we refined the site occupancy (G) of the crystallographic sites of these atoms. After this step, the values of the RED were already −0.71/+0.91e Å−3. However, refining the atomic displacement parameters (ADPs), we observed an unphysically large value of Biso for Te1 at the 4c position. Shifting it off the center (i.e., assuming this atom to occupy ≈ 25% of a 16e site with x = y = z ≈ 0.02), a reliable ADP could be obtained. The final values of the R-factors, atomic coordinates, Biso and occupancy obtained from the performed refinement are collected in Tables S1 and S2.† The experimental, theoretical and differential profiles corresponding to these are plotted in Fig. 2 (upper panel). The good agreement of the refined and experimental compositions, together with the improvement of the RI-factor by almost 30%, indicate the reliability of the obtained model.
![]() | ||
Fig. 2 Synchrotron PXRD patterns and Rietveld refinements for LT α- and HT β-In2Te3 together with peak positions, as well as experimental, theoretical and difference profiles. |
Since now the majority of the crystallographic positions shown in Table S2† cannot be derived via group–subgroup relation from the ZnS type, the latter model is obviously inconsistent with the HT β-In2Te3. However, by performing indexing of the PXRD patterns presented in Fig. S2†, we confirmed the earlier reported UCP aHT ≈ aLT/3 ≈ 6.2(1) Å as well as the face-centered cubic lattice with the above-mentioned SGs. Keeping in mind that the klassengleiche transformation with an index of 9 should work during the α → β phase transition in In2Te3, we attempted direct methods, assuming again the SG F3m for HT-modification. There, two In-atoms are localized at 4a and 4b and Te at 4c. Such a structural arrangement corresponds to the MgAgAs half-Heusler type.74 It should also be noted that in the prototype compound, the 4a-, 4b- and 4c-sites are occupied by Mg, As and Ag, respectively, which makes the discussion within the group–subgroup scheme less conclusive (Fig. S3†). Therefore, in the further considerations we refer to the MnPdTe compound, where Pd and Te are at the 4b and 4c sites.75 The R-factors of the refinements performed at different temperatures, the obtained atomic coordinates, and the Biso and occupancy values for HT-β-In2Te3 are shown in Table S3.† The experimental, theoretical and difference profiles for T = 323 K and T = 899 K are depicted in Fig. 2.
The shortest interatomic distances in the structures of α- and β-In2Te3 are collected in Tables S1 and S3.† Both structures are characterized by no bonding In–In and Te–Te contacts. In contrast, the In–Te distances are shortened by ≈5–12% in comparison with the sum of the covalent radii r(In) = 1.5 Å and r(Te) = 1.37 Å.76
The group–subgroup relation scheme between α- and β-In2Te3 modifications is presented in Fig. S3.† As expected, the majority of In- and Te-positions in the LT-polymorph are derived from the largely occupied 4b and 4c sites in the HT-structure. This finding differs from the previously reported model derived from the ZnS type, where both In-positions were deduced from the 4a site. In our case, the 24g position originating from this site is only ≈17% occupied, in agreement with the low occupancy in the HT-polymorph (cf. Tables S2 and S3†). Additionally, the presence of a Te1-atom at 4c in the LT-modification would suggest the initial 4a-site in β-In2Te3 to be occupied with a statistical In/Te mixed site occupancy. However, such a refinement from PXRD data is rather impossible due to the close atomic form factors of indium and tellurium. This problem could be addressed by neutron diffraction on high-quality crystals.77
The In1-atom at 4b in α-In2Te3 can only be derived from the group–subgroup scheme if a 4d position in the HT-polymorph would be occupied (Fig. S3†). The structure arrangement with the (dcba) Wyckoff sequence, SG F3m and a ≈ 6.5(3) Å, is known as the TiCuHg2 type.78 For better visualization, we use in the further discussions its ordered superstructure LiPtMgSb.79 This structural arrangement is obviously not suitable for β-In2Te3, since the 4d-site remained non-occupied therein (i.e., the final values of RED after refinement were −0.14/+0.23e Å−3), which raises a question about the correctness of the proposed group–subgroup scheme in Fig. S3.† However, accounting for G = 0.25 for the In1-atom in the LT-modification, one would expect the 4d site occupancy in β-In2Te3 to be lower than ≈3%, which is again undetectable within the refinement of the PXRD data. Obviously, as mentioned above, to unambiguously solve the crystal structures of α- and β-In2Te3, a temperature-dependent neutron diffraction study is strongly required.
In Fig. 3, the arrangement of polyhedra in the crystal structures of α- and β-In2Te3, assuming full occupancies of all crystallographic sites, in comparison with the LiPtMgSb type is presented. As is known, not being a centered structure [i.e., no atom at 4b (1/2 1/2 1/2)], ZnS reveals for this site an [□S4]-tetrahedron, [□Zn6]-octahedron and [□Zn8]-cube as the first, second and third coordination spheres, respectively.80 Centering such a unit cell by adding an atom into the 4b Wyckoff position, one would obtain the MgAgAs type (β-In2Te3) with an [In2HTTeHT4]-tetrahedron, [In2HTIn16HT]-octahedron and [In2HTIn2HT12]-cuboctahedron (Fig. S3† and 3b). Adding a further atom at 4d (LiPtMgSb type) results in a [MgPt4Li4]-cube as a first coordination sphere, whereas the second and third ones remain the same as in β-In2Te3 (MgAgAs type) (cf.Fig. 3a and b). Being closely related with both simple prototypes, the α-In2Te3 crystal structure reveals a similar arrangement of polyhedra. As one can see from Fig. 3c, there appears an array of corner-sharing [In5Te44]-tetrahedra (tan) typical for MgAgAs and ZnS types (not shown in Fig. 3a and b). Since the 4d-sites are unoccupied in the HT-modification, the [□Te34]-tetrahedra (red), [□In46]-octahedra (blue) and [□In512]-cuboctahedra (dark grey) remain empty, in contrast to the MgAgAs prototype. The same observation is made for unoccupied [□Te14Te24]-cubes and [□In46]-octahedra surrounding them, which resemble the structural units of the LiPtMgSb arrangement. Also, the third coordination sphere of this empty site is only a [□In24]-tetrahedron (dark grey) (instead of a cuboctahedron, as it is the case for LiPtMgSb), which is due to the unoccupied 16e site derived from the 4b one (Fig. S3†).
![]() | ||
Fig. 3 Structural arrangement and comparison of the LiMgPtSb (a), HT β- (b) and LT α- (c) In2Te3 structures. |
As the temperature increases, some weak reflections become systematically suppressed (e.g., 004, 135, 026 and 226) until they completely disappear at the transition temperature Tα→β ≈ 865 K (Fig. 1). The strong background increase in the PXRD pattern measured at ∼926 K indicates the beginning of the melting process. It becomes completed at Tmelt ≈ 950 K as no diffraction peaks are observed. These results agree well with our differential thermal analysis (DTA) (see below) and the reported In–Te phase diagrams (Tα→β and Tmelt range between 878–915 K and 938–952 K, respectively).64–66 The small offset between the structurally and thermally determined Tα→β might be attributed to Te-mass losses in the dynamic Ar-flow used during DTA (bulk).
The temperature dependencies of the unit-cell volumes, defined as aLT3 and 27aHT3 for the α and β phases, respectively, are depicted in Fig. 4. The thermal expansion of the LT-modification can be described by a 2nd-order polynomial function in the temperature range of 100–800 K:
Vα(T) = V0α + V1αT + V2αT2 | (1) |
![]() | (2) |
![]() | ||
Fig. 4 Temperature-dependent unit-cell volume V of α- and β-In2Te3 from 102 K to 974 K together with the fit to eqn (1) (red-dotted line). Inset: Near-phase-transition unit-cell volume variation ΔV estimated at 850 K. |
Parameters | α-In2Te3 |
---|---|
Fitted T-range | 100–864 K |
V 0 (Å3) | 6264(1) |
V 1 (Å3 K−1) | 0.165(6) |
V 2 (Å3 K−2) | 5.4(7)× 10−5 |
α V (10−5 K−1) | 2.8–4.0 |
γ XRD | 1.5–2.1 |
Unexpectedly, at T ≈ 820 K the UCP of α-In2Te3 starts to decrease, revealing a weak negative thermal expansion (i.e., reduction of the unit-cell volume ΔV ≈ –0.05%) (Fig. 4 inset). Only after the structural transformation is completed at T ≈ 865 K (no superstructural peaks are observed) does the UCP start to increase with temperature again. Such a “negative step” in the a3(T) is in contrast with the behavior of thermal expansion at the SPT (e.g., for the closely related In2S3 (ref. 81) as well as with the classical expectations), where a positive jump in this dependence should be the case.82,83 Nevertheless, an analogous effect has been also observed for TiGePt84 and Zn2−xMgxP2O7.85
Knowing the average speed of sound in α-In2Te3 (vS = 2560 ms−1)86 and the Dulong–Petit limit (cp) we calculate the Grüneisen parameter from the formula γXRD = αvνs2cp−1 (Table 1). The obtained value agrees with those observed for many semiconductors (i.e., γ ≈ 1–2)87 as well as with γ ≈ 2.188 found for the state-of-the-art thermoelectric material SnTe. All these suggest an enhanced phonon anharmonicity.89
The temperature dependencies of the isotropic atomic displacement parameters (ADPs) for the α- and β phases are shown in Fig. 5. In agreement with the theory proposed in ref. 90, we observe a small Biso dependence on T, together with their displacements upward by a constant value for the strongly statistically disordered sites (Table. S2†). On the other hand, for atoms at almost fully occupied crystallographic positions, the Biso(T) values for T > 400 K are increasing nearly linearly with different slopes, which is a signature of the so-called dynamic disorder.90 The latter is known to be mainly due to thermal motion of atoms, which can be treated as local vibrations of a quantized harmonic oscillator (Einstein model). Hence, for sufficiently high temperatures (i.e., hω ≪ 2kBT), Biso(T) is given as:
![]() | (3) |
Further, applying the definition of a cage-compound92–94 to α-In2Te3, we assume the Te-atoms form an anionic framework incorporating positively charged In-cations. Such a description can be also justified by the larger electronegativity (Pauling scale) of Te (χTe = 2.10) in comparison to that of indium (χIn = 1.78).76 This concerns the steepest slope of Biso(T) dependence for In5-atoms, which indicates them to potentially be a ‘rattler’.91 Thus, applying eqn (3), a characteristic ΘIn5E = 69(1) K (Fig. 5, upper panel) is deduced.
The temperature dependencies of the specific heat capacity in the cpT−1(T) representation near the Tα→β and Tmelt transitions are presented in Fig. 7 and the inset therein, respectively. The Tonset–Toffset ranges of both anomalies agree well with the DTA measurements (Table 2). The discrepancies in the assignment of Tmax can be explained by the asymmetries of the peaks in the cpT−1(T) dependencies.
α → β | Melting | |
---|---|---|
T max c p (K) | 882 | 956 |
T onset c p − Toffsetcp (K) | 865–905 | 923–970 |
T PXRDα→β (K) | 865 | — |
T PXRDmelt (K) | — | 950 |
ΔH (J mol−1) | ∼44.6 | 66× 103 |
ΔS (J mol−1 K−1) | 4.9 × 10−2 | ∼107 |
ΔS(R) | 6 × 10−3 | 12.9 |
Further, knowing the Dulong–Petit limit cDPpT−1 (see discussion below) and subtracting it as a peak baseline, we calculated the changes in entropy and enthalpy at the Tα→β and Tmelt transitions (Table 2) using ΔS = ∫cpT−1dT and ΔH = ∫cpdT, respectively. The drastic difference in the obtained ΔS(R) values (where R is the gas constant) would obviously indicate different orders of the transitions.83
To define the α → β structural phase transition in In2Te3 as being of 1st order, the following criteria are expected to be fulfilled: (i) the coexistence of both polymorphs at Tα→β, (ii) a discontinuous change in both unit-cell volume and ΔS, (iii) a sharp symmetric peak in cp(T) and (iv) ρ(T) hysteresis while performing the measurement in the heating and cooling regimes.81,82 The first and last criteria from this list have not been proven up to now, whereas points (ii) and (iii), as is clearly visible from the discussions above, are not fulfilled. Obviously, to unambiguously estimate the order of this transition, some additional studies are required.
![]() | (4) |
![]() | (5) |
![]() | (6) |
![]() | ||
Fig. 8 Temperature dependence of specific heat capacity for α-In2Te3 in the cp/T3 representation, including the fit (red line) to eqn (4). The individual contributions from eqn (5) and (6) are shown as dotted lines. |
Since the quantity of Debye terms is defined by the number of elemental constituents in the chemical formula (i.e., two for In2Te3) and that of Einstein modes by the amount of ‘rattlers’, we set i = 2 and j = 1 in the preliminary model. Further, accounting for the stoichiometric composition of the studied compound, we find the total number of modes to be Ntot = NDi + NEj = 15. Knowing from the analysis of the Biso(T) dependencies that only In5 could be a ‘rattler’ and thus, having 48-fold site symmetry with a site occupancy G = 0.87 (Table S2†), one could assume NE1 ≈ 3. Then NDi = 12 with ND1 ≈ 3 (for the not-‘rattling’ In-atom) and ND2 = 9 (for Te-atoms) should be the case. However, a model based on this concept failed in the description of the peak in the cp/T3(T) dependence, which is an indication of a much more complex phonon spectrum of the studied α-In2Te3. Such a situation was frequently reported in the literature.98–101 Therefore, in the next attempt of the fitting we assumed an additional ‘rattling’ contribution (i.e., j = 2), which could originate from the In3-atom due to the observed slope in the Biso(T) dependence. Removing all constrains on all NDi and NEj values in the fit, the best description is obtained for a model with ND1 = 4(1), ND2 = 9(1), NE1 = 0.7(1), NE2 = 1.1(2) and thus Ntot = 14.8. The corresponding Debye and Einstein temperatures were ΘD1 = 322(7) K, ΘD2 = 101(5) K, ΘE1 = 39(2) K and ΘE2 = 60(9) K, respectively. Such a result indicates an acceptable description of the phonon spectrum of the anionic framework and a model of the ‘rattling’ motion. Noteworthily, the ΘE2 value correlates well with the In5-atom Einstein temperature ΘXRDE = 69(1) K deduced from the ADP values, whereas ΘE1 is smaller by a factor of ≈3. Obviously, the model given by eqn (4) does not account for the structural disorder and thus, is not providing a complete understanding of the complex phonon spectrum of α-In2Te3.
A similar result (i.e., 2ΘE1 ≈ ΘE2) was obtained for the Sn24P19.4Br8 clathrate (cf. Table 1 in the reference), where two ‘rattlers’ are well established.102 This would again assume a remarkable difference between the Biso(T) dependencies for those atoms (not reported in ref. 102), which makes less sense in the view of the same physical effect (i.e., thermal motion with enhanced amplitudes within the voids with close volumes). This would bring us to a similar conclusion: the consideration of only the cp/T3(T) dependence provides just tentative insights into the ‘rattling’ problem.
Finally, the negligibly small Sommerfeld coefficient of the electronic specific heat capacity γ = 6.2(9) × 10−8 J mol−1 K−2 is in nice agreement with the semiconducting properties of the studied telluride (i.e., γ = 0 for no states at the Fermi level),83 thus indicating a minor metallic impurity (cf. Fig. S5†). The latter is also evidenced by an upturn in the LT diamagnetic susceptibility χ(T) of α-In2Te3 (Fig. S6†).
As is known, both the positions of modes in a Raman spectrum and the thermal expansion of a material are mainly related to the so-called anharmonic effects, which are due to the change in frequencies of lattice vibration with temperature.106 Therefore, measurement of T-dependent Raman spectra (Fig. 9 inset) allows the elucidation of underlying harmonic and anharmonic phonon processes. For this purpose, the Klemens–Balkanski model is applied. It defines the peak center W and full-width at half-maximum (FWHM) R of the strongest mode as follows:
![]() | (7) |
![]() | (8) |
![]() | ||
Fig. 10 Klemens–Balkanski analysis of α-In2Te3’s strongest feature for its peak center and FWHM within the 100–300 K range, including independent contributions to eqn (7) and (8), respectively. |
To describe the W(T) and R(T) behaviors, both AW,R and BW,R contributions were required (Fig. 10). This indicates the presence of both three-phonon (most common in crystalline materials) and four-phonon processes. However, the latter are contributing only ≈5% in total. Since W and R in α-In2Te3 reveal rather weak T-dependencies, the obtained anharmonic constants are smaller by approximately two-orders of magnitude than those observed for FeS2108 or PbCuSbS3.98
The Tauc plot [(αE)1/2vs. E] together with the linear projection to zero absorption is presented in Fig. 11. The s-like shape of the (αE)1/2-plot vs. the photon energy confirms α-In2Te3 to possess an indirect optical band gap with Eoptg = 1.03(1) eV. This value perfectly agrees with the earlier reported one (1.01 eV)55,56 for a single crystal grown using the Bridgman method. On the other hand, the Eoptg values found for both polycrystalline (1.13 eV)57 and amorphous (1.22 eV)58 specimens were slightly larger, which indicates the dependence on the sample microstructure. In this respect, we must again conclude that the crystal studied here is of good quality.
Interestingly, some thin films of α-In2Te3 are reported to possess optical gaps of the same magnitude, however with a direct transition.59–61 Here, we would like to emphasize that in all these works, the estimation of the Eoptg values and their natures was carried out via absorption spectroscopy. Importantly, the elipsometric method is much more sensitive in this respect, since it depends only on the polarization changes after interaction with the sample.109
![]() | (9) |
As is known, the appearance of two energy gaps (Eg) in the ρ(T) dependence is a signature of a doped semiconductor. In this class of materials, the intrinsic Eg is observed in the HT regime, whereas the LT one represents that between the impurity level and the corresponding band edge [e.g., valence band maximum (VBM) or conduction band minimum (CBM)]. In the T-region in-between, the charge-carrier concentration [n(T)] is considered to remain nearly temperature independent, whereas the mobility [μ(T)] should determine the shape of the ρ(T) curve.110 Noteworthily, the EArr(LT)g value for In2Te3 is lower by at least one order of magnitude than those occurring in doped semiconductors, as well as being much smaller than EArr(HT)g.83 Also, both n(T) and μ(T) (Fig. 13A and B) are nearly T-independent for 150–250 K and for T > 270 K they start to increase and to decrease, respectively. All these observations indicate that despite revealing a ρ(T)-shape similar to that of a doped semiconductor, In2Te3 cannot be considered as belonging to this class of compounds. Additionally, this finding would be in line with the declared purity of the crystal studied here. Obviously, the observed ρ(T)-shape is due to the high number of partially occupied positions in the crystal structure, which possibly can become a source of additional energy levels in close vicinity to the VBM/CBM. Here, we would also like to stress that the origins of such ρ(T) behavior could be of very different natures (e.g., charge density waves,111,112 topological band structures113 or simply remaining unclear114,115).
![]() | ||
Fig. 13 Temperature dependency of the charge-carrier concentration n(T) (A) and mobility μ(T) (B) of α-In2Te3 assuming a single band conduction mechanism. |
The temperature dependence of the Seebeck coefficient S(T) for In2Te3 is depicted in Fig. 12 (right scale). S(T) is negative in the whole studied T-range indicating electrons to be the dominant charge carriers. Also, n-type conductivity was earlier reported for the here-studied telluride from Hall effect measurements.63 However, there are numerous investigations where p-type transport is detected, namely in single crystals55,56 as well as in thin-films.59,61 Interestingly, in all these cases, a slight In-excess (i.e., In2+xTe3) was found. The off-stoichiometry of In2Te3 would be in line with the In–Te phase diagrams proposed in ref. 64–66.
The Seebeck coefficient of In2Te3 decreases with increasing temperature, revealing a well-defined anomaly centered at Tkink ≈ 450 K, coinciding with the maximum in ρ(T). This might be an indication of a switch or change in charge-carrier scattering mechanisms. However, to shed light on such behavior, HT Hall effect measurements would be strongly required. A further temperature increase leads to the appearance of a broad maximum in |S(T)| at Tmax ≈ 714 K. Such a behavior of |S(T)| allows the estimation of the energy gap by applying the Goldsmid–Sharp formula:116
ESg = 2e|Smax|Tmax | (10) |
![]() | (11) |
The values of the charge-carrier mobilities [μ(T)] observed in this work (Fig. 13B) are larger by a factor of ≈10 than those previously reported for α-In2Te3 at RT (i.e., 32–70 cm2 V−1 s−1).67 This fact stresses again the strong influence of the structural disorder. Since the μ(T) decrease for T ≳ 250 K is compensated by a simultaneous increase in n(T), a smooth variation (rather than a step-like change) is the case in ρ(T) (Fig. 12).
Further, we analyzed the μ(T)-dependence using the μ ∝ Tm power law. Here, m is an indication of different scattering mechanisms.119,120 Two T-regions were identified within such an approach: (i) μ ∝ T−3/2 for 100–250 K and (ii) μ ∝ T−5/2 for 250–290 K (red and green lines, respectively, in Fig. 13B). m = −3/2 and m = −5/2 stand for scattering of charge carriers on only acoustic or on both acoustic and optical phonons, respectively.
To fully understand the scattering mechanisms in α-In2Te3, we also analyzed its electrical transport properties within the single parabolic band (SPB) model.121 The obtained Pisarenko plot confirmed an excellent agreement between the measured S(n) and the theoretically predicted values (Fig. S7†). Also, a fair agreement is observed for the experimental and simulated μ(n) dependencies for T > 200 K (Fig. S8†), hinting towards the dominance of charge-carrier scattering on the acoustic phonons, as discussed above. Finally, this type of analysis indicates the thermoelectric (TE) power factor [PF(T) = ρ(T)−1S(T)2T] for the studied telluride to be close to the maximal expected semiconducting values (Fig. S9†), which means that its drastic increase (i.e., improvement of the TE efficiency) would be possible only while going towards a metallic type of tuning.
Having the hole mobilities μmin ≈ 210 and 1380 cm2 (V s)−1,55,56 which were reported for p-type single crystals at RT, together with our value μmaj ≈ 225 cm2 (V s)−1 (Fig. 13B), we again analyzed eqn (11). Calculating the μmaj/μmin-ratio to be ∼46/50 or ∼8/50 and assuming A ≈ 1/500, one can conclude that the effective masses of holes in α-In2Te3 are larger by a factor of 19–60 than those of electrons. The numbers obtained in such a way agree well with the values deduced from theoretical calculations for chemically related In2S3 and its doped variant In2S3−xSex122 as well as InSb.123 All of them also assume the electronic structure of α-In2Te3 to be anisotropic and to be characterized by a flat valence band just below the Fermi level EF, which seems to be in contradiction with the fact that the SPB model fairly describes the electrical transport.83 Obviously, to shed light on this problem, reasonable DFT simulations would be strongly required, which are still a challenge for state-of-the-art approaches, due to the large unit-cell volume and strong structural disorder in the studied telluride. For more details on this topic, see our discussion on the electronic structure calculation of hypothetically ordered α-In2Te3 in the ESI.†
![]() | (12) |
The phononic contribution (κph) obtained after subtraction of κel(T) from the total thermal conductivity is presented in Fig. 14. It is very low: κph ≤ 0.7 W m−1 K−1 (which is by a factor of ≈2 smaller than that reported for a polycrystalline sample at RT34) and thus indicates α-In2Te3 to possess potential for possible TE applications.125 To understand the underlying individual phonon scattering mechanisms, κph was further analyzed within the modified Debye–Callaway model:126
![]() | (13) |
τtot−1 = τB−1 + τPD−1 + τU−1 | (14) |
![]() | ||
Fig. 14 In2Te3 phononic contribution to the thermal conductivity κph together with the fit to eqn (13). Inset: Simulated smallest physically achievable κph by considering a glass-like state. |
Model126 | Fitted parameter | |
---|---|---|
τ B −1 | v S/D | D = 7.65(1) × 10−7 m |
τ PD −1 |
![]() |
Γ = 1.057(1) |
τ U −1 |
![]() |
γ th = 2.75(2) |
The τB−1 value deduced from the fit to eqn (13) indicates a rather large D = 7650(1) Å crystallite size and thus negligibly small phonon scattering on the grain boundaries in the studied telluride. This finding would be in contradiction with the low κph(T), since large D-values normally imply better conductivity.129 On the other hand, τPD−1 (calculated using the average volume per atom per unit cell ) suggests an enlarged point-defect scattering parameter Γ = 1.057(1), which should be the case in strongly disordered structures [reported also for Bi2Te3−xSex nanoplates (Γ = 0.18),130 nanograined Cu2Se (Γ = 0.38)131 and polycrystalline AgBiSe2 (Γ = 1.6)132]. Obviously, higher scattering on point defects compensates for the absence of that on grain boundaries in α-In2Te3. Finally, in the temperature range T ≳ 100 K, τU−1 (calculated using the average atomic mass
) dominates the thermal transport. From it, we deduce the average thermal Grüneisen parameter γth = 2.75(2) in agreement with γXRD (Table 1). Our values are similar to those observed for systems [e.g., InTe (γ ≈ 3.3),133 SnTe (γ ≈ 2.1),88 In2Se3 (γ ≈ 5.4),134etc.] with high phonon anharmonicity and thus low κph.
Assuming disruption of the long-range order in In2Te3 (i.e., an amorphous or glass-like state) we calculated the minimal possible phononic thermal conductivity κmin(T) of this compound, which is estimated as:
![]() | (15) |
Calculating the dimensionless TE figure-of-merit zT(T) = PF(T)T[κph(T) + κel(T)]−1 for In2Te3, we found that despite suitable low thermal conductivity, it varies between 10−5–10−2, which is by a few orders of magnitude lower than those observed for state-of-the-art thermoelectric materials.83 This situation is mainly due to the enhanced electrical resistivity caused by low charge-carrier concentration in this compound.
The temperature dependence of the electrical resistivity of α-In2Te3 reveals two regions with an activation-like decay and a well-defined maximum. Since the studied crystal is shown to be of a good quality with no impurities, we ascribe these effects to the high structural disorder and do not classify it as a doped semiconductor. This conclusion is additionally corroborated by the nearly temperature-independent behavior of the charge carriers concentration and their mobility. The Seebeck and Hall coefficients for α-In2Te3 are found to be negative in the whole studied T-range, indicating that electrons dominate electrical transport in this material. The analysis performed within the Goldsmid-Sharp approach showed the effective masses of holes to be by a factor of ≈19–60 larger than those of electrons, which assumes the electronic structure of α-In2Te3 to be highly anisotropic and to contain flat valence bands.
The T-dependent evolution of the atomic displacement parameter of the In5-atom at the 48h crystallographic site, combined with the appearance of a boson peak in the specific heat at low temperatures, indicates a ‘rattling’ effect in α-In2Te3. This is obviously one of the reasons for a very low thermal conductivity (κph ≤ 0.7 W m−1 K−1) in this compound. Further analysis performed within the Debye–Callaway model has shown extremely strong phonon scattering on point-defects, whereas temperature-dependent Raman spectroscopy indicated the occurrence of four-phonon decay processes. Both effects are again governing the low κph and nicely agree with the enhanced values of the Grüneisen parameters deduced from differential thermal (γth) and structural (γXRD) analyses. All these findings indicate stronger anharmonicity of phonon processes in α-In2Te3. The simulated minimum κph dependence confirms that the observed thermal conductivity cannot be significantly reduced by, e.g., nanostructuring, and thus it is an intrinsic effect.
Having a sample of an advanced quality (no transport agents were used for its synthesis), we shed light on the interplay between its strong structural disorder and the underlying electrical and thermal transport mechanisms. Obviously, when redistributing the numerous defects by varying the synthesis conditions of α-In2Te3, one can expect some enhancements in charge-carrier density and/or their mobilities, thus improving the electrical transport characteristics. To make such a tuning more predictable, additional theoretical studies concerning electronic and phononic structures are also strongly desired. Undoubtedly, α-In2Te3 is a promising system possessing a remarkable potential with respect to its further investigations.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta06357k |
This journal is © The Royal Society of Chemistry 2025 |