Next Article in Journal
In Vitro Comparison of Internal and Marginal Adaptation between Printed and Milled Onlays
Previous Article in Journal
A Simplified Method for the Preparation of Highly Conductive and Flexible Silk Nanofibrils/MXene Membrane
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improved Thermoelectric Performance of Sb2Te3 Nanosheets by Coating Pt Particles in Wide Medium-Temperature Zone

1
Beijing Institute of Technology, School of Optics & Photonics, Beijing Key Laboratory Precise Optoelectronics Measurement Institute, Beijing 100081, China
2
Beijing Institute of Technology, Yangtze Delta Region Academy, Jiaxing 314019, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(21), 6961; https://doi.org/10.3390/ma16216961
Submission received: 8 October 2023 / Revised: 22 October 2023 / Accepted: 26 October 2023 / Published: 30 October 2023

Abstract

:
The p-type Sb2Te3 alloy, a binary compound belonging to the V2VI3-based materials, has been widely used as a commercial material in the room-temperature zone. However, its low thermoelectric performance hinders its application in the low-medium temperature range. In this study, we prepared Sb2Te3 nanosheets coated with nanometer-sized Pt particles using a combination of solvothermal and photo-reduction methods. Our findings demonstrate that despite the adverse effects on certain properties, the addition of Pt particles to Sb2Te3 significantly improves the thermoelectric properties, primarily due to the enhanced electronic conductivity. The optimal ZT value reached 1.67 at 573 K for Sb2Te3 coated with 0.2 wt% Pt particles, and it remained above 1.0 within the temperature range of 333–573 K. These values represent a 47% and 49% increase, respectively, compared to the pure Sb2Te3 matrix. This enhancement in thermoelectric performance can be attributed to the presence of Pt metal particles, which effectively enhance carrier and phonon transport properties. Additionally, we conducted a Density Functional Theory (DFT) study to gain further insights into the underlying mechanisms. The results revealed that Sb2Te3 doped with Pt exhibited a doping level in the band structure, and a sharp rise in the Density of States (DOS) was observed. This sharp rise can be attributed to the presence of Pt atoms, which lead to enhanced electronic conductivity. In conclusion, our findings demonstrate that the incorporation of nanometer-sized Pt particles effectively improves the carrier and phonon transport properties of the Sb2Te3 alloy. This makes it a promising candidate for medium-temperature thermoelectric applications, as evidenced by the significant enhancement in thermoelectric performance achieved in this study.

1. Introduction

Thermoelectric materials have garnered significant attention as a promising and sustainable source of new energy due to their ability to directly convert heat into electricity through the Seebeck effect [1,2,3,4,5,6]. The performance of thermoelectric devices is commonly evaluated using the dimensionless figure of merit (ZT, ZT = S2σT/κ) where S, σ, T, and κ represent the Seebeck coefficient, electrical conductivity, absolute temperature in Kelvin, and total thermal conductivity, respectively [7,8]. The thermoelectric efficiency is directly proportional to the ZT value. A higher ZT value indicates a more efficient conversion of heat into electrical energy, leading to a higher thermoelectric efficiency [9]. The ZT value is a figure of merit that quantifies the thermoelectric efficiency of a material. It is an important parameter in assessing the performance of thermoelectric materials. The ZT value allows researchers and engineers to compare different materials and identify those with the highest thermoelectric efficiency. It serves as a crucial parameter in the development and optimization of thermoelectric materials for applications such as waste heat recovery, power generation, and solid-state cooling [10]. To achieve excellent thermoelectric properties and high efficiency, it is crucial to minimize thermal conductivity (κ) while maximizing the power factor (PF, PF = S2σ) in order to obtain a higher ZT value [11,12]. However, achieving substantial ZT values remains an immense challenge due to the intricate interplay between the Seebeck coefficient, electrical conductivity, carrier concentration, and electronic structure [13,14]. Total thermal conductivity is a key parameter in thermoelectric materials, comprising electronic thermal conductivity (κe) and lattice thermal conductivity (κl). κe relates to heat conduction by charge carriers, while κl pertains to heat conduction through lattice vibrations [15,16]. In thermoelectric materials, the goal is to maximize electrical conductivity and minimize electronic thermal conductivity to retain heat. Additionally, reducing lattice thermal conductivity through structural modifications inhibits heat transfer, maintaining a significant temperature difference. Optimizing total thermal conductivity enhances the ZT and improves energy conversion efficiency for practical applications like waste heat recovery and energy-efficient cooling.
Thermoelectric devices are commonly designed and optimized for operation within the following temperature ranges: low-temperature zone (<200 °C); medium-temperature zone (200–600 °C); high-temperature zone (>600 °C) [4]. In this paper, we demonstrated that Sb2Te3 coated with 0.2 wt% Pt particles can maintain ZT values above 1.0 within the temperature range of 333–573 K (60–300 °C). Some examples of room temperature thermoelectric applications include: energy harvesting (low-power electronic devices, sensors, or wireless communication modules, eliminating the need for batteries or extending their lifetime); climate control (such as in portable refrigeration, cooling seats in vehicles, or small-scale air conditioning for electronic enclosures); thermal management (manage heat dissipation and maintain stable operating temperatures, improving the reliability and performance of electronic components) [17]. Some examples of medium-temperature thermoelectric applications include: waste heat recovery (recover waste heat from high-temperature industrial processes, such as exhaust gases from power plants, furnaces, or industrial manufacturing plants); geothermal energy (use Earth’s heat to generate electricity); solar power generation (convert solar heat into electricity) [18].
V2VI3-type (V = Bi, Sb; VI = Te, Se) thermoelectric materials have gained significant attention due to their highly desirable thermoelectric properties. As a benchmark, both p-type and n-type Bi2Te3-based alloys demonstrate a ZT value of approximately one at room temperature, highlighting their excellent performance in converting heat differentials into electrical energy [19,20,21,22,23,24,25]. However, their application in the medium-temperature zone is restricted by the inherent characteristics of narrow band gap compounds. Various processing techniques have been employed to improve the power factor or reduce thermal conductivity, including nanotechnology (energy barrier filtering [26], minority carrier blocking [27], liquid-phase hot deformation [28], textured engineering [29]) and band engineering (resonant level [30], band convergence [31], electronic density of states distortion [32]).
Zheng et al. [33] demonstrated that by controlling the geometric sizes and introducing nano-sized Au particles, the relative contributions of electrons can be adjusted, leading to enhanced phonon scattering. This approach has been applied to improve the thermoelectric properties of Sb2Te3 nanosheets, resulting in a ZT value of 0.8 at 523 K. Unlike solvothermal methods that form larger metal particles, which greatly influence thermal conductivity, the synthesis of small-sized particles is necessary to primarily modify the electrical transport properties. Furthermore, DFT studies have proven to be effective in analyzing the thermoelectric properties of materials [34,35].
In this study, we aimed to enhance the thermoelectric performance of Sb2Te3 by developing a composite material consisting of Sb2Te3 hexagonal plates coated with Pt nanoparticles. The composite was synthesized using a solvothermal method and a photo-reduction method. By incorporating Pt nanoparticles, we successfully constructed a carrier channel within the material, resulting in a significant improvement in its electrical conductivity. Remarkably, this enhancement in electrical conductivity was achieved while maintaining the thermal conductivity of the material at nearly unchanged levels. As a result, the overall thermoelectric performance of the Sb2Te3-based composite was greatly improved. Specifically, the Sb2Te3 composite, incorporating 0.2 wt% Pt particles, exhibited a maximum electrical conductivity (σmax) of 1.38 × 105 S m−1 at 300 K. Furthermore, the composite demonstrated an impressive maximum figure of merit (ZTmax) value of 1.67 at 573 K. Sb2Te3 coated with 0.2 wt% Pt particles can maintain ZT values above 1.0 within the temperature range of 333–573 K (60–300 °C). This temperature zone covers a 200 K range (low-temperature zone and medium-temperature zone), making it suitable for a broader range of thermoelectric applications, including waste heat recovery and automotive exhaust systems. By operating within this wider temperature zone, thermoelectric devices can effectively harness waste heat and improve energy efficiency in various industrial and automotive settings.

2. Experiment

2.1. Synthesis of Sb2Te3 NCs

Potassium Tellurite (K2TeO3, 99.5%), antimony trichloride (SbCl3, 99.9%), and chloroplatinic acid (H2PtCl6) were purchased from Aladdin Company in Shanghai, China. Sodium hydroxide (NaOH, ≥98%), isopropanol ((CH3)2CHOH, ≥98%), absolute ethanol (CH3CH2OH, ≥98%), diethylene glycol (DEG, HOCH2CH2OCH2CH2OH, 99%), and polyvinylpyrrolidone (PVP, (C6H9NO)n, average mol wt ∼55,000) were purchased from Sinopharm Group. Acetone (CH3COCH3, ≥98%) was obtained from various sources. All chemicals were used without further purification. Firstly, 4 mmol SbCl3, 6 mmol K2TeO3, 10 mmol KOH, and 0.15 g PVP were dissolved at room temperature in 80 mL DEG in a 250 mL three-necked flask for 10 min. Then, under the action of magnetic stirring, slowly heated to 250 degrees Celsius for 6 h, during which DEG was added as a high boiling solvent and reducing agent [36], and PVP was added as a surface modifier to the solvent [37]. The final product Sb2Te3 hexagonal nanosheets were obtained from black solution by high-speed centrifugation. Subsequently, chloroplatinic acid was dissolved in deionized water to form a uniform solution with a mass solubility of 20 mg/mL. Next, the synthesized Sb2Te3 hexagonal tablets were dissolved in deionized water in a 100 mL beaker and then uniformly dispersed using ultrasound. Finally, a uniform solution containing Pt4+ ions was added in proportion to the aforementioned beaker. Among them, Pt was generated by photoreduction method. Under traditional Xe light irradiation, platinum chlorate was reduced to generate Pt, and the specific reaction was as follows:
H 2 PtCl 6 2 H + + PtCl 6 2
PtCl 6 2 + 4 e Pt + 6 Cl

2.2. Synthesis of Sb2Te3 NCs Coated Pt Nanoparticles

Chloroplatinic acid was dissolved in deionized water to form a uniform solution with a mass solubility of 20 mg/mL. The synthesized Sb2Te3 hexagonal sheets were dissolved into 100 mL beaker with deionized water, and then were dispersed uniformly by ultrasound. The dissolution and subsequent dispersion of the sheets in our synthesis process serve two purposes. Firstly, it allows the platinum ions to adhere more effectively to the surface of Sb2Te3, ensuring good interaction between the two materials. This improved adhesion helps to enhance the overall performance of the composite material. Secondly, the dissolution and dispersion process contribute to achieving a more uniform distribution of platinum nanoparticles on the surface of Sb2Te3. This uniformity is important for obtaining consistent and reliable properties throughout the material. By dissolving the sheets and dispersing them, we create conditions that facilitate the homogeneous distribution of platinum nanoparticles, leading to improved conductivity and other desired properties. A homogeneous solution containing Pt4+ ions was added to the above beaker in proportion. Pt nanoparticles were reduced and were attached to the surface of Sb2Te3 hexagonal sheet through traditional Xe light-reduction method under full spectrum conditions simultaneously. Then, the Sb2Te3-coated Pt nanoparticles were washed and collected by high-speed centrifugation. The final product was dried at 60 degrees for 6 h. These fine powders were consolidated to ensure high density under argon atmosphere by spark plasma sintering (SPS) under a pressure of 50 MPa at 523 K for 10 min. The surface carbon paper of sintered sample was removed through sandpaper of different meshes. Subsequently, the sintered bulk was cut into different sizes to meet different measurements.

2.3. Characterization and Thermoelectric Measurements

The phase composition of all nanocomposite specimens were investigated by powder X-ray diffraction analysis (XRD; SmartLab, Rigaku Corporation, Tokyo, Japan, with Cu Kα radiation, λ = 1.5406 Å). The scanning range of 2θ was 10~90 degrees, the scan speed was 6 degrees per minute, and the scan increment was 0.02 degrees. The lattice parameters of all samples were calculated by the Rietveld refinement method. The microstructural properties and morphology were further investigated by field emission scanning electron microscopy (FE-SEM) equipped with the X-ray energy dispersive spectroscopy (EDS) and microstructures were fully characterized by high resolution transmission electron microscopy (HRTEM) (JEOL-F200 instrument, with an acceleration voltage of 200 kV). Electrical conductivity (σ) and Seebeck coefficient (S) were attained simultaneously on a commercial system (LSR-3, Linseis, Bavaria, Germany). The thermal diffusivity coefficient (D) were measured for the disk sample on a laser flash system (Netzsch LFA 457, Netzsch, Selb, Germany). The heat capacity (Cp) was assumed to be the Dulong–Petit limit and to be temperature independent. The sample density (ρ) was determined by the Archimedes method, and then the total thermal conductivity was calculated using the formula of κ = DρCp. Hall coefficient (RH) was determined under a magnetic field of 6000 G by Van der Pauw method and the carrier concentration (nH = 1/(eRH)) of specimens was calculated. Then, the Hall carrier mobility (μH) was calculated by σ = eμHnH. The accuracies in measurements of electrical conductivity and Seebeck coefficient are ±2% and ±6%, respectively. The accuracies in measurements of thermal diffusivity coefficient and sample density are ±5% and ±3%, respectively. Finally, the uncertainty of ZT is around ±10%.

2.4. DFT Calculations

The first-principles density functional theory calculations are implemented in the Quantum ESPRESSO (QE) code [38,39] by using projected augmented wave (PAW) [40] and Perdew–Burke–Ernzerhof (PBE)-type generalized gradient approximation (GGA) [41]. The corresponding pseudopotential files are taken from the SSSP PBEsol Precision v1.3.0 [42]. The G-centered Monkhorst–Pack k-point mesh [43] is considered as 12 × 12 × 1. The spin–orbit coupling (SOC) has been taken into consideration. To obtain stable structures, the energy and force convergence thresholds were set to 10−6 Ry and 10−2 Ry. The electronic transport properties are calculated by the BoltzTraP code [44] based on Boltzmann transport theory. The views of structure and charge density are illustrated by VESTA software (3.5.7) [45].

3. Result and Discussion

Figure 1a,b displays the X-ray diffraction (XRD) patterns of Sb2Te3 and composite specimens containing Sb2Te3 coated with various weight percentages (x wt%) of Pt particles (x = 0, 0.2, 0.4, and 0.6). The main XRD peaks of all specimens correspond well with the standard PDF cards of Sb2Te3 (PDF#71-0393). Notably, a weak peak at the angle of 27°, which corresponds to Te (PDF#86-2269), is conspicuously absent in the X-ray diffraction patterns of all the specimens. This absence indicates that the excess Te, resulting from the deliquescence of antimony trichloride, appears as a second phase, as illustrated in Figure 1a,b. The deliquescence of antimony trichloride resulted in the formation of excessive Te as a second phase. This had an impact on the characterization analysis. Additionally, the uniformity of the Pt coating needs improvement, which may be related to the uniform dispersion of Sb2Te3 in the solution during the photoreduction process.
The XRD patterns of Sb2Te3 coated with Pt demonstrate no additional second phase, indicating that the presence of Pt nanoparticles does not alter the crystal structure of Sb2Te3. To quantitatively analyze the actual Te content, Rietveld refinements for the pure specimen were performed using the Generalized Structural Analysis System (GSAS-II), as depicted in Figure 1c. Concurrently, the lattice parameters a and c were determined for all samples through Rietveld refinements, as shown in Figure 1d. Among all of the specimens, the lattice parameter a of Sb2Te3 coated with 0.2 wt% Pt is higher, while the lattice parameter c is lower. It is important to note that the presence of Te precipitation in the second sample with relatively different Te content may introduce some errors, but it does not affect the overall experimental results.
The XRD analysis confirms that the composite specimens maintain the crystal structure of Sb2Te3, with the excess Te appearing as a second phase due to the deliquescence of antimony trichloride. The presence of Pt nanoparticles does not introduce any additional second phase. The quantitative analysis of Te content and lattice parameters further supports these findings, with the Sb2Te3 coated with 0.2 wt% Pt exhibiting distinctive values for lattice parameters a and c compared to the other specimens.
To gain a comprehensive understanding of the enhanced thermoelectric (TE) properties, a detailed microstructural analysis was conducted using Field Emission Scanning Electron Microscopy (FE-SEM) and High-Resolution Transmission Electron Microscopy (HRTEM) on Sb2Te3 coated with Pt nanoparticles.
FE-SEM observations revealed the morphology of pure Sb2Te3, as shown in Figure 2a, exhibiting a regular hexagonal sheet structure. The SEM micrographs further revealed that the Sb2Te3 nanosheets had dimensions of approximately 8 μm with a thickness of about 500 nm, as depicted in Figure 2d. In order to assess the uniformity, element mapping of Sb and Te was performed using Energy-Dispersive X-ray Spectroscopy (EDS), as shown in Figure 2b,c, confirming the good homogeneity of the material.
Figure 2e displays the morphology of Sb2Te3 coated with 0.2 wt% Pt nanoparticles. The surface of the hexagonal pieces exhibited a transition from a smooth state to a rough state, indicative of the presence of Pt nanoparticles. To obtain detailed information regarding the specific composition of the surface particles, HRTEM analysis was conducted at an acceleration voltage of 200 kV. Under low magnification conditions, the hexagonal sheet structure of the Sb2Te3 nanomaterials remained intact, as shown in Figure 3b. Element mapping of Sb and Te, depicted in Figure 3c,d, respectively, highlighted the homogeneity of the nanosheets. Moreover, element mapping of Pt within the same region confirmed the uniform coating of Pt particles on the surface of the hexagonal sheets, as illustrated in Figure 3e and Table 1.
The microstructural analysis conducted using FE-SEM and HRTEM provided valuable insights into the morphology and composition of Sb2Te3 coated with Pt nanoparticles. The observations confirmed the regular hexagonal sheet structure of Sb2Te3 and demonstrated the uniformity of the material. The presence of rough surfaces and the uniform distribution of Pt nanoparticles were observed, further supporting the enhanced TE properties observed in the composite material. SEM analysis confirmed successful synthesis of Sb2Te3 coated with 0.2 wt% Pt nanoparticles, uniformly adhered to the surface. The presence of Pt atoms enhanced electronic conductivity, promoting charge carrier transport. TEM observations confirmed a hexagonal sheet-like structure and material uniformity. Rough surfaces and a uniform distribution of Pt nanoparticles were observed. Pt presence significantly increased carrier concentration, enhancing conductivity.
The presence of S b T e antisite defects in Sb2Te3 nanosheets, synthesized using the solvothermal method, contributes to the generation of holes according to Equation (3). This leads to the material exhibiting p-type conductivity [46].
Sb 2 Te 3 Sb Te + 2 Te Te × + Sb Sb × + Te ( g ) + h
In Figure 4a, the relationship between the Seebeck coefficient and temperature for Sb2Te3 coated with Pt is presented. The measured Seebeck coefficient (S) of all Sb2Te3-coated Pt nanoparticle pellets demonstrates a consistent p-type behavior across the entire temperature range. This indicates that the addition of Pt particles does not alter the conductivity type of the base material. The values of the Seebeck coefficient increase with temperature until they reach a maximum value at a specific temperature (T ≈ 523 K). Subsequently, the Seebeck coefficient decreases with further increases in temperature for Sb2Te3 coated with 0 wt%, 0.4 wt%, and 0.6 wt% Pt within the measured temperature range. However, for Sb2Te3 coated with 0.2 wt% Pt, the Seebeck coefficient (S) exhibits a monotonic increase with temperature. The incorporation of Pt particles in Sb2Te3 does not change its p-type conductivity. The Seebeck coefficient shows temperature-dependent behavior, with different patterns observed for different Pt concentrations. The monotonic increase in the Seebeck coefficient with temperatures for Sb2Te3 coated with 0.2 wt% Pt suggests a unique thermoelectric response in that specific composition.
Based on the Mott formula [47,48,49,50] for a degenerated semiconductor, the Seebeck coefficient could be written as follows:
S = π 2 k b 2 T 3 e ln ( σ ( E ) ) E E = E F = π 2 k b 2 T 3 e 1 e p ( E ) E + 1 μ μ ( E ) E E = E F
In relaxation time approximation, the relation between carrier mobility, relaxation time, and energy could be written as μ ( E ) = e τ ( E ) m * and τ ( E ) = τ 0 E λ 1 / 2 ( τ 0 is a constant independent of parameter (λ)), then Mott Equation (4) could be expressed as
S = π 2 k b 2 T 3 e 1 e n ( E ) E + λ 1 / 2 E E = E F
Under conditions where acoustic phonon scattering is the major scattering mechanism, the single parabolic (SPB) model [51,52] is used to study the variation in the effective mass and the relation is written as:
S = 8 π 2 k b 2 3 e h 2 m * T ( π 3 n ) 2 / 3
where S is Seebeck coefficient, kb is Boltzmann’s constant, e is the electron charge, h is the Planck constant, m* is the effective mass, T is the temperature in Kelvin, and n is carrier concentration.
In order to investigate the variations in electrical transport properties and effective mass, measurements of carrier concentration, carrier mobility, and effective mass were conducted using the Van der Pauw method [53]. The results are summarized in Table 2. All samples exhibit p-type electrical transport behavior, consistent with the observed Seebeck coefficient. The addition of Pt particles has a notable effect on the carrier mobility (μ) at 300 K. The carrier mobility values decrease from 4.63 × 102 cm2/Vs, 3.63 × 102 cm2/Vs, and 2.79 × 102 cm2/Vs to 2.5 × 102 cm2/Vs with the incorporation of Pt particles. The decrease in carrier mobility can be attributed to the presence of Pt particles, which act as a second phase and impede carrier transfer.
It is worth noting that the carrier concentration of the composite samples initially increases from 0 to 0.2 wt% Pt, reaching a maximum value, and then decreases with a further increase in Pt particle content. The carrier concentrations of all the samples coated with Pt particles are significantly higher than those of the pure samples. The addition of a small amount of Pt particles effectively reduces the potential barrier between the nano-hexagonal sheets, facilitating carrier transfer. Low-energy carriers can pass through the interfacial potential barrier without annihilation, leading to an increase in carrier concentration.
The effective mass (m*) of Sb2Te3 coated with x wt% Pt particles at 300 K was calculated using the measured Seebeck coefficient (S) and carrier concentration (n) based on Equation (6). These values can vary due to the electronic band structure, impurities, and dopants. The presence of Pt atoms can influence effective mass by altering the band structure, carrier concentration, and scattering mechanisms. Dopants interact with electronic orbitals, causing modifications in effective mass values. In undoped materials, effective mass is determined by the intrinsic band structure, but dopants introduce changes. The effective mass values for Sb2Te3 nanosheets coated with x wt% Pt particles are 0.2646 me, 0.4691 me, 0.4162 me, and 0.2875 me, respectively. Notably, Sb2Te3 coated with 0.2 wt% Pt particles exhibits the highest effective mass compared to the other treated samples and the pure sample. This observation is consistent with the increase in electrical conductivity. The measurements of carrier concentration, carrier mobility, and effective mass reveal the impact of Pt particles on the electrical transport properties of Sb2Te3. The addition of Pt particles decreases carrier mobility but increases carrier concentration. The effective mass of the composite samples varies with the Pt particle content, with Sb2Te3 coated with 0.2 wt% Pt particles exhibiting the highest effective mass, aligning with the enhanced electrical conductivity observed in the material.
Figure 4b illustrates the variation in electrical conductivity with temperature for all the measured composite specimens. It is observed that the electrical conductivity decreases monotonically with increasing temperature throughout the measured temperature range, indicating that the specimens behave as degenerate semiconductors. However, the introduction of Pt nanoparticles leads to an increase in electrical conductivity. At 300 K, the electrical conductivity increases from 6.46 × 104 S/m to 1.38 × 105 S/m due to the coating of Pt nanoparticles. This increase can be attributed to the variation in carrier concentration. The Pt particles occupy interstitial sites within the nanosheets, acting as conductive channels and enhancing the carrier concentration and electrical conductivity.
Figure 4c presents the thermal conductivity measurements for all Sb2Te3 samples coated with different concentrations of Pt particles. The excellent thermal transport properties of the composites mainly stem from their textured structure, as depicted in Figure 2f. In comparison to the sample with x = 0 (approximately 0.5346 W/mK at 300 K), the bulk composites containing Pt particles exhibit a slight increase in thermal conductivity. Notably, the sample with x = 0.2 demonstrates the highest κ, ranging from 0.56 W/mK to 0.66 W/mK across the entire temperature range of measurement.
Figure 4b demonstrates the degenerate semiconductor behavior of the composite specimens, with electrical conductivity decreasing with increasing temperature. The coating of Pt nanoparticles contributes to an increase in electrical conductivity by enhancing carrier concentration. Figure 4c illustrates the thermal conductivity measurements, with the textured structure of the composites playing a significant role in their excellent thermal transport properties. The addition of Pt particles leads to a slight increase in thermal conductivity, with the sample coated with 0.2 wt% Pt exhibiting the highest thermal conductivity in the measured temperature range. Figure 4d displays the dimensionless figure of merit of the samples at different coated concentrations (x) as a function of temperature. The ZT value is a key parameter that indicates the effectiveness of a material for thermoelectric applications, considering the synergetic effects of its electrical and thermal properties. In this study, the ZT value exhibits significant improvement.
Remarkably, the ZT value remains above 1.0 within the temperature range of 333 K to 573 K. Specifically, for Sb2Te3 coated with 0.2 wt% Pt particles, the ZT value reaches 1.67 at 573 K, which is 47% higher than the ZT value of the pure Sb2Te3 matrix. This enhancement in ZT is a result of the combined improvements in electrical and thermal properties achieved by incorporating Pt nanoparticles. Furthermore, the average ZT value (ZTave) reaches 1.32 within the measured temperature range of 300 K to 573 K. This notable ZTave value positions the material as a promising candidate for medium-temperature TE applications. Figure 4d demonstrates the substantial enhancement in ZT values achieved by incorporating Pt nanoparticles. The ZT value remains above 1.0 over a wide temperature range, with the sample coated with 0.2 wt% Pt particles exhibiting the highest ZT value of 1.67 at 573 K. The average ZT value of 1.32 further highlights the material’s potential for medium-temperature TE applications.
A small amount of Pt doping has entered the matrix, which prompted us to conduct DFT calculations. Here, we focus on discussing the role of a small amount of Pt doping in the Sb2Te3 structure, rather than a Pt coating. The study involved constructing Sb2Te3 samples using a conventional cell with a 2 × 2 × 1 supercell, with the Pt atom positioned in the middle of the interlayer. The structures are visually depicted in Figure 5a. Following structural relaxation, the lattice constant for the Sb2Te3 conventional cell was determined to be 4.337 Å, while for the Sb2Te3 2 × 2 × 1 supercell with Pt, it was found to be 8.685 Å. These lattice constants closely align with experimental results, lending credibility to the study.
Band structure and corresponding Density of States (DOS) analyses were performed on these structures, as depicted in Figure 5b–e. The examination revealed that the Sb2Te3 conventional cell is a direct semiconductor with a bandgap of 0.15 eV. Both the conduction band minimum and valence band maximum are located at the gamma point. Upon Pt doping, a doping level is observed, with the valence band shifting towards the Fermi energy, indicating p-type doping. Conversely, the conduction band remains largely unaffected. In the DOS plots, a notable sharp increase is observed at around −0.25 eV in Figure 5e. The DOS refers to the number of available energy states per unit volume at a given energy level within the material. When there is a sharp rise in the DOS, it means that there is a higher density of energy states available for electrons to occupy. This can lead to a higher population of charge carriers (holes) in the material, effectively increasing the charge carrier concentration. Overall, the DOS profiles are enhanced due to the presence of Pt, contributing to the improved thermoelectric properties that were observed.
We conducted a theoretical analysis of the electronic transport properties, specifically the Seebeck coefficient and electronic conductivity, as depicted in Figure 6a,b. The carrier concentration for pristine Sb2Te3 was determined to be 8 × 1018 cm−3, while for Sb2Te3 with Pt doping, it increased to 5 × 1019 cm−3. The electronic relaxation time was set at 10−13 s. Notably, the introduction of Pt doping induces the emergence of additional electronic bands, resulting in a substantial improvement in in-plane electronic conductivity while causing a decrease in out-plane electronic conductivity. Regarding the Seebeck coefficient, the values in the in-plane and out-plane directions exhibit close proximity. However, with the addition of Pt, we observed a decrease in the Seebeck coefficient. However, the growth rate of electronic conductivity surpasses that of the Seebeck coefficient, resulting in an overall improvement in the thermoelectric properties. The temperature dependence of these electronic transport properties aligns with experimental observations, reflecting the semiconductor nature of the material. Furthermore, Figure 6c presents a 2D side view of the charge density. It demonstrates that the charge density is localized around the Pt atoms [54]. Moreover, the planar average potential shown in Figure 6d yields similar values throughout, except for regions around the Pt atom. This localization of charge density and the distinct potential values near the Pt atom can be attributed to the enhanced thermoelectric properties observed in the system.

4. Conclusions

In summary, a novel approach combining the solvothermal method with the photo-reduction method was employed to prepare p-type Sb2Te3 nanosheets incorporated with nanometer-sized Pt particles. Then, their thermoelectric transport properties were fully investigated. Those results indicate that nanometer-sized Pt particles can boost electrical conductivity remarkably and slightly improve thermal conductivity. As a result, a Sb2Te3-based composite system incorporated with 0.2 wt% Pt particles contributes a large σmax = 1.38 × 105 S m−1 at 300 K and σave = 8.7 × 104 S m−1. Furthermore, density functional theory (DFT) calculations supported the experimental results by demonstrating that Pt doping induces a sharp increase in the density of states (DOS), with charge density localized at the Pt atoms. This theoretical insight provides a deeper understanding of the underlying mechanisms behind the enhanced electrical conductivity observed in the composite system. Finally, a large ZTmax = 1.67 (at 573 K) and ZTave = 1.32 (in the measured temperature range of 300~573 K) are reached. The incorporation of nanometer-sized Pt particles into p-type Sb2Te3 nanosheets resulted in a substantial enhancement of electrical conductivity and a minor improvement in thermal conductivity. This, in turn, led to the development of a composite material with high electrical conductivity and superior thermoelectric performance. The combination of experimental results and theoretical insights underscores the efficacy of metal particle optimization in achieving enhanced electron and phonon transport properties, ultimately improving the thermoelectric performance of the material.

Author Contributions

Q.-L.G. and L.-Q.D. conceived the study. Q.-L.G. performed the calculations and analyzed the data. Q.-L.G. and Q.H. wrote, reviewed, and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JCJQ Project (Grant No. 2019-JCJQ-ZD-254), National Natural Science Foundation of China (Grant No. U1804261), National Key Research and Development Program of China (Grant No. 2020YFB200190-03, 2020YFB2009304).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data provided in this study could be released upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pei, Y.; Shi, X.; LaLonde, A.; Wang, H.; Chen, L.; Snyder, G.J. Convergence of Electronic Bands for High Performance Bulk Thermoelectrics. Nature 2011, 473, 66–69. [Google Scholar] [CrossRef] [PubMed]
  2. Biswas, K.; He, J.; Blum, I.D.; Wu, C.-I.; Hogan, T.P.; Seidman, D.N.; Dravid, V.P.; Kanatzidis, M.G. High-Performance Bulk Thermoelectrics with All-Scale Hierarchical Architectures. Nature 2012, 489, 414–418. [Google Scholar] [CrossRef]
  3. Tan, G.; Zhao, L.-D.; Kanatzidis, M.G. Rationally Designing High-Performance Bulk Thermoelectric Materials. Chem. Rev. 2016, 116, 12123–12149. [Google Scholar] [CrossRef]
  4. Shi, X.-L.; Zou, J.; Chen, Z.-G. Advanced Thermoelectric Design: From Materials and Structures to Devices. Chem. Rev. 2020, 120, 7399–7515. [Google Scholar] [CrossRef]
  5. Park, J.; Dylla, M.; Xia, Y.; Wood, M.; Snyder, G.J.; Jain, A. When Band Convergence Is Not Beneficial for Thermoelectrics. Nat. Commun. 2021, 12, 3425. [Google Scholar] [CrossRef] [PubMed]
  6. Zheng, Y.; Slade, T.J.; Hu, L.; Tan, X.Y.; Luo, Y.; Luo, Z.-Z.; Xu, J.; Yan, Q.; Kanatzidis, M.G. Defect Engineering in Thermoelectric Materials: What Have We Learned? Chem. Soc. Rev. 2021, 50, 9022–9054. [Google Scholar] [CrossRef]
  7. Zhao, L.-D.; Lo, S.-H.; Zhang, Y.; Sun, H.; Tan, G.; Uher, C.; Wolverton, C.; Dravid, V.P.; Kanatzidis, M.G. Ultralow Thermal Conductivity and High Thermoelectric Figure of Merit in Snse Crystals. Nature 2014, 508, 373–377. [Google Scholar] [CrossRef] [PubMed]
  8. Hinterleitner, B.; Knapp, I.; Poneder, M.; Shi, Y.; Müller, H.; Eguchi, G.; Eisenmenger-Sittner, C.; Stöger-Pollach, M.; Kakefuda, Y.; Kawamoto, N.; et al. Thermoelectric Performance of a Metastable Thin-Film Heusler Alloy. Nature 2019, 576, 85–90. [Google Scholar] [CrossRef]
  9. Snyder, G.J.; Ursell, T.S. Thermoelectric Efficiency and Compatibility. Phys. Rev. Lett. 2003, 91, 148301. [Google Scholar] [CrossRef]
  10. Zoui, M.A.; Bentouba, S.; Stocholm, J.G.; Bourouis, M. A Review on Thermoelectric Generators: Progress and Applications. Energies 2020, 13, 3606. [Google Scholar] [CrossRef]
  11. Song, S.; Mao, J.; Shuai, J.; Zhu, H.; Ren, Z.; Saparamadu, U.; Tang, Z.; Wang, B.; Ren, Z. Study on Anisotropy of N-Type Mg3sb2-Based Thermoelectric Materials. Appl. Phys. Lett. 2018, 112, 092103. [Google Scholar] [CrossRef]
  12. Zhong, Y.; Tang, J.; Liu, H.; Chen, Z.; Lin, L.; Ren, D.; Liu, B.; Ang, R. Optimized Strategies for Advancing N-Type Pbte Thermoelectrics: A Review. ACS Appl. Mater. Interfaces 2020, 12, 49323–49334. [Google Scholar] [CrossRef] [PubMed]
  13. Hu, L.P.; Zhu, T.J.; Yue, X.Q.; Liu, X.H.; Wang, Y.G.; Xu, Z.J.; Zhao, X.B. Enhanced Figure of Merit in Antimony Telluride Thermoelectric Materials by In–Ag Co-Alloying for Mid-Temperature Power Generation. Acta Mater. 2015, 85, 270–278. [Google Scholar] [CrossRef]
  14. Huang, S.; Liu, X.; Zheng, W.; Guo, J.; Xiong, R.; Wang, Z.; Shi, J. Dramatically Improving Thermoelectric Performance of Topological Half-Heusler Compound Luptsb Via Hydrostatic Pressure. J. Mater. Chem. A 2018, 6, 20069–20075. [Google Scholar] [CrossRef]
  15. Wan, C.; Wang, Y.; Wang, N.; Norimatsu, W.; Kusunoki, M.; Koumoto, K. Development of Novel Thermoelectric Materials by Reduction of Lattice Thermal Conductivity. Sci. Technol. Adv. Mater. 2010, 11, 044306. [Google Scholar] [CrossRef] [PubMed]
  16. Eivari, H.A.; Sohbatzadeh, Z.; Mele, P.; Assadi, M.H.N. Low Thermal Conductivity: Fundamentals and Theoretical Aspects in Thermoelectric Applications. Mater. Today Energy 2021, 21, 100744. [Google Scholar] [CrossRef]
  17. Soleimani, Z.; Zoras, S.; Ceranic, B.; Shahzad, S.; Cui, Y. A Review on Recent Developments of Thermoelectric Materials for Room-Temperature Applications. Sustain. Energy Technol. Assess. 2020, 37, 100604. [Google Scholar] [CrossRef]
  18. Dong, J.; Sun, F.-H.; Tang, H.; Pei, J.; Zhuang, H.-L.; Hu, H.-H.; Zhang, B.-P.; Pan, Y.; Li, J.-F. Medium-Temperature Thermoelectric Gete: Vacancy Suppression and Band Structure Engineering Leading to High Performance. Energy Environ. Sci. 2019, 12, 1396–1403. [Google Scholar] [CrossRef]
  19. Hong, M.; Chasapis, T.C.; Chen, Z.-G.; Yang, L.; Kanatzidis, M.G.; Snyder, G.J.; Zou, J. N-Type Bi2Te3−XSeX Nanoplates with Enhanced Thermoelectric Efficiency Driven by Wide-Frequency Phonon Scatterings and Synergistic Carrier Scatterings. ACS Nano 2016, 10, 4719–4727. [Google Scholar] [CrossRef]
  20. Liu, Y.; Zhang, Y.; Lim, K.H.; Ibáñez, M.; Ortega, S.; Li, M.; David, J.; Martí-Sánchez, 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]
  21. Du, B.; Lai, X.; Liu, Q.; Liu, H.; Wu, J.; Liu, J.; Zhang, Z.; Pei, Y.; Zhao, H.; Jian, J. Spark Plasma Sintered Bulk Nanocomposites of Bi2Te2.7Se0.3 Nanoplates Incorporated Ni Nanoparticles with Enhanced Thermoelectric Performance. ACS Appl. Mater. Interfaces 2019, 11, 31816–31823. [Google Scholar] [CrossRef] [PubMed]
  22. Vieira, E.M.F.; Pires, A.L.; Silva, J.P.B.; Magalhães, V.H.; Grilo, J.; Brito, F.P.; Silva, M.F.; Pereira, A.M.; Goncalves, L.M. High-Performance Μ-Thermoelectric Device Based on Bi2Te3/SB2Te3 P–N Junctions. ACS Appl. Mater. Interfaces 2019, 11, 38946–38954. [Google Scholar] [CrossRef]
  23. Hu, L.; Meng, F.; Zhou, Y.; Li, J.; Benton, A.; Li, J.; Liu, F.; Zhang, C.; Xie, H.; He, J. 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]
  24. Meroz, O.; Elkabets, N.; Gelbstein, Y. Enhanced Thermoelectric Properties of N-Type Bi2Te3−XSeX Alloys Following Melt-Spinning. ACS Appl. Energy Mater. 2020, 3, 2090–2095. [Google Scholar] [CrossRef]
  25. Wang, Y.; Liu, W.-D.; Gao, H.; Wang, L.-J.; Li, M.; Shi, X.-L.; Hong, M.; Wang, H.; Zou, J.; Chen, Z.-G. High Porosity in Nanostructured N-Type Bi2Te3 Obtaining Ultralow Lattice Thermal Conductivity. ACS Appl. Mater. Interfaces 2019, 11, 31237–31244. [Google Scholar] [CrossRef]
  26. Liu, S.; Li, H.; Han, W.; Sun, J.; Chen, G.; Chen, J.; Yang, X.; Chen, G.; Ma, F. Realizing the Interface Tuned Thermoelectric Transport Performance in Bi2Te3-Based Hierarchical Nanostructures. J. Phys. Chem. C 2019, 123, 23817–23825. [Google Scholar] [CrossRef]
  27. Bahk, J.-H.; Shakouri, A. Minority Carrier Blocking to Enhance the Thermoelectric Figure of Merit in Narrow-Band-Gap Semiconductors. Phys. Rev. B 2016, 93, 165209. [Google Scholar] [CrossRef]
  28. Wu, Y.; Yu, Y.; Zhang, Q.; Zhu, T.; Zhai, R.; Zhao, X. Liquid-Phase Hot Deformation to Enhance Thermoelectric Performance of N-Type Bismuth-Telluride-Based Solid Solutions. Adv. Sci. 2019, 6, 1901702. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, H.; Luo, G.; Tan, C.; Xiong, C.; Guo, Z.; Yin, Y.; Yu, B.; Xiao, Y.; Hu, H.; Liu, G.; et al. Phonon Engineering for Thermoelectric Enhancement of P-Type Bismuth Telluride by a Hot-Pressing Texture Method. ACS Appl. Mater. Interfaces 2020, 12, 31612–31618. [Google Scholar] [CrossRef] [PubMed]
  30. Pan, Z.; Wang, H. A Descriptive Model of Thermoelectric Transport in a Resonant System of Pbse Doped with Tl. J. Mater. Chem. A 2019, 7, 12859–12868. [Google Scholar] [CrossRef]
  31. Wang, S.; Sun, Y.; Yang, J.; Duan, B.; Wu, L.; Zhang, W.; Yang, J. High Thermoelectric Performance in Te-Free (Bi,Sb)2Se3via Structural Transition Induced Band Convergence and Chemical Bond Softening. Energy Environ. Sci. 2016, 9, 3436–3447. [Google Scholar] [CrossRef]
  32. Heremans, J.P.; Jovovic, V.; Toberer, E.S.; Saramat, A.; Kurosaki, K.; Charoenphakdee, A.; Yamanaka, S.; Snyder, G.J. Enhancement of Thermoelectric Efficiency in Pbte by Distortion of the Electronic Density of States. Science 2008, 321, 554–557. [Google Scholar] [CrossRef] [PubMed]
  33. Zheng, W.; Luo, Y.; Liu, Y.; Shi, J.; Xiong, R.; Wang, Z. Synergistical Tuning Interface Barrier and Phonon Propagation in Au–Sb2Te3 Nanoplate for Boosting Thermoelectric Performance. J. Phys. Chem. Lett. 2019, 10, 4903–4909. [Google Scholar] [CrossRef]
  34. Wang, N.; Li, M.; Xiao, H.; Gao, Z.; Liu, Z.; Zu, X.; Li, S.; Qiao, L. Band Degeneracy Enhanced Thermoelectric Performance in Layered Oxyselenides by First-Principles Calculations. NPJ Comput. Mater. 2021, 7, 18. [Google Scholar] [CrossRef]
  35. Cao, W.; Wang, Z.; Miao, L.; Shi, J.; Xiong, R. Extremely Anisotropic Thermoelectric Properties of Snse under Pressure. Energy Environ. Mater. 2023, 6, e12361. [Google Scholar] [CrossRef]
  36. Gu, Y.; Liu, X.; Huang, S.; Guo, J.; Bi, P.; Shi, J.; Zheng, W.; Wang, Z.; Xiong, R. Porosity Induced Thermoelectric Performance Optimization for Antimony Telluride. Ceram. Int. 2018, 44, 21421–21427. [Google Scholar] [CrossRef]
  37. Zheng, W.; Yang, D.; Wei, W.; Liu, F.; Tang, X.; Shi, J.; Wang, Z.; Xiong, R. Decoupling Electrical and Thermal Properties in Antimony Telluride Based Artificial Multilayer for High Thermoelectric Performance. Appl. Phys. Lett. 2015, 107, 203901. [Google Scholar] [CrossRef]
  38. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. Quantum Espresso: A Modular and Open-Source Software Project for Quantum Simulations of Materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef]
  39. Giannozzi, P.; Andreussi, O.; Brumme, T.; Bunau, O.; Nardelli, M.B.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Cococcioni, M.; et al. Advanced Capabilities for Materials Modelling with Quantum Espresso. J. Phys. Condens. Matter 2017, 29, 465901. [Google Scholar] [CrossRef]
  40. Blöchl, P.E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  41. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  42. Prandini, G.; Marrazzo, A.; Castelli, I.E.; Mounet, N.; Marzari, N. Precision and Efficiency in Solid-State Pseudopotential Calculations. NPJ Comput. Mater. 2018, 4, 72. [Google Scholar] [CrossRef]
  43. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  44. Madsen, G.K.H.; Singh, D.J. BoltzTraP: A Code for Calculating Band-Structure Dependent Quantities. Comput. Phys. Commun. 2006, 175, 67–71. [Google Scholar] [CrossRef]
  45. Momma, K.; Izumi, F. Vesta 3 for Three-Dimensional Visualization of Crystal, Volumetric and Morphology Data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  46. Fang, T.; Li, X.; Hu, C.; Zhang, Q.; Yang, J.; Zhang, W.; Zhao, X.; Singh, D.J.; Zhu, T. Complex Band Structures and Lattice Dynamics of Bi2Te3-Based Compounds and Solid Solutions. Adv. Funct. Mater. 2019, 29, 1900677. [Google Scholar] [CrossRef]
  47. Wang, X.-Y.; Wang, H.-J.; Xiang, B.; Fu, L.-W.; Zhu, H.; Chai, D.; Zhu, B.; Yu, Y.; Gao, N.; Huang, Z.-Y.; et al. Thermoelectric Performance of Sb2Te3-Based Alloys Is Improved by Introducing PN Junctions. ACS Appl. Mater. Interfaces 2018, 10, 23277–23284. [Google Scholar] [CrossRef] [PubMed]
  48. Qin, H.; Zhu, J.; Cui, B.; Xie, L.; Wang, W.; Yin, L.; Qin, D.; Cai, W.; Zhang, Q.; Sui, J. Achieving a High Average zT Value in Sb2Te3-Based Segmented Thermoelectric Materials. ACS Appl. Mater. Interfaces 2020, 12, 945–952. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, J.; Ming, H.; Li, D.; Qin, X.; Zhang, J.; Huang, L.; Song, C.; Wang, L. Ultralow Thermal Conductivity and High Thermoelectric Performance of N-Type Bi2Te2.7Se0.3-Based Composites Incorporated with Gaas Nanoinclusions. ACS Appl. Mater. Interfaces 2020, 12, 37155–37163. [Google Scholar] [CrossRef] [PubMed]
  50. Lee, C.-H.; Dharmaiah, P.; Kim, D.H.; Yoon, D.K.; Kim, T.H.; Song, S.H.; Hong, S.-J. Synergistic Optimization of the Thermoelectric and Mechanical Properties of Large-Size Homogeneous Bi0.5Sb1.5Te3 Bulk Samples Via Carrier Engineering for Efficient Energy Harvesting. ACS Appl. Mater. Interfaces 2022, 14, 10394–10406. [Google Scholar] [CrossRef]
  51. Ma, S.; Li, C.; Wei, P.; Zhu, W.; Nie, X.; Sang, X.; Zhang, Q.; Zhao, W. High-Pressure Synthesis and Excellent Thermoelectric Performance of Ni/BiTeSe Magnetic Nanocomposites. J. Mater. Chem. A 2020, 8, 4816–4826. [Google Scholar] [CrossRef]
  52. Musah, J.-D.; Guo, C.; Novitskii, A.; Serhiienko, I.; Adesina, A.E.; Khovaylo, V.; Wu, C.-M.L.; Zapien, J.A.; Roy, V.A.L. Ultralow Thermal Conductivity in Dual-Doped N-Type Bi2Te3 Material for Enhanced Thermoelectric Properties. Adv. Electron. Mater. 2021, 7, 2000910. [Google Scholar] [CrossRef]
  53. Budak, S.; Guner, S.; Minamisawa, R.A.; Muntele, C.I.; Ila, D. Thermoelectric Properties of Zn4Sb3/CeFe(4−x)CoxSb12 Nano-Layered Superlattices Modified by MeV Si Ion Beam. Appl. Surf. Sci. 2014, 310, 226–229. [Google Scholar] [CrossRef]
  54. Jain, A.; Shin, Y.; Persson, K.A. Computational Predictions of Energy Materials Using Density Functional Theory. Nat. Rev. Mater. 2016, 1, 15004. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of Sb2Te3 powders. (b) XRD patterns of Sb2Te3 bulks after SPS. (c) Refined patterns of Sb2Te3 powders (d) Lattice parameter a and c.
Figure 1. (a) XRD patterns of Sb2Te3 powders. (b) XRD patterns of Sb2Te3 bulks after SPS. (c) Refined patterns of Sb2Te3 powders (d) Lattice parameter a and c.
Materials 16 06961 g001
Figure 2. (a) FESEM micrograph of pure Sb2Te3. (b,c) EDS mappings. (d) Thickness of Sb2Te3 nanosheets. (e) Sb2Te3-coated 0.2 wt% Pt nanoparticles. (f) Fragment of Sb2Te3 bulk.
Figure 2. (a) FESEM micrograph of pure Sb2Te3. (b,c) EDS mappings. (d) Thickness of Sb2Te3 nanosheets. (e) Sb2Te3-coated 0.2 wt% Pt nanoparticles. (f) Fragment of Sb2Te3 bulk.
Materials 16 06961 g002
Figure 3. (a) High-magnification TEM image of Sb2Te3 coated Pt nanoparticles. (b) Low magnification TEM image of Sb2Te3 coated Pt nanoparticles. (ce) Corresponding EDs elemental mapping images of Sb, Te, Pt.
Figure 3. (a) High-magnification TEM image of Sb2Te3 coated Pt nanoparticles. (b) Low magnification TEM image of Sb2Te3 coated Pt nanoparticles. (ce) Corresponding EDs elemental mapping images of Sb, Te, Pt.
Materials 16 06961 g003
Figure 4. (a) Seebeck coefficient vs. Temperature. (b) Electrical conductivity vs. Temperature. (c) Thermal conductivity vs. Temperature. (d) ZT vs. Temperature.
Figure 4. (a) Seebeck coefficient vs. Temperature. (b) Electrical conductivity vs. Temperature. (c) Thermal conductivity vs. Temperature. (d) ZT vs. Temperature.
Materials 16 06961 g004
Figure 5. (a) Structural view of Sb2Te3 and Sb2Te3 2 × 2 × 1 supercell with Pt. Orange, green, and gray balls represent Sb, Te, and Pt atoms. The calculated band structure with SOC for (b) Sb2Te3 and (d) Sb2Te3 2 × 2 × 1 supercell with Pt, respectively. The calculated DOS with SOC for (c) Sb2Te3 and (e) Sb2Te3 2 × 2 × 1 supercell with Pt, respectively.
Figure 5. (a) Structural view of Sb2Te3 and Sb2Te3 2 × 2 × 1 supercell with Pt. Orange, green, and gray balls represent Sb, Te, and Pt atoms. The calculated band structure with SOC for (b) Sb2Te3 and (d) Sb2Te3 2 × 2 × 1 supercell with Pt, respectively. The calculated DOS with SOC for (c) Sb2Te3 and (e) Sb2Te3 2 × 2 × 1 supercell with Pt, respectively.
Materials 16 06961 g005
Figure 6. The calculated (a) Seebeck coefficients and (b) electronic conductivity as a function of temperature. The side view of (c) charge density and (d) planar average potential along c lattice.
Figure 6. The calculated (a) Seebeck coefficients and (b) electronic conductivity as a function of temperature. The side view of (c) charge density and (d) planar average potential along c lattice.
Materials 16 06961 g006
Table 1. The distribution ratio of Sb, Te, Pt elements.
Table 1. The distribution ratio of Sb, Te, Pt elements.
ElementMass (%)Atom (%)
Sb36.3437.74
Te61.2560.70
Pt2.411.56
Table 2. Carrier concentration, carrier mobility, and electrical conductivity at 300 K.
Table 2. Carrier concentration, carrier mobility, and electrical conductivity at 300 K.
SampleP/N TypeCarrier Concentration
(cm−3)
Carrier Mobility
(cm2/V/S)
Electrical Conductivity
(T = 300 K S/m)
ST-1P8.76 × 10184.63 × 1026.46 × 104
ST-2P2.37 × 10193.63 × 10213.8 × 104
ST-3P1.87 × 10192.79 × 1028.9 × 104
ST-4P9.8 × 10182.5 × 1023.93 × 104
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

Guan, Q.-L.; Dong, L.-Q.; Hao, Q. Improved Thermoelectric Performance of Sb2Te3 Nanosheets by Coating Pt Particles in Wide Medium-Temperature Zone. Materials 2023, 16, 6961. https://doi.org/10.3390/ma16216961

AMA Style

Guan Q-L, Dong L-Q, Hao Q. Improved Thermoelectric Performance of Sb2Te3 Nanosheets by Coating Pt Particles in Wide Medium-Temperature Zone. Materials. 2023; 16(21):6961. https://doi.org/10.3390/ma16216961

Chicago/Turabian Style

Guan, Qing-Ling, Li-Quan Dong, and Qun Hao. 2023. "Improved Thermoelectric Performance of Sb2Te3 Nanosheets by Coating Pt Particles in Wide Medium-Temperature Zone" Materials 16, no. 21: 6961. https://doi.org/10.3390/ma16216961

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