A recent study has shown the thermoelectric performance of Al-doped Mg2Si materials can be significantly enhanced at moderate pressure. To understand the cause of this phenomenon, we have performed in situ angle dispersive X-ray diffraction and infrared reflectivity measurements up to 17 GPa at room temperature. Contrary to previous experiment, using helium as a pressure transmission medium, no structural transformation was observed in pure Mg2Si. In contrast, a phase transition from cubic anti-fluorite (Fm-3m) to orthorhombic anti-cotunnite (Pnma) was observed in the Al-doped sample at 10 GPa. Infrared reflectivity measurements show the electrical conductivity increases with pressure and is further enhanced after the phase transition. The electron density of states at the Fermi level computed form density functional calculations predict a maximum thermoelectric power factor at 1.9 GPa, which is in good agreement with the experimental observation.

Doped Magnesium silicide (Mg2Si)-based alloys have been suggested as a candidate of a new generation of high performance and environmental friendly thermoelectric materials.1–4 Compared with other lead-based thermoelectric materials, Mg2Si-based alloys have the merits of being non-toxic, sustainable, and low cost. It is well known that the efficiency of thermoelectric materials may be increased through p- or n-type doping at ambient pressure.5–9 A dopant can increase the carrier concentration and mobility of the conduction electrons. Doping with heavy atoms can affect the lattice vibrations and help to lower the thermal conductivity by increasing the phonon-phonon scatterings. Pressure can also be used to enhance the thermoelectric power factor. Recently, Mg2Si nominally doped with 1% Al was compressed to 2–3 GPa, and the thermal power was found to increase significantly reaching a maximum value of 8 × 10−3 W/(K2 m). In this case, the increase in the thermoelectric efficiency is associated with an increase of electrical conductivity, and it was suggested that the Al-doped sample became metallic between 5 and 12 GPa. A Raman spectroscopy study also hinted at two possible structural phase transitions at 5–7 GPa and 11–12 GPa.10 To date, the structure and the cause for the enhancement of the thermal power of the lightly Al-doped Mg2Si are still not known. This is the focus of this investigation.

At ambient pressure and temperature, Mg2Si has a cubic anti-fluorite structure (space group Fm-3m) and is a semiconductor with a small indirect band gap of 0.6 eV.11 Using energy dispersive X-ray diffraction and silicone oil as the pressure medium, it was reported that Mg2Si underwent a structural transition to an orthorhombic anti-cotunnite structure with space group Pnma at 7.5 GPa. Further compression led to a hexagonal Ni2In-type P63/mmc structure at 21.3 GPa.12,13 The high pressure structural phase transition behavior of pure Mg2Si was re-investigated by Zhu et al. with an angle dispersive diffraction using NaCl as the pressure transmitting medium.14 A phase transition from the anti-fluorite structure to the monoclinic structure was reported at 11.1 GPa. No further phase transition was found up to 37.5 GPa.

The discrepancy between the two diffraction measurements is unsettling and raises the question on the true identity of compressed Mg2Si and the role of the pressure transmission medium to the crystal structure. This is further complicated by experimental resistivity measurements of un-doped Mg2Si, where three distinct regimes separated at 7, 12, and 22 GPa were observed, but Mg2Si remained a semiconductor with the hint of eventual metallization at pressure exceeding 22 GPa.15 Since theoretical calculations have predicted that the high pressure anti-cotunnite and Ni2In structures are metallic,16,17 the resistivity measurements clearly show there is no insulator → metal transition below 7 GPa. As will be shown and discussed below, and contrary to earlier reports,12,13 the present study reveals no structural phase transition in pure Mg2Si up to 18 GPa. However, a structural phase transition was found in Al-doped Mg2Si at 11 GPa.

Since the reflectivity is related to the dielectric function, frequency dependent optical conductivity can be extracted from the analysis. Static (dc) conductivities can be obtained by extrapolation to zero photon energy. In-situ high pressure far and mid-infrared reflectivity measurements have been performed on a nominal 1 at. % Al-doped Mg2Si sample and compared to the results from bulk measurements.10 The general trend that the electrical conductivity increases with pressure is the same as that found in previous measurements. Data analysis is complemented with theoretical density functional band structure calculations, which reveal a partially filled mid-gap band due to localized electronic interactions between the Al dopant with the surrounding host atoms is responsible for the enhancement of the thermopower.

The objective of this paper is to investigate the high pressure structure and the relationship to the transport properties of Al-doped Mg2Si. For comparison, we have also examined the structure of pure Mg2Si. For this purpose, room-temperature high resolution synchrotron angle dispersive synchrotron X-ray diffraction experiments up to 14.7 GPa using helium as the quasi-hydrostatic pressure-transmitting medium. Compared to silicone oil and NaCl, helium has the advantage of remaining in the liquid state at higher pressure and, therefore, able to maintain the isotropic strain of the sample in the diamond anvil cell (DAC).

The layout of the paper is as follows. First, details on the experimental procedure will be described. Diffraction results on pure Mg2Si will be then compared with earlier studies. New results on the structure and structural transformation of Al-doped Mg2Si under pressure will then be presented. This is followed by a discussion on the electrical properties extracted from the analysis of far and mid-IR reflectivity spectra. Comparison of the electrical conductivity with the variation of theoretical electronic density of states (DOS) is then made. The paper concludes with a proposed mechanism for the increase in thermopower under pressure.

Electron microprobe analysis: Powder Mg2Si with a purity of 99.5% was purchased from Alfa Products. Mg2Si doped with nominal 1 at. % of Al was synthesized by spark plasma sintering technique. To quantitatively examine the chemical compositions of Al-doped Mg2Si powder samples, elemental analysis was performed on a JEOL 8600 Superprobe electron microprobe analyzer operating at 15 kV and 50 nA. The beam diameter was 5 mm. Dwell time on the peak was 60 s. SPI metals were used as standards for each element. Random sites on a compressed powder sample were selected and analyzed to give an unbiased sampling of compositions.

High-pressure powder x-ray diffraction measurement: The ambient powder diffraction patterns for pure and 1% Al-doped Mg2Si were measured at the CMCF beamline, Canadian Light Source (CLS). High-pressure X-ray diffraction experiments were performed at Sector 20, Advanced Photon Source (APS) at the Argonne National Laboratory, using synchrotron radiation (λ = 0.47685 Å). A DAC was used to generate the pressure. The powder sample was placed in the hole of a stainless steel gasket between the diamond anvils and then loaded with helium as the pressure transmitting medium. A ruby sphere was placed with the powder sample with the pressure determined from the peak shift of the ruby R1 and R2 fluorescence lines.18 The maximum pressure studied was 14.7 GPa for pure Mg2Si and 16.6 GPa for Al-doped Mg2Si. The diffraction patterns were analyzed using the JANA 2006 software package and the lattice parameters were determined by Le Bail fit.19,20

High-pressure infrared reflectivity measurement: High-pressure infrared reflectance spectra on 1% Al-doped Mg2Si sample were measured at the side-station of the U2A beamline at the National Synchrotron Radiation Facility, Brookhaven National Laboratory. The mid-infrared spectra were recorded on a Bruker Vertex 80v FTIR spectrometer and a Hyperion 2000 IR microscope attached with a liquid nitrogen cooled HgCdTe detector.

The far-infrared spectra were recorded by a vacuum liquid helium cooled bolometer detector. Powder sample of Al-doped Mg2Si was prepared and loaded into a stainless steel gasket placed between two 300 μm culets of a Sintek mini type IIa diamond anvil cell. A stainless steel gasket was pre-indented in the diamond anvil cell, and a 100 μm hole was drilled in the center of the indentation to serve as a sample chamber. Prior to the loading of the sample, the reflected power intensities of air-diamond base (Id) and air-diamond culet (Ic) interfaces were measured at each pressure point. This process allowed us to minimize the error due to the intensity difference between diamond culet and base. After loading the sample and closing the DAC, the power intensity reflected from the air-diamond base (Id) was measured again, and subsequently, the power intensity reflected from the sample-diamond interface (Isd) was measured.

The reflectivity of sample-diamond interface (Rsd) was calculated as Rsd = (Isd/Id) × (Id/Ic) × (Ic/I0), where I0 is the power intensity reflected from gold foil. And Ic/Io is a constant of 0.185. All the spectral data were collected at a resolution of 4 cm−1 and accumulated for 512 scans. Once again, the fluorescence from a ruby crystal placed in the powder sample was used for pressure calibration.

Frequency dependent optical conductivity was obtained by Kramers−Kronig (K−K) analysis of the data obtained from normal incidence reflectivity measurements.21 The data were subsequently fitted using a variational K−K constrained dielectric function, as implemented in the RefFIT code.22 After the correction for the diamond contribution, the frequency dependent optical conductivity is derived from fitting to a Drude-Lorentz (DL) model, and the dc conductivity is obtained from the extrapolation to zero frequency.

Electronic structure calculations: First Principles electronic calculations were performed using density functional theory (DFT) within the Perdew-Burke-Ernzerh (PBE) of parameterization of the generalized gradient approximation (GGA) as implemented in the Vienna ab initio simulation package (VASP) code.23–26 To obtain accurate band gap energy of pure Mg2Si at ambient and high pressure, additional calculations were performed with the GW approximation (GWA), which utilized the one-particle Green's function and screened Coulomb interaction W to account for the effect of electrons to the electronic band structure.27,28 For all calculations, the projector-augmented wave (PAW)29,30 potentials constructed from the generalized PBE functional were used with 3s23p1, 3s23p2, and 2p63s2 as valence electrons for the Al, Si, and Mg atoms, respectively. A 4 × 4 × 4 k-points mesh was used for total density of states (TDOS) and projected density of states (PDOS) calculations. A 1 at. % Al doped Mg2Si was constructed from a 2 × 2 × 2 supercell of the crystal where one of the Mg atoms is replaced with Al (i.e., Mg63Si32Al).

Room-temperature high resolution synchrotron angle dispersive X-ray diffraction patterns of pure Mg2Si were measured up to 14.7 GPa (Figure 1) using helium as the pressure medium. It was shown in a recent report comparing argon and silicone oil as the pressure transmission medium.31,32 The linewidth of diffraction pattern in silicone oil is substantially broadened above 10 GPa. Compared to silicone oil and NaCl, helium has the advantage of remaining in the liquid state at higher pressure and, therefore, able to maintain the isotropic strain of the sample in the DAC. Furthermore, the diffraction patterns will be interfered by the diffraction lines of NaCl. It is unlikely the helium will occupy the empty sites in Mg2Si. Since Mg2Si is built from a FCC Si lattice with Mg in the octahedral sites, the only possible site that helium atoms may diffuse into is the tetrahedral site. However, the powder diffraction pattern of pure Mg2Si was measured independently without a medium. The cubic unit cell parameter determined at ambient pressure agrees well with the results obtained in the DAC extrapolated to zero pressure with helium as the medium. In addition, the lattice constant and volume fitted with the 3rd order Birch-Murnaghan equation of state (red lines) are also in reasonable agreement with results of theoretical calculations (Figure 2). Therefore, there is no evidence that the helium can incorporate in the crystal lattice.

FIG. 1.

Angle dispersive x-ray diffraction patterns of pure Mg2Si at selected pressures measured at room temperature.

FIG. 1.

Angle dispersive x-ray diffraction patterns of pure Mg2Si at selected pressures measured at room temperature.

Close modal
FIG. 2.

(a) Lattice parameters of pure Mg2Si as a function of pressure obtained in the present study and compared with previous works. (b) The volume per formula of pure Mg2Si as a function of pressure and compared with previous works. The red color lines are fits to 3rd order Birch-Murnaghan equation of state.

FIG. 2.

(a) Lattice parameters of pure Mg2Si as a function of pressure obtained in the present study and compared with previous works. (b) The volume per formula of pure Mg2Si as a function of pressure and compared with previous works. The red color lines are fits to 3rd order Birch-Murnaghan equation of state.

Close modal

It is important to point out that we observed no impurity peaks that can be attributed to MgO. The lattice constants were determined from full profile Le Bail fit to the diffraction peaks between 7° to 18° using the JANA 2006 software package. All the X-ray diffraction patterns can be indexed readily to the ambient pressure cubic anti-fluorite (CaF2) structure with space group Fm-3m. The Miller indices (hkl) of the Bragg peaks are identified as (111), (200), (220), (311), (222), and (400) reflections with increasing scattering angles. No structural phase transition was found up to 14.7 GPa, the highest pressure studied. All diffraction patterns display the same profile except shifts of the Bragg peaks to higher angles with increasing pressure. The derived lattice parameters and unit cell volumes as a function of pressure were reported in Figures 2(a) and 2(b), respectively. The pressure coefficient for the cubic lattice determined from the experiment is −0.045 Å/GPa. The lattice parameters are in good accord with previous results by Zhu et al. and Hao et al. within the overlapping pressure region up to 7.5 GPa.13 

To examine the electronic property of Mg2Si at high pressure, theoretical density functional calculations with the PBE function were performed at 0, 4, 8, and 11 GPa (Figure 3). The calculated band structures were corrected for electron correlation effect as GW-correction.40 Obviously, Mg2Si is a semiconductor with the small band gap over this pressure range. The GW band gaps are 0.62, 0.46, 0.41, and 0.30 eV at 0, 4, 8, and 11 GPa, respectively. The theoretical results agree with the resistivity measurements that show Mg2Si is an insulator at least up to 22 GPa. Together with the electrical measurements, structural transitions to the high pressure anti-cotunnite and Ni2In phases at 7 and 11 GPa can be ruled out.

FIG. 3.

GW band structures of pure Mg2Si at 0.1, 4, 8, and 11 GPa.

FIG. 3.

GW band structures of pure Mg2Si at 0.1, 4, 8, and 11 GPa.

Close modal

Experimental angle dispersive synchrotron radiation X-ray diffraction patterns of 1% Al-doped Mg2Si were measured as a function of pressure up to 16.6 GPa (Figure 4). Again crystal lattice parameters were extracted from full profile of the diffraction patterns. At 0.9 GPa, diffraction peaks located at 7.5°, 8.7°, 12.3°, 14.5°, 15.1°, and 17.5° are indexed to (111), (200), (220), (311), (222), and (400) reflections of the cubic anti-fluorite structure of Mg2Si. Below 10.5 GPa, the diffraction patterns of 1% Al-doped Mg2Si are similar to pristine Mg2Si. At low pressure, a small amount of Al dopant is not expected to affect the crystal lattice of the host significantly.

FIG. 4.

Room temperature angle dispersive x-ray diffraction patterns of 1% Al-doped Mg2Si at selected pressures.

FIG. 4.

Room temperature angle dispersive x-ray diffraction patterns of 1% Al-doped Mg2Si at selected pressures.

Close modal

The unit cell of Al-doped Mg2Si is slightly larger than the pure Mg2Si below 2 GPa. This suggests the Al dopants should occupy the Mg sites in the crystal structure. Therefore, under ambient conditions and low Al-doping concentration, Mg2Si is expected to be an n-type semiconductor. This assignment is in agreement with the negative value of Hall coefficient reported.10 The cubic structure is maintained up to 10.5 GPa. Interestingly, the derived pressure coefficient for the cubic lattice of 0.023 Å/GPa is substantially smaller than in the pure sample. This observation indicates that even at 1% doping the Al-Mg2Si crystal lattice is more resilient to compression. Figures 5(a) and 5(b) show the pressure dependence of the lattice parameters and the unit cell volumes (Figure 5). At ambient pressure, the cell parameter of the cubic anti-fluorite structure (Fm-3m) of 6.2878(9) Å is reduced to 6.0261(11) Å at 10.5 GPa. Concomitantly, the cell volume per formula unit decreased from 62.15(12) Å3 to 54.71(12) Å3. It is interesting to note that the ambient pressure lattice constant obtained here with high resolution synchrotron radiation is noticeably shorter than the earlier report of 6.396(1) Å.10 As will be discussed below (vide supra), this difference may have important implications on the transport properties. The experimental results conclusively indicated the anti-fluorite structure (Fm-3m) of Al-doped Mg2Si is stable up to 10.5 GPa.

FIG. 5.

(a) Lattice parameters of anti-fluorite (Fm-3m), and anti-cotunnite (Pnma) phases for Al-doped Mg2Si and pure Mg2Si as a function of pressure. (b) The volume per formula of Al-doped Mg2Si and pure Mg2Si as a function of pressure.

FIG. 5.

(a) Lattice parameters of anti-fluorite (Fm-3m), and anti-cotunnite (Pnma) phases for Al-doped Mg2Si and pure Mg2Si as a function of pressure. (b) The volume per formula of Al-doped Mg2Si and pure Mg2Si as a function of pressure.

Close modal

Above 11.9 GPa, three new diffraction peaks marked by arrows in Figure 4 have emerged, indicating a change in the crystal structure. The new peaks at 10.7°, 11.4°, and 13.5° can be indexed to an orthorhombic anti-cotunnite (Pnma) structure with a = 5.8282(20) Å, b = 4.6313(12) Å, and c = 6.6059(23) Å. The strong diffraction peaks at 7.5°, 8.7°, and 12.3° can still be assigned to the cubic anti-fluorite structure (Fm-3m), indicating that the orthorhombic anti-cotunnite (Pnma) structure co-exists with the cubic anti-fluorite structure (Fm-3m) at this pressure. The amount of the cubic structure decreased with increasing pressure and eventually vanished at 16.6 GPa. Concomitantly, the diffraction peaks around 10.7°, 11.4°, and 13.5°, corresponding to orthorhombic anti-cotunnite (Pnma) structure, increased gradually and became dominant. Therefore, the diffraction results show unambiguously a structural phase transition of Al-doped Mg2Si had occurred around 11.9 GPa and was completed at 16.6 GPa. This structural transformation is accompanied by a large volume reduction of almost 23%. The experimental equation of state (volume vs. pressure) shown in Figure 5 indicates that the anti-cotunnite phase is more compressible than the cubic structure.

In the previous high energy dispersive X-ray diffraction study of pure Mg2Si using silicone oil as the pressure medium, the same Fm-3m to Pnma transition was observed at 7.5 GPa where the volume collapsed from the cubic structure of 56.3 Å3/f.u to Pnma of 49.4 Å3/f.u, a reduction of 12.3%.13 The transition in Al-doped Mg2Si was observed at a higher pressure of 10 GPa, and the volume reduced by 18% from the cubic phase of 55 Å3/f.u to Pnma 45 Å3/f.u. Therefore, the volume changes are comparable.

The behavior of pressure-induced structural phase transition for Al-doped Mg2Si is different from pristine Mg2Si which shows no structural transition up to 14.7 GPa. The phase transformation is related to the Al dopant in the crystal structure. A larger atomic size of Al compared to Mg expanded the cubic crystal lattice at low pressure, resulting in a smaller pressure coefficient (vide supra). We speculate this mismatch in the atomic size helps to promote the structural transformation at high pressure.

The measured reflectivities from ambient to 14 GPa are shown in Figure 6. Even at ambient pressure, the IR reflectivity shows a Drude-like behavior at a low frequency implying a metallic-like behaviour. In Mg2Si, the valence bands are completely filled, the electrons provided by the Al dopants must occupy the conduction band. By definition, the partial occupation of the conduction band is a metal. A practical definition for a metal is that the electrical conductivity should decrease with temperature. Previous experiment10 shows that the electrical conductivity of Al-doped Mg2Si indeed decreases with temperature, an indication of a metallic-like behaviour. We recognize that doped Mg2Si are often referred as a semiconductor. However, we believe it is just a matter of semantic. This point is highlighted from a comparison of the reflectivity of pure and Al-doped Mg2Si at zero pressure (Figure 7). In comparison to the flat and featureless reflectivity, typical of a semiconductor of pure Mg2Si, the extrapolated reflectivity at zero frequency of the doped sample is almost 0.4. On closer examination (inset of Fig. 6), it is observed that the reflectivity rose gradually from ambient pressure to about 10.5 GPa. From 8.5 to 10.5 GPa, there is a noticeable jump in the reflectivity from 0.32 to 0.58. Beyond this pressure, the reflectivities increase rapidly and reach 0.75 at 600 cm−1 at 14.1 GPa. This is to be compared with reflectivity close to unity at zero frequency for a good metal, such as copper or aluminum. Qualitatively, the observed trend in the reflectivity parallels that of the bulk measurements reported previously.10 It was reported that the electrical resistivity decreases (conductivity increases) almost by one order of magnitude from 0 to 1 GPa but became more gradual from 1 to 8 GPa. However, the rapid drop (rise) in the electrical resistivity (conductivity) at low pressure reported earlier by quasi-4-probe method10 was not so obvious in the infrared reflectivity measurements.

FIG. 6.

Infrared reflectivity spectra of 1% Al doped Mg2Si under different pressures.

FIG. 6.

Infrared reflectivity spectra of 1% Al doped Mg2Si under different pressures.

Close modal
FIG. 7.

Comparison of the infrared reflectivity spectra of 1% Al-doped Mg2Si and pure Mg2Si under ambient conditions.

FIG. 7.

Comparison of the infrared reflectivity spectra of 1% Al-doped Mg2Si and pure Mg2Si under ambient conditions.

Close modal

The frequency dependent conductivity (optical conductivity) can be obtained by performing a K−K analysis of the reflectivity data using the RefFIT code.22 The procedure is as follows: the optical conductivity obtained from a variational K-K transformation is fitted to a DL model and the dc conductivity is estimated by extrapolation to zero frequency. A typical result shows in Figure 8, illustrating the quality of the fitting procedure. In the Figure, the original and the fitted infrared reflectivity spectra at 10.5 GPa are compared. The dc conductivities are obtained with this procedure (Figure 9), where an almost linear dependence with pressure. The dc conductivity is 80 S cm−1 at ambient pressure and reaches 300 S cm−1 at 14.1 GPa, a fourfold increase. It is apparent that there are two “plateau” regions in the dc conductivity at ca. 6 and 10 GPa. However, due to the simplicity of the Drude-Lorentz model and the limited low frequency range accessible by IR radiation, there is not enough firm data to make a definitive statement. Nevertheless, the results indicate that the dc conductivity of Al-doped Mg2Si can be improved significantly by increasing pressure.

FIG. 8.

A comparison of measured and fitted infrared reflectivity spectra of 1% Al doped Mg2Si at 10.5 GPa.

FIG. 8.

A comparison of measured and fitted infrared reflectivity spectra of 1% Al doped Mg2Si at 10.5 GPa.

Close modal
FIG. 9.

The change of the dc conductivities of 1% Al-doped Mg2Si changes with pressure.

FIG. 9.

The change of the dc conductivities of 1% Al-doped Mg2Si changes with pressure.

Close modal

In the free electron model, the electron scattering relaxation time (τ) is related to be the electrical conductivity. This parameter can be extracted from the fitting the reflectivity data to the Drude-Lorentz model.33 The relaxation times obtained at different pressures are summarized in Table I. The derived relaxation times of ca. 10−14 s is consistent with other doped semiconductors.

TABLE I.

Summary of dc conductivity (σ) and carrier relaxation time (τ) of 1% Al-doped Mg2Si changing with the function of pressure.

Pressure (GPa)σ (S cm−1)τ (s)
0.3 90 2.9 × 10−14 
0.6 81 6.1 × 10−14 
0.9 102 6.3 × 10−14 
2.0 134 7.1 × 10−14 
2.3 128 9.3 × 10−14 
3.3 135 7.3 × 10−14 
3.7 143 7.4 × 10−14 
4.3 141 8.0 × 10−14 
4.8 147 6.7 × 10−14 
5.3 143 6.3 × 10−14 
5.9 164 6.2 × 10−14 
6.8 193 6.7 × 10−14 
7.7 213 6.5 × 10−14 
8.5 213 4.6 × 10−14 
9.6 218 2.8 × 10−14 
10.5 229 2.5 × 10−14 
11.4 259 1.4 × 10−14 
12.5 284 0.7 × 10−14 
13.4 292 0.6 × 10−14 
14.1 293 0.4 × 10−14 
Pressure (GPa)σ (S cm−1)τ (s)
0.3 90 2.9 × 10−14 
0.6 81 6.1 × 10−14 
0.9 102 6.3 × 10−14 
2.0 134 7.1 × 10−14 
2.3 128 9.3 × 10−14 
3.3 135 7.3 × 10−14 
3.7 143 7.4 × 10−14 
4.3 141 8.0 × 10−14 
4.8 147 6.7 × 10−14 
5.3 143 6.3 × 10−14 
5.9 164 6.2 × 10−14 
6.8 193 6.7 × 10−14 
7.7 213 6.5 × 10−14 
8.5 213 4.6 × 10−14 
9.6 218 2.8 × 10−14 
10.5 229 2.5 × 10−14 
11.4 259 1.4 × 10−14 
12.5 284 0.7 × 10−14 
13.4 292 0.6 × 10−14 
14.1 293 0.4 × 10−14 

To understand the mechanism for the pressure enhancement of the thermoelectric power factor, theoretical density functional calculations were performed at several pressures using a Mg63Si32Al1 supercell model (vide supra). No GW corrections were needed since the model system is already metallic-like. The aim of the calculations was to investigate the total and projected DOS and how they affect the thermopower with pressure. The results of the calculations are shown in Figure 10.

FIG. 10.

Total density of states and projected density of states for Al-doped Mg2Si at (a) 0.1 GPa, (b) 0.9 GPa, (c) 1.9 GPa, (d) 3.3 GPa, (e) 4.9 GPa, (f) 6.4 GPa, (g) 8.4 GPa, and (h) 10.3 GPa.

FIG. 10.

Total density of states and projected density of states for Al-doped Mg2Si at (a) 0.1 GPa, (b) 0.9 GPa, (c) 1.9 GPa, (d) 3.3 GPa, (e) 4.9 GPa, (f) 6.4 GPa, (g) 8.4 GPa, and (h) 10.3 GPa.

Close modal

At zero pressure, Al-doped Mg2Si already has a metallic-like behaviour. The Al impurity produced a distinct narrow band with very high DOS located between the valence and conduction band of the Mg2Si host. The sharp feature is due to localized interactions between the Al dopant with Mg and Si electronic states of the host.

From 0.1 to 3.3 GPa, a localized energy band was predicted to situate between the original gap of pristine Mg2Si. As the pressure was increased, this electronic band broadened and became more “free-electron” like. This trend explains the observed increase in metal-like behaviour, as well as the higher electrical conductivity and infrared reflectivities. It is known that the thermoelectric power factor is dependent on both the density of states and the derivative at the Fermi energy.33,34 A significant increase in the thermoelectric power factor is usually caused by the existence of a sharp localized DOS at the Fermi level.35,36 The calculate DOS at the Fermi level N(Ef) was found to decrease from 19 electronic states/eV/spin at 0.1 GPa to 9 electronic states/eV/spin at 10.3 GPa. In comparison, the slope of the density of states (dN(E)/dE)|Ef increased from 561 states/eV2 at zero pressure to a maximum value of 604 states/eV2 at 1.9 GPa and then decreased rapidly to 104 states/eV2 at 10.3 GPa (Figure 11 and Table II). On theoretical ground, it is expected the effect of N(Ef) and (dN(E)/dE)|Ef will result in an initial increase in the thermopower to 1.9 GPa then followed by a rapid drop. The predicted trend is in qualitative agreement with the measured thermopower, which has a maximum at 2.3 GPa.10 

FIG. 11.

(a) The derivative of total density of states (dN(E)/dE)|Ef. (b) The magnitude of total density of states at the Fermi level N(Ef) for Al-doped Mg2Si as a function of pressure.

FIG. 11.

(a) The derivative of total density of states (dN(E)/dE)|Ef. (b) The magnitude of total density of states at the Fermi level N(Ef) for Al-doped Mg2Si as a function of pressure.

Close modal
TABLE II.

Pressure dependence of total density of states N(E) at Fermi energy (Ef) and the derivative of N(E) at Fermi energy (Ef) of 1% Al-doped Mg2Si.

Pressure (GPa)N(E) at E = EfdN(E)/dE at E = Ef
0.1 19 561 
0.9 19 591 
1.9 17 604 
3.3 12 455 
4.9 13 263 
6.4 10 212 
8.4 139 
10.3 104 
Pressure (GPa)N(E) at E = EfdN(E)/dE at E = Ef
0.1 19 561 
0.9 19 591 
1.9 17 604 
3.3 12 455 
4.9 13 263 
6.4 10 212 
8.4 139 
10.3 104 

Although the trend of increasing electrical conductivity with pressure is the same as obtained from IR reflectivities and four-probe conductivity measurements,10 the comparison of the absolute magnitude is less satisfactory. One contributing factor may be the infrared beam that only surveyed a very small region (20 × 20 μm2) and penetrated only a few microns into the sample surface. This process is different from the quasi-four-probe conductivity method, which is a bulk sensitive technique. At the sample surface, the atom density is less than the bulk, and the chemical bonding in the surface is also stronger.37,38 Therefore, there are fewer free electrons near the sample surface, thus reducing the electrical conductivity. Although plausible, this effect cannot satisfactory explain the discrepancy observed here. Since the dc conductivity was obtained from the extrapolation of the frequency dependent conductivity to zero frequency, it is not certain if the omission of reflectivity at very low energy (i.e., <200 cm−1) may have an effect. From past experience, we expect the dc conductivity derived from an IR measurement should agree within an order of the magnitude of the value obtained from the bulk technique.39,40 We found no systematic error in the experiment, nor in the treatment of the data. In a previous study, we have compared the dc conductivity derived from IR reflectivities to bulk measurements on doped Mg2Si and the agreements were favorable.39 We suspect the most likely source of the disagreement may be related to the differences in the concentration of the dopant on the surface. The Al-doped Mg2Si used in this study was synthesized by plasma spark sintering method. With this technique, this is difficult to control the precise stoichiometry and the homogeneity of the doped samples. As mentioned above, compared to the pure crystal, the cubic lattice constant of doped Mg2Si is expanded by the inclusion of the Al dopants. The unit cell parameter for the sample used in this study is 6.2878(9) Å, which is noticeably shorter than the reported lattice constant of 6.396(1) Å on a previous sample. A larger unit cell suggests the concentration of Al in the sample used in the earlier investigate is probably higher.

Since Al-doped Mg2Si is n-doped with the Al atoms providing electrons into the conduction band, a higher Al content will increase the carrier concentration and enhance the electrical conductivity. It is noteworthy that from the diffraction patterns, unlike in the previous study, the Al-doped Mg2Si sample used in this study is free of MgO and other impurities. Another possible source of the discrepancy may be due to non-uniform distribution of Al in the sample.

In a study of Sb and Bi doped-Mg2Si, it was found by high resolution transmission microscopy (TEM) that due to the limited solubility excess Sb and Bi atoms are present in the grain boundaries that may enhance the conductivity of the bulk sample.39 In comparison, IR measurements only examine a very small spot of the sample so this may produce different results. A careful characterization of both samples is critical to resolve the discrepancy. For this purpose, we examined the chemical compositions of the Al-doped Mg2Si sample using scanning electron microscopy (SEM). The chemical compositions of several randomly chosen spots on the surface of a compressed sample were analyzed (Figure 12). The results of the analysis are summarized in Table III, showing that the doping is highly non-uniform. The concentrations of the doped Al can vary between a minimum of 0.27% at site #4 to a maximum of 1.42% at site #6. The large inhomogeneity in the dopant concentrations will certainly affect the local electronic. Since infrared reflectivity only probes a small spot, we believe that this is the main reason for the discrepancy between the electrical conductivity determined from infrared reflectivities and the bulk measurements.

FIG. 12.

Scanning electron microscopy (SEM) image of Al-doped Mg2Si powder sample.

FIG. 12.

Scanning electron microscopy (SEM) image of Al-doped Mg2Si powder sample.

Close modal
TABLE III.

The chemical compositions of Al-doped Mg2Si powder sample were quantitatively determined by the electron microprobe analysis.

PositionSi (wt. %)Al (wt. %)Mg (wt. %)
35 0.48 63 
33 0.36 64 
35 0.57 64 
34 0.26 65 
35 0.39 65 
34 1.4 63 
32 0.62 65 
PositionSi (wt. %)Al (wt. %)Mg (wt. %)
35 0.48 63 
33 0.36 64 
35 0.57 64 
34 0.26 65 
35 0.39 65 
34 1.4 63 
32 0.62 65 

The structural transformation and electrical conductivity of a nominally 1 at. % Al-doped Mg2Si sample synthesized from spark plasma sintering have been investigated by angle dispersive synchrotron radiation X-ray diffraction and infrared reflectivity up to 16.6 GPa at room temperature. Contrary to two earlier reports, no structural transformation was observed in pure Mg2Si up to 15 GPa when helium was used as the quasi-hydrostatic pressure transmission medium. The electronic band structures of pure Mg2Si were calculated at selected pressures with density functional theory including the correction for electron correlation effect with the GW approximation. It is found that the band gap of Mg2Si does not close at 11 GPa. The theoretical result is consistent with electrical resistivity measurements, which show that Mg2Si remains an insulator at least up to 22 GPa. Both studies contradict the reported structural phase transition to metallic anti-cotunnite and Ni2In phases 7 and 11 GPa, respectively. However, in Al-doped Mg2Si, a structural phase transition from the cubic anti-fluorite to the anti-cotunnite structure was found to occur around 11.9 GPa. Infrared reflectivity show the Al-doped Mg2Si already has a metallic-like behaviour under ambient conditions.

The dc conductivities at high pressure were calculated from the analysis of the infrared reflectivity spectra employing the Drude-Lorentz model. In agreement with bulk four-probe measurements, the electrical conductivity was found to increase with pressure as the sample became more metallic-like. However, the values of the dc conductivity derived from IR reflectivities are consistently lower. We attribute this difference to the difference in the stoichiometry of the sample used in the experiment and non-uniform doping. Theoretical density functional calculations reveal the presence of a mid-gap localized electronic band below 3.3 GPa. The sharp mid-gap DOS is the result of hybridization between the electron orbitals of the host elements with the dopant. As the pressure is increased further, the localized band broadened and the profile became more free-electron like. The derivative of the electron density of states N(E) at the Fermi energy (Ef) is found to maximize at 1.9 GPa. Thus, the theory predicted a maximum thermoelectric power factor, which is in agreement with the experimental observation at 2–3 GPa.

The present study provides new results and insight on the high pressure structures and transport properties of pure and Al-doped Mg2Si. The diffraction experiments found no phase transition at 6–7 GPa, suggested by the Raman study.10 However, transformation to an orthorhombic anti-cotunnite (Pnma) structure is confirmed at 11 GPa. The information obtained here may help to further enhance the performance of Mg2Si-based thermoelectric materials.

Ambient pressure synchrotron X-ray diffraction experiments were performed at the CMCF beamline, Canadian Light Source, which was made possible by support from NSERC, NRC, CIHR, and the University of Saskatchewan. High pressure synchrotron X-ray diffraction experiments were performed at sector 20 at the Advanced Photon Source, supported by the U.S. Department of Energy—Basic Energy Sciences, the Canadian Light Source and its funding partners, the University of Washington, and the Advanced Photon Source. Use of the Advanced Photon Source, an Office of Science User Facility operated for the U.S. Department of Energy (DOE) Office of Science by Argonne National Laboratory, was supported by the U.S. DOE under Contract No. DE-AC02-06CH11357. The use of U2A beamline was supported by COMPRES under NSF Cooperative Agreement EAR 11-57758 and CDAC (DE-FC03 03N00144). The National Synchrotron Light Source, Brookhaven National Laboratory, was supported by the U.S. Department of Energy, Office of Science, Office of Basic Energy Sciences under Contract No. DE-AC02-98CH10886. The electron microprobe analysis was supported by the Microscopy Lab, Department of Geological Sciences, University of Saskatchewan. J.Z. and J.S.T. thank AUTO21 for a Research Grant and to T. Bonli for the SEM measurements and J. Smith of HPCAT, Advanced Photon Source, for assistance with the high pressure diffraction experiments. We thank S. Tkachev for the help in using the gas-loading system, which is supported by GSECARS and COMPRES. APS is a user facility operated for the DOE Office of Science by ANL under Contract No. DE-AC02-06CH11357.

1.
T.
Sakamoto
,
T.
Iida
,
S.
Kurosaki
,
K.
Yano
,
H.
Taguchi
,
K.
Nishio
, and
Y.
Takanashi
,
J. Electron. Mater.
40
,
629
634
(
2011
).
2.
S.
Battiston
,
S.
Fiameni
,
M.
Saleemi
,
S.
Boldrini
,
A.
Famengo
,
F.
Agresti
,
M.
Stingaciu
,
M. S.
Toprak
,
M.
Fabrizio
, and
S.
Barison
,
J. Electron. Mater.
42
,
1956
1959
(
2013
).
3.
T.
Sakamoto
,
T.
Iida
,
A.
Matsumoto
,
Y.
Honda
,
T.
Nemoto
,
J.
Sato
,
T.
Nakajima
,
H.
Taguchi
, and
Y.
Takanashi
,
J. Electron. Mater.
39
,
1708
1713
(
2010
).
4.
S.
You
,
K.
Park
,
I.
Kim
,
S.
Choi
,
W.
Seo
, and
S.
Kim
,
J. Electron. Mater.
41
,
1675
1679
(
2012
).
5.
J.
Tani
and
H.
Kido
,
Intermetallics
16
,
418
423
(
2008
).
6.
N.
Wang
,
H.
Chen
,
H.
He
,
W.
Norimatsu
,
M.
Kusunoki
, and
K.
Koumoto
,
Sci. Rep.
3
,
3449
(
2013
).
7.
Q.
Chen
,
Q.
Xie
,
F.
Zhao
,
D.
Cui
, and
X.
Li
,
IEEE Comput. Soc.
22
,
338
341
(
2009
).
8.
G. J.
Snyder
and
E. S.
Toberer
,
Nat. Mater.
7
,
105
114
(
2008
).
9.
G. H.
Kim
,
L.
Shao
,
K.
Zhang
, and
K. P.
Pipe
,
Nat. Mater.
12
,
719
723
(
2013
).
10.
N. V.
Morozova
,
S. V.
Ovsyannikov
,
I. V.
Korobeinikov
,
A. E.
Karkin
,
K.
Takarabe
,
Y.
Mori
,
S.
Nakamura
, and
V. V.
Shchennikov
,
J. Appl. Phys.
115
,
213705
(
2014
).
11.
M. Y.
Au–Yang
and
M. L.
Cohen
,
Phys. Rev.
178
,
1358
(
1969
).
12.
J.
Hao
,
Z.
Guo
, and
Q.
Jin
,
Solid State Commun.
150
,
2299
2302
(
2010
).
13.
J.
Hao
,
B.
Zou
,
P.
Zhu
,
C.
Gao
,
Y.
Li
,
D.
Liu
,
K.
Wang
,
W.
Lei
,
Q.
Cui
, and
G.
Zou
,
Solid State Commun.
149
,
689
692
(
2009
).
14.
F.
Zhu
,
X.
Wu
,
S.
Qin
, and
J.
Liu
,
Solid State Commun.
152
,
2160
2164
(
2012
).
15.
F.
Yu
,
J. X.
Sun
,
W.
Yang
,
R. G.
Tian
, and
G. F.
Ji
,
Solid State Commun.
150
,
620
(
2010
).
16.
W.
Ren
,
Y.
Han
,
C.
Liu
,
N.
Su
,
Y.
Li
,
B.
Ma
,
Y.
Ma
, and
C.
Gao
,
Solid State Commun.
152
,
440
442
(
2012
).
17.
B. H.
Yu
and
D.
Chen
,
Chin. Phys. B
20
,
030508
(
2011
).
18.
H. K.
Mao
,
J.
Xu
, and
P. M.
Bell
,
J. Geophys. Res.
91
,
4673
4676
, doi: (
1986
).
19.
V.
Petricek
,
M.
Dusek
, and
L.
Palatinus
,
Z. Kristallogr.
229
,
345
352
(
2014
).
20.
A.
Le Bail
,
H.
Duroy
, and
J. L.
Fourquet
,
Mater. Res. Bull.
23
,
447
452
(
1988
).
21.
R. L.
Kronig
,
J. Opt. Soc. Am. Rev. Sci. Instrum.
12
,
547
557
(
1926
).
22.
A. B.
Kuzmenko
,
Rev. Sci. Instrum.
76
,
083108
(
2005
).
23.
G.
Kresse
and
J.
Hafner
,
Phys. Rev. B
47
,
558
(
1993
).
24.
G.
Kresse
and
J.
Hafner
,
Phys. Rev. B
49
,
14251
(
1994
).
25.
G.
Kresse
and
J.
Furthmuller
,
Comput. Mater. Sci.
6
,
15
(
1996
).
26.
G.
Kresse
and
J.
Furthmuller
,
Phys. Rev. B
54
,
11169
(
1996
).
27.
L.
Hedin
,
Phys. Rev.
139
,
A796
A823
(
1965
).
28.
P.
Garcia–Gonzalez
and
R. W.
Godby
,
Phys. Rev. B
63
,
075112
(
2001
).
29.
P. E.
Blochl
,
Phys. Rev. B
50
,
17953
(
1994
).
30.
G.
Kresse
and
D.
Joubert
,
Phys. Rev. B
59
,
1758
(
1999
).
31.
S. J.
You
,
L. C.
Chen
, and
C. Q.
Jin
,
Chin. Phys. Lett.
26
,
096202
(
2009
).
32.
M.
Fox
,
Optical Properties of Solids
, 2nd ed. (
Oxford University Press
,
2010
).
33.
G. D.
Mahan
and
J. O.
Sofo
,
Proc. Natl. Acad. Sci. U. S. A.
93
,
7436
(
1996
).
34.
R.
Kim
,
S.
Datta
, and
M. S.
Lundstrom
,
J. Appl. Phys.
105
,
034506
(
2009
).
35.
J. P.
Heremans
,
V.
Jovovic
,
E. S.
Toberer
,
A.
Saramat
,
K.
Kurosaki
,
A.
Charoenphakdee
,
S.
Yamanaka
, and
G. J.
Snyder
,
Science
321
,
554
(
2008
).
36.
J. H.
Lee
,
J.
Wu
, and
J. C.
Grossman
,
Phys. Rev. Lett.
104
,
016602
(
2010
).
37.
D. M.
Larson
,
K. H.
Downing
, and
R. M.
Glaeser
,
J. Struct. Biol.
174
,
420
(
2011
).
38.
S.
Mafe
,
J. A.
Manzanares
, and
P.
Ramirez
,
Phys. Chem. Chem. Phys.
5
,
376
(
2003
).
39.
N.
Farahi
,
M.
VanZant
,
J.
Zhao
,
J. S.
Tse
,
S.
Prabhudev
,
G. A.
Botton
,
J. R.
Salvador
,
F.
Borondics
,
Z.
Liu
, and
H.
Kleinke
,
Dalton Trans.
43
,
14983
(
2014
).
40.
J. W. L.
Wong
,
A.
Mailman
,
K.
Lekin
,
S. M.
Winter
,
W.
Yong
,
J.
Zhao
,
S. V.
Garimella
,
J. S.
John
,
R. A.
Secco
,
S.
Desgreniers
,
Y.
Ohishi
,
F.
Borondics
, and
R. T.
Oakley
,
J. Am. Chem. Soc.
136
,
1070
(
2014
).