Henrik L.
Andersen
a,
Jiawei
Zhang
a,
Hao
Yin
b and
Bo B.
Iversen
*a
aCenter for Materials Crystallography, Department of Chemistry and iNANO, Aarhus University, DK-8000 Aarhus C, Denmark. E-mail: bo@chem.au.dk
bTEGnology ApS, Lundagervej 102, DK-8722, Hedensted, Denmark
First published on 26th December 2016
Undoped, anion-doped (Sb, Bi), and cation-doped (Ca, Zn) solid solutions of Mg2Si0.4Sn0.6 have been prepared by a commercially feasible large-scale solid state synthesis method. The compositional and structural stability of the prepared samples are investigated by high resolution synchrotron powder X-ray diffraction (PXRD) in the potential application temperature range of 300–750 K. Quantitative compositional and structural information are extracted from the multi-temperature PXRD data by the Rietveld method. Detailed analysis of the PXRD data reveals an irreversible thermally induced partial conversion of Mg2Si0.4Sn0.6 into a discrete Sn-rich Mg2Si1−xSnx-phase in the undoped and anion-doped samples. On the other hand, the cation-doped samples only undergo very minor compositional and structural changes with increasing temperature, indicating a stabilizing effect of Ca and Zn on the Mg2Si0.4Sn0.6 solid solution. The structural instability of the undoped and anion-doped samples is corroborated by the measured electrical resistivity as function of temperature in the same temperature range, in which a clear difference is observed between values during initial heating and subsequent cooling. In contrast, the resistivity data of the cation-doped samples exhibit good repeatability for two thermal cycles, confirming that cation doping greatly improves the thermal stability. This work highlights the importance of conducting multiple temperature cycles in the measurement of physical properties combined with a thorough structural characterization in studies of thermoelectric materials.
In recent years, Mg2Si1−xSnx-based compounds have attracted substantial interest as potential materials for thermoelectric energy conversion in the medium temperature range ∼500–900 K.6,8–13 Relatively large thermoelectric figure-of-merit, zT, values above 1 are often reported.14–24 In addition, the chemical constituents are abundant, low-cost and environmentally benign, which makes Mg2Si1−xSnx a suitable candidate for large-scale commercial applications. Mg2Si and Mg2Sn are known to form solid solutions of Mg2Si1−xSnx, but the presence and extent of a miscibility gap in the phase diagram remains a controversial topic.25,26 Thermodynamic modeling based on experimental data predicts a splitting into discrete Sn- and Si-rich phases in the 0.1 ≲ x ≲ 0.7 region below ∼1150 K.27 The miscibility gap narrows with increasing temperature, making the composition Mg2Si0.4Sn0.6 thermodynamically unstable at 300 K but stable above ∼700 K. However, there have been several reports of seemingly stable solid solutions (undoped and doped) being synthesized within the miscibility gap.11,12,15,28–32 The explanation may be slow kinetics related to the phase separation at ambient conditions. Despite the well-known structural instability and formation Mg2Si1−xSnx solid solutions in the system, repeated thermal cycling in the measurement of physical properties combined with meticulous structural characterization before and after thermal cycling, employing sufficiently resolved powder X-ray diffraction (PXRD) and Rietveld analysis, is rarely seen in the literature.
In this work, the influence of various anion (Sb, Bi) and cation (Ca, Zn) dopants on the structural stability and thermoelectric properties of Mg2Si0.4Sn0.6 is investigated. The samples have been prepared by relatively large-scale solid state syntheses which makes it a feasible route for commercialization. In addition, the study is focused on the thermal stability of the compound, which is another key aspect in real-life applications of functional materials. Characterization by multiple temperature high resolution synchrotron powder X-ray diffraction has been carried out, and the structural evolution with temperature is followed. Quantitative information is extracted from the PXRD patterns through in-depth analysis by the Rietveld method. The ability of the various dopants to stabilize the compound within the miscibility gap is investigated. In order to simultaneously eliminate and accentuate the influence of the thermal history of the samples, characterization of the various physical properties is carried out on fresh and identically sintered pellets from the same synthesis batch.
m space group illustrated in Fig. 1 using the VESTA software.36 Small amounts of Sn and MgO impurities were identified in certain samples and these were modeled in space groups I41/amd and Fm
m, respectively.37,38 The instrumental contribution to the peak profiles was determined by Rietveld refinement of data obtained from a CeO2 standard, and it was corrected for in the refinements. The Thompson–Cox–Hastings formulation of the pseudo-Voigt function was applied to describe the peak profiles. The residual peak broadening after instrumental correction was described by refinement of profile parameters related to isotropic microstrain [Gaussian (U) and Lorentzian (X) contribution].39 The background was modeled using a linear interpolation between a set of background points of refinable intensity. Atomic displacement parameters (ADPs) and atomic positions were held fixed while the scale factor, unit cell parameters, peak profile and background parameters were refined. Simultaneous refinement of the ADPs and atomic occupancies lead to unphysical results due to the broad and often irregular peak profiles. The ADPs were therefore fixed at values estimated by stoichiometrically weighted extrapolation from Rietveld refinements of multi-temperature high resolution PXRD data of the pure Mg2Si and Mg2Sn end members.40 The ADPs of the impurities were held fixed at 0.5 for Sn and typical room temperature literature values for MgO in the refinements at all temperatures.38 Fixing the ADPs facilitated constrained refinement of Sn and Si occupancies in the Mg2Si1−xSnx structure. From the refined parameters, the weight fraction W of phase i was calculated by the formula, Wi = [SiZiMiVi/ti]/sum(j)[SjZjMjVj/tj], where S is the scale factor, Z is the formula units in the unit cell, M is formula unit mass, V is the unit cell volume and t is the Brindley particle absorption contrast factor defined as,
, where m is the linear absorption coefficient, μ is the average linear absorption coefficient of the entire sample, D is the crystallite diameter and Vol is the crystallite volume.41
![]() | ||
| Fig. 1 The cubic Mg2Si1−xSnx crystal structure with green spheres representing Mg and blue spheres representing Si and Sn. | ||
Considering the diffraction patterns in their entirety may lead to the conclusion that the samples consist of a single Mg2Si1−xSnx phase (apart from the aforementioned impurities). However, a closer examination of the individual peaks, as illustrated in the magnification of the (111) peaks in Fig. 2, reveals that this is not the case. In the diffraction patterns of MGS_undoped, MGS_Sb, and MGS_Bi, a shoulder is visible on the left side of the (111) peak and the shoulder is consistently present on all the diffraction peaks of the main Mg2Si1−xSnx phase (MGS1), indicating the presence of a bimodal distribution of solid solutions between MGS1 and another discrete Mg2Si1−xSnx phase (MGS2). These structural subtleties may significantly influence the material properties.
The two Mg2Si1−xSnx phases make up the bulk (∼78–97%) of all the samples. In MGS_undoped, MGS_Sb, and MGS_Bi, the secondary MGS2 phase accounts for 6.2(1)%, 10.5(2)% and 5.3(1)% of the Mg2Si1−xSnx, while MGS_Ca and MGS_Zn only contain the main phase at room temperature. Considerable amounts of MgO impurity (4–15%) are present in all the samples and for MGS_undoped, MGS_Sb, and MGS_Ca, small amounts of Sn (∼1–2%) are found as well. Notably, the actual MGS1 and MgO contents in the MGS_Ca sample are slightly lower than indicated as the unknown impurity phase was not accounted for in the refinement. The accuracy of quantitative phase analysis from PXRD experiments is influenced by microabsorption.48 The effect arises when the sample contains phases with different absorption coefficients and/or crystallite size distributions. This causes reflections from heavily absorbing phases to be suppressed while those of lighter absorbers are effectively enhanced. The coarseness of the samples makes it difficult to determine the crystallite size which again makes it hard to determine an appropriate Brindley particle absorption contrast.49 The Brindley factor t was therefore set to 1 for all phases in the refinements. Since MgO has a relatively low linear absorption coefficient (1.70 cm−1) compared to Mg2Si0.4Sn0.6 (28.37 cm−1), Mg2Sn (44.70 cm−1) and Sn (128.23 cm−1) at the given X-ray energy, the weight fraction of MgO will be overestimated. On the other hand, the amount of Sn in the sample is underestimated. Consequently, the values presented here should only be used as an indication of the composition, and for relative comparison among samples.
![]() | ||
| Fig. 4 Refined unit cell parameters as a function of x in Mg2Si1−xSnx calculated from the refined atomic occupancies and plotted together with values reported by Søndergaard et al.31 The estimated standard deviations are smaller than the symbols. A linear fit to the data reported by Søndergaard et al. has been added to illustrate the validity of Vegard's law for this system. The target stoichiometry is indicated by the dashed black vertical line. | ||
In conclusion, it has been found that the stoichiometry of the main MGS1 phase, which makes up the bulk of the samples (>78%), is very close to the targeted value. At room temperature, the secondary Sn-rich MGS2 phase, which was observed in smaller amounts (<11%) in certain samples, seems to be the phase pure Mg2Sn (x = 1) in MGS_undoped and MGS_Bi, and Mg2Sn0.95(1)Si0.05(1) for MGS_Sb. Considering that the only impurities visible from the PXRD data are Sn and MgO, this leads to the question of where the remaining Si has gone since it is not observable in the crystalline part of the sample.
For MGS_undoped, MGS_Sb and MGS_Bi, the amount of Sn-rich MGS2 phase increases when the temperature increases. This is apparent from the initially subtle shoulder on the left of the peak which develops into a separate and easily distinguishable Bragg peak. However, for the cation-doped samples MGS_Ca and MGS_Zn, only slight changes are observed when heating. A hint of a shoulder indicating the formation of a Sn-rich MGS2 phase is observed when the temperature is increased above 600 K. It seems that the doping by Ca or Zn dramatically improves the thermal stability of the MGS1 phase. Subsequent cycling of the temperature down to 315 K and back to 750 K only introduces a minor change in these samples.
The data shown in Fig. 5 were collected on powder samples in an Argon atmosphere, whereas physical property data are collected on pressed pellets. It is possible that decomposition processes may not be as significant in pressed pellets e.g. due to the pressed material having much less exposed surfaces. However, this is not the case as shown in the ESI Fig. S8–S12,† which plots conventional PXRD data measured in reflection geometry on the top and bottom sides of the pressed pellets before and after thermal cycling (300–725 K). Indeed, the PXRD data of the pellets agree very well with the multi-temperature synchrontron PXRD data with significant peak splitting occurring both on the top and bottom sides in the MGS_undoped, MGS_Sb and MGS_Bi samples. For the MGS_Ca and MGS_Zn samples only subtle changes are observed in the PXRD patterns following thermal cycling confirming the stabilizing effect of Ca and Zn doping also on the pellets.
The refined weight fractions of MGS1 and MGS2 shown in Fig. 6(A) confirm the initial assessment from the (111) Bragg peaks in Fig. 5 showing a simultaneous decrease in MGS1 and increase of MGS2. The highest degree of phase separation takes place in MGS_undoped where MGS2 goes from initially accounting for 6.9% of the total Mg2Si1−xSnx content at 300 K to 28.6% at 750 K. A less dramatic, however still significant, evolution in phase composition occurs in MGS_Sb and MGS_Bi where MGS2 go from 11.3% to 22.7% and 6.3% to 20.0% of the total Mg2Si1−xSnx content respectively. In MGS_Ca and MGS_Zn no MGS2 phase was observed at 300 K, however a slight amount forms when heating the samples above 550 K. Consequently, MGS2 accounts for 3.1% of the Mg2Si1−xSnx content in the Ca-doped and 8.7% in the Zn-doped at 750 K. The phase compositions of all samples remain relatively stable during the subsequent heating cycles. As mentioned earlier, the lack of microabsorption correction in the weight fraction calculation causes the relative amount of MGS2 to be underestimated and the given percentages should thus only be used as estimates.
The compositional stability of the individual MGS1 and MGS2 phases in the bimodal solid solution systems was also determined. Fig. 6(B) shows the stoichiometry parameter, x, for the Mg2Si1−xSnx phases as a function of temperature. The parameter has been calculated from the refined Si and Sn occupancies on the atomic Wyckoff 4a site in the structure and represents the mean stoichiometry of the given group of solid solutions, i.e. MGS1 or MGS2-phase. The stoichiometries of the main MGS1 phases remain stable at their initial values slightly above the targeted x = 0.6 throughout the temperature scans for all five samples. For MGS_undoped the x of MGS2 decreases gradually in a somewhat linear fashion with temperature, starting at x = 1 at 300 K and reaching x = 0.88(1) at 750 K. For the Sb-doped sample a slight decrease from x = 0.95(1) to 0.91(1) is observed. The MGS2 stoichiometry in MGS_Bi and MGS_Ca remain stable at the maximum x = 1 value. For MGS_Zn the MGS2 stoichiometry drops relatively fast from 1 to 0.92(1) after the secondary phase appears at 550 K.
The evolution of the crystallographic unit cell lengths of MGS1 and MGS2 with temperature is shown in Fig. 6(C). For all the samples the unit cell length increases linearly with temperature due to the thermal expansion of the material. Slight variations are seen which are likely related to the changes in stoichiometry for the MGS2 phases.
Fig. 6(D) shows the evolution of the Stokes-Wilson upper-limit isotropic microstrain, found from the modeling of peak profiles in the refinements.39 The strain has been normalized by the final value attained after the three thermal cycles. Within the Stokes-Wilson approximation, the strain varies from crystallite to crystallite but takes on a constant value within each crystallite. This effectively means that the calculated strain can be used as a measure of the distribution of Mg2Si1−xSnx phases around the mean stoichiometry. For all samples, the peaks sharpen and the strain decreases indicating a phase separation towards specific discrete stoichiometries. Most dramatic is the sharpening of the MGS2 distributions. The abnormal behavior of MGS2 in the Ca-doped sample is caused by the presence of the unknown impurity which is not included in the refined model. Additional thermal cycling induces no further changes in the refined Stokes-Wilson upper-limit isotropic microstrain.
The results illustrate the importance of investigating the structural and compositional changes that take place when subjecting the material to elevated temperatures. The phase composition of the material changes dramatically with temperature for certain samples during the first heating while remaining relatively stable during additional cycling of the samples in the same temperature range. Performing structural and compositional characterization solely prior to thermal cycling thus may lead to false conclusions. In the case of the investigated undoped, Sb-doped, Bi-doped, Ca-doped and Zn-doped Mg2Si1−xSnx samples it is clearly more meaningful performing the characterization after thermal cycling. Furthermore, the results show the importance of performing several thermal cycles in the characterization of the thermoelectric properties.
![]() | ||
| Fig. 7 Temperature dependence of electrical resistivity during heating and cooling of the first and second thermal cycles. | ||
For MGS_undoped, MGS_Sb, and MGS_Bi, the resistivity undergoes a considerable decrease in the first heating curve but only a minor increase during the subsequent first cooling curve. The difference between the first heating and cooling curves is substantial, indicating an irreversible process for the first thermal cycle. The rapid decrease of the first heating curve can be well explained by the temperature-dependent structural evolution, i.e., the increasing amount of the secondary MGS2 phase with increasing temperature (see Fig. 6(A)). For the MGS_undoped sample, the resistivity data of the second thermal cycle are consistent with the first cooling curve; however, for the anion-doped samples MGS_Sb and MGS_Bi, the second cooling curve still cannot repeat the first cooling curve. The above results reveal that Sb-doped and Bi-doped samples are not thermally stable, highlighting the importance of conducting structural characterization after the property measurement.
For MGS_Ca, in spite of the slight difference of the first heating curve for the resistivity data in the temperature range of 300–400 K, the first cooling curve and second heating/cooling curve of the two thermal cycles show excellent consistency between each other. In addition, the resistivity data of the MGS_Zn sample is reproduced upon repeated heating and cooling for two thermal cycles. The good repeatability of the resistivity curves in MGS_Ca and MGS_Zn is consistent with the very slight change of structures with increasing temperature (see Fig. 6(A)). Thus, the cation doping by Ca or Zn greatly improves the thermal stability of MGS1 phase.
![]() | ||
| Fig. 8 Electronic transport properties of all samples from room temperature to 725 K during first cooling. (A) Hall carrier concentration. (B) Hall carrier mobility. (C) Electrical conductivity. | ||
The Hall carrier mobility μH was calculated from the resistivity, ρ, and carrier concentration, nH, as follows, μH = 1/(nHeρ), and is shown in Fig. 8(B). The dashed lines illustrate μH ∝ T−1 and μH ∝ T−3/2 temperature dependences emphasizing the boundaries for the applicability of acoustic phonon scattering theory.50 For the anion-doped samples, it seems that acoustic phonon scattering dominates the transport behavior at elevated temperature while alloy scattering (μH ∝ T−1/2) contributes more significantly around ambient conditions. The Ca- and Zn-doped samples demonstrate acoustic phonon dominated scattering behavior throughout almost the entire temperature interval. Notably, the phase stability of the Zn-doped sample gives it a significantly higher mobility below 550 K due to the absence of impurity scattering.
Fig. 8(C) shows the measured electrical conductivity as a function of temperature. Compared with the undoped sample, the electrical conductivity values of cation-doped samples are enhanced by a factor of ∼2.4 at 350 K and a factor of ∼1.47 at 725 K. However, for the anion-doped samples, the conductivity at 350 K is similar to that of MGS_undoped and exhibits about 33% enhancement. Heavy doping of Mg2Si0.4Sn0.6 usually results in a more metallic-like behavior i.e. larger absolute values which decrease somewhat linearly with temperature.22,30,31 However, the doped samples prepared by the present large-scale synthesis method, exhibit typical semiconductor-like behavior, where the electrical conductivity increases with increasing temperature.
Fig. 9(A) shows the measured Seebeck coefficient data of the cooling curve of the first thermal cycle and a curve generated from polynomial fits to the raw data points. The absolute value of the Seebeck coefficient decreases with increasing temperature for all samples, which is consistent with the trend of the temperature-dependent resistivity data. At lower temperatures MGS_Zn has the largest Seebeck coefficient of ∼400 μV K−1 at 375 K but it drops relatively fast with temperature and is overtaken by MGS_Sb and MGS_Bi above 500 K which both decrease in a slow linear fashion. The curves of MGS_undoped and MGS_Ca nearly coincide and follow a trend similar to MGS_Zn but with lower values.
![]() | ||
| Fig. 9 (A) Seebeck coefficient as a function of temperature. (B) Power factor as a function of temperature. | ||
Fig. 9(B) shows the power factor of all samples as a function of temperature. The power factor of the undoped sample exhibits a nearly temperature-independent behavior from 350 to 475 K and then a gentle decreasing trend in the higher temperature range of 475–725 K. The power factor values of MGS_Zn and MGS_Ca decrease as the temperature increases. However, the power factors of anion-doped samples MGS_Sb and MGS_Bi increase with increasing temperature and then decrease, attaining maximum values of 5.8 and 5.6 μW cm−1 K−2 at 650 K and 675 K, respectively.
![]() | ||
| Fig. 10 Thermal transport properties as a function of temperature (A) Total thermal conductivity. (B) Lorentz number. (C) Lattice thermal conductivity. | ||
The total thermal conductivity, κ, shown in Fig. 10(A), is the sum of a lattice component, κL, and an electronic component, κe, i.e. κ = κL + κe. The electronic contribution can be estimated based on the Wiedemann–Franz law, κe = LσT, where L is the Lorentz number, σ is the electrical conductivity and T is the temperature. The Lorentz number (Fig. 10(B)) has been calculated using the single parabolic band model and assuming an acoustic phonon scattering mechanism for the entire temperature range.51 Detailed information about the calculation of the Lorenz number can be found in the ESI.† The very low Hall carrier concentration means that the electronic contribution to the thermal conductivity becomes minor. This is clearly illustrated by the similarity of the lattice (Fig. 10(C)) and the total thermal conductivity curves, where only a slight difference at high temperature is observed. The total and lattice thermal conductivity values all show increasing trends at high temperatures, which may be attributed to the bipolar conduction effect.
The highest zT values attained in this study are significantly lower than the zT values reported in a number of similar investigations in the literature, where zT values above 1 are routinely obtained.14–24 This may be attributed to the challenges associated with optimizing thermoelectric performance in large-scale, commercially viable synthesis methods.
![]() | ||
| Fig. 12 High resolution synchrotron PXRD data presented in the size and style currently the norm in the thermoelectric literature. | ||
If one uses the plotting procedure in Fig. 12, one could conclude that all samples are phase pure. In fact, as shown in Fig. 3, the investigated samples contain up to around 20% impurity phases. In the ESI Fig. S1 and S2† PXRD data measured for less than 30 minutes on an in-house Rigaku SmartLab diffractometer using Cu radiation are shown. The data clearly show that even standard laboratory diffractometers can provide sufficient resolution to discern the important structural subtleties of the present samples. As discussed above interpretation of the thermoelectric transport data would be of very limited value if not coupled with a robust structural analysis. The lack of data quality and limited importance given to structural analysis in the thermoelectric literature may be a serious limitation for making real progress the field, since interpretation of the transport properties possibly is based on wrong assumptions. Relating physical properties to structural features by meticulous characterization and analysis is the key to truly designing and tailoring functional materials with specific properties.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c6qi00520a |
| This journal is © the Partner Organisations 2017 |