Next Article in Journal
A Swiss-Roll-Type Methanol Mini-Steam Reformer for Hydrogen Generation with High Efficiency and Long-Term Durability
Next Article in Special Issue
Towards Optimizing Width Modulation for Maximum Thermoelectric Efficiency
Previous Article in Journal
A Conformal Tri-Band Antenna for Flexible Devices and Body-Centric Wireless Communications
Previous Article in Special Issue
Study on the Targeted Improvement Mechanism of the Carrier Concentration and Mobility of BiCuSeO Ceramics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Composition Adjustment on the Thermoelectric Properties of Mg3Bi2-Based Thermoelectric Materials

1
Institute of Engineering and Technology, Hubei University of Science and Technology, Xianning 437100, China
2
School of Mechanical and Electrical Engineering, Wuhan Donghu University, Wuhan 430070, China
3
The State Key Laboratory of Refractories and Metallurgy, Wuhan University of Science and Technology, Wuhan 430081, China
*
Author to whom correspondence should be addressed.
Micromachines 2023, 14(10), 1844; https://doi.org/10.3390/mi14101844
Submission received: 31 August 2023 / Revised: 15 September 2023 / Accepted: 18 September 2023 / Published: 27 September 2023

Abstract

:
Thermoelectric materials are widely used in refrigeration chips, thermal power generation, catalysis and other fields. Mg3Bi2-based thermoelectric material is one of the most promising thermoelectric materials. Herein, the Mg3Bi2-based samples were prepared by high temperature synthesis, and the influence of Mg/Sb content on the electrical transport properties and semi-conductivity/semi-metallicity of the materials has been studied. The results indicate that the efficiency of introducing electrons from excess Mg prepared by high temperature synthesis is lower than that introduced by ball milling, due to the high vapor pressure of Mg. The doping of Sb/Te at the Bi site would make it easier for the material to change from p-type conduction to n-type conduction. With the increase in Mg content, the semi-conductivity of the material becomes weaker, the semi-metallicity becomes stronger, and the corresponding conductivity increases. With the increase in Sb content, the samples exhibit the opposite changes. The highest power factor of ~1.98 mWm−1K−2 is obtained from the Mg3.55Bi1.27Sb0.7Te0.03 sample.

1. Introduction

Thermoelectric materials can realize the mutual conversion of thermal energy and electrical energy and can reduce the work function and improve the catalytic efficiency of chemical reactions [1,2,3]. The conversion efficiency of thermoelectric materials is evaluated by dimensionless thermoelectric figures of merit (ZT = S2σT/κ, S, σ, T, κ is the Seebeck coefficient, the electrical conductivity, the absolute temperature, the thermal conductivity, respectively) [4,5]. In order to achieve high thermoelectric conversion efficiency, thermoelectric materials are supposed to exhibit a high power factor (PF = S2σ) and low thermal conductivity, which is “phonon-glass and electron-crystal” [6,7]. In accordance with this principle, the representative materials of Bi2Te3-based [8,9], PbTe-based [10,11] and SiGe-based [12,13] thermoelectric materials have been developed in the past several years, and have been used in refrigeration [14,15], waste-heat power generation [16,17], micro-electronic devices [18,19], etc. However, traditional thermoelectric materials have the problems of poisonous elements [20,21], high costs [22] and so on, which have limited their further applications.
The new type of Mg3Bi2-based thermoelectric material is low-cost, non-toxic, and displays excellent thermoelectric performances, which make it one of the most promising thermoelectric materials for large-scale commercial application [23,24,25,26,27]. Mg3Bi2 (P-3m1 space group)-based thermoelectric material exhibits a special layered structure, with the Mg+ layer and the (Mg2Bi2) layer alternately stacked along the c axis [28,29]. The binding force between the Mg+ layer and the (Mg2Bi2) layer is relatively weak, and Mg vacancies are easy to generate in the material, exhibiting intrinsic p-type conduction [30]. Through composition adjustment, such as Te doping, a donor effect can be introduced with an increase in the number of electrons, and the material can be adjusted to n-type conduction. The thermoelectric temperature difference realized by the devices made of n-type Mg3Bi2-based thermoelectric material is significantly higher than that of commercial n-type Bi2Te3-based thermoelectric devices at the same temperature.
Mg-Bi-Sb-based thermoelectric materials are the most promising new materials to replace commercial bismuth telluride thermoelectric materials. They have great advantages in thermoelectric performance, maximum cooling temperature difference and cost. It is difficult for high-energy ball milling to meet the requirements of mass industrial production, but the smelting method is easy to realize. Previous literature tends to use the ball milling method to prepare this kind of material [31], but there are some problems, such as non-repeatability and non-uniformity in large-scale industrial application [32]. Herein, a high-temperature reaction method was adopted, which is easy to realize for large-scale industrial application [33]. Based on prepared Mg3Bi2-based thermoelectric materials, a composition adjustment was adopted to turn the material to n-type conduction and improve the power factor by Mg excess and Sb/Te doping [1].

2. Experimental Section

2.1. Materials and Synthesis

Samples with the chemical composition of Mg3Bi2-based thermoelectric materials were synthesized by a high-temperature reaction method combined with resistance pressing sintering (RPS) [34,35]. The commercial Mg (99.99%), Bi (99.999%), Te (99.999%) and Sb (99.999%) were used as starting materials. The raw materials come from China National Pharmaceutical Group. The powder, weighed according to the stoichiometric ratio and mixed in a glove box, was put into a capped graphite mold, then put into a quartz tube filled with argon and annealed at 1173 K for two hours. The corresponding heating rate is 50 K/min, and the cooling method is furnace cooling. The obtained blocks were crushed into powder in a glove box and then sintered by resistance pressing sintering (RPS) under the axial compressive pressure of 50 MPa under an argon atmosphere at 873 K for 10 min. Machine demolding was used to avoid sample breakage. The inner cavity size of the graphite mold is Φ 20 × 13 mm. The heating rate of the sintering process is 50 K/min, the cooling rate is 50 K/min, the pressurization method is to evenly increase the pressure of 50 MPa within 12 min, and the pressure relief method is to evenly release the pressure within 12 min. After sintering to obtain the block, the graphite paper on the surface of the sample was ground off with sandpaper, and then the sample of the required size for testing was cut using wire cutting. Before testing, the sample would be washed clean with water and alcohol, and dried in a vacuum atmosphere in a drying oven for more than 48 h.

2.2. Measurements

The X-ray diffraction (XRD) patterns were collected with the PANalytical X’Pert Pro (Malvern, UK) diffractometer for phase identification. The microstructures of the samples were observed using field emission scanning electron microscopy (SEM, FEI, Nova 400 Nano, FEI, Eindhoven, The Netherlands) and the chemical compositions were analyzed using energy-dispersive X-ray spectroscopy (EDS). In order to observe the morphology of naturally formed grains, the fracture morphology of naturally broken samples was observed. For the measurement of electrical transport properties, the sintered bulks were cut into bar-shaped specimens with the dimensions of 3 mm × 3 mm × 15 mm along the radial direction. The electrical resistivity (ρ) and Seebeck coefficient (S) were measured by commercial equipment (ZEM-3, ADVANCE-RIKO, Yokohama, Japan) with an uncertainty of ~5%. For the measurement of the band gap between the valence band and the conduction band, the infrared spectroscopy was obtained by a Fourier transform infrared/Raman spectrometer (VERTEX 70, Bruker, Mannheim, Germany).

3. Results and Discussion

It can be seen from Figure 1a that the diffraction peaks of the Mg3Bi2-based series of samples prepared by high-temperature reactions correspond to the standard cards one by one, without any obviously differing peaks. The results indicate that smelting, combined with spark plasma sintering, is an effective method to prepare Mg-Bi-Sb-based thermoelectric materials. This method can promote the industrialization of Mg-Bi-Sb-based thermoelectric materials. Figure 1b shows the SEM pictures of the fracture morphology of the Mg3.4Bi2 sample. It can be seen that the grains of the sample exhibit a lamellar structure, similar to Bi2Te3. The EDS results show that the elements are evenly distributed without obvious component segregation.
The changed pattern of electrical conductivity (σ) varying with temperature of the Mg3+xBi2 (x = 0, 0.2, 0.4) samples is shown in Figure 2a. It can be seen that the electrical conductivity of all samples decreased with the increase in temperature. The electrical conductivity of Mg3Bi2 at room temperature was ~1720 S cm−1. With the increase in Mg content (x), the electrical conductivity decreased in the testing temperature range. In order to analyze the change in electrical properties, we tested the infrared spectrum of the samples and calculated the band gap by extrapolation as shown in Figure 3a [36,37]. With the increase in Mg content, the extrapolation curve exhibits a downward trend, and the corresponding band gap shows an upward trend. The wider the band gap, the more difficult it is for the carriers to jump over the forbidden band, which leads to a decrease in carrier concentration and electrical conductivity [38,39]. On the one hand, excess Mg enters vacancies, weakening the acceptor effect, which could reduce the number of hole carriers [40]. On the other hand, according to the band gap testing results, the widening of the band gap results in a decrease in the number of current carriers passing through the band gap, which also reduces the number of hole carriers. The corresponding modifications have been added to the article.
The Seebeck coefficient (S) of Mg3+xBi2 (x = 0, 0.2, 0.4) varying with temperature is shown in Figure 2b. It can be seen that the Seebeck coefficients of the samples are positive, exhibiting p-type conduction. This is related to the high vapor pressure of Mg, which results in easily-formed Mg vacancies during the process of high-temperature preparation [41]. Mg vacancies can enhance the acceptor effect and increase the hole concentration, so that most carriers in the material are holes [42,43]. With the increase in Mg content, the change in the Seebeck coefficient is not obvious. As shown in Equation (1):
S = ± k B e [ γ + 2 + ln 2 ( 2 π m * k B T ) 3 / 2 h 3 n ]
The Seebeck coefficient is mainly determined by the influence of the effective mass (m*) and carrier concentration (n). On the one hand, the Seebeck coefficient is related to the parameters regarding the carrier concentration (related to electrical conductivity), and on the other hand, it is also related to effective mass. Due to the increase in Mg, the electrical conductivity and the number of hole carriers decrease, possibly due to a decrease in effective mass, resulting in a small change in the Seebeck coefficient. The power factor (PF) of all Mg3+xBi2 (x = 0, 0.2, 0.4) samples increased with the increase in temperature, and the maximum value of ~0.73 mWm−1K−2 was obtained at 483 K.
Figure 2. The electrical conductivity (σ) (a) and the Seebeck coefficient (S) (b) varying with temperature for Mg3+xBi2 (x = 0, 0.2, 0.4) samples.
Figure 2. The electrical conductivity (σ) (a) and the Seebeck coefficient (S) (b) varying with temperature for Mg3+xBi2 (x = 0, 0.2, 0.4) samples.
Micromachines 14 01844 g002
Figure 3. The diagram of band gap calculation by extrapolation (a) and power factor (PF); (b) varying with temperature for Mg3+xBi2 (x = 0, 0.2, 0.4) samples.
Figure 3. The diagram of band gap calculation by extrapolation (a) and power factor (PF); (b) varying with temperature for Mg3+xBi2 (x = 0, 0.2, 0.4) samples.
Micromachines 14 01844 g003
For the ball milling method, when the excess content of Mg reached ~6.7% (0.2/3, corresponding to Mg3.2Bi2), the Seebeck coefficient adjusted to a negative value [1]. For the high temperature reaction method, due to the high temperature of the process, the Mg volatilized easily in the high vapor pressure, and the efficiency of introducing the electrons through Mg decreased, according to defect chemistry theory [44,45]. When the excess content of Mg reached ~13. 3% (0.4/3, corresponding to Mg3.4Bi2), the Seebeck coefficient was still positive. Therefore, we adopted 35% Sb and 1.5% Te to dope at Bi sites to introduce more electrons, followed by optimizing Mg content to adjust the half-metallicity/semi-conductivity and electrical properties of the material [1].
The electrical conductivity (σ) varying with temperature for the Mg3.45+xBi1.27Sb0.7Te0.03 (x = 0, 0.1, 0.2, 0.3) samples is shown in Figure 4. It can be seen that, with the increase in x from 0 to 0.2, the electrical conductivity increases in the testing temperature range. When x increases from 0.2 to 0.3, the corresponding electrical conductivity decreases, but x is still higher than zero. Through the log conversion of the relationship between the electrical conductivity data and the temperature (σ303K/σ)~(1/T), as shown in Figure 5a, it can be seen that with the increase inf Mg content, the slope of the curve decreases, the rake ratio decreases, the half-metallicity of the material increases and the semi-conductivity decreases [1].
Figure 4b shows the Seebeck coefficient (S) varying with temperature for the Mg3.45+xBi1.27Sb0.7Te0.03 (x = 0, 0.1, 0.2, 0.3) samples. The Seebeck coefficients of all samples are negative in the whole temperature range, exhibiting n-type conduction [46]. When x increases from 0 to 0.1, the absolute value of the Seebeck coefficient shows an upward trend, with the maximum value being ~−210 μVK−1 at 483 K. When x increases from 0.1 to 0.3, the absolute value of the Seebeck coefficient decreases. The power factor (PF) varying with temperature for the Mg3.45+xBi1.27Sb0.7Te0.03 (x = 0, 0.1, 0.2, 0.3) samples is shown in Figure 5b. The highest power factor at room temperature is ~1.27 mWm−1K−2, obtained from Mg3.65Bi1.27Sb0.7Te0.03. The highest power factor in the whole temperature range was ~1.98 mWm−1K−2 at 423 K, obtained from Mg3.55Bi1.27Sb0.7Te0.03. The actual composition of the highest Mg content of the sample (3.75) as detected by EDS was ~3.69. Compared to doping experiments with materials such as bismuth telluride, this deviation is significant due to the volatility of Mg. The detection results of the XRD equipment can indirectly reflect that there is no excess of the Mg element. Adding more Mg has two considerations. On the one hand, Mg is easily volatile and may be lost during high-temperature preparation. On the other hand, our goal is to make n-type materials, and we need to reduce the number of hole carriers. Adding more magnesium is beneficial for suppressing the acceptor effect, aimed at adjusting the material conductivity to n-type.
Figure 4. The electrical conductivity (σ) (a) and Seebeck coefficient (S) (b) varying with temperature for Mg3.45+xBi1.27Sb0.7Te0.03 (x = 0, 0.1, 0.2, 0.3) samples.
Figure 4. The electrical conductivity (σ) (a) and Seebeck coefficient (S) (b) varying with temperature for Mg3.45+xBi1.27Sb0.7Te0.03 (x = 0, 0.1, 0.2, 0.3) samples.
Micromachines 14 01844 g004
Figure 5. The temperature dependence of electrical conductivity (σ) (a) and power factor (PF) (b) varying with temperature for Mg3.45+xBi1.27Sb0.7Te0.03 (x = 0, 0.1, 0.2, 0.3) samples.
Figure 5. The temperature dependence of electrical conductivity (σ) (a) and power factor (PF) (b) varying with temperature for Mg3.45+xBi1.27Sb0.7Te0.03 (x = 0, 0.1, 0.2, 0.3) samples.
Micromachines 14 01844 g005
The electrical conductivity (σ) varying with temperature for the Mg3.55Bi(1.27+x)Sb(0.7−x)Te0.03 (x = 0, 0.2, 0.4, 0.6) samples is shown in Figure 6. The electrical conductivity of all samples decreased with the increase in temperature, indicating semiconductor-transporting behavior. With the increase in Sb content, the electrical conductivity increased in most temperature ranges. The relation of log (σ303K/σ)~(1/T), as shown in Figure 7a, means that with the increase in Sb content, the slope of the curve becomes steeper, corresponding to the semi-conductivity of the materials increasing, while the semi-metallicity weakens [1].
Figure 6b shows the Seebeck coefficient (S) varying with temperature for the Mg3.55Bi(1.27+x) Sb(0.7−x)Te0.03 (x = 0, 0.2, 0.4, 0.6) samples. When the Sb content is 0.1 (x = 0.6), the Seebeck coefficient of the sample is very small. When the Sb content is 0.3, 0.5 or 0.7 (x = 0.4, 0.2, 0), the absolute values of the Seebeck coefficient from ~303 K to ~350 K vary a little. In the relatively high temperature range, when the Sb content is 0.5 (x = 0.2), the absolute value of the Seebeck coefficient is higher, with the highest value of −156 μVK−1 being obtained at 483 K. In general, as the Sb content increases from 0.1 to 0.7 (from x = 0.6 to x = 0.1), the power factor shows an upward trend, with the highest power factor in the whole temperature range being ~0.82 mWm−1K−2 at 303 K, obtained from Mg3.55Bi1.27Sb0.7Te0.03.
Figure 6. The electrical conductivity (σ) (a) and Seebeck coefficient (S) (b) varying with temperature for Mg3.55Bi(1.27+x)Sb(0.7−x)Te0.03 (x = 0, 0.2, 0.4, 0.6) samples.
Figure 6. The electrical conductivity (σ) (a) and Seebeck coefficient (S) (b) varying with temperature for Mg3.55Bi(1.27+x)Sb(0.7−x)Te0.03 (x = 0, 0.2, 0.4, 0.6) samples.
Micromachines 14 01844 g006
Figure 7. The temperature dependence of electrical conductivity (σ) (a), and power factor (PF) (b) varying with temperature for Mg3.55Bi(1.27+x)Sb(0.7−x)Te0.03 (x = 0, 0.2, 0.4, 0.6) samples.
Figure 7. The temperature dependence of electrical conductivity (σ) (a), and power factor (PF) (b) varying with temperature for Mg3.55Bi(1.27+x)Sb(0.7−x)Te0.03 (x = 0, 0.2, 0.4, 0.6) samples.
Micromachines 14 01844 g007

4. Conclusions

Herein, Mg3Bi2-based thermoelectric materials were prepared by high temperature reaction. The effects of composition adjustment on their electrical transporting properties have been studied. The results indicate that the electrical conductivity decreased when the content of Mg increased, due to the filling of Mg vacancies and the expansion of the band gap, while the Seebeck coefficient does not change significantly. During the high temperature process, the efficiency of introducing electrons is lower than that of ball milling, due to the high vapor pressure of Mg. After doping Sb and Te at Bi sites, the material shows n-type conduction. On this basis, the effects of the content of Mg/Sb have been studied. It is found that with an increase in Mg content, the semi-conductivity of the material becomes weak, the semi-metallicity becomes strong, and the corresponding conductivity increases. With an increase in Sb content, the sample shows the opposite change. The highest power factor obtained from the Mg3.55Bi1.27Sb0.7Te0.03 sample is ~1.98 mWm−1K−2.

Author Contributions

All authors contributed to the study conception and design. Material preparations were performed by J.Y., Z.W., H.Z., X.L., X.S., A.Q., Y.P., W.H., H.W., L.M. and G.C. and data collection and analysis were performed by B.F., Y.L. and Y.T. The first draft of the manuscript was written by B.F. and all authors commented on previous versions of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The work is supported by the Natural Science Foundation of Hubei Province (2021CFB009), the guiding project of Hubei Province in 2022 (B2022313), Hundred Talents Plan Project of Xianning City in Hubei Province and the School Youth Fund of Wuhan Donghu University.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time as the data also forms part of an ongoing study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. DiSalvo, F.J. Thermoelectric cooling and power generation. Science 1999, 285, 703–706. [Google Scholar] [CrossRef] [PubMed]
  2. Achour, A.; Chen, K.; Reece, M.J.; Huang, Z. Tuning of catalytic activity by thermoelectric materials for carbon dioxide hydrogenation. Adv. Energ. Mater. 2018, 8, 1701430. [Google Scholar] [CrossRef]
  3. Shi, X.; Chen, L. Thermoelectric materials step up. Nat. Mater. 2016, 15, 691–692. [Google Scholar] [CrossRef] [PubMed]
  4. Snyder, G.J.; Ursell, T.S. Thermoelectric efficiency and compatibility. Phys. Rev. Lett. 2003, 91, 148301. [Google Scholar] [CrossRef] [PubMed]
  5. Kim, G.H.; Shao, L.; Zhang, K.; Pipe, K.P. Engineered doping of organic semiconductors for enhanced thermoelectric efficiency. Nat. Mater. 2013, 12, 719–723. [Google Scholar] [CrossRef]
  6. Snyder, G.J.; Christensen, M.; Nishibori, E.; Caillat, T.; Iversen, B.B. Disordered zinc in Zn4Sb3 with phonon-glass and electron-crystal thermoelectric properties. Nat. Mater. 2004, 3, 458–463. [Google Scholar] [CrossRef]
  7. Takahata, K.; Iguchi, Y.; Tanaka, D.; Itoh, T.; Terasaki, I. Low thermal conductivity of the layered oxide (Na, Ca)Co2O4: Another example of a phonon glass and an electron crystal. Phys. Rev. B 2000, 61, 12551. [Google Scholar] [CrossRef]
  8. Wang, Z.L.; Yokoyama, Y.; Onda, T.; Adachi, Y.; Chen, Z.C. Improved Thermoelectric Properties of Hot-Extruded Bi–Te–Se Bulk Materials with Cu Doping and Property Predictions via Machine Learning. Adv. Electron. Mater. 2019, 5, 1900079. [Google Scholar] [CrossRef]
  9. Kumari, M.; Sharma, Y.C. A Review on Recent Enhancement in Thermoelectric Properties in Telluride Compounds. J. Mater. Sci. 2019, 7, 12. [Google Scholar] [CrossRef]
  10. Chen, Z.; Jian, Z.; Li, W.; Chang, Y.; Ge, B.; Hanus, R.; Yang, J.; Chen, Y.; Huang, M.; Snyder, G.J.; et al. Lattice dislocations enhancing thermoelectric PbTe in addition to band convergence. Adv. Mater. 2017, 29, 1606768. [Google Scholar] [CrossRef]
  11. Pei, Y.; Lensch-Falk, J.; Toberer, E.S.; Medlin, D.L.; Snyder, G.J. High thermoelectric performance in PbTe due to large nanoscale Ag2Te precipitates and La doping. Adv. Funct. Mater. 2011, 21, 241–249. [Google Scholar] [CrossRef]
  12. Minnich, A.J.; Lee, H.; Wang, X.W.; Joshi, G.; Dresselhaus, M.S.; Ren, Z.F.; Chen, G.; Vashaee, D. Modeling study of thermoelectric SiGe nanocomposites. Phys. Rev. B 2009, 80, 155327. [Google Scholar] [CrossRef]
  13. Bathula, S.; Jayasimhadri, M.; Singh, N.; Srivastava, A.K.; Pulikkotil, J.; Dhar, A.; Budhani, R.C. Enhanced thermoelectric figure-of-merit in spark plasma sintered nanostructured n-type SiGe alloys. Appl. Phys. Lett. 2012, 101, 213902. [Google Scholar] [CrossRef]
  14. Goldsmid, H.J.; Douglas, R.W. The use of semiconductors in thermoelectric refrigeration. Brit. J. Appl. Phys. 1954, 5, 386. [Google Scholar] [CrossRef]
  15. Astrain, D.; Vian, J.G.; Domınguez, M. Increase of COP in the thermoelectric refrigeration by the optimization of heat dissipation. Appl. Therm. Eng. 2003, 23, 2183–2200. [Google Scholar] [CrossRef]
  16. Rosi, F.D. Thermoelectricity and thermoelectric power generation. Solid-State Electron. 1968, 11, 833–868. [Google Scholar] [CrossRef]
  17. Liu, W.; Jie, Q.; Kim, H.S.; Ren, Z. Current progress and future challenges in thermoelectric power generation: From materials to devices. Acta Mater. 2015, 87, 357–376. [Google Scholar] [CrossRef]
  18. Min, G.; Rowe, D.M. Cooling performance of integrated thermoelectric microcooler. Solid-State Electron. 1999, 43, 923–929. [Google Scholar] [CrossRef]
  19. Wang, W.; Jia, F.; Huang, Q.; Zhang, J. A new type of low power thermoelectric micro-generator fabricated by nanowire array thermoelectric material. Microelectron. Eng. 2005, 77, 223–229. [Google Scholar] [CrossRef]
  20. Beaumale, M.; Barbier, T.; Bréard, Y.; Guelou, G.; Powell, A.V.; Vaqueiro, P.; Guilmeau, E. Electron doping and phonon scattering in Ti1 + xS2 thermoelectric compounds. Acta Mater. 2014, 78, 86–92. [Google Scholar] [CrossRef]
  21. Zevalkink, A.; Zeier, W.G.; Pomrehn, G.; Schechtel, E.; Tremel, W.; Snyder, G.J. Thermoelectric properties of Sr3 GaSb3–a chain-forming Zintl compound. Energ. Environ. Sci. 2012, 5, 9121–9128. [Google Scholar] [CrossRef]
  22. Yazawa, K.; Shakouri, A. Cost-efficiency trade-off and the design of thermoelectric power generators. Environ. Sci. Technol. 2011, 45, 7548–7553. [Google Scholar] [CrossRef] [PubMed]
  23. Mao, J.; Zhu, H.; Ding, Z.; Liu, Z.; Gamage, G.A.; Chen, G.; Ren, Z. High thermoelectric cooling performance of n-type Mg3Bi2-based materials. Science 2019, 365, 495–498. [Google Scholar] [CrossRef]
  24. Zhang, J.; Song, L.; Mamakhel, A.; Jørgensen MR, V.; Iversen, B.B. High-performance low-cost n-type Se-doped Mg3Sb2-based Zintl compounds for thermoelectric application. Chem. Mater. 2017, 29, 5371–5383. [Google Scholar] [CrossRef]
  25. Zhang, J.; Song, L.; Pedersen, S.H.; Yin, H.; Hung, L.T.; Iversen, B.B. Discovery of high-performance low-cost n-type Mg3Sb2-based thermoelectric materials with multi-valley conduction bands. Nat. Commun. 2017, 8, 13901. [Google Scholar] [CrossRef]
  26. Li, A.; Fu, C.; Zhao, X.; Zhu, T. High-performance Mg3Sb2-xBix thermoelectrics: Progress and perspective. Research 2020, 2020, 1934848. [Google Scholar] [CrossRef] [PubMed]
  27. Liu, K.J.; Zhang, Z.W.; Chen, C.; Wei, L.H.; He, H.L.; Mao, J.; Zhang, Q. Entropy engineering in CaZn2Sb2–YbMg2Sb2 Zintl alloys for enhanced thermoelectric performance. Rare Met. 2022, 41, 2998–3004. [Google Scholar] [CrossRef]
  28. Ullah, M.; Murtaza, G.; Ramay, S.M.; Mahmood, A. Structural, electronic, optical and thermoelectric properties of Mg3X2 (X = N, P, As, Sb, Bi) compounds. Mater. Res. Bull. 2017, 91, 22–30. [Google Scholar] [CrossRef]
  29. Sedighi, M.; Nia, B.A.; Zarringhalam, H.; Moradian, R. Density functional theory study of the structural and electronic properties of Mg3Bi2 in hexagonal and cubic phases. Eur. Phys. J. Appl. Phys. 2013, 61, 10103. [Google Scholar] [CrossRef]
  30. Wood, M.; Kuo, J.J.; Imasato, K.; Snyder, G.J. Improvement of Low-Temperature zT in a Mg3Sb2–Mg3Bi2 Solid Solution via Mg-Vapor Annealing. Adv. Mater. 2019, 31, 1902337. [Google Scholar] [CrossRef]
  31. Kim, I.K. Fabrication of Mg3Sb2 and Mg3Bi2 Compounds and their composites by mechanical alloying. J. Korean Crys. Growth Crys. Technol. 2013, 23, 189–194. [Google Scholar] [CrossRef]
  32. Ajaal, T.; Smith, R.W.; Yen, W.T. The development and characterization of a ball mill for mechanical alloying. Can. Metal. Quart. 2002, 41, 7–14. [Google Scholar] [CrossRef]
  33. Yadav, G.G.; Susoreny, J.A.; Zhang, G.; Yang, H.; Wu, Y. Nanostructure-based thermoelectric conversion: An insight into the feasibility and sustainability for large-scale deployment. Nanoscale 2011, 3, 3555–3562. [Google Scholar] [CrossRef] [PubMed]
  34. Cai, X.; Rong, Z.; Yang, F.; Gan, Z.; Li, G. Improved thermoelectric properties of Bi2Te3 − xSex alloys by melt spinning and resistance pressing sintering. J. Phys. D Appl. Phys. 2014, 47, 115101. [Google Scholar] [CrossRef]
  35. Cai, X.; Rong, Z.; Yang, F.; Li, G.; Gan, Z. Resistance pressing sintering: A simple, economical and practical technique and its application to p-type (Bi, Sb)2Te3 thermoelectric materials. J. Alloys Compd. 2014, 607, 91–98. [Google Scholar]
  36. Liu, Y.; Ding, J.; Xu, B.; Lan, J.; Zheng, Y.; Zhan, B.; Zhang, B.; Lin, Y.; Nan, C. Enhanced thermoelectric performance of La-doped BiCuSeO by tuning band structure. Appl. Phys. Lett. 2015, 106, 233903. [Google Scholar] [CrossRef]
  37. Zou, P.; Xu, G.; Wang, S.; Chen, P.; Huang, F. Effect of high pressure sintering and annealing on microstructure and thermoelectric properties of nanocrystalline Bi2Te2.7Se0.3 doped with Gd. Prog. Nat. Sci. Mater. Int. 2014, 24, 210–217. [Google Scholar] [CrossRef]
  38. Altermatt, P.P.; Schenk, A.; Geelhaar, F.; Heiser, G. Reassessment of the intrinsic carrier density in crystalline silicon in view of band-gap narrowing. J. Appl. Phys. 2003, 93, 1598–1604. [Google Scholar] [CrossRef]
  39. Lu, J.G.; Fujita, S.; Kawaharamura, T.; Nishinaka, H.; Kamada, Y.; Ohshima, T.; Ye, Z.Z.; Zeng, Y.J.; Zhang, Y.Z.; Zhu, L.P.; et al. Carrier concentration dependence of band gap shift in n-type ZnO: Al films. J. Appl. Phys. 2007, 101, 083705. [Google Scholar] [CrossRef]
  40. Jeong, M.; Tak, J.Y.; Lee, S.; Seo, W.S.; Cho, H.K.; Lim, Y.S. Effects of Cu incorporation as an acceptor on the thermoelectric transport properties of CuxBi2Te2.7Se0.3compounds. J. Alloys Compd. 2017, 696, 213–219. [Google Scholar] [CrossRef]
  41. Song, S.W.; Mao, J.; Bordelon, M.; He, R.; Wang, Y.M.; Shuai, J.; Sun, J.; Lei, X.; Ren, Z.; Chen, S.; et al. Joint effect of magnesium and yttrium on enhancing thermoelectric properties of n-type Zintl Mg3+δY0.02Sb1.5Bi0.5. Mater. Today Phys. 2019, 8, 25–33. [Google Scholar] [CrossRef]
  42. Yuen, J.D.; Fan, J.; Seifter, J.; Lim, B.; Hufschmid, R.; Heeger, A.J.; Wudl, F. High performance weak donor–acceptor polymers in thin film transistors: Effect of the acceptor on electronic properties, ambipolar conductivity, mobility, and thermal stability. J. Am. Chem. Soc. 2011, 133, 20799–20807. [Google Scholar] [CrossRef] [PubMed]
  43. Wang, L.; Pan, C.; Liang, A.; Zhou, X.; Zhou, W.; Wan, T.; Wang, L. The effect of the backbone structure on the thermoelectric properties of donor–acceptor conjugated polymers. Polym. Chem. 2017, 8, 4644–4650. [Google Scholar] [CrossRef]
  44. Tamaki, H.; Sato, H.K.; Kanno, T. Isotropic Conduction Network and Defect Chemistry in Mg3+ δSb2-Based Layered Zintl Compounds with High Thermoelectric Performance. Adv. Mater. 2016, 28, 10182–10187. [Google Scholar] [CrossRef] [PubMed]
  45. Jiang, G.; He, J.; Zhu, T.; Fu, C.; Liu, X.; Hu, L.; Zhao, X. High performance Mg2 (Si, Sn) solid solutions: A point defect chemistry approach to enhancing thermoelectric properties. Adv. Funct. Mater. 2014, 24, 3776–3781. [Google Scholar] [CrossRef]
  46. Chen, L.D.; Kawahara, T.; Tang, X.F.; Goto, T.; Hirai, T.; Dyck, J.S.; Chen, W.; Uher, C. Anomalous barium filling fraction and n-type thermoelectric performance of BayCo4Sb12. J. Appl. Phys. 2001, 90, 1864–1868. [Google Scholar] [CrossRef]
Figure 1. (a) Room temperature powder XRD patterns for Mg3Bi2-based series of samples; (b) SEM image of Mg3.4Bi2; (c) the enlarged view of the frame selected area in (b) and corresponding EDS mappings of Bi and Mg.
Figure 1. (a) Room temperature powder XRD patterns for Mg3Bi2-based series of samples; (b) SEM image of Mg3.4Bi2; (c) the enlarged view of the frame selected area in (b) and corresponding EDS mappings of Bi and Mg.
Micromachines 14 01844 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, J.; Wang, Z.; Zhao, H.; Luo, X.; Han, W.; Wang, H.; Meng, L.; She, X.; Quan, A.; Peng, Y.; et al. Effect of Composition Adjustment on the Thermoelectric Properties of Mg3Bi2-Based Thermoelectric Materials. Micromachines 2023, 14, 1844. https://doi.org/10.3390/mi14101844

AMA Style

Yang J, Wang Z, Zhao H, Luo X, Han W, Wang H, Meng L, She X, Quan A, Peng Y, et al. Effect of Composition Adjustment on the Thermoelectric Properties of Mg3Bi2-Based Thermoelectric Materials. Micromachines. 2023; 14(10):1844. https://doi.org/10.3390/mi14101844

Chicago/Turabian Style

Yang, Jianbao, Zhibin Wang, Hong Zhao, Xinyu Luo, Wenyuan Han, Hao Wang, Linghao Meng, Xinqi She, Anlong Quan, Yixin Peng, and et al. 2023. "Effect of Composition Adjustment on the Thermoelectric Properties of Mg3Bi2-Based Thermoelectric Materials" Micromachines 14, no. 10: 1844. https://doi.org/10.3390/mi14101844

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop