Next Article in Journal
Incorporating Birch Bark Suberinic Acid Residue Powder into Structural Particleboards: Exploring Fractional Influence on Material Properties in Circular Economy Framework
Previous Article in Journal
Microstructure Evolution and Mechanical Properties of 16-Layer 2195 Al-Li Alloy Components Manufactured by Additive Friction Stir Deposition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modulation of BixSb2−xTe3 Alloy Application Temperature via Optimizing Material Composition

1
School of Materials Science and Engineering, Jiujiang University, Jiujiang 332005, China
2
Jiangxi Key Laboratory of Material Surface Engineering, Jiujiang University, Jiujiang 332005, China
3
School of Power and Mechanical Engineering, Wuhan University, Wuhan 430072, China
4
Nanjing Institute of Geography and Limnology, Chinese Academy of Sciences, Nanjing 210008, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(23), 5751; https://doi.org/10.3390/ma17235751
Submission received: 28 October 2024 / Revised: 11 November 2024 / Accepted: 15 November 2024 / Published: 24 November 2024

Abstract

:
Bi2Te3-based alloys are representatively commercialized thermoelectric materials for refrigeration and power generation. Refrigeration mainly utilizes thermoelectric properties near room temperature, while the power generation temperature is relatively high. However, it is difficult for bismuth telluride to maintain good thermoelectric properties throughout the entire temperature range of 300–500 K. Herein, a series of BixSb2−xTe3 alloys with different Bi contents were prepared by a simple preparation method and systematically investigated, and their best application temperature range was found. The Bi content can modulate carrier concentration and band gap, and the maximum dimensionless figure of merit (ZT) value of BixSb2−xTe3 can be achieved in the corresponding application temperature range. The maximum ZT of Bi0.3Sb1.7Te3 with a Bi content equal to 0.3 reaches 1.14 at 400 K, and the average ZT is 1.06 in the range of 300–500 K, which is suitable for both power generation and refrigeration. Therefore, power generation technologies with higher application temperatures should be selected from BixSb2−xTe3 materials with Bi content less than 0.3, and refrigeration technologies with lower application temperatures should be selected with Bi content greater than 0.3. This work provides experimental guidance for finding the composition of Bi2Te3-based alloys in scientific research and practical applications.

1. Introduction

Thermoelectric (TE) materials can directly mutually convert heat–electricity energy without any harmful substance emissions [1,2], which provides an eco-friendly pathway and promising applications for all solid-state refrigeration and power generation techniques based on the Peltier effect and Seebeck effect. The TE device conversion efficiency is closely related to the dimensionless figure of merit, ZT = α2σT/κ, where α2σ, T, and κ are the power factor (including Seebeck coefficient α and electrical conductivity σ), absolute temperature, and thermal conductivity, respectively [3,4,5]. According to the above formula, the performance of materials is proportional to the power factor and inversely proportional to the thermal conductivity. Therefore, the ideal TE materials require a relatively larger α2σ and lower κ [6,7]. Due to the intrinsically intricate coupling connection between the aforementioned parameters, the improvement of TE performance is a critical challenge [8,9]. In addition, a simple and high-efficiency synthesis in terms of economy is also crucial for practical commercial applications.
Bi2Te3 and related alloys are the most extensively studied TE materials and the only ones with commercial applications below 500 K. Among the numerous Bi2Te3 alloys, the zone-melting Bi2Te3 with a maximum ZT value of around 1.0 in approximately room temperature conditions is widely applied in TE refrigeration and waste heat recovery [10,11,12]. However, zone-melting Bi2Te3 has two disadvantages. For one thing, the weak van der Waals bonding between adjacent atom layers leads to inferior mechanical performance, resulting in material splitting during the cutting process and material waste. For another, the zone-melted Bi2Te3 maximum ZT is about 1.0 and occurs near room temperature, which is relatively low and limits their further commercial applications. Therefore, extensive endeavors have been devoted to heightening the TE properties and broadening the application temperature range of polycrystalline Bi2Te3. In the past several decades, carrier engineering [13,14,15] and nanostructure engineering [16,17,18] have been applied to control carrier and phonon transport properties, thus improving the TE performance of polycrystalline Bi2Te3. For p-type Bi2Te3 alloys, in particular, dense bulks with a maximum ZT value above 1.2 have been prepared by hot deformation [19,20,21], three-dimensional printing [22,23], melt spinning [24,25], cyclic SPS [26], liquid-phase sintering [27,28,29], or solvothermal synthesis [30,31]. Although the TE performance of polycrystalline p-type Bi2Te3 obtained by the above methods is better than that of zone-melted materials, it requires a long preparation time, complex operation, complex preparation equipment, or difficult synthesis control, which limits its commercial application.
Ball milling is a simple, well-known, and established preparation technique for obtaining Bi2Te3 powders and is recognized as an indispensable technology for synthesizing TE materials. The obtained p-type polycrystalline Bi2Te3 bulk performance is better than the TE performance achieved using the zone-melted method; the maximal ZT value reaches 1.4 at 373 K due to the enhanced phonon scattering caused by nanocrystallinity and a large reduction in the lattice thermal conductivity [32]. Subsequently, many studies have focused on ball milling to improve the TE properties of p-type Bi2Te3 alloys [33,34,35]. Although those materials possess high TE performance, there is still a key issue that needs further exploration: the bulk materials’ optimal composition for different application temperatures in the range of 300–500 K. The materials’ composition affects the formation of the antisite defect, which is closely related to the electrical transport properties of the material. Optimizing electrical transport properties is an effective measure to modulate the application temperature range of materials. This investigation focuses on p-type Bi2Te3-based alloys, so Bi, Sb, and Te were selected for the composition.
In this work, a series of p-type BixSb2−xTe3 bulk materials were prepared by a simple preparation method (ball milling and spark plasma sintering). The phase composition and TE properties of these bulk materials have been systematically investigated. The Bi content has a great influence on the electrical and thermal transport properties. As a result, the modulation of the Bi content can effectively adjust the optimal application temperature range of p-type BixSb2−xTe3. Although TE materials are slightly lower than the above methods, the p-type BixSb2−xTe3 with an adjustable application temperature range is more beneficial to promoting large-scale commercial applications.

2. Experimental Section

High-purity Bi (99.99%), Sb (99.99%), and Te (99.99%) powders were mixed according to the stoichiometric composition of p-type BixSb2−xTe3 (x = 0.2, 0.3, 0.4, 0.5, 0.6). The mixed powders were loaded into stainless-steel jars with stainless-steel balls. The jars were vacuum-sealed and filled with an argon gas mixture three times, and then set on a planetary ball-milling machine for 360 min at 450 rpm in an argon-protective atmosphere. The synthesized powders were loaded into a graphite die for spark plasma sintering (SPS) at 673 K with an axial pressure of 50 MPa for 5 min.
The phase purity and composition of synthesized p-type BixSb2−xTe3 materials were analyzed by X-ray diffraction (XRD, Cu , Bruker D8, Selb, Germany). The surface morphology and elemental distribution of the p-type BixSb2−xTe3 bulk were investigated by an electron probe micro analyzer (EPMA) (JEOL, JXA-8230, Kyoto, Japan). The rectangular specimens (about 2.5 × 3.5 × 11.5 mm) were cut from the sintered bulk materials and were used to measure electrical conductivity σ and Seebeck coefficient α by the four-probe method (Cryoall CTA-3/500). The thermal conductivity κ was calculated using the formula κ = DρCp. The disk specimens (about ϕ 6 × 1.5 mm) were cut for thermal diffusivity D measurement, which were measured by a Netzsch laserflash instrument (LFA-467, Selb, Germany). The sample density ρ was obtained by the Archimedes method. The specific heat capacity Cp is estimated by Dulong–Petit law. The carrier thermal conductivity was estimated by the Wiedemann–Franz law of κE = σLT, where L is the Lorentz number. The lattice thermal conductivity κL was calculated according to the equation κL = κκE. The Hall coefficient RH from 300 K to 500 K was collected on a self-made test system with the magnetic field varied in the range of ±2.0 T. The carrier concentration nH and Hall mobility μH were calculated according to the equations nH = 1/(eRH) and μH = σ/(enH), respectively.

3. Results and Discussion

3.1. Composition and Microstructures Analysis

The purity and lattice parameters of the sintered BixSb2−xTe3 (x = 0.2, 0.3, 0.4, 0.5, 0.6) polycrystalline materials were analyzed by XRD. As shown in Figure 1a, the major diffraction peaks of BixSb2−xTe3 materials can be well indexed to the standard pattern of JCPDS 49–1713, indicating that the main phase of the sintered materials is the Bi0.5Sb1.5Te3 phase with the rhombohedral structure. In order to analyze the influence of Bi content on phase structure, the XRD patterns of the sintered BixSb2−xTe3 materials were expanded in the range of 45–60° (Figure 1b). Obviously, the diffraction peaks near 51° and 58° move to a lower angle with an increase in Bi content. According to the Bragg equation 2dsinθ =λ (where d, θ, and λ are interplanar distance, Bragg angle, and XRD wavelength, respectively), when the λ remains unchanged, θ decreases and d increases, indicating that the lattice expands with increasing Bi contents. The lattice parameters of BixSb2−xTe3 were calculated according to XRD results (Figure 1c). With an increase in Bi replacing Sb content, the size along the a-axis and c-axis of the BixSb2−xTe3 unit cell increases gradually, resulting in an expansion of the lattice parameters along both directions. The lattice expansion is attributed to the larger atomic radius of Bi (RBi = 1.55 Å) in comparison with that of Sb (RBi = 1.45 Å). The above results indicate that all BixSb2−xTe3 alloys with different Bi contents are single-phase materials, and the Bi element successfully occupies the Sb site in the crystal structure, which is beneficial to explain the influence of Bi content on electrical and thermal transport properties of BixSb2−xTe3.
As shown in Figure 2a, the backscattered electron image of the Bi0.4Sb1.6Te3 material contains many micropores (black contrast area), which will be further analyzed in the following macrostructure analysis. Immediately afterwards, an electron probe micro analyzer (EPMA) equipped with wave-dispersive spectrometry (WDS) was utilized to investigate the elements’ distributions of Bi0.4Sb1.6Te3 materials. Except for some distinct shape and size areas, the Bi, Sb, and Te elements’ distributions are uniformly distributed in Bi0.4Sb1.6Te3 materials (Figure 2b–d). By comparing the distribution of the three elements, it was found that Bi and Sb are absent in these positions, while Te is enriched in the same position, revealing that the distinct shape and size areas are a Te impurity phase. However, no characteristic diffraction peaks of Te were detected in XRD results due to the low content, which was far less than the detection limit of XRD technology (about 1%). The analysis indicates that sintered materials are composed of the BixSb2−xTe3 main phase and the Te impurity phase.
The backscattered electron images of all the sintered BixSb2−xTe3 bulk polished surfaces were investigated by EPMA. It can be seen that a lot of black contrast is distributed over the gray contrast, where the black contrast and the gray contrast areas are micropores and BixSb2−xTe3 (Figure 3a–e), respectively. There are two main causes for forming micropores in bulk materials. One is that the adsorbed oxygen was not completely discharged during the sintering process, leading to the formation of micropores inside the bulk materials. Another is that liquefied Te was partially extruded, and then left many internal voids. A similar phenomenon was also reported in some studies in the literature [33,34,36]. Figure 3f shows the density and relative density of all the sintered BixSb2−xTe3. The density gradually increases with an increase in the Bi/Sb ratio due to the larger relative atomic mass of Bi than Sb. However, the relative density fluctuates between 97.2% and 97.9%, maintaining a small range of change, indicating that the formation of micropores during the sintering process is not significantly affected by the Bi/Sb ratio. All the BixSb2−xTe3 bulk materials contain about 2.5% micropores, which can enhance the phonon scattering and is beneficial in regulating phonon transport.

3.2. Thermoelectric Properties

The temperature-dependent electrical transport performances are presented for BixSb2−xTe3 (x = 0.2, 0.3, 0.4, 0.5, 0.6) materials in Figure 4. As shown in Figure 4a, the electrical conductivity σ significantly decreases with increasing Bi content at room temperature and slightly reduces at 500 K. The σ of BixSb2−xTe3 in the range of 0.2 ≤ x ≤ 0.5 monotonously decreases as the temperature increases from 300 K to 500 K, consistent with typical metallic behavior. The σ of BixSb2−xTe3 with x = 0.6 shows an opposite change in regularity, implying a typical semiconductor behavior. Therefore, the carrier concentration nH and Hall mobility μH were measured to uncover the opposite and inconsistent behavior. With increasing Bi content from 0.2 to 0.6, the room temperature nH decreases gradually from 4.81 × 1019 cm−3 to 0.35 × 1019 cm−3, which results from the change in defects in the Bi2Te3 alloy (Figure 4b). In fact, the antisite defects ( Sb Te and Bi Te ) and vacancy ( V Te ) are the dominant defects in p-type polycrystalline BixSb2−xTe3. The formation energy of antisite defect (Equation (1)) and vacancy (Equation (2)) has been reported in a previous publication [37].
E AS ( Sb Te ) < E AS ( Bi Te ) < E AS ( Sb Se ) < E AS ( Bi Se )
E V ( Sb Te ) > E V ( Bi Te ) > E V ( Sb Se ) > E V ( Bi Se )
Obviously, the formation energy of the Bi Te antisite defect is larger than that of the Sb Te antisite defect; thus, the concentration of antisite defects will decrease with increasing Bi content. The formation energy of vacancy reduces with the substitution of Sb by Bi, which boosts the vacancy formation. The antisite defect generates holes, and vacancy generates electrons. It is well known that the majority carrier of p-type Bi2Te3 alloy is holes, so nH is reduced with increasing Bi content at room temperature.
As shown in Figure 4b, the variation trend of nH over the entire temperature range is inconsistent. To uncover the variation trend, the band gaps (Eg) were estimated according to Equation (3):
Eg = 2maxTmax
where αmax and Tmax are the maximum value of the Seebeck coefficient and the corresponding temperature, respectively. The estimated calculation results are listed in Table 1. The estimated Eg values are decreased from 0.20 eV to 0.17 eV for the BixSb2−xTe3 sample (x = 0.3, 0.4, and 0.5). For the Bi0.2Sb1.8Te3 sample, αmax is achieved at 500 K and still increases with the testing temperature (Figure 4d). For the Bi0.6Sb1.4Te3 sample, αmax is achieved at 300 K, and an even larger value may be obtained if the testing temperature is below 300 K. The Eg values of Bi0.2Sb1.8Te3 and Bi0.6Sb1.4Te3 can be rationally deduced. The Eg values of Bi0.2Sb1.8Te3 will be larger than 0.20 eV and the Eg values of Bi0.6Sb1.4Te3 will be lower than 0.17 eV. Eg is the minimum energy for the carrier to transition from the valence band to the conduction band. If Eg is very low, carriers are more likely to be excited as the temperature increases, resulting in an intrinsic excitation. The diminished Eg values explicitly give a reasonable explanation for nH increasing above room temperature. Therefore, the nH of the BixSb2−xTe3 sample (x = 0.4, 0.5, and 0.6) gradually rises with testing temperature increases.
The μH values of the BixSb2−xTe3 samples are almost constant or slightly decreased except for the Bi0.6Sb1.4Te3 sample (Figure 4c). However, μH changes significantly with improving testing temperature, indicating that the carrier scattering mechanism has changed. As shown in the inset in Figure 4c, the trend of μH versus temperature exhibits a nearly T−1.5 exponential relationship in all testing temperatures for Bi0.2Sb1.8Te3, indicating that the carrier scattering mechanism is dominated by acoustic phonon scattering. The exponential relationship gradually decreases with increasing Bi content and reaches nearly T−2.5 for Bi0.6Sb1.4Te3, which should stem from the intrinsic excitation effect due to the decreasing Eg. nH and μH analyses can reasonably explain the σ change regularity.
The positive Seebeck coefficient α over the entire temperature range indicates that the dominant charge carriers of all samples are holes (i.e., p-type conduction behavior) (Figure 4d), which is consistent with the Hall coefficient (Figure 4e). At room temperature, the α of BixSb2−xTe3 significantly increases with the increasing Bi content except for Bi0.6Sb1.4Te3. As is well known, the α of the degenerate semiconductor can be expressed as Equation (4) (Mott equation):
α = 8 π 2 k B 2 T m * 3 e h 2 π 3 n 2 3
where kB is the Boltzmann constant, m* is the carrier effective mass, e is the elementary charge, h is the Planck constant, and n is the hole concentration for the p-type Bi2Te3 alloy. The m* at room temperature was calculated using the same method in our previously reported study (Table 1) [38]. According to the Mott equation, α is proportional to m* and inversely proportional to nH. The decreased nH is beneficial to α and the decreased m* is disadvantageous to α. Therefore, compared with BixSb2−xTe3 (x = 0.2, 0.3, 0.4, and 0.5), the less increase in α of Bi0.6Sb1.4Te3 stems from a significant decrease in m*, which may be related to the change in band structure caused by Bi content.
The α of BixSb2−xTe3 (x = 0.2, 0.3, 0.4, 0.5, and 0.6) samples exhibit the dissimilarity trend with increasing temperature. The α of Bi0.2Sb1.8Te3 gradually rises with the increase in temperature and Bi0.6Sb1.4Te3 shows an inverse trend. For BixSb2−xTe3 (x = 0.3, 0.4, and 0.5), α initially increases, reaching its maximum value, then decreases in the entire temperature range. Due to the gradual decreases in Eg and nH, the temperature of maximum α shifts gradually to a lower temperature, mainly originating from the activation of the effect of the minority carrier. With increasing Bi content, the power factor α2σ firstly rises and then falls due to the combined effect of α and σ (Figure 4f). For the Bi0.3Sb1.7Te3 sample, the highest α2σ value reaches 4.2 mW·m−1·K−2 at 300 K and gradually decreases with increasing temperature.
The thermal transport performance and dimensionless figure of merit (ZT) values of BixSb2−xTe3 (x = 0.2, 0.3, 0.4, 0.5, and 0.6) specimens are depicted in Figure 5. The thermal conductivity κ firstly drops and then rises with increasing Bi content at room temperature (Figure 5a). The Bi0.5Sb1.5Te3 sample has the lowest κ and is about 0.85 W·m−1·K−1 at 300 K, a 46% reduction as compared with that of Bi0.2Sb1.8Te3. The κ change trend of BixSb2−xTe3 with the increasing temperature is related to the evolutions of Eg and nH, which will be discussed later. The above results indicate that the Bi content can effectively impact the thermal transport properties. The Lorenz number was calculated using the same method in our pioneer work (Table 1) [39], and the obtained κE is shown in Figure 5b. It can be seen that κE gradually decreases with increasing test temperature, and the change trend is consistent with the change in σ. The lattice thermal conductivity κL was calculated by subtracting κE from κ. κL initially decreases and then increases with increasing Bi content at room temperature (Figure 5c). There are three aspects affecting the room temperature κL. Firstly, the enhancing cation site disorder degree with increasing Bi content results in a stronger mass and strain field fluctuation, and a decrease in κL. Secondly, the increasing formation energy antisite defect leads to a decrease in the concentration of antisite defect with increasing Bi content, which would weaken the defect phonon scattering, and κL will rise. Thirdly, a reduction in nH with an increase in Bi content would weaken the carrier–phonon scattering, giving rise to an increase in κL. The competition among the three mechanisms leads to the change trend of κL with increasing Bi content at room temperature. As the temperature increases, the κL of BixSb2−xTe3 (x = 0.2, 0.3, 0.4, and 0.5) initially decreases due to the enhanced Umklapp phonon scattering and then increases owing to the intensified intrinsic excitation. Noteworthily, the temperature of intrinsic excitation gradually shifts to a lower temperature with increasing content of Bi, which is derived from the decreases in Eg and nH. The κL of Bi0.6Sb1.4Te3 gradually increases with an increase in temperature, indicating that the intrinsic excitation temperature is lower than the room temperature.
The temperature-dependent ZT values of BixSb2−xTe3 (x = 0.2, 0.3, 0.4, 0.5, and 0.6) specimens are shown in Figure 5d. The ZT value initially increases and then decreases with increasing Bi content at room temperature, which displays an opposite variation trend of κ. As is known, TE materials can be used for refrigeration and power generation, and the corresponding utilization temperatures are not similar. Refrigeration requires a large ZT value near room temperature and power generation needs a large ZT value at higher temperatures because the high device conversion efficiency requires a large ZT value. The maximum ZT value corresponds to the pivotal temperature shift to lower temperatures with increasing Bi content. The corresponding maximum ZT values for BixSb2−xTe3 (x = 0.2, 0.3, 0.4, 0.5, and 0.6) specimens are 0.94 at 430 K, 1.14 at 400 K, 1.16 at 360 K, 1.08 at 300 K, and 0.24 at 300 K, respectively. Therefore, the power generation materials should be sought from Bi content near 0.2 (Bi0.2Sb1.8Te3) and 0.3 (Bi0.3Sb1.7Te3). The refrigeration materials should be sought from Bi content near 0.4 (Bi0.4Sb1.6Te3) and 0.5 (Bi0.5Sb1.5Te3). The highest average ZT values of Bi0.2Sb1.8Te3 and Bi0.3Sb1.7Te3 for power generation reached 0.93 and 1.05 in the range of 400–500 K. The highest average ZT values of Bi0.4Sb1.6Te3 and Bi0.5Sb1.5Te3 for refrigeration reached 1.14 and 0.96 in the range of 300–400 K. The ZT of this work approaches or exceeds the commercial zone-melting Bi2Te3 with a ZT value around 1.0 in approximately room temperature conditions. In particular, the highest average ZT of Bi0.3Sb1.7Te3 reaches 1.06 in the range of 300–500 K, which can be used for both refrigeration and power generation over the entire temperature range. Consequently, researchers can select suitable Bi2Te3-based material components based on the above analysis and further optimize them to obtain high-performance TE materials based on the corresponding utilization temperature range of the materials.

4. Conclusions

A series of BixSb2−xTe3 (x = 0.2, 0.3, 0.4, 0.5, and 0.6) materials were prepared via ball milling and spark plasma sintering. The influence of Bi content on the phase composition and TE performance was systematically investigated between 300 and 500 K. The cell parameters of BixSb2−xTe3 increase gradually with an increase in Bi replacing Sb content. The changing Bi content can optimize nH, increase α, greatly reduce κ, and adjust Eg. By modulating nH and Eg, the maximum ZT value of BixSb2−xTe3 can be achieved in the corresponding application temperature range. Power generation technologies with higher application temperatures should be selected from BixSb2−xTe3 materials with a Bi content of less than 0.3. Refrigeration technologies with lower application temperatures should be selected with Bi contents greater than 0.3. In particular, the maximum ZT of Bi0.3Sb1.7Te3 with a Bi content equal to 0.3 reaches 1.14 at 400 K, and the average ZT is 1.06 in the range of 300–500 K, which is suitable for both power generation and refrigeration. Therefore, the Bi0.3Sb1.7Te3 material is the most effective alloy over the entire temperature range. This work provides experimental guidance for finding the composition of Bi2Te3-based alloys in scientific research and practical applications.

Author Contributions

Conceptualization, D.D., P.T. and J.G.; methodology, M.C., X.R., Q.L. and J.G.; software, Q.L. and J.L.; validation, M.L. and Q.H.; formal analysis, X.R., M.C. and M.L.; data curation, J.L., D.D., Q.H. and P.T.; writing—original draft, S.M. and J.Y.; writing—review and editing, S.M. and J.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research and the APC were funded by the National Natural Science Foundation of China, grant numbers 52361038 and 52201256; the Jiangxi Provincial Natural Science Foundation, grant numbers 20232BAB204023 and 20202BABL204004; the Science and Technology Planning Program of Jiangxi Provincial Education Department, grant numbers GJJ211833, GJJ190930, and GJJ2401817; the Innovation and Entrepreneurship training Program for College students of Jiujiang University, grant number X202211843002; the Natural Science Foundation of Jiujiang Science and Technology Bureau, grant number S2022KXJJ001; the China postdoctoral Science Foundation, grant number 2024M753321; and the Jiangxi Provincial Key Laboratory of Surface Engineering, grant number 2024SSY05072.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zheng, Y.; Zhang, Q.; Su, X.L.; Xie, H.Y.; Shu, S.C.; Chen, T.L.; Tan, G.J.; Yan, Y.G.; Tang, X.F.; Uher, C.; et al. Mechanically robust BiSbTe alloys with superior thermoelectric performance: A case study of stable hierarchical nanostructured thermoelectric materials. Adv. Energy Mater. 2014, 5, 1401391. [Google Scholar] [CrossRef]
  2. Li, L.; Wei, P.; Yang, M.J.; Zhu, W.T.; Nie, X.L.; Zhao, W.Y.; Zhang, Q.J. Strengthened interlayer interaction and improved room-temperature thermoelectric performance of Ag-doped n-type Bi2Te2.7Se0.3. Sci. China Mater. 2023, 66, 3651–3658. [Google Scholar] [CrossRef]
  3. Zhao, W.Y.; Liu, Z.Y.; Sun, Z.G.; Zhang, Q.J.; Wei, P.; Mu, X.; Zhou, H.Y.; Li, C.C.; Ma, S.F.; He, D.Q.; et al. Superparamagnetic enhancement of thermoelectric performance. Nature 2017, 549, 247–251. [Google Scholar] [CrossRef]
  4. Ma, S.F.; Li, C.C.; Wei, P.; Zhu, W.T.; Nie, X.L.; Sang, X.H.; Zhan, Q.J.; Zhao, W.Y. High-pressure synthesis and excellent thermoelectric performance of Ni/BiTeSe magnetic nanocomposites. J. Mater. Chem. A 2020, 8, 4816–4826. [Google Scholar] [CrossRef]
  5. Yi, L.B.; Xu, H.W.; Yang, H.B.; Huang, S.L.; Yang, H.; Li, Y.N.; Zhang, Q.; Guo, Z.; Hu, H.Y.; Sun, P.; et al. Design of Bi2Te3-based thermoelectric generator in a widely applicable system. J. Power Sources 2023, 559, 232661. [Google Scholar] [CrossRef]
  6. Zhao, L.D.; Dravid, V.P.; Kanatzidis, M.G. The panoscopic approach to high performance thermoelectrics. Energy Environ. Sci. 2014, 7, 251–268. [Google Scholar] [CrossRef]
  7. He, J.Q.; Girard, S.N.; Kanatzidis, M.G.; Dravid, V.P. Microstructure-Lattice Thermal Conductivity Correlation in Nanostructured PbTe0.7S0.3Thermoelectric Materials. Adv. Funct. Mater. 2010, 20, 764–772. [Google Scholar] [CrossRef]
  8. Liu, X.D.; Yu, J.C.; Wang, B.; Maji, K.; Alvarez-Ruiz, D.T.; Guilmeau, E.; Freer, R. Enhancing the thermoelectric properties of Nb-doped TiO2-based ceramics through in-situ synthesis of β-Sn inclusions at grain boundaries. J. Eur. Ceram. Soc. 2023, 43, 2523–2533. [Google Scholar] [CrossRef]
  9. Yu, Y.; He, D.S.; Zhang, S.; Cojocaru Mirédin, O.; Schwarz, T.; Stoffers, A.; Wang, X.Y.; Zheng, S.; Zhu, B.; Scheu, C.; et al. Simultaneous optimization of electrical and thermal transport properties of Bi0.5Sb1.5Te3 thermoelectric alloy by twin boundary engineering. Nano Energy 2017, 37, 203–213. [Google Scholar] [CrossRef]
  10. Zhang, Q.; Zhai, R.S.; Fang, T.; Xia, K.Y.; Wu, Y.H.; Liu, F.; Zhao, X.B.; Zhu, T.J. Low-cost p-type Bi2Te2.7Se0.3 zone-melted thermoelectric materials for solid-state refrigeration. J. Alloys Compd. 2020, 831, 154732. [Google Scholar] [CrossRef]
  11. Huang, H.W.; Li, J.; Chen, S.; Zhang, Z.K.; Yan, Y.G.; Su, X.L.; Tang, X.F. Anisotropic thermoelectric transport properties of Bi0.5Sb1.5Te2.96+x zone melted ingots. J. Solid State Chem. 2020, 288, 121433. [Google Scholar] [CrossRef]
  12. Jiang, J.; Chen, L.D.; Yao, Q.; Bai, S.Q.; Wang, Q. Effect of TeI4 content on the thermoelectric properties of n-type Bi–Te–Se crystals prepared by zone melting. Mater. Chem. Phys. 2005, 92, 39–42. [Google Scholar] [CrossRef]
  13. Zhu, B.; Wan, W.; Cui, J.; He, J.Q. Point defect engineering: Co-doping synergy realizing superior performance in n-type Bi2Te3 thermoelectric materials. Small 2021, 17, e2101328. [Google Scholar] [CrossRef]
  14. Hu, L.P.; Meng, F.C.; Zhou, Y.J.; Li, J.B.; Benton, A.; Li, J.Q.; Liu, F.S.; Zhang, C.H.; Xie, H.P.; He, J. Leveraging deep levels in narrow bandgap leveraging deep levels in narrow bandgap Bi0.5Sb1.5Te3 for record-high zTave near room temperature. Adv. Funct. Mater. 2020, 30, 2005202. [Google Scholar] [CrossRef]
  15. Liu, F.; Zhang, M.; Nan, P.F.; Zheng, X.; Li, Y.Z.; Wu, K.; Han, Z.K.; Ge, B.H.; Zhao, X.B.; Fu, C.G.; et al. Unraveling the origin of donor-like effect in bismuth–telluride-based thermoelectric materials. Small Sci. 2023, 2300082. [Google Scholar] [CrossRef]
  16. Zhu, Y.K.; Guo, J.; Zhang, Y.X.; Cai, J.F.; Chen, L.; Liang, H.; Gu, S.W.; Feng, J.; Ge, Z.H. Ultralow lattice thermal conductivity and enhanced power generation efficiency realized in Bi2Te2.7Se0.3/Bi2S3 nanocomposites. Acta Mater. 2021, 218, 117230. [Google Scholar] [CrossRef]
  17. Li, S.; Peng, C.; Wang, C.; Chen, Y.; Li, L.; Yang, G.; Cheng, Z.; Wang, J. In situ generation of flower-like and microspherical dendrites to improve thermoelectric properties of p-type Bi0.46Sb1.54Te3. Mater. Today Phys. 2022, 23, 100633. [Google Scholar] [CrossRef]
  18. Yang, G.S.; Sang, L.N.; Yun, F.F.; Mitchell, D.R.G.; Casillas, G.; Ye, N.; See, K.; Pei, J.; Wang, X.G.; Li, J.F.; et al. Significant enhancement of thermoelectric figure of merit in BiSbTe-based composites by incorporating carbon microfiber. Adv. Funct. Mater. 2021, 31, 2008851. [Google Scholar] [CrossRef]
  19. Zhu, T.J.; Xu, Z.J.; He, J.; Shen, J.J.; Zhu, S.; Hu, L.P.; Tritt, T.M.; Zhao, X.B. Hot deformation induced bulk nanostructuring of unidirectionally grown p-type (Bi,Sb)2Te3 thermoelectric materials. J. Mater. Chem. A 2013, 1, 11589. [Google Scholar] [CrossRef]
  20. Xu, Z.J.; Hu, L.P.; Ying, P.J.; Zhao, X.B.; Zhu, T.J. Enhanced thermoelectric and mechanical properties of zone melted p-type (Bi,Sb)2Te3 thermoelectric materials by hot deformation. Acta Mater. 2015, 84, 385–392. [Google Scholar] [CrossRef]
  21. Tan, C.; Tan, X.J.; Yu, B.; Liu, G.Q.; Wang, H.X.; Luo, G.Q.; Xu, J.T.; Wu, Q.S.; Liang, B.; Jiang, J. Synergistically optimized thermoelectric performance in Bi0.48Sb1.52Te3 by hot deformation and Cu doping. ACS Appl. Energ. Mater. 2019, 2, 6714–6719. [Google Scholar] [CrossRef]
  22. Yang, S.E.; Han, H.; Son, J.S. Recent progress in 3D printing of Bi2Te3-based thermoelectric materials and devices. J. Phys. Energy 2024, 6, 022003. [Google Scholar] [CrossRef]
  23. Hu, Q.J.; Luo, D.; Guo, J.B.; Qiu, W.B. 3D printing of Bi2Te3-based thermoelectric materials with high performance and shape controllability. ACS Appl. Mater. Interfaces 2023, 15, 38623–38632. [Google Scholar] [CrossRef] [PubMed]
  24. Xie, W.J.; Tang, X.F.; Yan, Y.G.; Zhang, Q.J.; Tritt, T.M. Unique nanostructures and enhanced thermoelectric performance of melt-spun BiSbTe alloys. Appl. Phys. Lett. 2009, 94, 102111. [Google Scholar] [CrossRef]
  25. Deng, R.G.; Su, X.L.; Hao, S.Q.; Zheng, Z.; Zhang, M.; Xie, H.Y.; Liu, W.; Yan, Y.G.; Wolverton, C.; Uher, C.; et al. High thermoelectric performance in Bi0.46Sb1.54Te3 nanostructured with ZnTe. Energy Environ. Sci. 2018, 11, 1520–1535. [Google Scholar] [CrossRef]
  26. Zhuang, H.L.; Pei, J.; Cai, B.; Dong, J.F.; Hu, H.H.; Sun, F.H.; Pan, Y.; Snyder, G.J.; Li, J.F. Thermoelectric performance enhancement in BiSbTe alloy by microstructure modulation via cyclic spark plasma sintering with liquid phase. Adv. Funct. Mater. 2021, 31, 2009681. [Google Scholar] [CrossRef]
  27. Kim, S.; Lee, K.H.; Mun, H.A.; Kim, H.S.; Hwang, S.W.; Roh, J.W.; Yang, D.J.; Shin, W.H.; Li, X.S.; Lee, Y.H.; et al. Dense dislocation arrays embedded in grain boundaries for high-performance bulk thermoelectrics. Science 2015, 348, 109–114. [Google Scholar] [CrossRef]
  28. Witting, I.T.; Grovogui, J.A.; Dravid, V.P.; Snyder, G.J. Thermoelectric transport enhancement of Te-rich bismuth antimony telluride (Bi0.5Sb1.5Te3+x) through controlled porosity. J. Mater. 2020, 6, 532–544. [Google Scholar] [CrossRef]
  29. Zhang, C.H.; Maria de la, M.; Li, Z.; Belarre, F.J.; Arbiol, J.; Khor, K.A.; Poletti, D.; Zhu, B.B.; Yan, Q.Y.; Xiong, Q.H. Enhanced thermoelectric performance of solution-derived bismuth telluride based nanocomposites via liquid-phase Sintering. Nano Energy 2016, 30, 630–638. [Google Scholar] [CrossRef]
  30. Wu, X.K.; Cao, W.; Gu, Y.X.; Wang, Z.Y.; Hou, Y.; Yu, T.; Chen, Z.Q.; Jin, D.; Liu, Y.; Yu, H.Y.; et al. Realizing high thermoelectric performance in Bi0.4Sb1.6Te3 nanosheets by doping Sn element. ACS Appl. Energ. Mater. 2022, 5, 12614–12621. [Google Scholar] [CrossRef]
  31. Liu, Y.; Zhang, Y.; Ortega, S.; Ibanez, M.; Lim, K.H.; Grau-Carbonell, A.; Marti-Sanchez, S.; Ng, K.M.; Arbiol, J.; Kovalenko, M.V.; et al. Crystallographically textured nanomaterials produced from the liquid phase sintering of BixSb2- xTe3 nanocrystal building blocks. Nano Lett. 2018, 18, 2557–2563. [Google Scholar] [CrossRef] [PubMed]
  32. Poudel, B.; Hao, Q.; Ma, Y.; Lan, Y.C.; Minnich, A.; Yu, B.; Yan, X.; Wang, D.Z.; Muto, A.; Vashaee, D.; et al. High-thermoelectric performance of nanostructured bismuth antimony telluride bulk alloys. Science 2008, 320, 634–638. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, D.; Zhu, B.R.; Feng, J.H.; Ling, Y.F.; Zhou, J.; Qiu, G.J.; Zhou, M.H.; Li, J.; Hou, X.F.; Ren, B.G.; et al. High thermoelectric performance of p-Type Bi0.4Sb1.6Te3+x synthesized by plasma-assisted ball milling. ACS Appl. Mater. Interfaces 2022, 14, 54044–54050. [Google Scholar] [CrossRef]
  34. Pan, Y.; Qiu, Y.; Witting, I.; Zhang, L.G.; Fu, C.G.; Li, J.W.; Huang, Y.; Sun, F.H.; He, J.Q.; Snyder, G.J.; et al. Synergistic modulation of mobility and thermal conductivity in (Bi,Sb)2Te3 towards high thermoelectric performance. Energy Environ. Sci. 2019, 12, 624–630. [Google Scholar] [CrossRef]
  35. Cai, B.; Zhuang, H.L.; Pei, J.; Su, B.; Li, J.W.; Hu, H.H.; Jiang, Y.L.; Li, J.F. Spark plasma sintered Bi-Sb-Te alloys derived from ingot scrap: Maximizing thermoelectric performance by tailoring their composition and optimizing sintering time. Nano Energy 2021, 85, 106040. [Google Scholar] [CrossRef]
  36. Liang, H.; Lou, Q.; Zhu, Y.K.; Guo, J.; Wang, Z.Y.; Gu, S.W.; Yu, W.; Feng, J.; He, J.Q.; Ge, Z.H. Highly enhanced thermoelectric and mechanical properties of Bi-Sb-Te compounds by carrier modulation and microstructure adjustment. ACS Appl. Mater. Interfaces 2021, 13, 45589–45599. [Google Scholar] [CrossRef]
  37. Hu, L.P.; Zhu, T.J.; Liu, X.H.; Zhao, X.B. Point defect engineering of high-performance Bismuth-Telluride-Based thermoelectric materials. Adv. Funct. Mater. 2014, 24, 5211–5218. [Google Scholar] [CrossRef]
  38. Ma, S.F.; Zeng, L.J.; Du, D.M.; Cao, M.; Lin, M.; Hua, Q.X.; Luo, Q.; Tang, P.; Guan, J.Z.; Yu, J. Grain boundary engineering simultaneously optimized thermoelectric and mechanical properties of BiSbTe alloys. J. Power Sources 2024, 618, 235191. [Google Scholar] [CrossRef]
  39. Ma, S.F.; Li, C.C.; Cui, W.J.; Sang, X.H.; Wei, P.; Zhu, W.T.; Nie, X.L.; Sun, F.H.; Zhao, W.Y.; Zhang, Q.J. Magneto-enhanced electro-thermal conversion performance. Sci. China Mater. 2021, 64, 2835–2845. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of sintered BixSb2−xTe3 materials. (b) The expanded view of XRD in the range of 45–60°. (c) Lattice parameters of sintered BixSb2−xTe3 materials.
Figure 1. (a) XRD patterns of sintered BixSb2−xTe3 materials. (b) The expanded view of XRD in the range of 45–60°. (c) Lattice parameters of sintered BixSb2−xTe3 materials.
Materials 17 05751 g001
Figure 2. (a) Backscattered electron image of Bi0.4Sb1.6Te3. (bd) WDS elemental mapping of Bi, Sb, and Te elements, respectively.
Figure 2. (a) Backscattered electron image of Bi0.4Sb1.6Te3. (bd) WDS elemental mapping of Bi, Sb, and Te elements, respectively.
Materials 17 05751 g002
Figure 3. (ae) Backscattered electron images of BixSb2−xTe3 materials. (f) Density of BixSb2−xTe3 materials.
Figure 3. (ae) Backscattered electron images of BixSb2−xTe3 materials. (f) Density of BixSb2−xTe3 materials.
Materials 17 05751 g003
Figure 4. Temperature dependencies of (a) electrical conductivity, (b) carrier concentration, (c) Hall mobility, (d) Seebeck coefficient, (e) Hall coefficient, and (f) a power factor for BixSb2−xTe3 materials. The inset in (c) is log(μH) versus log(T).
Figure 4. Temperature dependencies of (a) electrical conductivity, (b) carrier concentration, (c) Hall mobility, (d) Seebeck coefficient, (e) Hall coefficient, and (f) a power factor for BixSb2−xTe3 materials. The inset in (c) is log(μH) versus log(T).
Materials 17 05751 g004
Figure 5. Temperature dependence of (a) thermal conductivity, (b) carrier thermal conductivity, (c) lattice thermal conductivity, and (d) ZT for BixSb2−xTe3 materials.
Figure 5. Temperature dependence of (a) thermal conductivity, (b) carrier thermal conductivity, (c) lattice thermal conductivity, and (d) ZT for BixSb2−xTe3 materials.
Materials 17 05751 g005
Table 1. Hall mobility, carrier concentration, Seebeck coefficient, carrier effective mass, band gap, and Lorenz number of BixSb2−xTe3 at room temperature.
Table 1. Hall mobility, carrier concentration, Seebeck coefficient, carrier effective mass, band gap, and Lorenz number of BixSb2−xTe3 at room temperature.
SamplesμH
(cm2/V·s)
nH
(1019 cm−3)
α
(μV/K)
m*/m0Eg (eV)L
(10−8 W·Ω/K2)
Bi0.2Sb1.8Te3260.444.81141.791.05-1.75
Bi0.3Sb1.7Te3245.803.35178.001.150.201.66
Bi0.4Sb1.6Te3238.622.27213.641.210.181.60
Bi0.5Sb1.5Te3262.481.12260.961.120.171.55
Bi0.6Sb1.4Te3177.970.35274.380.58-1.54
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

Ma, S.; Li, J.; Du, D.; Ruan, X.; Cao, M.; Lin, M.; Hua, Q.; Luo, Q.; Tang, P.; Guan, J.; et al. Modulation of BixSb2−xTe3 Alloy Application Temperature via Optimizing Material Composition. Materials 2024, 17, 5751. https://doi.org/10.3390/ma17235751

AMA Style

Ma S, Li J, Du D, Ruan X, Cao M, Lin M, Hua Q, Luo Q, Tang P, Guan J, et al. Modulation of BixSb2−xTe3 Alloy Application Temperature via Optimizing Material Composition. Materials. 2024; 17(23):5751. https://doi.org/10.3390/ma17235751

Chicago/Turabian Style

Ma, Shifang, Jianan Li, Daming Du, Xuefeng Ruan, Ming Cao, Ming Lin, Qiongxin Hua, Qi Luo, Ping Tang, Jinzhao Guan, and et al. 2024. "Modulation of BixSb2−xTe3 Alloy Application Temperature via Optimizing Material Composition" Materials 17, no. 23: 5751. https://doi.org/10.3390/ma17235751

APA Style

Ma, S., Li, J., Du, D., Ruan, X., Cao, M., Lin, M., Hua, Q., Luo, Q., Tang, P., Guan, J., & Yu, J. (2024). Modulation of BixSb2−xTe3 Alloy Application Temperature via Optimizing Material Composition. Materials, 17(23), 5751. https://doi.org/10.3390/ma17235751

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