Next Article in Journal
The Enhanced Hydrogen Storage Capacity of Carbon Fibers: The Effect of Hollow Porous Structure and Surface Modification
Previous Article in Journal
Conformal Self-Assembly of Nanospheres for Light-Enhanced Airtightness Monitoring and Room-Temperature Gas Sensing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Thermoelectric Performance of n-Type Bi2Se3 Nanosheets through Sn Doping

1
Catalonia Energy Research Institute—IREC, Sant Adrià de Besòs, 08930 Barcelona, Spain
2
Catalan Institute of Nanoscience and Nanotechnology (ICN2), CSIC and BIST, Campus UAB, Bellaterra, 08193 Barcelona, Spain
3
Istituto Italiano di Tecnologia, Via Morego 30, 16163 Genova, Italy
4
ICREA, Pg. Lluis Companys 23, 08010 Barcelona, Spain
5
Institute of Energy Technologies, Department of Chemical Engineering and Barcelona Research Center in Multiscale Science and Engineering, Universitat Politècnica de Catalunya, EEBE, 08019 Barcelona, Spain
6
Institute of Science and Technology Austria (IST Austria), Am Campus 1, 3400 Klosterneuburg, Austria
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(7), 1827; https://doi.org/10.3390/nano11071827
Submission received: 22 June 2021 / Revised: 3 July 2021 / Accepted: 12 July 2021 / Published: 14 July 2021

Abstract

:
The cost-effective conversion of low-grade heat into electricity using thermoelectric devices requires developing alternative materials and material processing technologies able to reduce the currently high device manufacturing costs. In this direction, thermoelectric materials that do not rely on rare or toxic elements such as tellurium or lead need to be produced using high-throughput technologies not involving high temperatures and long processes. Bi2Se3 is an obvious possible Te-free alternative to Bi2Te3 for ambient temperature thermoelectric applications, but its performance is still low for practical applications, and additional efforts toward finding proper dopants are required. Here, we report a scalable method to produce Bi2Se3 nanosheets at low synthesis temperatures. We studied the influence of different dopants on the thermoelectric properties of this material. Among the elements tested, we demonstrated that Sn doping resulted in the best performance. Sn incorporation resulted in a significant improvement to the Bi2Se3 Seebeck coefficient and a reduction in the thermal conductivity in the direction of the hot-press axis, resulting in an overall 60% improvement in the thermoelectric figure of merit of Bi2Se3.

Graphical Abstract

1. Introduction

Thermoelectric (TE) devices that directly and reversibly convert heat into electricity find unlimited applications [1,2,3,4,5,6], but their real implementation is hampered by their low cost-effectiveness. The efficiency of energy conversion of a TE device is in part determined by the transport properties of the TE material: Seebeck coefficient S, electrical conductivity σ, and thermal conductivity κ. These material properties are generally grouped into a dimensionless figure of merit, ZT = S2σT/κ, where T is the absolute temperature. Efficient TE materials are characterized by high power factors, S2σ, and low thermal conductivities, κ. However, the strong correlation between these parameters makes the optimization of the material performance extremely difficult. Several strategies have been developed to maximize ZT, including engineering the electronic band structure of the TE material through doping [7], the use of energy filtering interphases [8], and the reduction in lattice thermal conductivity through the introduction of abundant grain boundaries [9].
Commercial devices use large amounts of highly crystalline Bi2Te3-based alloys as the active TE material, which accounts for a significant part of the total cost of the device. To reduce the device cost, it is critical to develop low-cost materials not relying on scarce Te and low-cost processing strategies not based on high-temperature crystallization from high purity melts. Bi2Se3 presents the same structure and similar properties as Bi2Te3, and, thus, it is an obvious possible Te-free alternative to Bi2Te3. Numerous attempts to improve the TE figure of merit of this material have been reported, but with moderate success [10,11,12,13,14,15,16,17,18].
To maximize the TE performance of a material, one first critical point is to optimize its charge carrier concentration. Three main strategies can be used in this direction: tuning of the material stoichiometry, the introduction of atomic impurities, and/or the introduction of secondary phases able to inject the proper charge carriers to the host material [19,20,21,22,23,24,25,26,27,28]. Besides optimizing the charge carrier concentration, it is fundamental to minimize the thermal conductivity using multiscale phonon scattering centers. In this direction, the introduction of atomic impurities and abundant grain boundaries are highly effective. Besides, energy barriers at grain boundaries can also introduce energy- and charge sign-dependent charge carrier scattering, which could result in significant improvements to the Seebeck coefficient [29].
Here, we present a low temperature and high yield, solution-based strategy to prepare Bi2Se3 nanomaterials doped with different elements. The influence of different dopants on the TE properties of the material was tested. From the results obtained, we demonstrated Sn to have the highest potential to yield Bi2Se3-based materials with improved performance. Thus, we further analyzed the location and chemical state of this element within Bi2Se3 and its effect on the material transport properties.

2. Materials and Methods

2.1. Chemicals and Solvents

Polyvinylpyrrolidone (PVP, (C6H9NO)n, AMW ~55,000), indium(III) acetate (InC6H9O6, 99.99%), and bismuth(III) nitrate pentahydrate (Bi(NO3)3·5H2O, ≥99.99%) were purchased from Sigma-Aldrich. Sodium selenite (Na2SeO3, ≥98%), copper (II) nitrate trihydrate (Cu(NO3)2·3H2O, 99%), silver nitrate (AgNO3, ≥99.9%), ethylene glycol (EG, HOCH2CH2OH, 99%), lead (II) acetate trihydrate (PbC4H6O4·3H2O), tin(II) chloride anhydrous (SnCl2, 98%), and potassium hydroxide (KOH, ≥98%) were acquired from Fisher. Analytical grade ethanol and acetone were obtained from various sources.

2.2. Synthesis of Bi2Se3

Bi2Se3 particles were produced following the approach developed by Liu et al., with slight modifications [1]. Bi2Se3 (5 mmol), Bi(NO3)3·5H2O (10 mmol, 4.851 g), Na2SeO3 (15 mmol, 2.594 g), KOH (50 mmol, 2.806 g), and PVP (0.5 g) were first dissolved in a three-neck flask containing EG (200 mL) under an Ar atmosphere at an ambient temperature for 0.5 h. The solution was then heated to 180 °C and kept at this temperature for 3 h. Immediately after the reaction was completed, the solution was allowed to cool naturally to room temperature by removing the heating mantle. The solution was divided into several centrifuge tubes and acetone was added to collect the solid product by centrifugation. In the next step, ethanol was introduced to re-disperse the particles, and acetone was added to precipitate them again. This step was repeated twice. Next, purified Bi2Se3 particles were dried under a vacuum overnight at room temperature and kept in the glovebox for posterior use. Around 3 g of particles were obtained per batch.

2.3. Synthesis of Bi2−xMxSe3

The procedure used to produce Bi2−xMxSe3 was the same used to produce Bi2Se3, but the proper amount of bismuth nitrate was replaced with the corresponding metal precursor, as listed in the chemical section above.

2.4. Nanomaterial Consolidation

Dried Bi2Se3 particles were first annealed for 1 h at 350 °C under an Ar flow in a tubular furnace. Then, the annealed material was introduced into a graphite die and consolidated into cylinders (Ø 10 mm × 10 mm) at 480 °C and 50 MPa of pressure for 4 min with a custom-made hot press. The relative densities of the compacted pellets were measured by Archimedes’ method. To measure transport properties in the two relevant directions, cylindrical pellets were cut into ca. 8 × 6 × 1 mm rectangular bars along the pressure axis, and 1 mm thick disks along the perpendicular to the pressure axis.

2.5. Structural and Chemical Characterization

The field emission scanning electron microscope (SEM) was used to determine the nanoparticle morphology on an Auriga Zeiss. An Oxford energy dispersive X-ray spectrometer (EDX) was used to measure the material composition at 20.0 kV. X-ray diffraction (XRD) was performed on a Bruker AXS D8 Advance diffractometer. The crystal structure of the samples was analyzed under the 200 keV Tecnai F20 field emission microscope equipped with a Gatan quantum image filter. X-ray photoelectron spectroscopy (XPS) was performed on a Specs system with the material inside the chamber at a pressure below 10−7 Pa. Data processing was carried out using the CasaXPS program.

2.6. Performance Characterization of Bulk Nanomaterial

Seebeck coefficients and resistivities were measured by the static direct current method and by the standard four-probe method, respectively, in an LSR-3 Linseis system under helium. All samples were tested for at least three heating and cooling cycles. Considering the system accuracy and measurement accuracy, the measurement error of the conductivity and Seebeck coefficient was estimated to be about 4%. Thermal diffusivity (λ), constant pressure heat capacity (Cp), and density of the material (ρ) were used to obtain the thermal conductivities (κtotal), where κtotal = λCpρ. The thermal diffusivities of the samples were measured by a Xenon Flash Apparatus XFA600, which has an estimated error of ca. 5%. We used a constant as the heat capacity (Cp), which was estimated from empirical formulas by the Dulong–Petit limit (3R law). Under a magnetic field of 0.6 T, the Hall charge carrier concentrations and mobilities at room temperature were measured by the Van der Pauw method. The figures in this article do not have error bars in order to avoid cluttering the plots.

3. Results and Discussion

Following the above synthesis method, Bi2Se3 nanosheets grouped into flower-like particles were produced, as shown by the SEM micrographs (Figure 1a,b). Bi2Se3 shows a screw dislocation growth mechanism (Figure 1c). This layer-by-layer growth mechanism is usually due to the low supersaturation in the synthesis conditions, which inhibits dendritic growth. Axial screw dislocation is the self-sustaining spiral growth of nanosheets around the axis [15,23].
The XRD analysis showed that the crystal structure of the obtained nanosheets matched the rhombohedral Bi2Se3 phase (Figure 1d, JCPDS No. 00-033-0214). Figure 1e shows the layered rhombohedral crystal structure of Bi2Se3, which consists of five covalently bonded atomic planes of Se-Bi-Se-Bi-Se. Quintuple layers are weakly bonded by van der Waals. The HRTEM characterization showed that the material has good crystallinity and has a crystal phase consistent with the Bi2Se3 rhombohedral phase (space group = R3-MH) with a = b = 4.1340 Å and c = 28.6300 Å (Figure 1f,g).
Nanosheets were annealed at 350 °C, and the annealed powder was hot-pressed into cylindrical pellets with a diameter of 10 mm and a height of 10 mm in a glovebox filled with argon. Samples were hot-pressed for four minutes at 480 °C and 50 MPa, and then naturally cooled to an ambient temperature. The relative density of the cylinders produced by this process was about 93% of the theoretical value. From the consolidated cylinder, a rectangular bar of 8 × 6 × 1 mm was cut longitudinally, and a 1 mm thick pellet was cut transversely (Figure 2). These samples were used to measure the TE properties of materials parallel to the pressure axis (//) and perpendicular to the pressure axis (⊥).
Figure 3 displays the electrical conductivity (σ), Seebeck coefficient (S), and power factor (PF = S2σ) of the Bi2−xMxSe3 (M = Sn, Cu, Ag, Pb, In) pellets perpendicular to the press direction. The electrical conductivity of the undoped material decreased with temperature, which pointed to a degenerated semiconductor behavior. Compared to the undoped material, the electrical conductivity of all the doped samples was slightly lower, except for the Ag-doped samples. On the other hand, the samples doped with Sn and to a minor extent, Pb displayed an increase in the absolute value of the Seebeck coefficient. Overall, the highest power factors were obtained with the Sn doping. Thus, we decided to further study the effect of this element.
Figure 4 displays the electrical conductivity, Seebeck coefficient, and power factor of Bi2-xSnxSe3 materials containing different amounts of Sn. We observed that the electrical conductivity decreased and the absolute value of the Seebeck coefficient increased with the amount of Sn up to a certain Sn concentration. The highest power factors were finally obtained for Bi1.93Sn0.07Se3. The decrease in the electrical conductivity and increase in the absolute value of the Seebeck coefficient denoted a reduction in the charge carrier concentration, which points to the presence of Sn2+ ions instead of Sn4+ at Bi3+ sites.
Figure 5 shows representative SEM micrographs of the Bi1.93Sn0.07Se3 particles, which had similar sizes but did not display the flower-like morphology observed from Bi2Se3. As shown in Figure 6, the XRD characterization confirmed that the Bi1.93Sn0.07Se3 particles maintained their rhombohedral structure. With the introduction of Sn, the XRD peaks shifted to lower 2θ angles, suggesting the incorporation of Sn within the Bi2Se3 lattice [5].
The HRTEM characterization confirmed the rhombohedral phase of the Bi1.93Sn0.07Se3 particles (Figure 7a). EELS chemical composition maps obtained from the red squared region in the HAADF STEM micrograph shown in Figure 7b displayed a homogeneous distribution of Sn, Bi and Se within the Bi1.93Sn0.07Se3 nanosheet.
Bi 4f, Se 3d and Sn 3d high-resolution XPS spectra of Bi1.93Sn0.07Se3 particles are displayed in Figure 7c. The high-resolution Bi 4f spectrum was fitted with two doublets, associated with Bi3+ within a Bi2Se3 chemical environment (Bi 4f7/2 at 158.2 eV) and Bi within a more electronegative environment, as it could be Bi2O3 or Bi2SeO2 (Bi 4f7/2 at 159.4 eV) [30]. This oxidation of the particles surface is related to their transport and handling in the air [4]. The high-resolution Se 3d XPS spectrum was fitted with three doublets, corresponding to Se within the Bi2Se3 (Se 3d5/2 at 53.6 eV), selenium in an elemental or Bi2SeO2 environment (Se 3d5/2 at 55.2 eV), probably arising from the partial oxidation of the material surface, and Se in a SeO2 chemical environment (Se 3d5/2 at 58.6 eV) [30,31]. Finally, the high-resolution Sn 3d XPS spectrum was difficult to fit owing to the small amount of this element, but the broadness of the peaks pointed to the presence of at least two Sn chemical states that should be tentatively assigned to Snx+ within a SnSe and an oxidized environment [31]. According to the electrical properties measured at the introduction of Sn, we tentatively assigned the Snx+ component within SnSe to Sn2+.
Top-view and cross-section SEM micrographs of Bi2Se3 and Bi1.93Sn0.07Se3 (Figure 8) showed the final pellets presenting a laminar microstructure with an evident preferential orientation of the material layers. Bi2Se3 displayed larger and thinner layers than Bi1.93Sn0.07Se3. When comparing the XRD patterns of the sample held in two normal directions, parallel and perpendicular to the pressure axis (Figure 8e,f), we observed that the relative XRD peak intensity clearly differed. This result confirmed the preferential crystallographic orientation of the hot-pressed materials, with the [001] crystallographic direction oriented parallel to the pressure axis.
Figure 9 displays the TE properties of Bi2Se3 and Bi1.93Sn0.07Se3 measured in the two directions. We observed σ > σ//, as expected from the higher charge carrier mobilities in the ab crystal plane compared with the c direction, and the extended size of the crystal domains in the direction normal to the pressure direction within the layered pellets. On the other hand, similar Seebeck coefficients were obtained in both directions, pointing out that this parameter has little relationship with grain boundary scattering in this material.
When comparing the TE properties of Bi2Se3 and Bi1.93Sn0.07Se3, we observed σ to slightly increase and σ// to slightly decrease with the introduction of Sn, which could be in part related to the thinner and larger material layers observed within the layered Bi2Se3 pellets compared with Bi1.93Sn0.07Se3. On the other hand, with the introduction of Sn, S increased in both directions, which resulted in a higher PF in both directions for the Bi1.93Sn0.07Se3 sample.
Table 1 displays the Hall charge carrier concentration and mobility of Bi2Se3 and Bi1.93Se0.07Se3 pellets at room temperature. As expected from the electrical conductivity and Seebeck coefficient measurements, we observed the Sn introduction to result in a decrease in the charge carrier concentration, from nH = 1.8 × 1019 cm−3 for Bi2Se3 to nH = 1.3 × 1019 cm−3 for Bi1.93Se0.07Se3. This result is consistent with the presence of Sn2+ replacing Bi3+ ions, thus trapping a free electron. We believe this optimization of the charge carrier concentration to be at the origin of the higher PF obtained with the introduction of Sn. We further calculated the effective mass of the Bi2Se3 and Bi1.93Sn0.07Se3 materials using a single parabolic band (SPB) model. For detailed calculations, the carrier transport property analysis included [32]:
The Seebeck coefficient,
S ( η ) = κ B e [ ( r + 5 / 2 ) · F r + 3 / 2 ( η ) ( r + 3 / 2 ) · F r + 1 / 2 ( η ) η ]  
The Hall carrier concentration,
n H = 1 e · R H = ( 2 m * · κ B T ) 3 / 2 3 π 2 3 · ( r + 3 / 2 ) 2 · F r + 1 / 2 2 ( η ) ( 2 r + 3 / 2 ) · F 2 r + 1 / 2 ( η )  
The Hall mobility,
μ H = [ e π 4 2 ( κ B T ) 3 / 2 C l E d e f 2 ( m * ) 5 / 2 ] ( 2 r + 3 / 2 ) · F 2 r + 1 / 2 ( η ) ( r + 3 / 2 ) 2 · F r + 1 / 2 ( η )  
where F x ( η ) = 0 ε x 1 + e ( ε η ) d ε   is the Fermi integral.
In the above equations, S, μH, η, κB, e, r, RH, , Cl, Edef, and m* are the Seebeck coefficient, the carrier mobility, the reduced Fermi level, the Boltzmann constant, the electron charge, the carrier scattering factor (r = −1/2 for acoustic phonon scattering), the Hall coefficient, the reduced plank constant, the elastic constant for longitudinal vibrations, the deformation potential coefficient, and the density of state effective mass, respectively. Using the experimental S, the reduced Fermi level η can be obtained by Equation (1). Substituting the estimated η and the measured “nH” and “μH” into Equations (2) and (3), the effective mass “m*” can be calculated. The results show that m* increased slightly with the Sn doping (Table 1).
Consistent with the layered structure of the pellets and their crystallographic texture, the thermal conductivity of the materials measured in the direction normal to the pressure axis was much higher than in the pressure axis, κ > κ//. The introduction of Sn resulted in lower κ but higher κ// values, which is related to the different microstructure of the layers, with thicker and smaller Bi1.93Se0.07Se3 plates compared to those within the Bi2Se3 pellet.
Overall, the Bi1.93Sn0.07Se3 samples measured in the pellet plane displayed the highest PF and ZT values, 0.65 m Wm−1 K−2 and 0.41, respectively. These values are among the highest published for Bi2Se3-based materials, as displayed in Table 2.

4. Conclusions

We report here a large-scale method to synthesize n-type Bi2Se3 nanosheets with different dopants. Samples doped with Sn provided the highest TE performances. The composition was optimized at Bi1.93Se0.07Se3. Upon Sn doping, we observed an increase in the Seebeck coefficient and the power factor compared to pure Bi2Se3. As a result, a power factor up to 0.65 m Wm−1K−2 and a ZT of 0.41 were obtained for the Bi1.93Sn0.07Se3 pellet measured in the direction perpendicular to the pressure axis, which represents a 60% increase over pure Bi2Se3.

Author Contributions

Conceptualization, M.L.; methodology, M.L. and Y.L.; software, M.L. and Y.Z. (Yong Zuo); formal analysis, M.L., J.A. and J.L.; data curation, T.Z., Y.Z. (Yu Zhang) and K.X.; writing—original draft preparation, M.L.; writing—review and editing, A.C.; supervision, Y.L., A.C. All authors have read and agreed to the published version of the manuscript.

Funding

All sources of funding are acknowledged in the acknowledgements section.

Data Availability Statement

Data are contained within the article.

Acknowledgments

M.L., Y.Z., T.Z. and K.X. thank the China Scholarship Council for their scholarship support. Y.L. acknowledges funding from the European Union’s Horizon 2020 research and innovation program under the Marie Sklodowska-Curie grant agreement No. 754411. J.L. thanks the ICREA Academia program and projects MICINN/FEDER RTI2018-093996-B-C31 and G.C. 2017 SGR 128. ICN2 acknowledges funding from the Generalitat de Catalunya 2017 SGR 327 and the Spanish MINECO ENE2017-85087-C3.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, Y.; Zhang, Y.; Lim, K.H.; Ibáñez, M.; Ortega, S.; Li, M.; David, J.; Martí-Sànchez, S.; Ng, K.M.; Arbiol, J.; et al. High Thermoelectric Performance in Crystallographically Textured n type Bi2Te3-xSex Produced from Asymmetric Colloidal Nanocrystals. ACS Nano 2018, 12, 7174–7184. [Google Scholar] [CrossRef] [Green Version]
  2. Liu, Y.; Zhang, Y.; Ortega, S.; Ibáñez, M.; Lim, K.H.; Grau-Carbonell, A.; Martí-Sánchez, 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] [Green Version]
  3. Zhang, Y.; Liu, Y.; Lim, K.H.; Xing, C.; Li, M.; Zhang, T.; Tang, P.; Arbiol, J.; Llorca, J.; Ng, K.M.; et al. Tin Diselenide Molecular Precursor for Solution-Processable Thermoelectric Materials. Angew. Chem. 2018, 130, 17309–17314. [Google Scholar] [CrossRef]
  4. Li, M.; Liu, Y.; Zhang, Y.; Han, X.; Zhang, T.; Zuo, Y.; Xie, C.; Xiao, K.; Arbiol, J.; Llorca, J.; et al. Effect of the Annealing Atmosphere on Crystal Phase and Thermoelectric Properties of Copper Sulfide. ACS Nano 2021, 15, 4967–4978. [Google Scholar] [CrossRef] [PubMed]
  5. Li, M.; Liu, Y.; Zhang, Y.; Zuo, Y.; Li, J.; Lim, K.H.; Cadavid, D.; Ng, K.M.; Cabot, A. Crystallographically textured SnSe nanomaterials produced from the liquid phase sintering of nanocrystals. Dalt. Trans. 2019, 48, 3641–3647. [Google Scholar] [CrossRef] [PubMed]
  6. Ortega, S.; Ibáñez, M.; Liu, Y.; Zhang, Y.; Kovalenko, M.V.; Cadavid, D.; Cabot, A. Bottom-up engineering of thermoelectric nanomaterials and devices from solution-processed nanoparticle building blocks. Chem. Soc. Rev. 2017, 46, 3510–3528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Ibañez, M.; Hasler, R.; Genc, A.; Liu, Y.; Kuster, B.; Schuster, M.; Dobrozhan, O.; Cadavid, D.; Arbiol, J.; Cabot, A.; et al. Ligand-mediated band engineering in bottom-up assembled SnTe nanocomposites for thermoelectric energy conversion. J. Am. Chem. Soc. 2019, 141, 8025–8029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Zhang, Y.; Xing, C.; Liu, Y.; Li, M.; Xiao, K.; Guardia, P.; Lee, S.; Han, X.; Ostovari Moghaddam, A.; Josep Roa, J.; et al. Influence of copper telluride nanodomains on the transport properties of n-type bismuth telluride. Chem. Eng. J. 2021, 418, 129374. [Google Scholar] [CrossRef]
  9. Ibáñez, M.; Genç, A.; Hasler, R.; Liu, Y.; Dobrozhan, O.; Nazarenko, O.; De La Mata, M.; Arbiol, J.; Cabot, A.; Kovalenko, M.V. Tuning transport properties in thermoelectric nanocomposites through inorganic ligands and heterostructured building blocks. ACS Nano 2019, 13, 6572–6580. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, S.; Sun, Y.; Yang, J.; Duan, B.; Wu, L.; Zhang, W.; Yang, J. High thermoelectric performance in Te-free (Bi,Sb)2Se3: Via structural transition induced band convergence and chemical bond softening. Energy Environ. Sci. 2016, 9, 3436–3447. [Google Scholar] [CrossRef]
  11. Min, Y.; Park, G.; Kim, B.; Giri, A.; Zeng, J.; Roh, J.W.; Kim, S.I.; Lee, K.H.; Jeong, U. Synthesis of Multishell Nanoplates by Consecutive Epitaxial Growth of Bi2Se3 and Bi2Te3 Nanoplates and Enhanced Thermoelectric Properties. ACS Nano 2015, 9, 6843–6853. [Google Scholar] [CrossRef]
  12. Min, Y.; Roh, J.W.; Yang, H.; Park, M.; Kim, S.I.; Hwang, S.; Lee, S.M.; Lee, K.H.; Jeong, U. Surfactant-free scalable synthesis of Bi2Te3 and Bi2Se3 nanoflakes and enhanced thermoelectric properties of Their Nanocomposites. Adv. Mater. 2013, 25, 1425–1429. [Google Scholar] [CrossRef]
  13. Liu, W.; Lukas, K.C.; McEnaney, K.; Lee, S.; Zhang, Q.; Opeil, C.P.; Chen, G.; Ren, Z. Studies on the Bi2Te3-Bi2Se3-Bi2S3 system for mid-temperature thermoelectric energy conversion. Energy Environ. Sci. 2013, 6, 552–560. [Google Scholar] [CrossRef]
  14. Le, P.H.; Wu, K.H.; Luo, C.W.; Leu, J. Growth and characterization of topological insulator Bi2Se3 thin films on SrTiO3 using pulsed laser deposition. Thin Solid Films 2013, 534, 659–665. [Google Scholar] [CrossRef]
  15. Morin, S.A.; Forticaux, A.; Bierman, M.J.; Jin, S. Screw Dislocation-Driven Growth of 2D Nanoplates. Nano Lett. 2011, 11, 4449–4455. [Google Scholar] [CrossRef]
  16. Janíček, P.; Drašar, Č.; Beneš, L.; Lošták, P. Thermoelectric properties of Tl-doped Bi2Se3 single crystals. Cryst. Res. Technol. 2009, 44. [Google Scholar] [CrossRef]
  17. Cui, H.; Liu, H.; Li, X.; Wang, J.; Han, F.; Zhang, X.; Boughton, R.I. Synthesis of Bi2Se3 thermoelectric nanosheets and nanotubes through hydrothermal co-reduction method. J. Solid State Chem. 2004, 177, 4001–4006. [Google Scholar] [CrossRef]
  18. Bauer, C.; Veremchuk, I.; Kunze, C.; Benad, A.; Dzhagan, V.M.; Haubold, D.; Pohl, D.; Schierning, G.; Nielsch, K.; Lesnyak, V.; et al. Heterostructured Bismuth Telluride Selenide Nanosheets for Enhanced Thermoelectric Performance. Small Sci. 2021, 1, 2000021. [Google Scholar] [CrossRef]
  19. Zhang, J.; Peng, Z.; Soni, A.; Zhao, Y.; Xiong, Y.; Peng, B.; Wang, J.; Dresselhaus, M.S.; Xiong, Q. Raman spectroscopy of few-quintuple layer topological insulator Bi2Se3 nanoplatelets. Nano Lett. 2011, 11, 2407–2414. [Google Scholar] [CrossRef]
  20. Kadel, K.; Kumari, L.; Li, W.Z.; Huang, J.Y.; Provencio, P.P. Synthesis and Thermoelectric Properties of Bi2Se3 Nanostructures. Nanoscale Res. Lett. 2011, 6, 1–7. [Google Scholar] [CrossRef] [Green Version]
  21. Sun, Y.; Cheng, H.; Gao, S.; Liu, Q.; Sun, Z.; Xiao, C.; Wu, C.; Wei, S.; Xie, Y. Atomically thick bismuth selenide freestanding single layers achieving enhanced thermoelectric energy harvesting. J. Am. Chem. Soc. 2012, 134, 20294–20297. [Google Scholar] [CrossRef]
  22. Pei, Y.; Wang, H.; Snyder, G.J. Band engineering of thermoelectric materials. Adv. Mater. 2012, 24, 6125–6135. [Google Scholar] [CrossRef]
  23. Zhuang, A.; Li, J.J.; Wang, Y.C.; Wen, X.; Lin, Y.; Xiang, B.; Wang, X.; Zeng, J. Screw-dislocation-driven bidirectional spiral growth of Bi2Se3 nanoplates. Angew. Chem. 2014, 53, 6425–6429. [Google Scholar] [CrossRef] [PubMed]
  24. Kim, S.J.; We, J.H.; Cho, B.J. A wearable thermoelectric generator fabricated on a glass fabric. Energy Environ. Sci. 2014, 7, 1959–1965. [Google Scholar] [CrossRef]
  25. Saeed, Y.; Singh, N.; Schwingenschlögl, U. Thickness and strain effects on the thermoelectric transport in nanostructured Bi2Se3. Appl. Phys. Lett. 2014, 104, 2–7. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, F.; Wang, X. Mechanisms in the solution growth of free-standing two-dimensional inorganic nanomaterials. Nanoscale 2014, 6, 6398–6414. [Google Scholar] [CrossRef] [PubMed]
  27. He, W.; Zhang, G.; Zhang, X.; Ji, J.; Li, G.; Zhao, X. Recent development and application of thermoelectric generator and cooler. Appl. Energy 2015, 143, 1–25. [Google Scholar] [CrossRef]
  28. Hong, M.; Chen, Z.G.; Yang, L.; Han, G.; Zou, J. Enhanced Thermoelectric Performance of Ultrathin Bi2Se3 Nanosheets through Thickness Control. Adv. Electron. Mater. 2015, 1, 1–9. [Google Scholar] [CrossRef]
  29. Gayner, C.; Amouyal, Y. Energy Filtering of Charge Carriers: Current Trends, Challenges, and Prospects for Thermoelectric Materials. Adv. Funct. Mater. 2020, 30, 1901789. [Google Scholar] [CrossRef]
  30. Green, A.J.; Dey, S.; An, Y.Q.; O’Brien, B.; O’Mullane, S.; Thiel, B.; Diebold, A.C. Surface oxidation of the topological insulator Bi2Se3. J. Vac. Sci. Technol. A Vacuum Surfaces Films 2016, 34, 061403. [Google Scholar] [CrossRef] [Green Version]
  31. Das, L.; Guleria, A.; Adhikari, S. Aqueous phase one-pot green synthesis of SnSe nanosheets in a protein matrix: Negligible cytotoxicity and room temperature emission in the visible region. RSC Adv. 2015, 5, 61390–61397. [Google Scholar] [CrossRef] [Green Version]
  32. Shen, J.; Chen, Z.; Lin, S.; Zheng, L.; Li, W.; Pei, Y. Single Parabolic Band Behavior of Thermoelectric P-Type CuGaTe2. J. Mater. Chem. C 2015, 4, 209–214. [Google Scholar] [CrossRef]
  33. Luo, Z.Z.; Cai, S.; Hao, S.; Bailey, T.P.; Hu, X.; Hanus, R.; Ma, R.; Tan, G.; Chica, D.G.; Snyder, G.J.; et al. Ultralow Thermal Conductivity and High-Temperature Thermoelectric Performance in n-Type K2.5Bi8.5Se14. Chem. Mater. 2019, 31, 5943–5952. [Google Scholar] [CrossRef]
  34. Samanta, M.; Pal, K.; Pal, P.; Waghmare, U.V.; Biswas, K. Localized Vibrations of Bi Bilayer Leading to Ultralow Lattice Thermal Conductivity and High Thermoelectric Performance in Weak Topological Insulator n-Type BiSe. J. Am. Chem. Soc. 2018, 140, 5866–5872. [Google Scholar] [CrossRef]
  35. Jia, F.; Liu, Y.Y.; Zhang, Y.F.; Shu, X.; Chen, L.; Wu, L.M. Bi8Se7: Delocalized Interlayer π-Bond Interactions Enhancing Carrier Mobility and Thermoelectric Performance near Room Temperature. J. Am. Chem. Soc. 2020, 142, 12536–12543. [Google Scholar] [CrossRef] [PubMed]
  36. Zong, P.A.; Zhang, P.; Yin, S.; Huang, Y.; Wang, Y.; Wan, C. Fabrication and Characterization of a Hybrid Bi2Se3/Organic Superlattice for Thermoelectric Energy Conversion. Adv. Electron. Mater. 2019, 5, 1–8. [Google Scholar] [CrossRef]
  37. Dun, C.; Hewitt, C.A.; Huang, H.; Xu, J.; Zhou, C.; Huang, W.; Cui, Y.; Zhou, W.; Jiang, Q.; Carroll, D.L. Flexible n-type thermoelectric films based on Cu-doped Bi2Se3 nanoplate and Polyvinylidene Fluoride composite with decoupled Seebeck coefficient and electrical conductivity. Nano Energy 2015, 18, 306–314. [Google Scholar] [CrossRef]
Figure 1. (a,b) SEM micrographs of the Bi2Se3 particles. (c) Scheme of the screw dislocation growth mechanism. (d) The XRD pattern of the Bi2Se3 particles. (e) The layered rhombohedral crystal structure of Bi2Se3. (f) A low-resolution TEM image. (g) An HRTEM micrograph of the Bi2Se3 particles and its corresponding power spectrum. From the crystallographic domain, the Bi2Se3 lattice fringe distances were measured to be 0.205 nm, 0.205 nm, and 0.204 nm, at 60.96° and 120.86°, respectively, which could be illuminated as the rhombohedral Bi2Se3 phase, visualized along the [0001] zone axis.
Figure 1. (a,b) SEM micrographs of the Bi2Se3 particles. (c) Scheme of the screw dislocation growth mechanism. (d) The XRD pattern of the Bi2Se3 particles. (e) The layered rhombohedral crystal structure of Bi2Se3. (f) A low-resolution TEM image. (g) An HRTEM micrograph of the Bi2Se3 particles and its corresponding power spectrum. From the crystallographic domain, the Bi2Se3 lattice fringe distances were measured to be 0.205 nm, 0.205 nm, and 0.204 nm, at 60.96° and 120.86°, respectively, which could be illuminated as the rhombohedral Bi2Se3 phase, visualized along the [0001] zone axis.
Nanomaterials 11 01827 g001
Figure 2. (a) Photograph of the consolidated Bi2Se3 cylinder. (b,c) Scheme and images of the samples used to measure the material transport properties in the directions parallel (b) and vertical (c) to the pressure axis.
Figure 2. (a) Photograph of the consolidated Bi2Se3 cylinder. (b,c) Scheme and images of the samples used to measure the material transport properties in the directions parallel (b) and vertical (c) to the pressure axis.
Nanomaterials 11 01827 g002
Figure 3. TE properties of Bi2−xMxSe3 (M = Sn, Cu, Ag, Pb, In) samples measured perpendicular to the press direction: (a) electrical conductivity, σ; (b) Seebeck coefficient, S; (c) power factor, PF.
Figure 3. TE properties of Bi2−xMxSe3 (M = Sn, Cu, Ag, Pb, In) samples measured perpendicular to the press direction: (a) electrical conductivity, σ; (b) Seebeck coefficient, S; (c) power factor, PF.
Nanomaterials 11 01827 g003
Figure 4. Thermoelectric properties of Bi2−xSnxSe3 perpendicular to the press direction: (a) electrical conductivity, σ; (b) Seebeck coefficient, S; (c) power factor, PF.
Figure 4. Thermoelectric properties of Bi2−xSnxSe3 perpendicular to the press direction: (a) electrical conductivity, σ; (b) Seebeck coefficient, S; (c) power factor, PF.
Nanomaterials 11 01827 g004
Figure 5. SEM micrographs of Bi1.93Sn0.07Se3 nanosheets with different magnifications; (a) magnification 5.00 K ×, (b) magnification 10.00 K ×.
Figure 5. SEM micrographs of Bi1.93Sn0.07Se3 nanosheets with different magnifications; (a) magnification 5.00 K ×, (b) magnification 10.00 K ×.
Nanomaterials 11 01827 g005
Figure 6. (a) XRD patterns of Bi2Se3 and Bi1.93Sn0.07Se3; (b) expansion diagram of the area corresponding to the diffraction peak of Bi2Se3 (006).
Figure 6. (a) XRD patterns of Bi2Se3 and Bi1.93Sn0.07Se3; (b) expansion diagram of the area corresponding to the diffraction peak of Bi2Se3 (006).
Nanomaterials 11 01827 g006
Figure 7. (a) An HRTEM micrograph of a Bi1.93Sn0.07Se3 nanosheet, detailed information of the orange squared area and its corresponding power spectrum. Bi2Se3 lattice fringe distances were measured to be 0.370 nm, 0.349 nm, 0.202 nm, and 0.351 nm, at 62.89°, 92.61° and 122.99°, respectively, which could be illuminated as the rhombohedral Bi2Se3 phase, along its [0001] zone axis. (b) EELS chemical composition maps obtained from the red squared area of the STEM micrograph. Individual Bi N2,3-edges at 679 eV (red), Sn M4,5-edges at 485 eV (green), Se M2,3-edges at 162 eV (blue) and composites of Bi-Sn, Bi-Se as well as Bi-Sn-Se. (c) Bi 4f, Se 3d and Sn 3d high-resolution XPS spectra obtained from Bi1.93Sn0.07Se3 nanosheets.
Figure 7. (a) An HRTEM micrograph of a Bi1.93Sn0.07Se3 nanosheet, detailed information of the orange squared area and its corresponding power spectrum. Bi2Se3 lattice fringe distances were measured to be 0.370 nm, 0.349 nm, 0.202 nm, and 0.351 nm, at 62.89°, 92.61° and 122.99°, respectively, which could be illuminated as the rhombohedral Bi2Se3 phase, along its [0001] zone axis. (b) EELS chemical composition maps obtained from the red squared area of the STEM micrograph. Individual Bi N2,3-edges at 679 eV (red), Sn M4,5-edges at 485 eV (green), Se M2,3-edges at 162 eV (blue) and composites of Bi-Sn, Bi-Se as well as Bi-Sn-Se. (c) Bi 4f, Se 3d and Sn 3d high-resolution XPS spectra obtained from Bi1.93Sn0.07Se3 nanosheets.
Nanomaterials 11 01827 g007
Figure 8. (a,b) Cross-section and top-view SEM micrograph of Bi2Se3. (c,d) Cross-section and top-view SEM micrograph of Bi1.93Sn0.07Se3. (e,f) XRD patterns of Bi2Se3 and Bi1.93Sn0.07Se3 measured in two perpendicular directions, along (//) and perpendicular (⊥) to the pressure axis.
Figure 8. (a,b) Cross-section and top-view SEM micrograph of Bi2Se3. (c,d) Cross-section and top-view SEM micrograph of Bi1.93Sn0.07Se3. (e,f) XRD patterns of Bi2Se3 and Bi1.93Sn0.07Se3 measured in two perpendicular directions, along (//) and perpendicular (⊥) to the pressure axis.
Nanomaterials 11 01827 g008
Figure 9. TE properties of Bi2Se3 and Bi1.93Sn0.07Se3 parallel (//) and vertical (⊥) to the pressure axis: (a) electrical conductivity, σ; (b) Seebeck coefficient, S; (c) power factor, PF; (d) total thermal conductivity, κ; (e) lattice thermal conductivity, κL; and (f) the TE figure of merit, ZT.
Figure 9. TE properties of Bi2Se3 and Bi1.93Sn0.07Se3 parallel (//) and vertical (⊥) to the pressure axis: (a) electrical conductivity, σ; (b) Seebeck coefficient, S; (c) power factor, PF; (d) total thermal conductivity, κ; (e) lattice thermal conductivity, κL; and (f) the TE figure of merit, ZT.
Nanomaterials 11 01827 g009
Table 1. Room temperature transport properties and m* of the Bi2Se3 and Bi1.93Se0.07Se3 pellets.
Table 1. Room temperature transport properties and m* of the Bi2Se3 and Bi1.93Se0.07Se3 pellets.
MaterialsnH [1019 cm−3] μH [cm2 V−1 S−1]m*/m0
Bi2Se31.77178.60.27
Bi1.93Sn0.07Se31.34239.30.29
Table 2. A comprehensive summary of the thermoelectric performance of the previously reported Bi2Se3.
Table 2. A comprehensive summary of the thermoelectric performance of the previously reported Bi2Se3.
ZT at Room TemperatureZTmax
Materialsσ (104S/m)S (μV/K)κ (W/mK)ZTσ (104 S/m)S (μV/K)κ (W/mK)ZT
K2.5Bi8.5Se14 [33]0.4−1000.60.051.8−1600.381 (873 K)
Bi0.7Sb0.3Se [34]4.5−1380.50.454.2−1600.70.8 (425 k)
Bi5.6Sb2.4Se7 [35] 4.8−168.50.730.55 (300 K)
Bi2Se3 [28]2−1200.50.171.95−1550.40.48 (427 K)
Bi2Se3 [21]1.78−900.420.13−120.70.490.35 (400 K)
Bi2Se3 HA0.11DMSO0.06 [36]14.8−801.520.187
Bi2Se3 [12]2.5−600.550.052−1000.60.18 (480 K)
Cu0.1Bi2Se3 [37] 1.46−840.320.1 (290 K)
Bi2Se3 [20]0.212−1150.750.0110.6755−1500.830.096 (523 K)
This work5.87−10110.1954.72−1160.910.41 (577 K)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, M.; Zhang, Y.; Zhang, T.; Zuo, Y.; Xiao, K.; Arbiol, J.; Llorca, J.; Liu, Y.; Cabot, A. Enhanced Thermoelectric Performance of n-Type Bi2Se3 Nanosheets through Sn Doping. Nanomaterials 2021, 11, 1827. https://doi.org/10.3390/nano11071827

AMA Style

Li M, Zhang Y, Zhang T, Zuo Y, Xiao K, Arbiol J, Llorca J, Liu Y, Cabot A. Enhanced Thermoelectric Performance of n-Type Bi2Se3 Nanosheets through Sn Doping. Nanomaterials. 2021; 11(7):1827. https://doi.org/10.3390/nano11071827

Chicago/Turabian Style

Li, Mengyao, Yu Zhang, Ting Zhang, Yong Zuo, Ke Xiao, Jordi Arbiol, Jordi Llorca, Yu Liu, and Andreu Cabot. 2021. "Enhanced Thermoelectric Performance of n-Type Bi2Se3 Nanosheets through Sn Doping" Nanomaterials 11, no. 7: 1827. https://doi.org/10.3390/nano11071827

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