• Journal of Inorganic Materials
  • Vol. 34, Issue 3, 328 (2019)
Wei WANG1, Shi-Jie LUO1, Cong XIAN1, Qun XIAO1, Yang YANG2, Yun OU3, Yun-Ya LIU1, Shu-Hong XIE4, [in Chinese]1, [in Chinese]1, [in Chinese]1, [in Chinese]1, [in Chinese]2, [in Chinese]3, [in Chinese]1, and [in Chinese]4
Author Affiliations
  • 11. Hunan Provincial Key Laboratory of Thin Film Materials and Devices, School of Materials Science and Engineering, Xiangtan University, Xiangtan 411105, China
  • 22. Department of Mechanical Engineering, University of Washington, Seattle WA 98195-2600, USA
  • 33. Hunan Provincial Key Laboratory of Health Maintenance for Mechanical Equipment, Hunan University of Science and Technology, Xiangtan 411201, China
  • 44. Key Laboratory of Low Dimensional Materials and Application Technology of Ministry of Education, Xiangtan University, Xiangtan 411105, China
  • show less
    DOI: 10.15541/jim20180261 Cite this Article
    Wei WANG, Shi-Jie LUO, Cong XIAN, Qun XIAO, Yang YANG, Yun OU, Yun-Ya LIU, Shu-Hong XIE, [in Chinese], [in Chinese], [in Chinese], [in Chinese], [in Chinese], [in Chinese], [in Chinese], [in Chinese]. Enhanced Thermoelectric Properties of Hydrothermal Synthesized BiCl3/Bi2S3 Composites[J]. Journal of Inorganic Materials, 2019, 34(3): 328 Copy Citation Text show less

    Abstract

    Hierarchical spherical Bi2S3 particles with nanorod were synthesized by hydrothermal method, and then BiCl3/Bi2S3 composite powders with different molar ratios were consolidated into bulk samples by spark plasma sintering (SPS) technique. The addition of BiCl3 with appropriate amount not only increased the electrical conductivity, but also decreased the thermal conductivity of Bi2S3. The Bi2S3 sample doped with 0.5mol% BiCl3 shows a maximum electrical conductivity of 45.1 S·cm-1 at 762 K, which is much higher than that of pure Bi2S3 at 762 K (12.9 S·cm-1). The minimum thermal conductivity is 0.31 W·m-1·K-1 for the Bi2S3 sample doped with 0.25mol% BiCl3 at 762 K, which is lower than that of pure Bi2S3 (0.47 W·m-1·K-1)at the same temperature. The maximum ZT value of 0.63 at 762 K was achieved by Bi2S3 doped with 0.25mol% BiCl3, which is almost two times higher than that of pure Bi2S3(0.22).

    Thermoelectric (TE) materials have attracted more and more attention as they can convert heat into electricity directly[1,2,3]. The performance of thermoelectric materials depends on the figure of merit ZT, which is given by $ZT={{S}^{2}}\sigma T/({{\kappa }_{\text{e}}}+{{\kappa }_{\text{lat}}})$, where S, σ, T, κe and κlat are Seebeck coefficient, electrical conductivity, absolute temperature, carrier thermal conductivity and lattice thermal conductivity, respectively[1]. It is obvious that high efficiency thermoelectric conversion requires high power factor (${{S}^{2}}\sigma $) and low thermal conductivity simultaneously. However, the parameters mentioned above are coupled together and depend on the carrier concentrations, it is difficult to control the parameters independently. As such, many techniques have been used to enhance the ZT, such as nanostructures[4], layered thermoelectric materials[5,6,7], atomic substitution in alloys[8,9]. At the same time, many methods and theories were developed for analyzing the thermoelectric properties of composite materials, such as SThM[10], mesomechanics combined phase field simulation[11], and nonlinear asymptotic homogenization theories[12]. Thermoelectric materials are also combined with other energy technologies to achieve more application prospect, such as thermoelectric hybrid battery system[13].

    Among thermoelectric materials, the Pb-Te based[14,15]and Bi-Te based[16,17] compounds are the most promising candidates for TE applications. However, the rareness and high expense of tellurium have limited their application. Thus it is necessary to develop alternative materials that are abundant and cheaper. Bismuth sulfide (Bi2S3) has recently attracted much attention for its potential application in thermoelectric field.Bi2S3 has low thermal conductivity and high Seebeck coefficient at room temperature, though its electrical resistivity is high[18,19,20]. Thus the key point for improving the thermoelectric performance of Bi2S3 is to reduce the electrical resistivity[21]. In order to reduce the electrical resistivity of Bi2S3, several experimental methods have been reported, such as adding one-dimensional nanorods into Bi2S3 films[17,22], and Cu or Ag doping[23,24]. In this study, the TE properties of Bi2S3 are enhanced by doping with BiCl3, for BiCl3 as an electron donor dopant can improve the carrier content and increase the electrical conductivity[25].

    1 Experimental procedure

    Bi(NO3)3·5H2O and thiourea were used as the raw materials to synthesize the Bi2S3 powders. In the typical hydrothermal synthesis, 1.5 g Bi(NO3)3·5H2O and 1.5 g thiourea were dissolved in 40 mL deionized water with continuous stirring, and then transferred to a Teflon-lined stainless-steel autoclave with a capacity of 50 mL. After that the hybrid solution was maintained at 180 ℃ for 24 h, then it was air-cooled to room temperature. The synthesized precipitates of Bi2S3 were filtered and washed with distilled water and ethanol, and dried in air at 60 ℃. Afterwards, appropriate amounts of commercial BiCl3 (99.99%, Alfa Aesar) and Bi2S3 powders with different molar ratios were mixed in 40 mL ethylalcohol with stirring for 2 h, and the molar ratios of BiCl3 were controlled to be 0, 0.25%, 0.5% and 1.0%, respectively. They were sonicated for 30 min, and then the sample was dried in vacuum oven at 60 ℃. Different composite powders were loaded into a graphite die with an inner diameter of 12.6 mm and then sintered at 723 K for 10 min at a heating rate of 100 K/min under an axial compressive stress of 45 MPa in vacuum by using a spark plasma sintering (SPS) system (SPS-211LX). The SPSed specimens were disk-shaped with dimensions of 12.6 mm× 4 mm.

    The phase structure was analyzed by X-ray diffraction (XRD) with graphite monochromatized CuKα radiation (λ=0.15418 nm) (D/max-rA). The morphologies of the powders and the bulk samples after SPS were observed by field-emission scanning electron microscopy (FESEM, LEO-1525), and transmission electron microscopy (TEM, JEM-2100).

    Thermoelectric properties were measured with specimen surface perpendicular to the pressing direction of SPS. Seebeck coefficient and electrical resistivity were measured by Seebeck coefficient/electric resistance measuring system (ZEM-3). The thermal conductivity was calculated from the product of measured thermal diffusivity, specific heat and density. The thermal diffusivity was measured by the flash method (LFA427, NETZSCH). Before measurement, the sample was coated with a thin layer of graphite by graphite spray (Graphite33) to improve the thermal homogeneity of the sample. The specific heat capacity Cp was determined by differential scanning calorimetry (DSC 404F3, NETZSCH), which is in the range of 0.24-0.29 J·g-1·K-1 between room temperature and 750 K. The density was measured by Archimedes method. The thermal conductivity was calculated via the equation κ=ρDCP (κis the thermal conductivity, ρ is the density, D is the thermal diffusivity, and Cp is the specific heat capacity). Hall coefficients were measured on a home-built system with magnetic fields in the range of 0-1.25 T, utilizing a simple four-contact Hall-bar geometry in both negative and positive polarity to eliminate Joule resistive errors.

    2 Results and discussion

    Figure 1(a) shows the XRD patterns of BiCl3/Bi2S3 composite powders with different xmol% (x= 0, 0.25, 0.5, and 1.0) BiCl3. The results show that all the patterns correspond to the orthorhombic Bi2S3, indicating that the main phase of all the samples is Bi2S3 without preferred orientation, and no obvious impurity phases are observed as the content of BiCl3 no more than 1.0mol%. The enlarged patterns of 2θ in the range of 24°-32.5° are shown in Fig. 1(b). The ionic radius of Cl- is 0.181 nm, which is slightly smaller than that of S2-(0.184 nm), so the substitution of S2- with Cl- inducesa slight shrinkage of the unit cell, which results in the refined right shift of the major diffraction peaks with BiCl3 doping[25].According to the bond-order-length-strength (BOLS) theory, the fine nanocrystal can effectively decrease the average lattice constant of material[26], thus the shift of the major diffraction peaks for the doped materials is a combined result of both grain size and solid solubility. The major diffraction peaks should shift towards more right with increasing the solidsolubility of BiCl3, however,correspondingaverage grain size change simultaneously. The final result is thatthe sample doped with 0.25mol% BiCl3has the largest right shift among all specimens, as shown in the inset of Fig. 1(b), which is primarily due to the smallest grain size, and later SEM results will reveal it.

    XRD patterns (a) and the enlarged patterns of 2θ in the range of 24°-32.5°(b) with peaks of crystal plane (211) in the inset for BiCl3/Bi2S3 powders with different xmol% (x= 0, 0.25, 0.5, and 1.0) BiCl3

    Figure 1.XRD patterns (a) and the enlarged patterns of 2θ in the range of 24°-32.5°(b) with peaks of crystal plane (211) in the inset for BiCl3/Bi2S3 powders with different xmol% (x= 0, 0.25, 0.5, and 1.0) BiCl3

    The FESEM images of Bi2S3 powders and fractured surfaces of sintered samples of Bi2S3 doped with xmol% BiCl3 (x=0, 0.25, 0.5, 1.0) and TEM images of Bi2S3 doped with xmol% (x= 0, 1.0) BiCl3after SPS are shown in Fig. 2. It is observed in Fig. 2(a) that the major morphology of Bi2S3 powders synthesized by the hydrothermal method are spherical with diameter around 3-4 μm and the spherical particles are hierarchical, consisting of fine Bi2S3 nanorods with diameter in the range of 100-200 nm and length in the range of 1-3 μm[27]. Figure 2(b) shows the morphology of bulk Bi2S3 after SPS, showing that the grain size of the nanorods keeps stable, and the SPS sample is compact without obvious crack, though there are some small holes. The relative density of Bi2S3 is 89% as shown in Table 1, the theoretical density of Bi2S3 is 6.81 g·cm-3[28]. Figure 2(c-e) show the morphologies of Bi2S3 doped with different BiCl3. It is obvious that BiCl3 doping changes the morphologiesof Bi2S3 due to the activation role of BiCl3 by forming a substitution solid solution with Bi2S3[25]. In Fig. 2(c), the grain size of Bi2S3 doped with 0.25mol% BiCl3 is inhomogeneous, many fine particles are mixed with micrometer grains, and small pores are uniformly distributed in bulk. It is observed that the grain size of Bi2S3doped with 0.5mol% BiCl3 in Fig. 2(d) is similar to that ofBi2S3 doped with 1.0mol% BiCl3 in Fig. 2(e). Figure 2(f) shows the TEM image of Bi2S3 powders after SPS, which demonstrates that nanometer grain size is preserved after SPS. The orthorhombic structure of Bi2S3 is also confirmed by HRTEM, as shown in Fig. 2(g), where the lattice spacing is measured to be 0.421 nm along (220) plane and 0.357 nm along (130) plane, respectively. Figure 2(h) shows the TEM image of Bi2S3 doped with 1.0mol% BiCl3 powder after SPS, which confirms that partial nanorod structure still remains, and the grain size is similar to pure Bi2S3. SEM image of Bi2S3 doped with 1.0mol% BiCl3 bulk after SPS and corresponding elemental mappings of Bi, S and Cl are shown in Fig.3. Elemental mappingsrevealthat the distribution of Cl element is uniform, which indicates that Bi2S3 doped with BiCl3 is a substitutional solid solution as BiCl3 powders dispersed uniformly in Bi2S3 powder by alcohol ultrasonication.

    SEM images of Bi2S3 powders (a), fractured surfaces of SPSed samples of Bi2S3 doped with xmol% BiCl3 ((b) x=0, (c) x=0.25, (d) x=0.5, (e) x=1.0); TEM images of sintered samples of Bi2S3 doped with xmol% BiCl3 ((f) x=0, (h) x=1.0); HRTEM image of Bi2S3 powder after SPS (g)

    Figure 2.SEM images of Bi2S3 powders (a), fractured surfaces of SPSed samples of Bi2S3 doped with xmol% BiCl3 ((b) x=0, (c) x=0.25, (d) x=0.5, (e) x=1.0); TEM images of sintered samples of Bi2S3 doped with xmol% BiCl3 ((f) x=0, (h) x=1.0); HRTEM image of Bi2S3 powder after SPS (g)

    SEM image of the fractured surfaces of Bi2S3 doped with 1.0mol% BiCl3 bulk after SPS (a), corresponding elemental mappings of Bi, S and Cl (b-d)

    Figure 3.SEM image of the fractured surfaces of Bi2S3 doped with 1.0mol% BiCl3 bulk after SPS (a), corresponding elemental mappings of Bi, S and Cl (b-d)

    Table Infomation Is Not Enable

    Figure 4 shows the temperature dependence of thermoelectric performances for BiCl3/Bi2S3 composite samples. The negative Seebeck coefficients in Fig. 4(a) indicates that the composites are n-type semiconductors and the major carriers are electrons. The Seebeck coefficient of pure Bi2S3 sample is about -442.0 μV·K-1 at 336 K and the corresponding values of BiCl3/Bi2S3 composites are in the range of -322.0 - -247.3 μV·K-1. The values of carrier concentration at room temperature are displayed in Table 2, which shows that the carrier concentration increases monotonically with increasing BiCl3 content. It is obvious that near room temperature Seebeck coefficient de-creases with increasing amount of BiCl3 due to the increased carrier concentration with addition of BiCl3[19].It is shown that the Seebeck coefficient of pure Bi2S3 decreases with increasing temperature owing to the decreasing carrier mobility with increasing temperature. But for the BiCl3/Bi2S3 composite samples, Seebeck coefficient increases at lower temperature and then decreases at higher temperature, which is probably due to the bipolar effect caused by the band gap reduction at a higher temperature[29,30]. Figure 4(b) shows that the electrical conductivity (σ) as a whole increases with increasing temperature, and displays a semiconducting behavior. The σ of pure Bi2S3 is in the range of 1.0 S·cm-1 to 12.9 S·cm-1 at 336-762 K. The maximum σof 45.1 S·cm-1is present in Bi2S3 doped with 0.5mol% BiCl3 at 762 K. However, σ of Bi2S3 doped with 1.0mol% BiCl3 increases at first and then decreases with rising temperature.According to the formula σ=neμ, σ of semiconductor is proportional to carrier concentration (n), electron charge(e) and electron mobility (μ). With the increase of temperature, more electrons are excited and the carrier concentration increases, but the electron mobility decreases due to the phonon scattering[25], which leads to increased electrical conductivity at lower temperatures and then decreased electrical conductivity at a higher temperature for the sample doped with 1.0mol% BiCl3. It is observed that the change of σwith different doping concentration is irregular, which is probably due to the complex absorption of hydrothermally synthesized powders. The power factor (S2σ) is shown in Fig. 4(c). The pure Bi2S3 shows a minimum power factor of 19.8 μW·m-1·K-2 around 336 K, and the corresponding value is 110.0 μW·m-1·K-2 for Bi2S3 doped with 0.25mol% BiCl3, and the highest power factor reaches 350.2 µW·m-1·K-2 at 762 K for the Bi2S3 doped with 0.5mol% BiCl3. It is obvious that the doped BiCl3 enhances the power factor of Bi2S3.

    Temperature dependence of thermoelectric performances for BiCl3/Bi2S3 composite samples (a) Seebeck coefficient; (b) Electrical conductivity; (c) Power factor; (d) Total thermal conductivity; (e) Lattice thermal conductivity; (f) Figure of merit (ZT)

    Figure 4.Temperature dependence of thermoelectric performances for BiCl3/Bi2S3 composite samples
    (a) Seebeck coefficient; (b) Electrical conductivity; (c) Power factor; (d) Total thermal conductivity; (e) Lattice thermal conductivity; (f) Figure of merit (ZT)

    Table Infomation Is Not Enable

    The thermal conductivity (κ) of the samples decreases with increasing temperatures, as shown in Fig. 4(d), which is attributed to the strong phonon scattering at high temperatures. The composite sample has a lower κ than that of the pure Bi2S3 as the second phase BiCl3 enhances the phonon scattering[23]. Bi2S3 doped with 0.5mol% BiCl3 has the highest σand κ at higher temperatures probably due to the higher carrier concentration, lattice deformation and larger grain size as shown in Fig 2(d). Thethermal conductivity of Bi2S3 doped with 1.0mol% BiCl3 is a little lower than that of Bi2S3 doped with 0.5mol% BiCl3, probably owing to a high doping fraction introducing more lattice defects, which is an effective approach for decreasing the lattice thermal conductivity, thusachieving a lower κlat[31]. The minimum κ is 0.31 W·m-1·K-1 at 762 K for Bi2S3 doped with 0.25mol% BiCl3 due to the formation of many fine grains, which is 34% lower than that of pure Bi2S3(0.47 W·m-1·K-1). The temperature dependence of lattice thermal conductivity (κlat) is shown in Fig. 4(e). The ratio of lattice thermal conductivity (κlat) to κtot indicates that κtot is dominated by phonon transport. It is observed in Fig. 4(f) that all the samples show an increased ZT value with temperature increasing, which is attributed to the increased σand decreased κ. It is observed in Table 1 that the maximal ZT of 0.63 at 762 K is obtained in Bi2S3 doped with 0.25mol% BiCl3, which is a higher ZT compared with the reported Bi2S3-based material[25, 32-34].

    3 Conclusions

    n-type BiCl3/Bi2S3 composite samples were fabricated by hydrothermal method combined with SPS technique. The addition of BiCl3 effectively increased the electrical conductivity and decreased the thermal conductivity of Bi2S3. Bi2S3 doped with 0.5mol% BiCl3 shows a maximum electrical conductivity of 45.1 S·cm-1 at 762 K, which is more than twice higher than that of pure Bi2S3(12.9 S·cm-1), and Bi2S3 doped with 0.25mol% BiCl3 achieves the minimum thermal conductivity of 0.31 W·m-1·K-1 at 762 K, more than 30% decrease as compared with pure Bi2S3. Due to the higher electrical conductivity and lower thermal conductivity, a maximum ZT value of 0.63 is achieved at 762 K for Bi2S3 doped with 0.25mol% BiCl3, a significant enhancement compared to that of pure Bi2S3 (0.22).

    References

    [1] Y PEI, X SHI, A LALONDE et al. Convergence of electronic bands for high performance bulk thermoelectrics. Nature, 473, 66-69(2011).

    [2] E BELL L. Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science, 321, 1457-1461(2008).

    [3] J DISALVO F. Thermoelectric cooling and power generation. Science, 285, 703-706(1999).

    [4] S LIU W, F LI J, D ZHAO L et al. High-performance nanostructured thermoelectric materials. NPG Asia Materials, 2, 152-158(2010).

    [5] J SNYDER G, S TOBERER E. Complex thermoelectric materials. Nature Materials, 7, 105-114(2008).

    [6] Y MA F, H LEI C, Y YANG et al. Is thermoelectric conversion efficiency of a composite bounded by its constituents. Applied Physics Letters, 102, 053905-1-4(2013).

    [7] Y MA F, Y YANG, H XIE S et al. On the effective thermoelectric properties of layered heterogeneous medium. Journal of Applied Physics, 111, 013510-1-7(2012).

    [8] J ZIDE, W KIM, A GOSSARD et al. Thermal conductivity reduction and thermoelectric figure of merit increase by embedding nanoparticles in crystalline semiconductors. Physical Review Letters, 96, 045901-1-4(2006).

    [9] H MUTA, K KUROSAKI, A IEDA et al. Substitution effect on the thermoelectric properties of alkaline earth titanate. MaterialsLetters, 58, 3868-3871(2004).

    [10] F MA, S WANG, E NASR ESFAHANI et al. Quantitative nanoscale mapping of three-phase thermal conductivities in filled skutterudites via scanning thermal microscopy. National Science Review, 5, 59-69(2017).

    [11] L CHEN, Y LIU, J LI. Precipitate morphologies of pseudobinary Sb2Te3-PbTe thermoelectric compounds. Acta Materialia, 65, 308-315(2014).

    [12] Y MA F, Y YANG, H LEI C et al. Nonlinear asymptotic homogenization and the effective behavior of layered thermoelectric composites. Journal of the Mechanics and Physics of Solids, 61, 1768-1783(2013).

    [13] J JIANG, Y NIU, C WANG et al. Hybrid thermoelectric battery electrode FeS2 study. Nano Energy, 45, 432-438(2018).

    [14] H DUGHAISH Z. Lead telluride as a thermoelectric material for thermoelectric power generation. Physica B: Condensed Matter, 322, 205-223(2002).

    [15] V JOVOVIC, S TOBERER E, P HEREMANS J et al. Enhancement of thermoelectric efficiency in PbTe by distortion of the electronic density of states. Science, 321, 554-557(2008).

    [16] B POUDEL, Q HAO, Y MA et al. High-thermoelectric performance of nanostructured bismuth antimony telluride bulk alloys. Science, 320, 634-638(2008).

    [17] B ZHAO X, J ZHU T, Q CAO Y et al. Syntheses and thermoelectric properties of Bi2Te3/Sb2Te3 bulk nanocomposites with laminated nanostructure. Applied Physics Letters, 92, 143106-1-3(2008).

    [18] A SEGURA, J MARTINEZ-PASTOR, A CANTARERO et al. Transport properties of bismuth sulfide single crystals. Physical Review B, 35, 9586-9590(1987).

    [19] K BISWAS, D ZHAO L, G KANATZIDIS M. Tellurium-free thermoelectric: the anisotropic n-type semiconductor Bi2S3. Advanced Energy Materials, 2, 634-638(2012).

    [20] F GUO C, W LIU, M YAO et al. Bi2S3 nanonetwork as precursor for improved thermoelectric performance. Nano Energy, 4, 113-122(2014).

    [21] C UHER, B CHEN, L IORDANIDIS et al. Transport properties of Bi2S3 and the ternary bismuth sulfides KBi6.33S10 and K2Bi8S13. Chemistry of Materials, 9, 1655-1658(1997).

    [22] D CHEN L, Q YAO, C LIUFU S et al. Assembly of one-dimensional nanorods into Bi2S3 films with enhanced thermoelectric transport properties. Applied Physics Letters, 90, 112106-1-3(2007).

    [23] P ZHANG B, H GE Z, Y LIU et al. Nanostructured Bi2-xCuxS3 bulk materials with enhanced thermoelectric performance. Physical Chemistry Chemical Physics, 14, 4475-4481(2012).

    [24] P ZHANG B, Q YU Y, H GE Z et al. Thermoelectric properties of Ag-doped bismuth sulfide polycrystals prepared by mechanical alloying and spark plasma sintering. Materials Chemistry and Physics, 131, 216-222(2011).

    [25] X DU, F CAI, X WANG. Enhanced thermoelectric performance of chloride doped bismuth sulfide prepared by mechanical alloying and spark plasma sintering. Journal of Alloys and Compounds, 587, 6-9(2014).

    [26] F ZHOU Z, X YANG X, Y WANG et al. Raman spectroscopy determination of the Debye temperature and atomic cohesive energy of CdS, CdSe, Bi2Se3,Sb2Te3 nanostructures. Journal of Applied Physics, 112, 083508-1-6(2012).

    [27] S THONGTEM, A PHURUANGRAT, T THONGTEM. Characterization of Bi2S3 nanorods and nano-structured flowers prepared by a hydrothermal method. Materials Letters, 63, 1496-1498(2009).

    [28] P ZHANG B, D ZHAO L, S LIU W et al. Enhanced thermoelectric properties of bismuth sulfide polycrystals prepared by mechanical alloying and spark plasma sintering. Journal of Solid State Chemistry, 181, 3278-3282(2008).

    [29] D ZHAO L, H LO S, Y ZHANG et al. Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Nature, 508, 373-377(2014).

    [30] M ZHOU, J ZHAO, M HAN Y et al. Thermoelectric performance of SnS and SnS-SnSe solid solution. Journal of Materials Chemistry A, 3, 4555-4559(2015).

    [31] G KANATZIDIS M, D ZHAO L, G TAN. Rationally designing high-performance bulk thermoelectric materials. Chemical Reviews, 116, 12123-12149(2016).

    [32] P QIN, S HE D, H GE Z et al. Highly enhanced thermoelectric properties of Bi/Bi2S3 nanocomposites. ACS Applied Materials & Interfaces, 9, 4828-4834(2017).

    [33] J YAN, G LIU, J YANG et al. Enhanced the thermoelectric properties of n-type Bi2S3 polycrystalline by iodine doping. Journal of Alloys and Compounds, 728, 351-356(2017).

    [34] P ZHANG B, H GE Z, J ZHANG L et al. Fabrication and properties of Bi2S3-xSex thermoelectric polycrystals. Solid State Communications, 162, 48-52(2013).

    Wei WANG, Shi-Jie LUO, Cong XIAN, Qun XIAO, Yang YANG, Yun OU, Yun-Ya LIU, Shu-Hong XIE, [in Chinese], [in Chinese], [in Chinese], [in Chinese], [in Chinese], [in Chinese], [in Chinese], [in Chinese]. Enhanced Thermoelectric Properties of Hydrothermal Synthesized BiCl3/Bi2S3 Composites[J]. Journal of Inorganic Materials, 2019, 34(3): 328
    Download Citation