MATERIALS TRANSACTIONS
Online ISSN : 1347-5320
Print ISSN : 1345-9678
ISSN-L : 1345-9678
Synthesis of High-Density Bulk Tin Monoxide and Its Thermoelectric Properties
Shimpei KuwaharaSora-at TanusilpYuji OhishiHiroaki MutaShinsuke YamanakaKen Kurosaki
Author information
JOURNAL FREE ACCESS FULL-TEXT HTML

2018 Volume 59 Issue 7 Pages 1022-1029

Details
Abstract

SnSe exhibits exceptionally high thermoelectric (TE) figure of merit zT mainly due to its ultralow lattice thermal conductivity (κlat) [L.-D. Zhao et al.: Nature 508 (2014) 373.]. It is considered that strong lattice anharmonicity caused by the lone pair electrons of Sn2+ results in the ultralow κlat. Here, we focus on SnO because it has the lone pair electrons of Sn2+ like SnSe. Bulk samples of SnO were synthesized by low-temperature high-pressure spark plasma sintering and their TE properties were examined. The present study revealed that SnO exhibits very low κlat (1.44 Wm−1K−1 at 573 K) compared with SnO2 which has no lone pair electrons. The Grüneisen parameter (γ) of SnO was evaluated to be 1.70 and this high γ leads to large lattice anharmonicity and thereby low κlat. Even though SnO has low κlat, the zT values were significantly low compared with SnSe. The maximum zT value of SnO was 0.00141 at 573 K. Since the main reason of this low zT is its non-optimized carrier concentration, the zT of SnO can be enhanced through the carrier concentration optimization.

1. Introduction

Thermoelectric (TE) materials can convert heat into electricity and vice versa. The conversion efficiency of TEs is mainly determined by the materials’ properties called dimensionless figure of merit, zT = S2σT/κ, where S is the Seebeck coefficient, σ is the electrical conductivity, T is the absolute temperature, and κ is the total thermal conductivity (κ = κlat + κel, sum of the contributions from lattice and electronic).1) Thus, to enhance the zT, high electrical properties, i.e. high S2σ, as well as low κ are required. Since there exists a trade-off relationship among S and σ or κel depending on the carrier concentration, the most effective way to enhance zT is the reduction of κlat, which is nearly independent of the carrier concentration.

The reduction of κlat with keeping high electrical properties has been demonstrated in many TE materials, including filled-skutterudites24) and nanostructured materials.510) In the filled-skutterudites, AM4X12, the A atoms called filler elements are bonded weakly to the other atoms and rattles, leading to strong scattering of phonons, resulting in low κlat.11) In the nanostructured materials, the nanoscale intermediates scatter phonons much more effectively than carriers, when the size of the structure is smaller than the phonon mean free path but larger than the carrier mean free path, leading to a reduction in κlat without a negative effect on the electrical properties.12,13) On the other hand, a new concept of lattice anharmonicity to reduce κlat has been proposed recently.1420) For example, an extremely low κlat (0.23 ± 0.03 Wm−1K−1 at 973 K) was obtained in SnSe,21) and it has been considered that this significantly reduced κlat is mainly due to the high rate of lattice anharmonicity induced by the lone pair electrons of Sn2+. This anharmonicity is driven by a stereochemically active lone pair electrons produced by the interaction between Sn (5p) and antibonding Sn (5s)-X (p) states.22) Owing to this quite low κlat, SnSe exhibits record-high zT value of 2.6 ± 0.3 at 923 K.21) However, despite having very high zT, usability of SnSe is limited by several disadvantages of using Se, because Se is highly toxic and small production scale.23,24)

Here, we focus on SnO which has lone pair electrons of Sn2+ like SnSe as O is in the same group 16 elements as Se.22) A fine bulk sample of SnO is hard to be synthesized because SnO is decomposed into SnO2 and Sn between 700 and 1050 K.25) It is reported that the following disproportionation reactions of SnO occur in a N2 atmosphere:   

\begin{equation*} \text{4SnO}(s) \to \text{Sn$_{3}$O$_{4}$}(s) + \text{Sn}(l) \end{equation*}
  
\begin{equation} \text{Sn$_{3}$O$_{4}$}(s) \to \text{2SnO$_{2}$}(s) + \text{Sn}(l), \end{equation} (1)

This is why the physical properties of bulk SnO have been hardly measured so far. In this paper, we report on the temperature-dependent κlat and the Grüneisen parameter (γ) at room temperature of SnO and discuss the relationship among the lone pair electrons, anharmonicity, and κlat. γ can be used as a measure of anharmonicity. For further discussion of the influence of the lone pair electrons on the κlat, we compared the κlat and γ values of SnO with those of SnO2 which has no lone pair electrons. Figures 1(a) and 1(b) show the crystal structures of SnO and SnO2, respectively. SnO has a tetragonal structure with the space group P4/nmm,26) while SnO2 has a tetragonal structure with the space group P42/mnm.27) Unlike SnO2 with a three dimensional structure, SnO has a layered structure where the lone pair electrons of Sn2+ above the apexes of SnO4 square-based pyramids reside in the interstitial space between the layers.28) Such a structural feature is expected to be responsible for strong lattice anharmonicity.

Fig. 1

Three-dimensional crystal structures of (a) SnO and (b) SnO2.

2. Experimental Procedure

SnO (200 mesh, 99.9%, Furuuchi Chemical Co. Ltd.) powders were used as raw materials to prepare the bulk samples. To determine the sintering conditions, high-temperature stability of SnO was checked by the thermogravimetry/differential thermal analysis (TG-DTA) measurement both in air and an Ar atmosphere with a heating rate of 20 Kmin−1. The SnO powders were placed into a tungsten-carbide die, followed by spark-plasma-sintering (SPS) in an Ar flow atmosphere under four different pressure/temperature conditions (#1 500 MPa/573 K, #2 500 MPa/598 K, #3 500 MPa/623 K, and #4 300 MPa/673 K). The powders used for the SPS condition #2 were ball-milled at a rotational speed of 300 rpm for 10 min before the sintering process. In addition to SnO, commercially-available SnO2 tablets (ϕ 10.0 mm, height 5.0 mm, 99.99%, 88.5%T.D., Kojundo Chemical Laboratory Co., Ltd.) were prepared and their physical properties were characterized to compare the data with those of SnO. The phase state and microstructure of the samples were evaluated at room temperature by powder X-ray diffraction analysis (XRD; Ultima IV, Rigaku Co.) with Cu Kα radiation and scanning electron microscopy (SEM, JSM-6500F, JEOL) equipped with an energy dispersive spectrometer (EDS). κ was calculated by κ = αCPd, where α, CP, and d are the thermal diffusivity, specific heat capacity at constant pressure, and density, respectively. α parallel to the SPS (pressing) direction was measured in the temperature range from room temperature to 673 K in an Ar atmosphere using a light flash apparatus (Netzsch LFA-467). The CP data were obtained from the literatures.29,30) d was calculated from the measured weight and dimensions of the bulk samples. κlat was calculated by subtracting κel from κ, where κel was estimated by the Wiedemann-Franz law, i.e. κel = LTσ (L is the Lorenz number for nondegenerate limit: 1.5 × 10−8 WΩK−2). The volumetric thermal expansion coefficient (αV) was evaluated by two methods, high temperature XRD analysis in a He atmosphere and direct measurement using a dilatometer in an Ar flow atmosphere. The αV perpendicular to the press direction was measured using a dilatometer. The αV of SnO and SnO2 were evaluated in the temperature ranges from room temperature to 673 K and 973 K, respectively. The longitudinal (vl) and transverse speed of sound (vt) were measured by an ultrasonic pulse-echo method at room temperature in air using 10 MHz sound wave echogenic transducers. These properties were measured in the parallel to the press direction. S and σ of the bulk samples of SnO were simultaneously measured in the temperature range from room temperature to 573 K using a commercial apparatus (Advance Riko ZEM-3) under a He atmosphere. These properties were measured in the perpendicular to the press direction. The Hall coefficient (RH) was measured by the Van der Pauw technique with a Hall measurement system (Toyo Resitest8300) at room temperature in vacuum under an applied magnetic field of 0.5 T. The Hall carrier concentration (nH) and Hall mobility (μH) were calculated from the expressions nH = 1/|RH|e (e: elementary electric charge) and μH = |RH|σ, respectively.

3. Results and Discussion

It was confirmed from the XRD analysis that the as-received SnO powders and SnO2 tablet are single phase materials with no remarkable impurities (Fig. 2). Figures 3(a) and 3(b) show the TG-DTA curves of as-received SnO powders performed in an Ar-flow atmosphere and in air, respectively. As can be seen in Fig. 3(a), SnO exhibits good thermal stability in an Ar-flow atmosphere up to around 1000 K. On the other hand, SnO oxidizes in air at around 700 K. In case of the complete oxidation of SnO to SnO2, the mass change is around 12%. However, in the present case, the mass change is only 1.6%, which can be confirmed from the TG curve in Fig. 3(b). This result means that the oxidation occurs only at the surface of the SnO powders and thus SnO2 layers act as barriers for farther oxidation. From the TG-DTA results, we determined the SPS conditions.

Fig. 2

Powder XRD patterns of as-received SnO powders and SnO2 tablet, together with the literature data of SnO31) and SnO2.32)

Fig. 3

TG-DTA curves of as-received SnO powders, (a) in an Ar-flow atmosphere and (b) in air.

The powder XRD patterns of the bulk samples of SnO synthesized by SPS are shown in Fig. 4, together with the literature data of SnO,31) SnO2,32) and Sn.33) It can be confirmed that the samples #1 (573 K, 500 MPa) and #2 (598 K, 500 MPa) are almost pure SnO, while the samples #3 (623 K, 500 MPa) and #4 (673 K, 300 MPa) contain SnO2 clearly as an impurity phase. At 623 K and above, the sample is stable during the TG-DTA measurement while it decomposes during the SPS process. It has been reported that the decomposition rate of SnO into SnO2 and Sn remarkably increases under high pressure (2–15 GPa).34) Therefore, the decomposition rate would also increase under the SPS pressure (500 MPa), which leads to the decomposition of the sample during the SPS process. Table 1 summarizes the lattice parameters and densities of the samples. Among the SnO samples, the lattice parameters are slightly different each other, likely due to the slight differences in the O/Sn ratio. The sample #2 exhibits the highest relative density of 90.2%, thus, we used this sample for physical properties measurements. The bulk XRD pattern of sample #2 for the perpendicular plane to the SPS (pressing) direction is shown in Fig. 4. It is confirmed that the intensity of the (002) peak of the bulk sample is slightly higher than that of the powder sample, meaning that the bulk sample is slightly oriented to the [001] direction. Although the bulk sample is slightly oriented to the [001] direction, the physical properties are evaluated under the assumption that the sample is not anisotropic but isotropic. The appearance of the bulk sample of SnO (sample #2) and as-received SnO2 tablet is shown in Figs. 5(a) and 5(b), respectively. Due to the difference in the band gap energy (Eg), SnO (Eg = 0.73 eV35)) is black colored, while SnO2 (Eg = 3.64 eV36)) is white. Figure 6 shows SEM images and the EDX mapping images of the bulk sample of SnO (sample #2) and as-received SnO2 tablet. Both samples are homogeneous and no remarkable cracks and pores. SnO2 tablet has slightly smaller grains (< ∼5 µm) than those of the SnO bulk sample. The white area in the SEM image of SnO may correspond to SnO2 existed as the impurity phase.

Fig. 4

Powder XRD patterns of the bulk samples of SnO synthesized by SPS and bulk XRD pattern of sample #2 for the perpendicular plane to the SPS (pressing) direction, together with the literature data of SnO,31) SnO2,32) and Sn.33)

Table 1 Lattice parameters and relative densities of the samples. The literature data of the lattice parameters of SnO31) and SnO232) are shown for comparison.
Fig. 5

Appearance of the bulk samples of (a) SnO (sample #2) and (b) as-received SnO2 tablet.

Fig. 6

SEM and EDX mapping images of (top three images) the bulk sample of SnO (sample #2) and (bottom three images) as-received SnO2 tablet.

Figure 7 shows the temperature dependences of the lattice thermal conductivities (κlat) of SnO and SnO2. Since σ values of both SnO and SnO2 are very small, κel can be neglected, i.e., κ ≈ κlat. In the present study, the samples contain pores, thus the Maxwell-Eucken equation was adopted for the correction of porosity in κlat. The equation is expressed as follows:37)   

\begin{equation} \kappa_{0} = \kappa_{\text{P}}\frac{(1 + \beta P)}{(1 - P)}, \end{equation} (2)
where κ0 is the thermal conductivity with no pores, κP is the thermal conductivity of a sample with porosity P, and β is a constant depending on the shape of the pores (here, β = 0.538)). The density-corrected κlat are plotted along with the measured data. As can be confirmed in Fig. 7, SnO has much smaller κlat values than SnO2.

Fig. 7

Temperature dependence of the lattice thermal conductivity (κlat) of the bulk sample of SnO (sample #2) and as-received SnO2 tablet. Density-corrected data are shown in lines.

To understand the origin of the low κlat of SnO compared with SnO2, the κlat is analyzed by using various physical properties obtained in the present study. Figure 8 shows the temperature dependences of the lattice parameters and lattice volume of SnO and SnO2 obtained by the HT-XRD measurement. It can be confirmed that all the parameters of both SnO and SnO2 increase linearly with increasing temperature. The average linear thermal expansion coefficient (αL) and the average volumetric thermal expansion coefficient (αV) are determined by the following equations:   

\begin{equation} \alpha_{\text{L}} = \frac{1}{l_{0}}\frac{dL}{dT}, \end{equation} (3)
  
\begin{equation} \alpha_{\text{V}} = \frac{1}{V_{0}}\frac{dV}{dT}, \end{equation} (4)
where l0 and V0 are the lattice parameter and lattice volume at room temperature, respectively. The values of αL along the a-axis (αa) and c-axis (αc) and αV are summarized in Table 2. The αV of SnO (9.97 × 10−5 K−1) is clearly larger than that of SnO2 (1.89 × 10−5 K−1), showing that larger anharmonicity of SnO than that of SnO2. As for SnO, the αc (5.21 × 10−5 K−1) is larger than the αa (2.33 × 10−5 K−1), which is consistent with the fact that SnO is bounded by a strong covalent bond in the a-axis direction, while by a weak van der Waals force in the c-axis direction.39) As shown in Fig. 1, SnO has a layered structure in the c-axis direction, while SnO2 does not have such structure.

Fig. 8

Temperature dependences of the lattice parameters a and c, and lattice volume V of SnO (top three figures) and SnO2 (bottom three figures).

Table 2 Linear thermal expansion coefficient along the a-axis αa, linear thermal expansion coefficient along the c-axis αc, linear thermal expansion coefficient αL, volumetric thermal expansion coefficient αV, longitudinal speed of sound vl, transverse speed of sound vt, average speed of sound vm, Debye temperature θD, bulk modulus B, and Grüneisen parameter γ of SnO and SnO2. The thermal expansion data are the average values in the temperature range between 300 and 700 K for SnO and 300 and 1000 K for SnO2. The data of the speed of sound are the density-corrected ones. The literature data for SnO232,42,43) are shown for comparison.

To check the validity of the values of αL and αV, we evaluated the parameters by direct measurements using a dilatometer (TD5000SA, Bruker AXS Inc.) under an Ar atmosphere. The temperature range was set as 300–673 K for SnO and 300–973 K for SnO2. In the dilatometer measurement, the αL can be defined by the following equation:   

\begin{equation} \alpha_{\text{L}} = \frac{1}{L_{0}}\frac{\Delta L}{\Delta T}, \end{equation} (5)
where L0 is the initial length, ΔL is the change of the length, and ΔT is the change of the temperature. The αV can be calculated using the following relationship:   
\begin{equation} \alpha_{\text{V}} = 3\alpha_{\text{L}}, \end{equation} (6)
Figure 9 shows the results of the dilatometer measurement. It can be confirmed that SnO shows larger thermal expansion than SnO2. The αL and αV determined by the dilatometer measurements are summarized in Table 2. The αV of SnO and SnO2 are 5.57 × 10−5 K−1 and 1.76 × 10−5 K−1, respectively. These smaller values than those determined by the HT-XRD measurements would be caused by the effect of porosity. It has been suggested that the porosity suppressed the thermal expansion of bulk samples.40)

Fig. 9

Linear thermal expansion of SnO and SnO2, measured by using a dilatometer.

In order to obtain accurate values of the sound velocity, the porosity correction was performed by using the measured sound velocities for three bulk samples of SnO with different porosity. The value corresponding to the sample with zero porosity was evaluated by linear extrapolation.41) Since it was difficult to prepare several samples of SnO2 with different porosity, the porosity correction for SnO2 was done by using the same parameter obtained for SnO. The values obtained by the porosity correction of the longitudinal speed of sound (vl) and transverse speed of sound (vt) are summarized in Table 2, together with the literature data of SnO2.32,42,43) The average speed of sound (vm), Debye temperature (θD), and bulk modulus (B) were calculated using the following equations:44,45)   

\begin{equation} v_{\text{m}}^{-3} = \frac{(v_{\text{l}}^{-3} + 2v_{\text{t}}^{-3})}{3}, \end{equation} (7)
  
\begin{equation} B = d_{0}\left(v_{\text{l}}^{2} - \frac{4}{3}v_{\text{t}}^{2}\right), \end{equation} (8)
  
\begin{equation} \theta_{\text{D}} = v_{\text{m}}\left(\frac{h}{k_{\text{B}}}\right)\left(\frac{3nN_{\text{A}}d_{\text{0}}}{4\pi M_{\text{W}}}\right)^{\frac{1}{3}}, \end{equation} (9)
where d0 is the theoretical density, h is the Planck constant, kB is the Boltzmann constant, n is the number of atoms in the molecule, NA is the Avogadro constant, and MW is the molecular weight. The calculated values of the vm, B, and θD of SnO are summarized in Table 2, together with the literature data for SnO2. It was confirmed that vm, B, and θD of SnO are smaller than those of SnO2.

The Grüneisen parameter (γ) which is related to the anharmonicity can be calculated as follows:46)   

\begin{equation} \gamma = \frac{\alpha_{\text{V}}BV_{\text{M}}}{C_{\text{V}}}, \end{equation} (10)
where VM is the molar volume, CV is the specific heat at constant volume which can be estimated by using Dulong-Petit law (CV = 3nR where R is the gas constant), and αV is the volumetric thermal expansion coefficient evaluated by the HT-XRD method. Lattice anharmonicity is closely related with αV, viz., large αV means large anharmonicity. Therefore, the Grüneisen parameter (γ) can provide an indication of degree of anharmonicity as γ is proportional to αV according to eq. (10). As summarized in Table 2, the calculated γ values were 1.70 for SnO and 1.27 for SnO2. SnO with lone pair electrons clearly exhibits high γ value than SnO2 with no lone pair electrons.

In order to find which parameter contributes to the difference in the κlat of SnO and SnO2, we analyzed the κlat based on the following equation:47)   

\begin{equation} \kappa_{\text{lat}} = A\frac{M\delta\theta_{\text{D}}{}^{3}}{N^{2/3}\gamma^{2}}\frac{1}{T}, \end{equation} (11)
where A is a constant independent of the material, M is the average molar mass per atom, δ3 is the average volume per atom, N is the number of atoms in the unit cell. At high temperatures above θD where Umklapp phonon-phonon scattering is dominant, the κlat can be calculated by eq. (11). Furthermore, the Grüneisen parameter (γ) is related to the relaxation time for Umklapp phonon-phonon scattering (τU), i.e., τU is proportional to γ−2.48) Figure 10 shows the ratios of the parameters (SnO/SnO2) of individual contributions for determination of κlat, where the ratio below 1 means that the parameter contributes to the lower κlat of SnO than that of SnO2. As can be seen in Fig. 10, the SnO/SnO2 ratios for γ and θD are less than 1, meaning that these parameters are dominant contributors to lowering κlat of SnO. Since the ratio for θD is smaller than that for γ, it can be said that the θD is more predominant than γ to reduce the κlat of SnO.

Fig. 10

Ratios of the parameters (SnO/SnO2) of individual contributions for determination of κlat.

Figure 11 shows the temperature dependences of (a) σ, (b) S, (c) power factor S2σ, and (d) zT of the bulk sample of SnO (sample #2). As can be seen in Fig. 11, these parameters increase with increasing temperature. The S values are positive, indicating that the SnO sample is p-type. The maximum values of the power factor and zT are 3.76 µWm−1K−2 @ 523 K and 0.00141 @ 573 K, respectively. Very recently, Miller et al. reported the experimental results on the TE properties of polycrystalline SnO bulk samples.49) These data are summarized in Table 3, together with the data of our sample. The nH of our sample (2.8 × 1018 cm−3 at room temperature) is clearly larger than the reported value (1.0 × 1016 cm−3 at room temperature). The difference in the nH between our sample and the sample reported in Ref. 49 would be caused by the difference in the amount of impurities existed in the samples. Miller et al. synthesized the SnO powders through a precipitation reaction method using SnCl2 as a precursor, where high-purity SnO powders compared with the commercial grade could be obtained. Here, we used commercially-available SnO powders (99.9%), and impurities existed in the powders would act as hole dopants and increase the nH. Due to this large nH, our sample shows small S and low μH compared with those reported in the literature. We will try to enhance the TE properties of SnO through the carrier concentration optimization. The results will be reported in near future.

Fig. 11

Temperature dependences of (a) electrical conductivity σ, (b) Seebeck coefficient S, (c) power factor S2σ, and (d) dimensionless figure of merit zT of the bulk sample of SnO (sample #2).

Table 3 Results of the Hall effect measurements and the Seebeck coefficient of the bulk sample of SnO (sample #2). The literature data for SnO49 are shown for comparison. All the data are obtained at room temperature.

4. Conclusion

It is considered that SnO can be a good TE material because it has lone pair electrons and thereby has large anharmonicity and low κlat. However, a fine bulk sample of SnO has been never synthesized because SnO is easily decomposed into SnO2 and Sn even at 700 K. In the present study, we succeeded in synthetizing high-density pure bulk SnO samples by using high-pressure and low-temperature SPS (500 MPa, 598 K). It was revealed that SnO has much smaller κlat values (1.44 Wm−1K−1 at 573 K) than SnO2 which has no lone pair electrons. It was confirmed that not only large Grüneisen parameter (γ = 1.70) but also low Debye temperature (θD = 335 K) contributes to this low κlat of SnO. Furthermore, we reported the TE properties of SnO, where the maximum values of the power factor and zT are 3.76 µWm−1K−2 @ 523 K and 0.00141 @ 573 K, respectively. The TE properties can be enhanced through the carrier concentration optimization.

Acknowledgments

This work was supported in part by JST, PRESTO Grant Number JPMJPR15R1.

REFERENCES
 
© 2018 The Thermoelectrics Society of Japan
feedback
Top