J.-B. Vaneya,
J.-C. Crivellob,
C. Morinb,
G. Delaizirc,
J. Carreaudc,
A. Piarristeguyd,
J. Monnierb,
E. Allenob,
A. Pradeld,
E. B. Lopese,
A. P. Gonçalvese,
A. Dauschera,
C. Candolfia and
B. Lenoir*a
aInstitut Jean Lamour, UMR 7198 CNRS – Université de Lorraine, Nancy, France. E-mail: bertrand.lenoir@univ-lorraine.fr
bInstitut de Chimie et des Matériaux Paris Est (ICMPE), UMR 7182 CNRS-Université Paris-Est Créteil, Thiais, France
cScience des Procédés Céramiques et de Traitements de Surface (SPCTS), UMR CNRS 7315-Université de Limoges, France
dInstitut Charles Gerhardt (ICG), UMR 5253 CNRS-Université de Montpellier, France
eC2TN, Instituto Superior Técnico, Universidade de Lisboa, Lisboa, Portugal
First published on 23rd May 2016
β-As2Te3 belongs to the prominent family of Bi2Te3-based materials, which show excellent thermoelectric properties near room temperature. In this study, we report a joint theoretical and experimental investigation of its electronic and thermal properties at low temperatures (5–300 K). These results are complemented by specific heat measurements (1.8–300 K) that provide further experimental evidence of the first order lattice distortion undergone by β-As2Te3 near 190 K. Data taken on cooling and heating across this transition show that the lattice distortion has little influence on the electronic properties and further evidence a weak hysteretic behavior. Although first-principles calculations predict a semiconducting ground state, these measurements show that β-As2Te3 behaves as a degenerate p-type semiconductor with a high carrier concentration of 1020 cm−3 at 300 K likely due to intrinsic defects. Calculations of the vibrational properties indicate that the extremely low lattice thermal conductivity values (0.8 W m−1 K−1 at 300 K) mainly originate from low-energy Te optical modes that limit the energy window of the acoustic branches. This limited ability to transport heat combined with a relatively large band gap suggest that high thermoelectric efficiency could be achieved in this compound when appropriately doped.
Despite strong efforts devoted to the search for materials able to operate at ambient temperature, the semiconducting (Bi,Sb)2X3-based alloys (X = Se, Te) still remain to date the most prominent class of thermoelectric materials for room-temperature thermoelectric applications.1,2 These compounds crystallize in a rhombohedral unit cell leading to isotropic transport properties in both single-crystalline and in some polycrystalline samples depending on the synthesis process. A large number of substituting elements can be used to alter their transport properties resulting in a wide variety of physical phenomena,23,24 in addition to their Dirac-like conducting surface states classifying these compounds as 3D topological insulators.25–27
Recently, our initial investigation on the high-temperature thermoelectric properties of a new member of this family, namely β-As2Te3, evidenced extremely low lattice thermal conductivity approaching the amorphous limit (0.5 W m−1 K−1 at 423 K).28 This property, combined with enhanced power factors due to Sn doping, resulted in a peak ZT of 0.65 at 423 K. Further, joint laboratory X-ray and neutron diffraction experiments carried out on β-As2Te3 revealed a reversible first-order phase transition to a distorted monoclinic phase (space group P21/m) upon cooling below 210 K.29 Another consequence of this subtle structural transition from the β phase to the new β′ phase is a fourfold modulation of the atomic positions along the b axis (Fig. 1). This structural instability was backed up by first-principles calculations of the formation energies indicating that the β′ phase is more stable than the β polymorph at 0 K.29 Yet, the low-temperature physical properties of β-As2Te3 as well as the influence of the lattice distortion on the transport remain so far unexplored.
![]() | ||
Fig. 1 Comparison of the rhombohedral (R![]() |
Herein, we provide a detailed investigation of the transport and thermodynamic properties of β-As2Te3 below 300 K. Because first-principle calculations are essential for modeling the transport and for guiding future efforts to improve the thermoelectric properties of this material, our experimental findings are complemented by calculations of the electronic and phonon densities of states and dispersion curves of both the β and β′ phases, which are discussed in light of prior theoretical studies carried out on β-As2Te3.
Phonon calculations were performed by a supercell approach with the finite displacement method within both the harmonic (HA) and quasi-harmonic approximation (QHA) using the Phonopy code.34–36 Details concerning the unit cells containing from 60 to 80 atoms, which have been considered, are summarized in Table 1.
α C2/m (12) | β R![]() |
β′ P21/m (11) | β′ C2/m (12) | |
---|---|---|---|---|
a (Å) | 14.962 (14.337) | 4.102 (4.047) | 7.087 (6.99) | 7.088 |
b (Å) | 4.071 (4.015) | — | 16.366 (16.24) | 4.092 |
c (Å) | 10.130 (9.887) | 29.745 (29.498) | 10.475 (10.25) | 10.463 |
β (°) | 95.579 (95.06) | — | 103.016 (103.4) | 103.044 |
Volume (Å3 atom−1) | 30.701 (28.456) | 28.893 (27.893) | 29.592 (28.297) | 29.560 |
Theoretical band gap (eV) | 0.46 | 0.30 | 0.39 | 0.39 |
Experimental band gap (eV) | 0.43 (ref. 42) | — | — | — |
Phonon supercell | 2 × 2 × 2 (80 at) | 2 × 2 × 1 (60 at) | 1 × 1 × 1 (80 at) | 2 × 2 × 2 (80 at) |
The Hall resistivity was determined from measurements of the transverse electrical resistivity ρxy under magnetic fields μ0H ranging between −1 and +1 T. The data were corrected for slight misalignment of the contacts by applying the formula ρH = [ρxy(μ0H) − ρxy(−μ0H)]/2. The Hall coefficient RH was determined from the slope of the ρH(μ0H) data in the limit μ0H → 0. The Hall carrier concentration p and mobility μH were estimated within a single-band model with a Hall factor rH equal to 1 that yields the relations p = rH/RHe = 1/RHe and μH = RH/ρ. Specific heat measurements were carried out on cooling and heating between 1.8 and 300 K using the dedicated option of the PPMS. A sample of approximately 20 mg was glued on the sample holder with a tiny amount of Apiezon N grease. The experimental uncertainty on the transport, Hall effect and specific heat measurements is estimated to be 5%, 5% and 3%, respectively.
The valence bands, extending over ∼5.5 eV, are separated from this region by a gap of about 2 eV. The valence bands are flat and mainly dominated by the 4p-As and the 5p-Te states. As shown in the magnified Fig. 2c, the valence band maximum and conduction band minimum are located along the D–E direction resulting in a slightly indirect band gap of 0.39 eV (Δk = 0.05 × 2πa* where a* is the reciprocal lattice vector). The directional effective mass of holes along the D–E direction m*D–E is estimated to 0.31m0 where m0 is the free electron mass. The Fermi level has been conventionally set at the top of the valence bands. The magnitude of the band gap in the β′ phase is slightly higher than that obtained in the β phase which features a direct band gap of 0.30 eV (Fig. S6 and S8 in ESI†). This value is higher than that calculated in ref. 38 where a direct band gap of 0.24 eV was found at the Γ point. A comparison of these findings with the literature indicates a rather large discrepancy among the results reported for the β phase. While our value agrees with that obtained by Pal et al.39 who found a band gap of 0.35 eV with calculation based on all-electron full potential linearized augmented plane wave (FP-LAPW) technique, prior values derived theoretically (0.12 eV in ref. 37 and 0.22 eV in ref. 40) are significantly lower. These differences in the computed band gaps are most likely due to the different methodologies used, both the direct or indirect nature of the band gap and its value being strongly sensitive to details of the calculations.
In the prior study of Pal et al.,39 a compressive strain applied along the c-axis of the crystal structure was used to predict the transition from a trivial band insulator to a topological insulating state in β-As2Te3. Following this approach, we compared the evolution of the band gap under an isotropic stress for β-As2Te3 and β′-As2Te3 (Fig. S16 and S17 in ESI†). For both phases, the uniaxial compression leads to the inversion of the top of the valence band and the bottom of the conduction band at ΔV/V ∼ −7%, accompanied by a metallic state transition. This band inversion and parity reversal of bulk electronic bands are characteristic of an electronic topological phase transition. Although the isotropic stress shows some deviations compared to the uniaxial strain (inversion at ΔV/V ∼ −5% in ref. 39), our results demonstrate a similar behavior of the β and β′ phases. Hence, the lattice distortion does not preclude the observation of a topological state under pressure in this compound.
The phonon dispersion curves and the partial phonon DOS (PDOS) of β′-As2Te3 calculated in space group C2/m are depicted in Fig. 3a. Because of the mass ratio between As and Te, the top of the optical branches of As is located at higher frequencies (6.0 THz) with respect to that of the Te atoms (4.2 THz). The phonon dispersion curves of the three polymorphic As2Te3 phases exhibit fairly soft modes since the highest frequencies do not exceed 6.0 THz (∼200 cm−1 or ∼25 meV, see Fig. S12 to S15 in ESI†). Due to the mass ratio, the center of gravity of the PDOS of Te appears at lower frequencies compared to As, regardless of the investigated structures. As shown in Fig. 3b, the PDOS of Te can be decomposed into the 2a and 4i contributions. The Te atoms in the 2a site (Te1) contribute mainly at energies above 3 THz where the observed large PDOS corresponds to interactions with the 6 neighboring As atoms at about 3.0 Å. According to the coordination environment, the PDOS of Te in the 4i site (Te2) exhibits two main structures: one centered at low energies around 2.1 THz which corresponds to the interaction with three other Te2 atoms on the same site at ∼3.6 Å, i.e. across the van der Waals interspace, and another peak at higher frequencies (above 2.6 THz) corresponding to the interaction with the closest As atoms (∼2.8 Å). The acoustic phonon bands are limited to very low energies suggesting a poor ability of the β′ phase to conduct heat. This result is consistent with the very low thermal conductivity values of the β phase measured above 300 K (ref. 28) and, as we shall see below, with the measurements at low temperatures.
![]() | ||
Fig. 3 (a) Phonon dispersion curves and (b) partial phonon DOS of the β′-As2Te3 phase calculated in the irreducible C2/m Brillouin zone. |
The phonon dispersion curves of β′-As2Te3 calculated in the P21/m space group feature negative frequencies (see Fig. S15 in ESI†), while those calculated in the space group C2/m are only positive. This important difference indicates that, theoretically, the β′-As2Te3 phase in P21/m should be unstable at 0 K. In fact, as explained in our prior study,29 starting with the initial parameters of the P21/m space group, the DFT relaxation converges to the loss of the modulation along the b axis and hence, to a description of the crystal structure in the space group C2/m, stable at 0 K as shown in Fig. 3a. Although these DFT calculations suggest a higher symmetry at 0 K, we emphasize that the β′-As2Te3 in P21/m is likely stabilized by the phonon contribution at finite temperatures. In order to compare with specific heat measurements, the thermodynamic properties of β-As2Te3 and β′-As2Te3 were only computed in the space groups Rm and C2/m, respectively, within the quasi-harmonic approximation.
In addition to a monotonously increasing specific heat with increasing temperature as expected from a phonon contribution, the β-to-β′ phase transition shows up as a pronounced hump between 160 K and 210 K, peaking at 185 K. In agreement with our calculations, the similar specific heat values, below and above the transition, indicate nearly identical vibrational spectra due to the close relation between the crystal structures of the β and β′ phases. The relaxation curve, monitored across the transition (not shown) does not exhibit signs of shoulders, which would be direct evidence for a thermal arrest in agreement with the first-order nature of the phase transition. This is another consequence of the small crystallographic differences existing between both phases. The latent heat ΔH and entropy ΔS were calculated by integrating the Cp(T) and Cp(T)/T data, respectively, leading to respective values of ∼323 J mol−1 K−1 and 1.3 J mol−1 K−2, the latter being consistent with that estimated from differential scanning calorimetry measurements.29
Upon cooling, ρ(T) decreases from 300 to 200 K where a steplike increase is observed with a visible hysteresis between the cooling and heating data in the parallel direction. The −dρ/dT data (Fig. 5b) evidence that the transition extends between 210 K and 160 K with maxima at 186 and 188 K for the cooling and heating runs, respectively. This anomaly is consistent with the first-order phase transition revealed in our prior study.29 ρ(T) remains nearly constant from 160 K down to 2 K with a shallow minimum visible near 15 K in the parallel direction while in the perpendicular direction, ρ(T) smoothly decreases down to 25 K where it tends to level off. Upon warming, the ρ values measured in the parallel direction are higher than those taken during cooling with no sign of merging up to 300 K. This effect is however less discernable in the perpendicular direction to within experimental uncertainty. Similar behaviors have already been observed in other systems such as in Gd5(Si/Ge)4 where a first-order martensitic transition develops between 250 and 300 K.44 The slight difference in ρ(T) during the thermal cycle had not only been attributed to the transition itself but also to microcracks in the sample volume that form in the course of the transformation. Although the present transition does not show a martensitic character, internal stresses that appear during the transition may be responsible for the observed increase in ρ. This possibility seems corroborated by the data collected over three thermal cycles on the sample cut parallel to the pressing direction. As shown in Fig. 5c, a systematic change in the electrical resistivity ratio ρ300K/ρ5K is observed during cycles. The 2 K hysteresis observed during the first cycle remains nearly unchanged during the following cycles (Fig. 5d).
Measurement of the Hall coefficient reveals a dominant hole-like signal in the whole temperature range in agreement with our electronic band structure calculations. While the p-type character of β-As2Te3 remains unaffected across the transition, the changeover to the β′ phase is clearly visible as a pronounced minimum (Fig. 6). Likewise α-As2Te3 (ref. 42), the anisotropic crystal structure of β-As2Te3 results in a galvanomagnetic tensor with several independent components.45 This property could in principle give rise to distinct Hall coefficients in the parallel and perpendicular directions. However, in the present case, the anisotropy in the RH values amounts to at most 6% at 300 K, which is at the border of the experimental uncertainty. The Hall hole concentration p can be thus estimated to a good approximation with a single-carrier relation. The RH value yields an estimated p value of 9 × 1019 cm−3 at 300 K, which shows only little variations with temperature consistent with the idea that β-As2Te3 is intrinsically heavily-doped.
![]() | ||
Fig. 6 Hall coefficient RH as a function of the temperature measured parallel (![]() ![]() |
In agreement with the sign of the Hall coefficient, the thermopower (Fig. 7) is positive, indicative of holes as the dominant charge carriers. This result differs from that obtained by Scheidemantel et al.41 who measured negative thermopower values. However, this discrepancy is likely explained by the fact that their measurement was done under different conditions i.e. by subjecting α-As2Te3 to high pressures up to 10 GPa, the α phase transforming into the β phase at 8 GPa.41 As shown in Fig. 7, α varies linearly with temperature as typically observed in nonmagnetic metals and heavily-doped semiconductors. The transition is visible between 170 and 190 K as a small jump in the α values. Within experimental uncertainty, no difference between the data taken on cooling and heating is observed nor between the data measured in the parallel or in the perpendicular directions. This conclusion holds true regardless of the number of cycles performed (not shown) indicating that the hysteresis is too weak to be clearly observable in the α(T) data. This observation is consistent with the low-temperature X-ray and neutron diffraction experiments revealing that only subtle differences exist between the β and β′ phases.29 Because the thermopower is sensitive to changes in the electronic band structure near the chemical potential, these results are also consistent with our first-principles calculations indicating that the phase transition has little influence on the overall electronic properties of β-As2Te3.
As shown in Fig. 8a, the thermal conductivity is also influenced by the first-order transition. Noteworthy, the absence of dielectric maximum at low temperatures, featuring the “glass-like behavior” of κ, might be ascribed to the distorted structure of β′ alone (which might not be the case for the β phase, should it be stable below 180 K). The κ values, nearly constant between 200 and 300 K (0.8 W m−1 K−1), drop significantly below 200 K to 0.6 W m−1 K−1. The difference between the cooling and heating data at 300 K reflects almost entirely the difference observed in the ρ values that impact the electronic contribution κe to κ. This contribution was estimated using the Wiedemann–Franz relation κe = LT/ρ where L is the Lorenz number. As a first approximation, we used a single-parabolic band model assuming acoustic phonon scattering as the main mechanism of charge carrier diffusion to estimate the Lorenz number. This approach was extended down to 5 K to determine the temperature dependence of L that exhibits a room-temperature value of 1.9 × 10−8 V2 K−2 and tends to the degenerate value of 2.45 × 10−8 V2 K−2 in the limit T → 0 K. The resulting temperature dependence of the lattice thermal conductivity κL = κ − κe is shown in Fig. 8b. For sake of clarity, only the κL values estimated from the data taken upon cooling are shown, the values upon heating being nearly identical. The κL values are extremely low reaching 0.8 and 0.6 W m−1 K−1 at 300 K in the perpendicular and parallel directions, respectively. The β-to-β′ transition only weakly impacts the κL values in agreement with the trend previously revealed by the calculations of the phonon properties. These values are even lower than those measured in the Bi2Te3 analogues but are similar to those achieved in monoclinic α-As2Te3 and in several other chalcogenides such as SnSe or In4Se3.16,46–50 There is a large body of evidence indicating that the lattice thermal conductivity roughly scales inversely with the unit cell size, large unit cells favoring very low κL values. However, this general guideline breaks down in the present case, the simple crystal structure of β-As2Te3 suggesting the presence of an efficient phonon scattering mechanism in this material. To estimate the minimum lattice thermal conductivity κmin and hence, to determine whether or not further reduction may be achieved, we used the formula derived by Cahill et al.51,52 for disordered crystals. κmin depends on the longitudinal and transverse sound velocities vl and vt, respectively. Using the values derived in ref. 38 (vl = 3953 m s−1 and vt = 2357 m s−1), this approach yields κmin ≈ 0.4 W m−1 K−1 at 300 K (Fig. 8b). Near this temperature, the lattice thermal conductivity of β-As2Te3 approaches κmin in the parallel direction suggesting only little room for further optimization.
![]() | ||
Fig. 8 (a) Temperature dependence of the total thermal conductivity of β-As2Te3 measured parallel (circle symbol) and perpendicular (square symbol) to the pressing direction for cooling (in blue) and heating (in red) runs as indicated by the black arrows. (b) Lattice thermal conductivity as a function of temperature for the data collected upon cooling parallel (circle symbol) and perpendicular (square symbol) to the pressing direction. The black solid curve stands for the minimum lattice thermal conductivity estimated by the model developed by Cahill and Pohl.49,50 |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra01770c |
This journal is © The Royal Society of Chemistry 2016 |