Dorra Ibrahima,
Shantanu Misraa,
Sylvie Migota,
Jaafar Ghanbajaa,
Anne Dauschera,
Bernard Malamana,
Christopher Semprimoschnigb,
Christophe Candolfi*a and
Bertrand Lenoir*a
aInstitut Jean Lamour, UMR 7198 CNRS – Université de Lorraine, 2 allée André Guinier-Campus ARTEM, BP 50840, 54011 Nancy Cedex, France. E-mail: christophe.candolfi@univ-lorraine.fr; bertrand.lenoir@univ-lorraine.fr
bEuropean Space Agency, ESTEC, P. O. Box 299, Keplerlaan 1, 2200 AG Noordwijk, The Netherlands
First published on 6th February 2020
Because the binary chalcogenide SnTe is an interesting Pb-free alternative to the state-of-the-art thermoelectric material PbTe, significant efforts were devoted to the optimization of its thermoelectric properties over the last few years. Here, we show that saturation-annealing treatments performed at 823, 873 or 973 K under Sn-rich conditions provide a successful strategy to prepare polycrystalline samples with a controlled concentration of Sn vacancies. Both scanning transmission electron microscopy and Mössbauer spectroscopy demonstrate the absence of Sn-rich areas at the grain boundaries in the saturation-annealed samples. Transport property measurements, performed over a wide range of temperatures (5–800 K), show that this technique enables achieving thermoelectric performances at 800 K similar to those obtained using Sn self-compensation. The three saturation annealing temperatures result in comparable transport properties across the entire temperature range due to similar hole concentrations ranging between 1.0 and 1.5 × 1020 cm−3 at 300 K. As equally observed in samples prepared by other synthetic routes, the temperature dependence of the Hall mobility evidences that charge transport is strongly affected by point-defect scattering caused by the random distribution of Sn vacancies.
As a Pb-free analogue to PbTe, SnTe has been for a long time overlooked due to its inherent off-stoichiometry associated with Sn vacancies that result in highly-degenerate transport properties and poor thermoelectric performances.11–33 As in most chalcogenide semiconductors, this defect chemistry depends sensitively on the conditions employed during the growth process and post-treatments, should the samples be in single-crystalline or polycrystalline form.34–37 A careful control of the defect concentration is pivotal in ensuring good reproducibility and meaningful comparison of the transport data between different series of samples.
Saturation annealing is the experimental tool of choice to control deviations from the ideal stoichiometry in semiconductors and to systematically probe the maximum deviation at a given temperature. In addition, this technique allows avoiding the presence of secondary phases which may have a detrimental impact on the transport properties. While this strategy has been widely applied on most state-of-the-art thermoelectric materials in single-crystalline form such as Bi2Te3 or PbTe,2,38–41 this approach has been only scarcely considered on polycrystalline compounds for which the presence of grain boundaries can prevent diffusion of the elements. Here, we demonstrate that this technique can be successfully applied on polycrystalline SnTe. Varying the saturation annealing temperature from 973 down to 823 K under Sn-rich conditions enables controlling the Sn vacancy concentration and hence, the transport properties. In particular, samples with hole concentrations equivalent to those achieved by intentionally introducing Sn excess can be obtained, but without the presence of elemental Sn at the grain boundaries, as demonstrated by scanning transmission electron microscopy and Mössbauer spectroscopy.
The obtained ingots were ground into fine powders before being consolidated by spark plasma sintering in graphite dies at 773 K for 5 min under a pressure of 65 MPa. The experimental density of the cylindrical pellets was determined by weight and dimensions to be higher than 95% of the theoretical density from X-ray diffraction data. Subsequently, the cylindrical pellets were cut with a diamond wire saw into cylinders (10 mm in diameter and 1 mm thickness) and parallelepiped-shaped samples (1.5 × 2 × 7 mm3).
The saturation annealing process was carried out by placing these cut samples in a silica sample holder of 5 cm long on one end while a powdered source was placed on the other end. This sample holder was then placed in a silica tube sealed under secondary vacuum (Fig. 1). The composition of the source used here (Sn55Te45) was chosen to probe the Sn-rich solidus line according to the Sn–Te binary phase diagram. The source was synthesized by a direct reaction of the elements at high temperature in sealed silica tubes. During the saturation-annealing process, the solid, liquid and gaseous phases are all present. According to the Gibbs phase rule under constant temperature and pressure, the variance or degree of freedom of this system F = C − P + 2, where C is the number of components and P is the number of phases, is equal to 1 (C = 2 and P = 3 in the present case). Subjecting the samples to this annealing process at a fixed temperature thus determines their final chemical composition which will correspond to the composition of the solidus on the Sn-rich side of the phase diagram. This technique allows mass exchange between the source and the samples in the vapor phase and can be used to fix the defect concentration which remains constant and homogeneous within the samples. This annealing procedure was performed at 823, 873 and 973 K in a horizontal furnace. The duration of this process has been estimated by considering the self-diffusion profile of Sn in SnTe,42,43 yielding saturation times of 21, 10 and 7 days at 823, 873 and 973 K, respectively.
High-resolution transmission electron microscopy (HRTEM) and scanning transmission electron microscopy (STEM) experiments were carried out using a JEOL ARM 200F – cold FEG TEM/STEM microscope with an accelerating voltage of 200 keV equipped with double Cs-correctors. The observed thin slice of a sample saturation-annealed at 973 K was prepared by the dual ion beam (FIB) – scanning electron microscope system using the in situ lift-out technique. STEM images were taken in high-angle annular dark-field (HAADF) and annular bright-field (ABF) modes.
119Sn Mössbauer spectra were recorded at 4.2 K for the three saturation-annealed samples, in order to exalt the response of a possible Sn0 occurrence, in transmission geometry with a spectrometer operating in the constant-acceleration mode. The polycrystalline absorber with natural abundance of 119Sn used for these experiments had a thickness of 10 mg cm−2. The source Ba119SnO3 with a nominal strength of 10 mCi was kept at 300 K to collect the Mössbauer spectra. A palladium foil of 0.5 mm in thickness was used as a critical absorber for Sn X-rays. The velocity was calibrated at 300 K against a 12 μm thick Fe foil. The 119Sn isomer shifts IS reported herein are referred to BaSnO3 at 300 K. The Mössbauer spectra were fitted by a least-squares method assuming Lorentzian-like peak shapes.
High-temperature measurements (300–800 K) of the electrical resistivity and thermopower were performed using a ZEM-3 apparatus (Ulvac-Riko) on other bar-shaped samples. The thermal diffusivity a was measured on the circular annealed pellets (∼10 mm diameter, ∼1 mm thickness) up to the same temperature with a Netzsch laser flash instrument (LFA 467) under continuous argon flow. The measured thermal diffusivity was used to calculate the total thermal conductivity κ via the formula κ = aCpd where Cp is the specific heat and d is the experimental density. Cp was estimated using the Dulong–Petit law 3NR where N is the number of atoms per formula unit and R is the gas constant. The temperature-dependence of the density was not taken into account in the present case.
Because HRTEM experiments were conducted on one thin slice, we further tried to unveil the possible presence of elemental Sn by 119Sn Mössbauer spectroscopy which is a bulk-sensitive probe. As shown in Fig. 3, the Mössbauer spectra measured at 4.2 K on the three saturation-annealed samples do not show any evident signal of elemental Sn0, in contrast to the spectrum collected on a Sn1.03Te sample prepared by melt-quenching where a slight but detectable asymmetry of the main peak was a clear indication of the presence of a Sn0 phase.45 This absence is a strong evidence that mass exchange between the samples and the source has occurred during the saturation annealing process. These results demonstrate that this method is equally applicable to control the stoichiometry of polycrystalline SnTe without resorting to additional Sn in the nominal composition, the excess of which is located at the grain boundaries.
Fig. 4 Temperature dependence of (a) the electrical resistivity ρ and (b) thermopower α of the samples annealed at 823, 873 and 973 K. For comparison purposes, data collected on a sample prepared by melt-quenching (MQ) have been added (ref. 45). The color-coded symbols are similar in both panels. |
These conclusions hold true when considering the evolution of the thermopower in Fig. 4b. While the MQ sample shows a quasi-linear temperature dependence consistent with a diffusive regime in a strongly degenerate system, the α(T) curves of the annealed samples display more complex variations with temperature. Our prior study has shown that this complex evolution of α(T) is strongly sensitive to the synthesis conditions used, or equivalently, to the hole concentration.45 In particular, the maximum in α that exists in all samples near 35 K has been attributed to a phonon-drag contribution, similarly to what has been observed in PbTe.50,51 Interestingly, the behavior of α(T) below 300 K in the saturation annealed samples differs from that observed in polycrystalline SnTe prepared by conventional synthesis routes.45 The phonon-drag peak near 35 K is followed by a decrease in the α values upon further heating up to about 200–250 K where α(T) starts increasing again with temperature. The strong superlinear variation observed above 250 K likely reflects the temperature-dependent density-of-states effective mass of holes which varies with temperature up to 800 K.16 Both the low-temperature peak α value and the temperature at which α(T) reaches its minimum are sensitive to the annealing temperature, as evidenced by clear differences between the samples annealed at 873 and 973 K.
The marked differences in the electronic transport between the MQ and the annealed samples suggests that the concentration of Sn vacancies is drastically reduced by annealing under Sn-rich conditions. This is confirmed by Hall effect measurements indicating that the hole concentration pH decreases from 1.4 to 1.1 × 1020 cm−3 at 300 K with decreasing the annealing temperature (Table 1 and Fig. 5a). These values are one order of magnitude lower than that measured in the MQ sample (1.1 × 1021 cm−3 at 300 K), confirming the strong reduction in the Sn vacancy concentration. Thus, the saturation annealing process provides a reliable method to control precisely the chemical potential in SnTe through the Sn vacancy concentration, achieved over a large temperature window ranging from 823 up to 973 K.
Sample | pH (cm−3) | δ (Sn1−δTe) | Composition | at% Te | ath (Å) | aexp (Å) |
---|---|---|---|---|---|---|
SnTe MQ | 1.10 × 1021 | 0.0176 | Sn0.9824Te | 50.88 | — | 6.3120 |
SA 823 K | 1.09 × 1020 | 0.0034 | Sn0.9966Te | 50.16 | 6.3221 | 6.3262 |
SA 873 K | 1.12 × 1020 | 0.0035 | Sn0.9965Te | 50.17 | 6.3218 | 6.3258 |
SA 973 K | 1.38 × 1020 | 0.0044 | Sn0.9956Te | 50.22 | 6.3200 | 6.3247 |
Fig. 5 (a) Hall hole concentration pH and (b) Hall mobility μH as a function of temperature for the samples annealed at 823, 873 and 973 K. For comparison purposes, data collected on a sample prepared by melt-quenching (MQ) have been added (ref. 45). The color-coded symbols are similar in both panels. |
The temperature dependence of the hole mobility μH, shown in Fig. 5b, confirms the physical picture that emerged from data collected on samples prepared by various synthesis techniques.45 The μH data vary with temperature above 15 K following a T−1/2 law indicating that hole transport is limited by alloy scattering as a result of the random distribution of Sn vacancies within the samples. The absence of elemental Sn in the annealed samples demonstrates that this behavior is not affected by impurities and thus, is an intrinsic characteristic of SnTe. The smoother variation observed in the MQ sample has been attributed to the possible influence of electron–electron scattering shown to play a role in heavily-doped PbTe compounds.50
If we make the reasonable assumption that each Sn vacancy provides two holes and that the Hall factor rH is close to unity in the range of hole concentrations measured in the present case,34,35,45 the pH values can be used to estimate the deviations from the ideal 1:1 stoichiometry by applying successively the two relations and [VSn] = Sδ. In these relations, [VSn] is the concentration of Sn vacancies, S = 1.58 × 1022 cm−3 is the number of Sn sites in the crystal structure and δ represents the deviation from the ideal stoichiometry when considering the chemical formula Sn1−δTe. The results, gathered in Table 1, are in agreement with the relation proposed by Brebrick (ref. 51) that relates the percentage of Te, y, in the chemical formula to the lattice parameter a = 6.3278 Å − 3.54(y − 0.5) where y = 1/(2 − δ). Moreover, the very good agreement with the deviations in the phase diagram determined by various authors (ref. 47–49; Fig. 6) shows that saturation annealing provides a direct access to the maximum deviation at the annealing temperature considered. Furthermore, the pH values are also consistent with the literature data obtained on similarly post-treated single crystals (Fig. 7).52 This further demonstrates that saturation annealing is also effective in polycrystalline SnTe despite the presence of grain boundaries that may have prevented diffusion of the elements from the surface towards the core of the bulk samples.
Fig. 6 Portion of the Sn–Te phase diagram showing the deviations from the ideal stoichiometry in SnTe determined by various authors (ref. 47–49) and in this work. |
Fig. 7 Saturation-annealing temperature as a function of the Hall hole concentration pH measured at 77 K. The literature data obtained on single crystals treated by saturation annealing on both the Sn-rich and Te-rich sides of the solidus line are shown in black (ref. 52). |
Fig. 8 shows the temperature dependence of the total thermal conductivity of all the samples. The monotonic increase in the κ values observed in the MQ sample is significantly influenced by saturation annealing. In the three annealed samples, a well-defined Umklapp peak dominates the low-temperature data near 30 K. The magnitude of this peak trends with the Sn vacancy concentration estimated from pH, the lowest pH values being synonymous with the highest peak value. The sensitivity of the Umklapp peak to the vacancy concentration is not restricted to the annealed samples and has been also observed in other SnTe samples irrespective of the synthesis technique employed.45 These variations are due to the Sn vacancies that act as efficient point-defect scattering centers which directly affects the phonon transport at low temperatures. This is confirmed by a fit of the low-temperature κL(T) data to the Debye–Callaway model (Fig. 9; the equations used can be found in the ESI†) which shows that the parameter A describing point-defect scattering scales as δ(1 − δ) (see Table 2 for the values of the fitting parameters).
Fig. 8 Temperature dependence of the total thermal conductivity κ of the samples treated at 823, 873 and 973 K. For comparison purposes, data collected on a sample prepared by melt-quenching (MQ) have been added (ref. 45). |
Fig. 9 Fit of the lattice thermal conductivity κL of the saturation-annealed samples to the Debye–Callaway model represented as a solid black curve. κL has been obtained by subtracting the electronic contribution κe estimated using the Wiedemann–Franz law κe = LT/ρ. The temperature-dependence of the Lorenz number has been calculated assuming a single-parabolic band model which provides a reasonable approximation below 300 K for these compounds (see ref. 45). |
Sample | L (μm) | A (10−42 s3) | B (10−18 s K−1) |
---|---|---|---|
SnTe (MQ) | 1.5 | 42 | 1.3 |
SA 823 K | 1.8 | 7.3 | 5.8 |
SA 873 K | 1.5 | 6.8 | 7.0 |
SA 973 K | 1.5 | 10 | 8.5 |
Fig. 10 shows the evolution of the ZT values as a function of temperature. The peak ZT value achieved in this series (0.55 at 800 K) is similar for the three annealed samples to within experimental uncertainty and slightly higher than that obtained in the MQ sample for which a maximum value of 0.4 is obtained at 800 K. The saturation annealing method thus yields samples with thermoelectric performances comparable to those achieved by introducing excess Sn in the nominal composition (maximum ZT of 0.65 at 800 K),11,33 but without the presence of elemental Sn at the grain boundaries.
Fig. 10 Dimensionless thermoelectric figure of merit ZT as a function of temperature of the samples annealed at 823, 873 and 973 K. For comparison purposes, data collected on a sample prepared by melt-quenching (MQ) have been added (ref. 45). |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c9ra10841f |
This journal is © The Royal Society of Chemistry 2020 |