Next Article in Journal
Comparative Analysis of Energy Efficiency in High-Voltage Ozone Generators: Resonant Versus Non-Resonant Systems
Previous Article in Journal
Accuracy Performance of Open-Core Inductive Voltage Transformers at Higher Frequencies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress of Thermoelectric Materials—A Review

1
School of Energy and Environment, Southeast University, Nanjing 210096, China
2
Engineering Research Center for Building Energy Environment & Equipment, Ministry of Education, Nanjing 210096, China
*
Author to whom correspondence should be addressed.
Energies 2025, 18(8), 2122; https://doi.org/10.3390/en18082122
Submission received: 13 March 2025 / Revised: 31 March 2025 / Accepted: 15 April 2025 / Published: 21 April 2025
(This article belongs to the Section D1: Advanced Energy Materials)

Abstract

:
Thermoelectric materials are functional materials that directly convert thermal energy into electrical energy or vice versa, and due to their inherent properties, they hold significant potential in the field of energy conversion. In this review, we examine several fundamental strategies aimed at enhancing the conversion efficiency, classification, preparation methods, and applications of thermoelectric materials. First, we introduce an important parameter for evaluating the performance of thermoelectric materials, the dimensionless quality factor ZT, and present the theory of electroacoustic transport in thermoelectric materials, which provides the foundation for enhancing the performance of thermoelectric materials. Second, strategies for optimizing electroacoustic transport properties, carrier concentration, energy band engineering, phonon engineering, and entropy engineering are summarized, emphasizing that energy band engineering presents numerous possibilities for enhancing thermoelectric material performance by tuning the carrier effective mass, energy band convergence, and energy band resonance. By analyzing the importance of various optimization strategies, it is concluded that co-optimization is the primary method for improving the performance of thermoelectric materials in the future. In addition, an overview of the currently available thermoelectric materials is provided, including two categories, classical thermoelectric materials and novel thermoelectric materials, along with a highlight of two thermoelectric material preparation techniques. Finally, the principles of thermoelectric technology are illustrated, its applications in various fields are discussed, problems in the current research are analyzed, and future trends are outlined. Overall, this paper provides a comprehensive summary of optimization strategies, material classifications, and applications, offering valuable references and insights for the researchers in this field, with the aim of further advancing the development of thermoelectric material science.

1. Introduction

In recent years, the growing concern over the energy crisis and environmental pollution has garnered increasing attention. Therefore, finding renewable energy sources or improving the efficiency of traditional fossil fuel usage is one of the effective solutions to the aforementioned challenges. Thermoelectric materials are functional materials that rely on the Seebeck effect and the Peltier effect, as shown in Figure 1, to directly convert thermal energy into electrical energy or vice versa [1,2]. In the Seebeck effect, a temperature gradient is established when P-type and N-type thermoelectric materials are connected in a closed loop with their ends maintained at different temperatures. Under thermal excitation, charge carriers (holes in P-type and electrons in N-type materials) migrate from the high-temperature region to the low-temperature region, resulting in the generation of a potential difference, referred to as the Seebeck voltage. This enables an electric current to circulate within the closed loop, facilitating the direct conversion of thermal energy into electrical energy for thermoelectric power generation. In the Peltier effect, when an external power supply provides a current, it flows through the junctions of the P-type and N-type materials, leading to heat absorption or release. On one side (T − ΔT), the system absorbs heat due to carrier redistribution, resulting in cooling; on the other side (T), heat is released, causing heating. Thermoelectric performance is commonly described by the dimensionless figure of merit (ZT), as shown in Equation (1), where S represents the Seebeck coefficient, denotes the electrical conductivity, K is the thermal conductivity, and T is the absolute temperature. The power factor is given by S2, and the total thermal conductivity (K) consists of electronic thermal conductivity and lattice thermal conductivity. In general, high-performance thermoelectric materials must demonstrate both a high power factor and low lattice thermal conductivity.
Z T = S 2 K T                                              
In the 1950s, driven by rapid advancements in semiconductor physics, the researchers shifted from studying metals with low thermoelectric efficiency to focusing on semi-conductor materials. This led to the development of classic semiconductor thermoelectric materials, such as bismuth telluride (Bi2Te3) and lead antimony (PbTe). However, due to their relatively low performance and high cost, the practical applications of these materials remain limited. In 1998, Slack [3] introduced the concept of “phonon glass-electron crystal” at the Materials Research Society conference in Boston, proposing that high-performance thermoelectric materials should integrate the electronic properties of crystals with the thermal properties of glasses. Based on this concept and the rapid development of technologies and disciplines, such as materials science, manufacturing processes, and materials physics in recent years, thermoelectric materials have made significant progress in terms of the fundamental principles, performance enhancement of traditional materials, and development of new high-performance thermoelectric materials. To date, although the thermoelectric performance of materials has consistently achieved new highs, the strong coupling between electrical and thermal transport properties continues to limit the ZT value, which has not yet reached the ideal range. Therefore, the researchers have proposed a series of strategies based on the microscopic crystal structure of thermoelectric materials, including carrier concentration optimization, band engineering, defect engineering, and phonon engineering, to synergistically enhance the electrical and thermal transport properties, with the goal of producing materials with a high thermoelectric performance.
Based on the inherent properties of thermoelectric materials, thermoelectric devices are widely applied in the fields of waste heat recovery, solar thermoelectric power generation, and solid-state refrigeration. Compared to conventional power generation and refrigeration methods, thermoelectric conversion technologies based on thermoelectric materials provide advantages such as noiseless operation, lack of moving parts, and environmental friendliness [4,5]. Although thermoelectric conversion technologies offer numerous advantages in energy conversion, their efficiency still falls significantly short when compared to traditional technologies. As a result, efforts are needed in areas such as developing high-performance materials, optimizing device design, and selecting appropriate application scenarios.
This paper begins by discussing the performance optimization of thermoelectric materials and reviewing the impact of improving their electrical and thermal transport properties on the overall performance. Next, we summarize the classification and preparation methods of thermoelectric materials, briefly discussing the advantages and disadvantages of various preparation techniques and material types. Finally, we highlight the practical applications of thermoelectric materials and outline their future trends and development directions.

2. Optimization of Thermoelectric Material Performance

In this section, we summarize the key strategies proposed over the past few decades to optimize the performance of thermoelectric materials: carrier concentration optimization [6,7,8], band engineering [9,10,11], phonon engineering [12,13], and entropy engineering [14], as shown in Table 1. These strategies enhance the ZT value by improving the electrical and thermal transport properties of thermoelectric materials. Typically, the carrier concentration in thermoelectric materials directly influences the Seebeck coefficient (S), electrical conductivity (σ), and electronic thermal conductivity (KE), as shown in Figure 2; the relationship among these parameters is described by the following equation [15]:
S = 8 π 2 k B 2 3 e h 2 m d * T π 3 n 2 3
σ = n e μ
K = K E + K L = L n e μ T + K L
where md* is the effective mass of the density of states, n is the carrier concentration, μ is the carrier mobility, KB is the Boltzmann constant, and h is Planck’s constant. The thermal conductivity (K) of the material consists primarily of electronic thermal conductivity and lattice thermal conductivity (KL). As shown in Figure 2, S, n, and σ are interrelated. This coupling indicates that for thermoelectrics with high S, σ, and low K, a compromise in carrier concentration (n) is required, with semiconductor thermoelectrics performing optimally when n is in the range of 1019~1020 cm3. As shown in Equations (3) and (4), both electrical conductivity and thermal conductivity increase with increasing the carrier concentration, which results in a reduction in the material’s performance. The primary reason is that high thermal conductivity facilitates heat transfer from the hot side to the cold side, thereby reducing the temperature gradient and diminishing the Seebeck effect’s effectiveness. Furthermore, according to Equation (2), as the carrier concentration increases, the Seebeck coefficient (S) decreases, consequently impacting the ZT value of the material. Therefore, the coordinated control of carrier concentration represents an effective strategy for optimizing thermoelectric materials.
Table 1. Strategies for enhancing the performance of thermoelectric materials.
Table 1. Strategies for enhancing the performance of thermoelectric materials.
MaterialsStrategiesZTT(K)Ref.
Cu2CoTi3S8carrier concentration0.2670[16]
Cu3Sb0.92Mn0.06Sn0.02Se4carrier concentration0.74673[17]
Ti0.3Zr0.35Hf0.35CoSb1−xSnxcarrier concentration0.8380[18]
dithienothiophenephonon engineering1.48300[19]
CuxBi0.5Sb1.5Te3phonon engineering1.34400[20]
Nb0.8Ti0.2FeSbphonon engineering0.9973[21]
FeNb1−xTixSbband engineering1.11100[22]
α-MgAgSbband engineering2.0575[23]
Al0.04Sn0.96Seband engineering0.84823[24]

2.1. Optimization of Carrier Concentration

Currently, carrier concentration is typically optimized through the doping of impurity atoms, as shown in Figure 3. Figure 3a,b,e depict the trends of carrier concentration and carrier mobility with temperature for different samples. The carrier concentration either increases slightly or remains stable with temperature, which may be related to the doping concentration of the material or the solid solution effect. The mobility remains generally stable with temperature, but the black curve shows an increase with temperature and tends to saturate within a specific temperature range, possibly due to high-quality crystals or special doping. As the concentration of dopant atoms increases, the carrier concentration of the material increases, while the Hall mobility decreases, as shown in Figure 3c. However, in contrast to Figure 3c, in Figure 3f, the Hall mobility of the material increases, suggesting that additional doping introduces a distinct scattering mechanism with differing results. The carrier concentration of the material influences the power factor of the thermoelectric material, as shown in Figure 3d. The typical parabolic shape indicates that the power factor peaks within a specific carrier concentration range, suggesting the presence of an optimal carrier concentration. For example, doping erbium (Er) atoms into GeTe [25] results in an intrinsic carrier concentration of approximately 1021 cm3, which exceeds the optimal carrier concentration range (0.5–2 × 1020 cm3), leading to a lower ZT value. Doping results indicate that the carrier concentration in GeTe decreases with the increasing Er atom content. The primary reason is that in GeTe, holes are the dominant charge carriers, and Er atoms have one more valence electron than Ge atoms. Doping with Er atoms reduces the hole concentration in GeTe, resulting in a decrease in the overall carrier concentration as the Er content increases. Additionally, other experimental results have shown that doping is an effective method for adjusting the carrier concentration in thermoelectric materials [26,27,28,29].
In general, the carrier concentration is proportional to T3/2 and increases with temperature. Pei et al. [32] doped excess Ag atoms into PbTe/AgTe composite materials, and the results indicate that as the temperature increases from 300 K to 750 K, the carrier concentration increases by an order of magnitude. In the composite material, the solubility of Ag atoms increases with temperature, leading to the generation of additional charge carriers. As a result, the carrier concentration in the composite increases.

2.2. Band Engineering

The electrical and thermal transport properties of thermoelectric materials are governed by the material’s band structure. Therefore, modifying the band structure can enhance the material’s performance. Doping with atoms can increase the bandgap of the material’s band structure, raising the carrier transition temperature to higher values, which helps maintain the carrier concentration in the conduction band within the optimal range [33,34], as shown in Figure 4a,b. For a specific thermoelectric material, in addition to adjusting the bandgap width, increasing the degeneracy of the energy bands can also enhance the effective mass of the density of states, thereby improving the material’s Seebeck coefficient, as shown in Figure 4d,e. Previous studies have shown that, without inducing scattering between energy valleys, increasing the degeneracy of the energy bands significantly enhances the thermoelectric performance of the material [35,36,37]. For example, in the PbTe1−xSex alloy (Nv = 6, degeneracy) [38], doping increases the degeneracy of the energy bands to 12. When the temperature reaches 850 K, the material achieves a ZT value of 1.8.
In semiconductor materials, the energy levels of impurity atoms introduced by doping are typically located within the bandgap of the semiconductor. In certain cases, the energy levels of impurity atoms are located within the conduction or valence band of the semiconductor, forming resonance levels. This results in changes in the effective mass of the energy bands, which can, in some cases, enhance the performance of thermoelectric materials [41]. In Figure 4f, the energy band diagram of Mg24Sb15Te1 reveals a possible localized resonance energy level in a region where some energy bands cross or overlap, particularly between the conduction and valence bands. In this region, the energy bands may become denser or exhibit distinct peaks, suggesting that electrons may form resonance states near this energy level. In 2008, Joseph et al. [42] first demonstrated that Tl introduces resonance levels in the valence band of PbTe. Following this, resonance levels were also observed in other materials [43,44,45]. Currently, band engineering is a widely used strategy for enhancing the performance of thermoelectric materials.

2.3. Phonon Engineering

Phonon engineering involves reducing the lattice thermal conductivity of the crystal through phonon scattering, thereby enhancing the ZT value, as shown in Figure 5. Figure 5a illustrates a three-dimensional lattice structure featuring various microscopic defects and structural features, primarily nanoprecipitates, nanopores, and grain boundaries, which significantly affect phonon propagation. The figure illustrates how nanostructures and defects influence phonon propagation within a material and its thermal conductivity, with nanoprecipitates, pores, and grain boundaries scattering phonons at different frequencies, particularly at low- and mid-frequency ranges. This effect can significantly reduce the material’s thermal conductivity and enhance the properties of thermoelectric materials. Figure 5b–e show SEM images of the structural features. As shown in Equations (2)–(4), lattice thermal conductivity is the sole independent physical parameter. Currently, lattice thermal conductivity is primarily reduced by increasing the types and concentrations of defects in crystal. In thermoelectric materials, alloying to introduce point defects is an effective approach for reducing lattice thermal conductivity [46,47]. For example, in Sb2Se3, substituting Bi atoms for Sb atoms and Te atoms for Se atoms creates point defects at the atomic scale in the lattice. The ZT value of the resulting n-type thermoelectric material increased from 0.21 to 0.45 [4].
Dislocations, grain boundaries, phase boundaries, and nano-deposits are additional defects that can significantly reduce the lattice thermal conductivity of thermoelectric materials. These have been incorporated into materials, such as Bi2Te3, Sb2Se3, and Mg3(Sb, Bi)2 [49,50].

3. Classification and Fabrication Techniques of Thermoelectric Materials

Since the discovery of the thermoelectric effect in the 19th century, the research on thermoelectric materials has evolved from metals to semiconductors, with semiconductor thermoelectric materials undergoing rapid development in the early 21st century. Researchers have developed numerous high-performance thermoelectric materials to date, as shown in Table 2. Table 2 lists some of the high-performance thermoelectric materials (ZT > 1) for the last three years. These high-ZT-value thermoelectric materials provide significant advantages, including improved thermal energy conversion efficiency, reduced heat loss, and enhanced thermoelectric properties, thereby expanding their application potential across a wider range of scenarios. Based on their development stages, thermoelectric materials can be broadly classified into classical and novel types.
Table 2. Selected thermoelectric materials with ZT > 1 in the last 3 years.
Table 2. Selected thermoelectric materials with ZT > 1 in the last 3 years.
SamplesT (K)ZTRef.
SnSeS7003.07[51]
(Ca0.85Ba0.15)0.995Na0.005Mg1.85Cd0.15Bi28731.30[52]
Ge0.93Bi0.03Pb0.04Te6702.14[31]
Ge0.93Ti0.01Bi0.06Te0.01Cu6232.30[53]
CoGe2/Ge0.85Sb0.10Te7752.20[54]
(Ge0.89Pb0.08Bi0.03Te)0.97(HgTe)0.036502.30[55]
Bi2(Te,Se)33751.20[56]
AgSbTe26431.70[57]
Cu3SbS47731.30[58]
Na0.99Cd0.995Ag0.005Sb6731.41[59]
CuIn7Se118731.23[60]
PbSnS24731.20[61]
Sn0.71Ge0.2Mn0.07In0.02Te8731.64[62]
EMIM:DCA3303.10[63]
Sn0.78Sb0.16Te(MgB2)0.098501.22[64]
Mg3.2Bi1.998−xSbxTe0.002Cu0.0053481.10[65]
GeSb2Te46731.00[66]
Mg3(Sb,Bi)27731.82[67]
AgMnGePbSbTe57502.64[68]
Nb0.75Ti0.25FeSb9731.21[69]
(Nb, Hf)FeSb9731.47[70]

3.1. Classical Thermoelectric Materials

Classical thermoelectric materials are the most widely used and longest-researched types, primarily including bismuth telluride (Bi2Te3), lead telluride (PbTe), and germanium-silicon (GeSi).
Bismuth telluride materials are among the earliest-studied thermoelectric materials, with the research dating back to the mid-1950s. Due to their excellent thermoelectric performance near room temperature, they have been widely industrialized, particularly in solid-state refrigeration and temperature control applications. Bismuth telluride has a rhombohedral crystal structure, with lattice constants a = b = 0.4385 nm and c = 3.0497 nm. Since the fundamental unit of the bismuth telluride crystal structure consists of a layered arrangement of ionic and covalent bonds, it exhibits relatively low thermal conductivity. Bismuth telluride possesses a complex band structure, and due to the strong orbital coupling effect of Bi atoms, it exhibits a small bandgap and high band degeneracy [71,72]. N-type and P-type thermoelectric materials based on bismuth telluride have been extensively studied, with ZT values typically ranging from 1.0 to 1.6 [73,74], as shown in Figure 6. Recently, Shi et al. [75] proposed a doping method using NaBiS2 to increase the solubility of Na in p-type BST alloys. The BST + 4.0 wt% NaBiS2 sample achieved a peak ZT of 1.44 at 373 K. However, the extremely low abundance of tellurium in the Earth’s crust is one of the main factors limiting the widespread application of bismuth telluride alloys.
Lead telluride compounds are among the earliest-studied thermoelectric materials for the mid-temperature range. In the 1950s, international institutions, primarily NASA, studied its thermoelectric properties and applied it in radioactive isotope thermoelectric power-generation devices, Figure 7. Lead telluride adopts a NaCl-type crystal structure, with lattice constants a = b = c = 0.6446 nm, and is classified as a direct-bandgap semiconductor [95]. Recently, significant progress has been made in the research of lead telluride thermoelectric materials, owing to the advances in material physics theory and improvements in fabrication techniques [96,97]. Currently, the main research direction focuses on adjusting the electrical and thermal transport properties of lead telluride-based thermoelectric materials by combining methods such as element doping, solid solution, and nanocomposites. Maxim et al. [98] reported that nanostructured lead telluride outperforms bulk lead telluride, with a thermoelectric conversion efficiency 10–14% higher and a ZT value of 1.34. The presence of nanostructures increases phonon scattering, thereby reducing thermal conductivity and enhancing conversion efficiency.
Silicon and germanium are elements in the same group, both exhibiting a diamond crystal structure and showing good solubility in one another. The lattice constant, band structure, and other physical properties of germanium–silicon semiconductor materials can be tuned by adjusting the elemental ratio. The research on germanium–silicon thermoelectric materials began in the 1960s. Over the years, the researchers have optimized their thermoelectric performance through doping, nanostructuring, and the fabrication of nanocomposites. Currently, the ZT values of N-type and P-type germanium–silicon thermoelectric materials have increased from their initial values of 1.0 and 0.7, respectively, to 1.5 [100] and 1.3 [101]. Lee et al. [102] proposed that the minimum thermal conductivity of germanium–silicon thermoelectric materials has not yet reached the theoretical limit of the alloy, as scattering effects can only be fully realized when silicon and germanium atoms are randomly distributed in space, which allows the material’s thermal conductivity to achieve its minimum value. Therefore, there is still potential for the further enhancement of the ZT value of germanium–silicon thermoelectric materials.

3.2. Novel Thermoelectric Materials

Since the turn of the 21st century, the research on thermoelectric materials has made remarkable progress, owing to the rapid advancements in science and technology. Not only have the thermoelectric properties of traditional materials greatly improved, but new classes of thermoelectric materials, such as tellurides, half-Heusler alloys, and skutterudite, have also been identified.
Telluride thermoelectric materials are considered promising alternatives to traditional lead telluride thermoelectric materials due to their non-toxicity, environmental friendliness, and alignment with sustainable development principles. These materials include tin telluride (SnTe), germanium telluride (GeTe), and others. Tin telluride and germanium telluride thermoelectric materials, like lead telluride, belong to the IV-VI compound family. Both exhibit a rhombohedral crystal structure at room temperature. In tin telluride, the high intrinsic Sn vacancy concentration results in a higher carrier concentration, leading to a poorer thermoelectric performance. Consequently, the current research primarily focuses on optimizing the material’s performance. Tanveer et al. [103] significantly enhanced the Seebeck coefficient and power factor of tin telluride thermoelectric materials by coordinating the control of band degeneracy and resonance levels. Consequently, the ZT value of the material reached 1.85 at 823 K. Due to its numerous similarities to lead telluride, germanium telluride has been the subject of thermoelectric property research since the 1960s. Owing to the rapid advancements in science and technology, the thermoelectric performance of germanium telluride has improved significantly in recent years. By employing various optimization strategies and methods, thermoelectric materials with a ZT value exceeding 2 have been achieved [93,104]. The existing reports show that the research on P-type germanium telluride is advancing rapidly, and the development of high-performance N-type germanium telluride thermoelectric materials is one of the main research directions for the future. Manisha et al. [105] synthesized N-type germanium telluride via the solid solution method, but its ZT value was limited to 0.6. Therefore, understanding the defects in germanium telluride and applying effective strategies are crucial for developing high-performance N-type germanium telluride.
Skutterudite has attracted significant attention from the researchers due to its “electron crystal–phonon glass” crystal structure, making it a natural high-performance thermoelectric material. For example, CoSb3 has a body-centered cubic crystal structure, with each unit cell containing 8 CoSb3 units, composed of 32 atoms. The unit cell also contains two Sb atoms, creating voids that can be occupied by other elements to form filled Skutterudite materials [106,107]. Due to the unique crystal structure of CoSb3, its thermoelectric performance can be enhanced through methods such as impurity atom doping, solid solution alloys, and nanostructuring. Han et al. [108] synthesized N-type InxCeγCo4Sb12 thermoelectric materials by combining solution spinning with spark plasma sintering (SPS) techniques. The lattice thermal conductivity was reduced to 0.56 W/(m·K), while the ZT value reached 1.45. Xun et al. [109] doped three different atoms into the voids of Skutterudite, thus introducing multiphonon scattering and reducing the material’s thermal conductivity. At 850 K, the ZT value of the material reached 1.7.
Half-Heusler alloys are widely used in high-temperature applications due to their unique structure and properties. They typically have a ZnS-type crystal structure, with transition metals filling the octahedral voids, or a NaCl-type structure, with transition metals occupying the tetrahedral voids. The performance of half-Heusler alloys is influenced by the valence electron count of their constituent elements. When the valence electron count is 8 or 18, the material demonstrates a good thermoelectric performance. In the half-Heusler thermoelectric material system, MNiSn (M = Ti, Zr, Hf) is considered the best N-type material. Although it has a high Seebeck coefficient, its high thermal conductivity limits further development. Alloying, grain refinement, and doping are effective techniques for reducing thermal conductivity [110,111]. RFeSb (R = V, Nb) is the best P-type thermoelectric material, and band engineering improves its thermoelectric performance, providing a foundation for the development of half-Heusler thermoelectric devices [112,113].

3.3. Techniques for Material Preparation

Various methods have been developed for the synthesis and preparation of thermoelectric materials, such as solid-state reaction, solution methods [114,115], high-temperature and high-pressure methods [116], melt spinning [117,118], and chemical vapor deposition. This paper specifically discusses the preparation techniques for bulk materials (solid-state reaction) and flexible materials (vapor deposition).

3.3.1. Solid-State Reaction Technique

The solid-state reaction method involves a chemical reaction between solid materials to form new solid phases. It is one of the most commonly used methods for synthesizing thermoelectric materials. This method generally involves mixing two or more solid raw materials and reacting them at high temperatures to synthesize the desired material. It primarily involves raw material selection and mixing, high-temperature reactions, and cooling. The method offers several advantages, such as not requiring complex equipment and being relatively simple to operate. Chaithanya et al. [119]. synthesized Cu2-xBixSe material using the solid-phase reaction method, and the experimental results demonstrate that, at room temperature, Cu2-xBixSe exhibits a monoclinic crystal structure. The highest power factor of 1474 µW·m1·K2 was achieved at 700 K for the Cu1.988Bi0.012Se sample, demonstrating great potential for the development of high-performance Cu2Se-based thermoelectric materials. Qin et al. [120] synthesized SbxFeTe2 (x = 0.05, 0.1, 0.3, 0.5, 0.7, 0.9) using the high-temperature solid-state method. The study showed that excess Sb led to the formation of an Sb2Te3 secondary phase, resulting in a multiphase composite structure consisting of FeTe2 and Sb2Te3. These structures significantly reduced the resistivity of the sample. At 873 K, the sample with 0.9 Sb content exhibited a resistivity of 0.86 mΩ·cm. Additionally, at 473 K, the ZT value of the Sb0.7FeTe2 sample was 0.25. Bai et al. [104] employed a vacuum solid-state reaction method to synthesize the compound XTe2 (x = Fe, Co, Ni), which exhibits abundant micropores and a complex crystal structure. In this material, FeTe2 exhibited a maximum Seebeck coefficient of 47.24 μV/K at 305 K. NiTe2 exhibited the lowest resistivity of 1.45 mΩ·cm at 567 K. The CoTe2 sample achieved a maximum power factor of 334.05 μW·m1·K2 at 570 K. The thermal conductivity of the pyrite samples followed the trend: NiTe2 > CoTe2 > FeTe2, with FeTe2 exhibiting the lowest thermal conductivity of 2.29 W/(m·K) at 599 K.
Bulk materials synthesized by the solid-state method typically exhibit a high Seebeck coefficient and good electrical properties, but their mechanical processability and deformability are limited. These materials are inherently brittle and prone to cracking under tensile, compressive, or bending stresses, which makes fabricating flexible thermoelectric devices challenging. Flexible materials prepared by vapor deposition typically exhibit good mechanical properties and deformability, offering advantages such as flexibility, light weight, and environmental friendliness. Therefore, vapor deposition plays a key role in the development of thermoelectric materials.

3.3.2. Vapor Deposition Technique

The vapor deposition method involves the reaction of vapor-phase materials at the gas–solid interface to form thin film materials. It is primarily classified into chemical vapor deposition (CVD) and physical vapor deposition (PVD). This method produces high-purity material films with uniform thickness, and the strong adhesion between the film and the substrate ensures stability in diverse applications. Jin et al. [121] designed and synthesized SnSe single crystals using the vertical vapor deposition method. The crystals were found to have a stoichiometric ratio of 1:1 and a space group of Pnma at room temperature. The characterization results demonstrated that SnSe single crystals exhibited an electrical conductivity σ = 39.6 S cm−1 and a Seebeck coefficient of S = 566 μVK−1 at around 580 K. By calculating the figure of merit, ZT has the largest value of ∼1.0 at around 800 K. Chuai et al. [122] prepared bismuth telluride thermoelectric thin film materials using plasma-enhanced chemical vapor deposition. The experimental results demonstrated that the material exhibited a 90% increase in transparency in the mid-infrared region and room-temperature mobility of 2094 cm2/V·s. Furthermore, the thermoelectric thin film material exhibited a decrease in room temperature conductivity and a significant increase in the Seebeck coefficient with increasing selenium content.
Although flexible thermoelectric materials prepared by vapor deposition offer many advantages, their thermoelectric efficiency is lower than that of bulk materials synthesized by the solid-state reaction method, due to the effects of their structure and properties. This limits their broader application.

4. Principles and Applications of Thermoelectric Conversion Technologies

4.1. Thermoelectric Power Generation

The principle of thermoelectric power generation is based on the direct conversion of the thermoelectric effect, which primarily relies on the temperature difference to generate voltage in the thermoelectric material. The charge carriers within the material migrate from the hot side to the cold side, generating a potential difference that becomes the output voltage [123,124], as shown in Figure 8. A thermoelectric generator mainly consists of thermoelectric materials, heat sources, and coolers. Thermoelectric power generation technology holds broad application potential, including waste heat recovery [125], space exploration [126,127], and portable power sources [128,129], among others. For thermoelectric devices, the power generation efficiency is defined as the ratio of the output power to the heat absorbed at the hot end, as shown in Equation (6). The refrigeration efficiency is defined as the ratio of heat absorbed to input electric power, with the maximum coefficient of performance (COPmax) shown in Equations (5) and (6) [130].
= T h T c T h 1 + z T 1 1 + z T + T h T c
C O P m a x = T c T h T c 1 + z T T h T c 1 + z T + 1
where Th is the hot-end temperature, Tc is the cold-end temperature, and ZT is the average dimensionless thermoelectric figure of merit.

4.2. Thermoelectric Refrigeration

Thermoelectric refrigeration is based on the Peltier effect, which describes the relationship between an electric current and the temperature change in the material. When an electric current flows through the junction of two conductors or semiconductors, heat is either absorbed or emitted. This means that at one contact point, heat is absorbed, causing a temperature drop, while at the other contact point, heat is emitted, causing a temperature rise [132,133]. A thermoelectric cooling system typically consists of four components: thermoelectric elements, a power supply, heat exchangers, and cooling surfaces. In contrast to traditional cooling methods, thermoelectric refrigeration offers advantages such as the absence of moving parts, environmental friendliness, and precise control. Thermoelectric cooling has broad applications in various fields, including electronic device cooling, portable refrigeration, and automobiles and other transportation vehicles.

4.3. Applications

System integration is a critical aspect of applying thermoelectric technology, involving the efficient combination of thermoelectric modules with other system components. Proper system integration can significantly improve the overall efficiency and reliability of thermoelectric materials. Selecting an appropriate heat source is a critical aspect of system integration. The ideal heat source should maintain a stable temperature gradient to ensure the efficient operation of the thermoelectric module. The design of the thermoelectric module is critical for ensuring the system operates efficiently. The design process includes material selection [134,135], geometric design [136], thermal interface materials [137,138], and heat dissipation design. In some systems, phase change materials (PCMs) can be integrated with thermoelectric modules. This integration stores excess thermal energy during peak load periods and releases it when demand increases, thereby enhancing the energy efficiency and stability of system [139].

4.3.1. Waste Heat Recovery

Studies have reported on converting industrial waste heat, geothermal energy, and waste heat from civilian sectors into electrical energy based on thermoelectric power generation principles, as shown in Figure 9. In the 1990s, Matsuura et al. [140] studied a project that generated megawatt-scale electrical power from industrial waste heat. The project used 98 °C circulating water as the heat source and 25 °C cooling water as the cold reservoir, achieving a power generation efficiency in the range of 8–10%. In 2017, the researchers from the Naval Engineering University of China developed an industrial wastewater heat recovery device, achieving a recovery efficiency of 1.28% and an investment payback period of approximately 8 years [141]. In the civilian sector, for example, waste heat from waste incineration, Japanese scholars developed kilowatt-scale waste heat power generation systems, resulting in the effective utilization of waste.
Solar energy and other natural heat sources are renewable and environmentally friendly sources of energy. Harnessing these natural heat sources for power generation to promote the transformation of energy utilization is in line with the principles of sustainable development. Among these, solar thermal power generation is the earliest-studied method. In 1954, Maria [142] studied the energy conversion efficiency of a solar thermoelectric generator using 25 pairs of ZnSb thermocouples, achieving an efficiency of 3.35%. Since then, the research on solar thermal power generation has gained increasing attention from scholars. After years of development, the research on solar thermal power generation now primarily focuses on integrating it with solar photovoltaic power generation to create a photovoltaic–thermoelectric hybrid power system [143,144]. Although the photovoltaic–thermoelectric hybrid system is an ideal energy conversion technology, it faces several key challenges that hinder its development, such as temperature matching between photovoltaic and thermoelectric devices, heat transfer at the interface, device selection, and energy conversion under non-steady-state conditions. Ge et al. [145] investigated the impact of TEG structural parameters on the efficiency of hybrid systems through mathematical modeling of PV–thermoelectric hybrid systems. The results showed that an optimal structural parameter (the ratio of the cross-sectional ratio to the height) with a value of 0.36 mm−1 existed for the TEG, enabling the system efficiency to reach a maximum of 41.73% under simulated conditions.
Figure 9. (a) Assembling of the Te-free TE module; (b) performance of the Te-free segmented modules [146]. Schematic of an AEATEG with the counter-current flow: (c) overall structure, (d) sectional drawing in the axial direction, (e) sectional drawing in the radial direction, and (f) the structure of a single ATEG [147].
Figure 9. (a) Assembling of the Te-free TE module; (b) performance of the Te-free segmented modules [146]. Schematic of an AEATEG with the counter-current flow: (c) overall structure, (d) sectional drawing in the axial direction, (e) sectional drawing in the radial direction, and (f) the structure of a single ATEG [147].
Energies 18 02122 g009

4.3.2. Solid-State Refrigeration

Solid-state refrigeration is a cooling method that relies on thermoelectric materials and thermoelectric conversion technology, as shown in Figure 10. Compared to traditional compression-based systems, solid-state refrigeration technology offers numerous distinct advantages and application potential. At the core of solid-state refrigeration technology lies high-performance thermoelectric materials. Bi2Te3 is one of the most extensively studied thermoelectric refrigeration materials. Thermoelectric refrigeration devices and systems based on Bi2Te3 have been commercialized, including thermoelectric air conditioners and refrigerators. In the future, developing new, efficient thermoelectric materials to improve the efficiency and performance of solid-state refrigeration will be one of the primary research directions. Conversely, combining solid-state refrigeration technology with other cooling technologies could result in more efficient and flexible refrigeration solutions.
In recent years, with the advent of new preparation technologies and optimization strategies, the thermoelectric performance of Bi2Te3-based refrigeration materials has consequently improved. Wu et al. [150] incorporated Bi2Fe4O9 magnetic nanoparticles into Bi2Te3-based thermoelectric refrigeration materials, where these nanoparticles effectively scatter low-energy charge carriers, resulting in enhanced thermoelectric properties. At 393 K, the thermoelectric figure of merit of the sample reached 1.1, representing a 13% improvement compared to its unmodified state. Liu et al. [151] optimized the thermoelectric properties at room temperature by constructing multiscale interfacial phases in LaFeSi (LFS)/BiSbT(BST) thermo-electromagnetically cooled materials. The results indicate that, at the atomic scale, the interfacial reaction between LFS and BST leads to the formation of (Fe,Co)(Sb,Te)2 micrograins and LaTe2 nanograins, which create a low-mismatch phase boundary with the LFS matrix. Compared to other thermoelectric materials, a favorable trade-off TE performance is achieved in LFS/BST composites, simultaneously demonstrating a high-TE performance. The 20% LFS/BST composite exhibits a room-temperature ZT of 0.46, along with a large maximum magnetic entropy change and relative cooling power of 0.81 J kg−1 K−1 and 44.83 J kg−1, respectively.

5. Conclusions and Challenges

From the above discussion, it can be concluded that the current priority in the field of thermoelectric materials remains the development of thermoelectric materials with high ZT values to overcome low energy-conversion efficiencies, thereby establishing thermoelectric technology as a promising approach for energy utilization with extensive applications. The main strategies for enhancing the ZT value of thermoelectric materials include optimizing the carrier concentration, energy band engineering, phonon engineering, and entropy engineering, through which the electron and phonon transport properties of thermoelectric materials are optimized to achieve an enhanced thermoelectric performance. In the past decade, the ZT values of conventional thermoelectric materials have not only been increased to approximately 1.5, but new thermoelectric materials with ZT values exceeding 2 have also been developed.
Although there has been a rapid development of thermoelectric materials, the current research in this field continues to face several challenges, which not only limit their practical applications but also hinder the advancement of thermoelectric technology. The primary challenges include:
  • Lower conversion efficiency. The conversion efficiency of thermoelectric materials is a critical issue, and the relatively low efficiency exhibited by current thermoelectric materials significantly restricts their broader application. Currently, although some thermoelectric materials exhibit an excellent performance within specific temperature ranges, the overall conversion efficiency has not yet reached a practical level.
  • Poor high-temperature performance. Thermoelectric materials for high-temperature use typically have larger bandgaps, which complicates the balance between electrical conductivity and thermal conductivity. Most existing materials exhibit a poor performance in high-temperature environments, limiting their application in high-temperature conditions. However, a wider bandgap also results in a more complicated coupling effect between electrical transport properties (electrical conductivity and carrier mobility) and thermal transport properties (electronic and lattice thermal conductivities). This interaction makes the optimization of material properties more challenging; furthermore, most existing materials exhibit an inadequate performance at elevated temperatures due to insufficient thermal stability, decreased carrier mobility, and degradation of lattice structures, thus restricting their practical applications.
  • The challenge of multicriteria optimization. The parameters governing the performance of thermoelectric materials are interdependent, requiring that multiple criteria (thermal conductivity, electrical conductivity, and Seebeck coefficient) be simultaneously satisfied during the design process. Optimizing material properties while simultaneously satisfying these criteria remains a complex theoretical and practical challenge, and the existing theoretical models and experimental approaches have not yet provided comprehensive solutions.
  • Difficulty in converting technology. Although progress has been made in the basic research, translating these findings into practical applications remains challenging. Many promising thermoelectric materials perform well under laboratory conditions, but ensuring their consistency and repeatability in industrial production remains a challenge.
Although the research on thermoelectric materials is confronted with numerous challenges, advances may be achieved in overcoming these issues as new theoretical approaches, materials, and technologies continue to emerge. In particular, significant breakthroughs may be achieved through advancements in the multidisciplinary research, automated material design, and computational materials science. In summary, the development of high-performance thermoelectric materials is confronted with numerous significant challenges; however, with continued research and innovation, these materials are anticipated to achieve greater efficiency and broader applications in the future.

Author Contributions

J.W.: Conceptualization, Writing—original draft, Writing—review and editing, Formal analysis, and Methodology. Y.Y.: Supervision and Funding acquisition. C.C.: Writing—review and editing. M.C.: Writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 52476002).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yang, L.; Chen, Z.; Dargusch, M.S.; Zou, J. High performance thermoelectric materials: Progress and their applications. Adv. Energy Mater. 2018, 8, 1701797. [Google Scholar] [CrossRef]
  2. Deng, T.; Qiu, P.; Yin, T.; Li, Z.; Yang, J.; Wei, T.; Shi, X. High-Throughput Strategies in the Discovery of Thermoelectric Materials. Adv. Mater. 2024, 36, e2311278. [Google Scholar] [CrossRef]
  3. Tritt, T.M. Holey and unholey semiconductors. Science 1999, 283, 804–805. [Google Scholar] [CrossRef]
  4. Tian, Z.; Jiang, Q.; Li, J.; Kang, H.; Chen, Z.; Guo, E.; Wang, T. Achieving n-and p-type thermoelectric materials with the identical chemical composition BiSbTe1.5Se1.5 by defect structure engineering. Chem. Eng. J. 2024, 494, 152954. [Google Scholar] [CrossRef]
  5. Ivanov, A.A.; Kaplar, E.P.; Prilepo, Y.P.; Murav’ev, V.V.; Ustinov, V.S. Progress in the Research on Promising High-Performance Thermoelectric Materials. Nanobiotechnol. Rep. 2021, 16, 268–281. [Google Scholar] [CrossRef]
  6. Hou, Z.H.; Qian, X.; Cui, Q.J.; Wang, S.F.; Zhao, L.D. Strategies to advance thermoelectric performance of PbSe and PbS materials. Rare Met. 2024, 43, 4099–4114. [Google Scholar] [CrossRef]
  7. Zhang, D.; Zhong, R.; Gao, S.; Yang, L.; Xu, F.; He, P.; Liu, G.; San, X.; Yang, J.; Luo, Y.; et al. Reinforcing bond covalency for high thermoelectric performance in Cu3SbSe4-based thermoelectric material. Sci. China Mater. 2023, 66, 3644–3650. [Google Scholar] [CrossRef]
  8. Hu, L.; Zhang, Q.; Shan, Z.; Wang, L.; Zheng, Y.; Fan, J. Synergy of grain size and texture effect for high-performance Mg3Sb2-based thermoelectric materials. Scr. Mater. 2023, 235, 115629. [Google Scholar] [CrossRef]
  9. Jiang, Q.; Li, G.; Wang, X.; Kang, H.; Chen, Z.; Guo, E.; Wang, T. Enhanced thermoelectric properties for eco-friendly CaTiO3 by band sharpening and atomic-scale defect phonon scattering. Mater. Today Energy 2024, 44, 101655. [Google Scholar] [CrossRef]
  10. Anita; Gupta, V.; Pandey, A. Improvement in thermoelectric properties of Bi-Mg co-doped SnTe via band engineering and nanostructuring. J. Mater. Sci. Mater. Electron. 2024, 35, 1–16. [Google Scholar] [CrossRef]
  11. Jiang, Q.; Long, H.; Zeng, X.; Wang, B.; Yang, B.; Yi, J.; Luo, Y.; Yang, J.; Ye, H.; Liu, Y. A simple in-situ PZT oxide’s decomposition: Realizing synergistic tailoring of electrical and thermal transport properties of BiCuSeO thermoelectric ceramics through band and phonon engineering. Ceram. Int. 2024, 50, 35985–35992. [Google Scholar] [CrossRef]
  12. Vijay, V.; Harish, S.; Archana, J.; Navaneethan, M. Synergistic effect of grain boundaries and phonon engineering in Sb substituted Bi2Se3 nanostructures for thermoelectric applications. J. Colloid Interface Sci. 2022, 612, 97–110. [Google Scholar] [CrossRef] [PubMed]
  13. Li, J.W.; Liu, W.; Xu, W.; Zhuang, H.L.; Han, Z.; Jiang, F.; Zhang, P.; Hu, H.; Gao, H.; Jiang, Y.; et al. Bi-deficiency leading to high-performance in Mg3(Sb, Bi)2-based thermoelectric materials. Adv. Mater. 2023, 35, 2209119. [Google Scholar] [CrossRef] [PubMed]
  14. Tang, Q.; Jiang, B.; Wang, K.; Wang, W.; Jia, B.; Ding, T.; Huang, Z.; Lin, Y.; He, J. High-entropy thermoelectric materials. Joule 2024, 8, 1641–1666. [Google Scholar] [CrossRef]
  15. Lu, N.; Li, L.; Liu, M. A review of carrier thermoelectric-transport theory in organic semiconductors. Phys. Chem. Chem. Phys. 2016, 18, 19503–19525. [Google Scholar] [CrossRef]
  16. Hashikuni, K.; Suekuni, K.; Watanabe, K.; Bouyrie, Y.; Ohta, M.; Ohtaki, M.; Takabatake, T. Carrier concentration tuning in thermoelectric thiospinel Cu2CoTi3S8 by oxidative extraction of copper. J. Solid State Chem. 2018, 259, 5–10. [Google Scholar] [CrossRef]
  17. Wei, S.; Yu, L.; Zhang, Z.; Ji, Z.; Luo, S.; Liang, J.; Song, W.; Zheng, S. Enhancing the effective mass and covalent bond strength of Cu3SbSe4-based thermoelectric materials by Mn/Sn co-doping. Mater. Today Phys. 2023, 38, 101260. [Google Scholar] [CrossRef]
  18. Rausch, E.; Balke, B.; Deschauer, T.; Ouardi, S.; Felser, C. Charge carrier concentration optimization of thermoelectric p-type half-Heusler compounds. APL Mater. 2015, 3, 041516. [Google Scholar] [CrossRef]
  19. Mi, X.Y.; Yu, X.; Yao, K.L.; Huang, X.; Yang, N.; Lu, J.T. Enhancing the thermoelectric figure of merit by low-dimensional electrical transport in phonon-glass crystals. Nano Lett. 2015, 15, 5229–5234. [Google Scholar] [CrossRef]
  20. Yoon, J.S.; Song, J.M.; Rahman, J.U.; Lee, S.; Seo, W.S.; Lee, K.H.; Kim, S.; Kim, H.S.; Kim, S.I.; Shin, W.H. High thermoelectric performance of melt-spun CuxBi0.5Sb1.5Te3 by synergetic effect of carrier tuning and phonon engineering. Acta Mater. 2018, 158, 289–296. [Google Scholar] [CrossRef]
  21. Tan, C.; Wang, H.; Yao, J.; Chen, T.; Wang, L.; Sun, Y.; Khan, M.; Wang, H.; Wang, C. Synchronously enhanced thermoelectric and mechanical properties of Ti doped NbFeSb based half-heusler alloys by carrier regulation and phonon engineering. J. Eur. Ceram. Soc. 2022, 42, 7010–7016. [Google Scholar] [CrossRef]
  22. Fu, C.; Zhu, T.; Liu, Y.; Xie, H.; Zhao, X. Band engineering of high performance p-type FeNbSb based half-Heusler thermoelectric materials for figure of merit zT > 1. Energy Environ. Sci. 2015, 8, 216–220. [Google Scholar] [CrossRef]
  23. Tan, X.; Wang, L.; Shao, H.; Yue, S.; Xu, J.; Liu, G.; Jiang, H.; Jiang, J. Improving Thermoelectric Performance of α-MgAgSb by Theoretical Band Engineering Design. Adv. Energy Mater. 2017, 7, 1700076. [Google Scholar] [CrossRef]
  24. Xin, N.; Li, Y.; Shen, H.; Shen, L.; Tang, G. Realizing high thermoelectric performance in hot-pressed polycrystalline AlxSn1-xSe through band engineering tuning. J. Mater. 2022, 8, 475–488. [Google Scholar] [CrossRef]
  25. Li, X.; Liu, M.; Guo, M.; Niu, C.; He, H.; Liu, Z.; Zhu, Y.; Dong, X.; Cai, W.; Guo, F.; et al. Tailoring band structure and Ge precipitates through Er and Sb/Bi co-doping to realize high thermoelectric performance in GeTe. Chem. Eng. J. 2023, 474, 145820. [Google Scholar] [CrossRef]
  26. Kim, S.; Kihoi, S.K.; Kim, H.; Kahiu, J.N.; Lee, H.S. Synergetic effect of Bi and Al co-doping in GeTe-based thermoelectric materials leading to optimized carrier concentration tuning and high ZT. J. Alloys Compd. 2024, 970, 172574. [Google Scholar] [CrossRef]
  27. Cui, S.; Wang, C.; Hao, M.; Huang, X.; Wang, C.; Wang, X.; Wang, Y. Enhancing thermoelectric performance of p-type BixSb2−xTe3 by optimizing carrier concentration. J. Eur. Ceram. Soc. 2025, 45, 116942. [Google Scholar] [CrossRef]
  28. Dharmaiah, P.; Heo, M.; Nagarjuna, C.; Jung, S.-J.; Won, S.O.; Lee, K.H.; Kim, S.K.; Kim, J.-S.; Ahn, B.; Kim, H.-S.; et al. Enhancement of thermoelectric properties in p-type ZnSb alloys through Cu-doping. J. Alloys Compd. 2024, 1004, 175739. [Google Scholar] [CrossRef]
  29. Ren, Y.; Jiang, Q.; Yang, J.; Luo, Y.; Zhang, D.; Cheng, Y.; Zhou, Z. Enhanced thermoelectric performance of MnTe via Cu doping with optimized carrier concentration. J. Mater. 2016, 2, 172–178. [Google Scholar] [CrossRef]
  30. Liu, R.; Lan, J.-L.; Tan, X.; Liu, Y.-C.; Ren, G.-K.; Liu, C.; Zhou, Z.-F.; Nan, C.-W.; Lin, Y.-H. Carrier concentration optimization for thermoelectric performance enhancement in n-type Bi2O2Se. J. Eur. Ceram. Soc. 2018, 38, 2742–2746. [Google Scholar] [CrossRef]
  31. Liu, M.; Guo, M.; Zhu, J.; Zeng, X.; Chen, H.; Yuan, D.; Zhang, Q.; Cai, F.; Guo, F.; Zhu, Y.; et al. High-Performance CaMg2Bi2-Based Thermoelectric Materials Driven by Lattice Softening and Orbital Alignment via Cadmium Doping. Adv. Funct. Mater. 2024, 34, 2316075. [Google Scholar] [CrossRef]
  32. Pei, Y.; May, A.F.; Snyder, G.J. Self-Tuning the carrier concentration of PbTe/Ag2Te composites with excess ag for high thermoelectric performance. Adv. Energy Mater. 2011, 1, 291–296. [Google Scholar] [CrossRef]
  33. Zhu, T.; Liu, Y.; Fu, C.; Heremans, J.P.; Snyder, J.G.; Zhao, X. Compromise and synergy in high-efficiency thermoelectric materials. Adv. Mater. 2017, 29, 1605884. [Google Scholar] [CrossRef] [PubMed]
  34. Pei, Y.; LaLonde, A.D.; Heinz, N.A.; Snyder, G.J. High thermoelectric figure of merit in PbTe alloys demonstrated in PbTe–CdTe. Adv. Energy Mater. 2012, 2, 670–675. [Google Scholar] [CrossRef]
  35. Zhang, J.; Liu, R.; Cheng, N.; Zhang, Y.; Yang, J.; Uher, C.; Shi, X.; Chen, L.; Zhang, W. High-performance pseudocubic thermoelectric materials from non-cubic chalcopyrite compounds. Adv. Mater. 2014, 26, 3848–3853. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, H.; Gibbs, Z.M.; Takagiwa, Y.; Snyder, G.J. Tuning bands of PbSe for better thermoelectric efficiency. Energy Environ. Sci. 2014, 7, 804–811. [Google Scholar] [CrossRef]
  37. Tan, G.; Shi, F.; Hao, S.; Chi, H.; Bailey, T.P.; Zhao, L.D.; Uher, C.; Wolverton, C.; Dravid, V.P.; Kanatzidis, M.G.; et al. Valence Band Modification and High Thermoelectric Performance in SnTe Heavily Alloyed with MnTe. J. Am. Chem. Soc. 2015, 137, 11507–11516. [Google Scholar] [CrossRef] [PubMed]
  38. Li, A.; Hu, C.; He, B.; Yao, M.; Fu, C.; Wang, Y.; Zhao, X.; Felser, C.; Zhu, T. Demonstration of valley anisotropy utilized to enhance the thermoelectric power factor. Nat. Commun. 2021, 12, 1–9. [Google Scholar] [CrossRef]
  39. Qian, X.; Guo, H.R.; Lyu, J.X.; Ding, B.F.; San, X.Y.; Zhang, X.; Wang, J.L.; Wang, S.F. Enhancing thermoelectric performance of p-type SnTe through manipulating energy band structures and decreasing electronic thermal conductivity. Rare Met. 2024, 43, 3232–3241. [Google Scholar] [CrossRef]
  40. Yu, L.; Shi, X.L.; Mao, Y.; Liu, W.D.; Ji, Z.; Wei, S.; Zhang, Z.; Song, W.; Zheng, S.; Chen, Z.G. Simultaneously boosting thermoelectric and mechanical properties of n-type Mg3Sb1.5Bi0.5-based zintls through energy-band and defect engineering. ACS Nano 2024, 18, 1678–1689. [Google Scholar] [CrossRef]
  41. 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]
  42. Mao, J.; Liu, Z.; Zhou, J.; Zhu, H.; Zhang, Q.; Chen, G.; Ren, Z. Advances in thermoelectrics. Adv. Phys. 2018, 67, 69–147. [Google Scholar] [CrossRef]
  43. 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]
  44. Zhang, Q.; Wang, H.; Liu, W.; Wang, H.; Yu, B.; Zhang, Q.; Tian, Z.; Ni, G.; Lee, S.; Esfarjani, K.; et al. Enhancement of thermoelectric figure-of-merit by resonant states of aluminium doping in lead selenide. Energy Environ. Sci. 2012, 5, 5246–5251. [Google Scholar] [CrossRef]
  45. Jaworski, C.M.; Kulbachinskii, V.; Heremans, J.P. Resonant level formed by tin in Bi2Te3 and the enhancement of room-temperature thermoelectric power. Phys. Rev. B-Condens. Matter Mater. Phys. 2009, 80, 233201. [Google Scholar] [CrossRef]
  46. Zhang, Q.; Liao, B.; Lan, Y.; Lukas, K.; Liu, W.; Esfarjani, K.; Opeil, C.; Broido, D.; Chen, G.; Ren, Z. High thermoelectric performance by resonant dopant indium in nanostructured SnTe. Proc. Natl. Acad. Sci. USA 2013, 110, 13261–13266. [Google Scholar] [CrossRef]
  47. Callaway, J.; von Baeyer, H.C. Effect of Point Imperfections on Lattice Thermal Conductivity. Phys. Rev. B 1960, 120, 1149–1154. [Google Scholar] [CrossRef]
  48. Wang, L.; Shi, X.; Li, L.; Hong, M.; Lin, B.; Miao, P.; Ding, J.; Yuan, N.; Zheng, S.; Chen, Z. Zinc Doping Induces Enhanced Thermoelectric Performance of Solvothermal SnTe. Chem.–Asian J. 2024, 19, e202400130. [Google Scholar] [CrossRef]
  49. Klemens, P.G. Thermal resistance due to point defects at high temperatures. Phys. Rev. B 1960, 119, 507–509. [Google Scholar] [CrossRef]
  50. Chen, Z.; Jian, Z.; Li, W.; Chang, Y.; Ge, B.; Hanus, R.; Yang, J.; Chen, Y.; Huang, M.; Snyder, G.J.; et al. Lattice dislocations enhancing thermoelectric PbTe in addition to band convergence. Adv. Mater. 2017, 29, 1606768. [Google Scholar] [CrossRef]
  51. Chen, Z.; Ge, B.; Li, W.; Lin, S.; Shen, J.; Chang, Y.; Hanus, R.; Snyder, G.J.; Pei, Y. Vacancy-induced dislocations within grains for high-performance PbSe thermoelectrics. Nat. Commun. 2017, 8, 13828. [Google Scholar] [CrossRef] [PubMed]
  52. Zhu, B.C.; Bao, L.; Zeng, L.; Fang, W.Y.; Liu, C.J. First-principles calculations of the thermoelectric properties of 2D SnSeS thermoelectric materials. Mater. Today Commun. 2025, 42, 111593. [Google Scholar] [CrossRef]
  53. Zhang, M.; Gao, Z.; Lou, Q.; Zhu, Q.; Wang, J.; Han, Z.; Fu, C.; Zhu, T. Achieving High Carrier Mobility And Thermal Stability in Plainified Rhombohedral GeTe Thermoelectric Materials with ZT > 2. Adv. Funct. Mater. 2023, 34, 2307864. [Google Scholar] [CrossRef]
  54. Yin, L.C.; Liu, W.D.; Li, M.; Wang, D.Z.; Wu, H.; Wang, Y.; Zhang, L.; Shi, X.L.; Liu, Q.; Chen, Z.G. Interstitial Cu: An effective strategy for high carrier mobility and high thermoelectric performance in GeTe. Adv. Funct. Mater. 2023, 33, 2301750. [Google Scholar] [CrossRef]
  55. Hu, M.H.; Li, M.; Wang, D.Z.; Yin, L.C.; Wu, H.; Liu, W.D.; Shi, X.L.; Wang, Y.; Liu, Q.; Chen, Z.G. Revisiting Cobalt Dopability in GeTe System to Design Modulation-Doped Thermoelectrics. Adv. Funct. Mater. 2025, 2421837. [Google Scholar] [CrossRef]
  56. Guo, Z.; Wu, G.; Yuan, M.; Li, H.; Tan, X.; Cai, J.; Zhang, Z.; Shuai, J.; Liu, G.; Jiang, J. High Thermoelectric Performance and Energy Conversion in Hg-doped GeTe. Adv. Funct. Mater. 2024, 2419431. [Google Scholar] [CrossRef]
  57. Zhu, H.; Deng, Z.; Qu, Y.; He, P.; Geng, H. Cu-In Co-Doping and Layered Directional Sintering for High Thermoelectric Efficiency and Mechanical Strength in Bi2 (Te,Se)3. Adv. Funct. Mater. 2025, 2422007. [Google Scholar] [CrossRef]
  58. Kim, J.H.; Yun, J.H.; Cha, S.; Byeon, S.; Park, J.; Jin, H.; Kim, S.; Kim, S.J.; Park, J.; Jang, J. Enhancement of Phase Stability and Thermoelectric Performance of Meta-Stable AgSbTe2 by Thermal Cycling Process. Adv. Funct. Mater. 2024, 34, 2404886. [Google Scholar] [CrossRef]
  59. Zhang, D.; Wang, X.; Wu, H.; Huang, Y.; Zheng, S.; Zhang, B.; Fu, H.; Cheng, Z.; Jiang, P.; Han, G.; et al. High Thermoelectric Performance in Earth-Abundant Cu3SbS4 by Promoting Doping Efficiency via Rational Vacancy Design. Adv. Funct. Mater. 2023, 33, 2214163. [Google Scholar] [CrossRef]
  60. Liu, K.; Chen, C.; Cheng, J.; Ma, X.; Li, J.; Bao, X.; Li, H.; Zhang, Q.; Chen, Y. Realizing a High Thermoelectric Conversion Efficiency in Zintl-phase NaCdSb via Suppressing the Intrinsic Carrier Excitation. Adv. Funct. Mater. 2024, 2419145. [Google Scholar] [CrossRef]
  61. Yang, Z.; Zi, P.; Liu, K.; Luo, H.; Bai, H.; Chen, S.; Wu, J.; Su, X.; Uher, C.; Zhang, Q.; et al. Ultralow Lattice Thermal Conductivity and Extraordinary Thermoelectric Performance in Highly Disordered CuIn7Se11 Layered Compound. Adv. Funct. Mater. 2024, 34, 2306849. [Google Scholar] [CrossRef]
  62. Zhan, S.; Bai, S.; Qin, B.; Zhu, Y.; Wang, S.; Liu, D.; Hong, T.; Gao, X.; Zheng, L.; Wen, Y.; et al. High Carrier Mobility Promotes In-Plane Thermoelectric Performance of n-Type PbSnS2 Crystals. Adv. Funct. Mater. 2024, 34, 2406428. [Google Scholar] [CrossRef]
  63. Moshwan, R.; Zhang, M.; Li, M.; Liu, S.; Li, N.; Cao, T.; Liu, W.; Shi, X.; Chen, Z. Compromising Configurational Entropy Leading to Exceptional Thermoelectric Properties in SnTe-Based Materials. Adv. Funct. Mater. 2024, 2418291. [Google Scholar] [CrossRef]
  64. Qian, Q.; Cheng, H.; Le, Q.; Ouyang, J. Great Enhancement in the Thermopower of Ionic Liquid by a Metal-Organic Framework. Adv. Funct. Mater. 2023, 33, 2303311. [Google Scholar] [CrossRef]
  65. Yang, H.; Wu, L.; Feng, X.; Huang, X.; Wang, H.; Duan, B.; Li, G.; Zhai, P.; Zhang, Q. Constructing Coated Grain Nanocomposites and Intracrystalline Precipitates to Simultaneously Improve the Thermoelectric and Mechanical Properties of SnTe by MgB2 and Sb Co-Doping. Adv. Funct. Mater. 2024, 34, 2316344. [Google Scholar] [CrossRef]
  66. Cho, H.; Back, S.Y.; Sato, N.; Liu, Z.; Gao, W.; Wang, L.; Nguyen, H.D.; Kawamoto, N.; Mori, T. Outstanding Room-Temperature Thermoelectric Performance of n-type Mg3Bi2-Based Compounds Through Synergistically Combined Band Engineering Approaches. Adv. Funct. Mater. 2024, 34, 2407017. [Google Scholar] [CrossRef]
  67. Chen, P.; Wu, H.; Zhang, B.; Zhou, Z.; Zheng, S.; Dai, L.; Huo, Y.; Zhang, D.; Yan, Y.; Peng, K.; et al. Intrinsically Low Lattice Thermal Conductivity and Anisotropic Thermoelectric Performance in In-doped GeSb2Te4 Single Crystals. Adv. Funct. Mater. 2023, 33, 2211281. [Google Scholar] [CrossRef]
  68. Ma, Z.; Luo, Y.; Dong, J.; Liu, Y.; Zhang, D.; Li, W.; Li, C.; Wei, Y.; Jiang, Q.; Li, X.; et al. Synergistic Performance of Thermoelectric and Mechanical in Nanotwinned High-Entropy Semiconductors AgMnGePbSbTe5. Adv. Mater. 2024, 36, e2407982. [Google Scholar] [CrossRef]
  69. Chen, J.; Dong, Z.; Li, Q.; Ge, B.; Zhang, J.; Zhang, Y.; Luo, J. Enhanced Thermoelectric Performance in Vacancy-Filling Heuslers due to Kondo-Like Effect. Adv. Mater. 2024, 36, 2405858. [Google Scholar] [CrossRef]
  70. Li, W.; Poudel, B.; Kishore, R.A.; Nozariasbmarz, A.; Liu, N.; Zhang, Y.; Priya, S. Toward high conversion efficiency of thermoelectric modules through synergistical optimization of layered materials. Adv. Mater. 2023, 35, e2210407. [Google Scholar] [CrossRef]
  71. Solco, S.F.D.; Saglik, K.; Zhang, D.; Tan, X.Y.; Zhu, Q.; Liu, H.; Suwardi, A.; Cao, J. Thermoelectric performance enhancement of Mg2Si-based silicides synthesized in nitrogen atmosphere. Mater. Res. Express 2024, 11, 125505. [Google Scholar] [CrossRef]
  72. Kuroda, K.; Ye, M.; Kimura, A.; Eremeev, S.V.; Krasovskii, E.E.; Chulkov, E.V.; Ueda, Y.; Miyamoto, K.; Okuda, T.; Shimada, K.; et al. Experimental Realization of a Three-Dimensional Topological Insulator Phase in Ternary Chalcogenide TlBiSe2. Phys. Rev. Lett. 2010, 105, 146801. [Google Scholar] [CrossRef] [PubMed]
  73. Zhu, T.; Hu, L.; Zhao, X.; He, J. New insights into intrinsic point defects in V2VI3 thermoelectric materials. Adv. Sci. 2016, 3, 1600004. [Google Scholar] [CrossRef] [PubMed]
  74. Shi, Q.; Li, J.; Zhao, X.; Chen, Y.; Zhang, F.; Zhong, Y.; Ang, R. Comprehensive insight into p-type Bi2Te3-based thermoelectrics near room temperature. ACS Appl. Mater. Interfaces 2022, 14, 49425–49445. [Google Scholar] [CrossRef] [PubMed]
  75. Shi, Y.C.; Yang, J.; Wang, Y.; Li, Z.G.; Zhong, T.Y.; Ge, Z.H.; Feng, J.; He, J. Highly effective solid solution towards high thermoelectric and mechanical performance in Bi–Sb–Te alloys via Trojan doping. Energy Environ. Sci. 2024, 17, 2326–2335. [Google Scholar] [CrossRef]
  76. Fanguo, L. Preparation and Thermoelectric Properties of Cux/Bi2Te3 Composites. Compos. Mech. Comput. Appl. Int. J. 2024, 15, 35–43. [Google Scholar] [CrossRef]
  77. Fang, W.; Wu, F. The impact of varying EDTA dosage on the flower-like morphology and thermoelectric properties of Ce-doped Bi2Te3. Sci. Rep. 2024, 14, 21855. [Google Scholar]
  78. Lee, N.; Ye, S.W.; Park, J.; Seo, S.; Kang, M.; Kim, H.-S.; Kim, S.-I.; Cho, J.Y.; Nam, W.H.; Roh, J.W. The Electrical Conduction-Type Transition in Se-Free Bi2Te3 Thermoelectric Materials by Decoration of Ni Nanoparticles. ACS Appl. Energy Mater. 2024, 7, 8807–8813. [Google Scholar] [CrossRef]
  79. Lee, S.; Park, G.M.; Kim, Y.; Lee, S.H.; Jung, S.J.; Hong, J.; Kim, S.C.; Won, S.O.; Lee, A.S.; Chung, Y.J.; et al. Unlocking the potential of porous Bi2Te3-based thermoelectrics using precise interface engineering through atomic layer deposition. ACS Appl. Mater. Interfaces 2024, 16, 17683–17691. [Google Scholar] [CrossRef]
  80. Wang, X.; Shang, H.; Gu, H.; Chen, Y.; Zhang, Z.; Zou, Q.; Zhang, L.; Feng, C.; Li, G.; Ding, F. High-Performance p-Type Bi2Te3-Based Thermoelectric Materials Enabled via Regulating Bi–Te Ratio. ACS Appl. Mater. Interfaces 2024, 16, 11678–11685. [Google Scholar] [CrossRef]
  81. Aminorroaya Yamini, S.; Santos, R.; Fortulan, R.; Gazder, A.A.; Malhotra, A.; Vashaee, D.; Serhiienko, I.; Mori, T. Room-temperature thermoelectric performance of n-type multiphase pseudobinary Bi2Te3–Bi2S3 compounds: Synergic effects of phonon scattering and energy filtering. ACS Appl. Mater. Interfaces 2023, 15, 19220–19229. [Google Scholar] [CrossRef] [PubMed]
  82. Li, Y.Z.; Zhang, Q.; Liu, K.; Lin, Y.J.; Lin, N.; Yu, Y.; Liu, F.; Zhao, X.B.; Ge, B.H. Multi-scale hierarchical microstructure modulation towards high room temperature thermoelectric performance in n-type Bi2Te3-based alloys. Mater. Today Nano 2023, 22, 100340. [Google Scholar] [CrossRef]
  83. Lee, S.; Jung, S.J.; Park, G.M.; Na, M.Y.; Kim, K.C.; Hong, J.; Lee, A.S.; Baek, S.H.; Kim, H.; Park, T.J.; et al. Selective Dissolution-Derived Nanoporous Design of Impurity-Free Bi2Te3 Alloys with High Thermoelectric Performance. Small 2023, 19, 2205202. [Google Scholar] [CrossRef] [PubMed]
  84. Lin, F.H.; Liu, C.J. Reaction mechanism of room-temperature synthesized n-Bi2Te3 and high thermoelectric performance p-(Bi,Sb)2Te3–Te using nanosized precursor methods: A novel and energy-efficient production method. Ceram. Int. 2023, 49, 26077–26083. [Google Scholar] [CrossRef]
  85. Khan, J.S.; Akram, R.; Shah, A.A.; Hussain, M.; Rafique, S.; ur Rehman, A.; Khurshid, T.; Karim, K. Enhanced zT due to non-stoichiometric induced defects for bismuth telluride thermoelectric materials. Kuwait J. Sci. 2023, 50, 231–237. [Google Scholar] [CrossRef]
  86. Kwon, O.; Kim, M.; Lee, J.; Lee, S.; Kim, J.; Kim, M.; Kim, J. High Thermoelectric Performance of Bi0.3Sb1.7Te3/PEDOT: PSS Flexible Composite via Cation Cross-Linking. J. Alloys Compd. 2025, 1014, 178789. [Google Scholar] [CrossRef]
  87. Meang, E.J.; Shin, Y.J.; Park, K.H.; Chung, J.; Kim, H.; Lee, H.S. Scalable synthesis and enhanced thermoelectric properties of Cu-doped and Se-substituted Bi2Te3-based materials via high-pressure sintering. Mater. Today Commun. 2025, 111830. [Google Scholar] [CrossRef]
  88. Chen, S.; Luo, T.; Li, J.; Liao, L.; Liu, Y.; Yang, Z.; Zhong, S.; Wu, J.; Su, X.; Yan, Y.; et al. Synergistic regulation of texture and low-angle grain boundaries enables mechanically robust p-type Bi2Te3-based compounds with superior thermoelectric performance. Mater. Today Energy 2025, 48, 101788. [Google Scholar] [CrossRef]
  89. Niu, X.; Li, Z.; Fan, Z.; Wu, X.; Zhao, Q.; Wang, S.; Zhu, H.; Li, G.; Zhao, H. Vacancy Regulation to Achieve N-Type High Thermoelectric Performance PbSe through Titanium-Incorporation. Mater. Today Phys. 2025, 101675. [Google Scholar] [CrossRef]
  90. Guo, F.; Sun, Y.; Qin, H.; Zhu, Y.; Ge, Z.; Liu, Z.; Cai, W.; Sui, J. BiSbTe alloy with high thermoelectric and mechanical performance for power generation. Scr. Mater. 2022, 218, 114801. [Google Scholar] [CrossRef]
  91. Lou, L.Y.; Yang, J.; Zhu, Y.K.; Liang, H.; Zhang, Y.X.; Feng, J.; He, J.; Ge, Z.H.; Zhao, L.D. Tunable electrical conductivity and simultaneously enhanced thermoelectric and mechanical properties in n-type Bi2Te3. Adv. Sci. 2022, 9, 2203250. [Google Scholar] [CrossRef] [PubMed]
  92. Liu, Z.; Gao, W.; Oshima, H.; Nagase, K.; Lee, C.-H.; Mori, T. Maximizing the performance of n-type Mg3Bi2 based materials for room-temperature power generation and thermoelectric cooling. Nat. Commun. 2022, 13, 1–9. [Google Scholar] [CrossRef]
  93. Kumar Choudhary, K.; Rathore, V.; Chandra Dixit, R.; Kaurav, N. Phonon Scattering Mechanism for Size-Dependent Thermoelectric Properties of Bi2Te3 Nanoparticles. ChemistrySelect 2022, 7, e202202503. [Google Scholar] [CrossRef]
  94. Zhu, B.; Su, X.; Shu, S.; Luo, Y.; Tan, X.Y.; Sun, J.; Sun, D.; Zhang, H.; Zhang, Q.; Suwardi, A.; et al. Cold-sintered Bi2Te3-based materials for engineering nanograined thermoelectrics. ACS Appl. Energy Mater. 2022, 5, 2002–2010. [Google Scholar] [CrossRef]
  95. Androulakis, J.; Todorov, I.; Chung, D.-Y.; Ballikaya, S.; Wang, G.; Uher, C.; Kanatzidis, M. Thermoelectric enhancement in PbTe with K or Na codoping from tuning the interaction of the light- and heavy-hole valence bands. Phys. Rev. B 2010, 82, 115209. [Google Scholar] [CrossRef]
  96. Yu, X.; Li-Dong, Z. Charge and phonon transport in PbTe-based thermoelectric materials. Npj Quantum Mater. 2018, 3, 55. [Google Scholar]
  97. Shtern, Y.; Sherchenkov, A.; Shtern, M.; Rogachev, M.; Pepelyaev, D. Challenges and perspective recent trends of enhancing the efficiency of thermoelectric materials on the basis of PbTe. Mater. Today Commun. 2023, 37, 107083. [Google Scholar] [CrossRef]
  98. Shtern, M.; Sherchenkov, A.; Shtern, Y.; Borgardt, N.; Rogachev, M.; Yakubov, A.; Babich, A.; Pepelyaev, D.; Voloshchuk, I.; Zaytseva, Y.; et al. Mechanical properties and thermal stability of nanostructured thermoelectric materials on the basis of PbTe and GeTe. J. Alloys Compd. 2023, 946, 169364. [Google Scholar] [CrossRef]
  99. Sauerschnig, P.; Jood, P.; Ohta, M. Challenges and Progress in Contact Development for PbTe-based Thermoelectrics. Chemnanomat 2023, 9, e2200560. [Google Scholar] [CrossRef]
  100. Bathula, S.; Jayasimhadri, M.; Singh, N.; Srivastava, A.K.; Pulikkotil, J.; Dhar, A.; Budhani, R.C. Enhanced thermoelectric figure-of-merit in spark plasma sintered nanostructured n-type SiGe alloys. Appl. Phys. Lett. 2012, 101, 213902. [Google Scholar] [CrossRef]
  101. Yu, B.; Zebarjadi, M.; Wang, H.; Lukas, K.; Wang, H.; Wang, D.; Opeil, C.; Dresselhaus, M.; Chen, G.; Ren, Z. Enhancement of thermoelectric properties by modulation-doping in silicon germanium alloy nanocomposites. Nano Lett. 2012, 12, 2077–2082. [Google Scholar] [CrossRef] [PubMed]
  102. Lee, Y.; Pak, A.J.; Hwang, G.S. What is the thermal conductivity limit of silicon germanium alloys? Phys. Chem. Chem. Phys. 2016, 18, 19544–19548. [Google Scholar] [CrossRef]
  103. Hussain, T.; Li, X.; Danish, M.H.; Rehman, M.U.; Zhang, J.; Li, D.; Chen, G.; Tang, G. Realizing high thermoelectric performance in eco-friendly SnTe via synergistic resonance levels, band convergence and endotaxial nanostructuring with Cu2Te. Nano Energy 2020, 73. [Google Scholar] [CrossRef]
  104. Hong, M.; Chen, Z.G. Chemistry in advancing thermoelectric GeTe materials. Acc. Chem. Res. 2022, 55, 3178–3190. [Google Scholar] [CrossRef] [PubMed]
  105. Samanta, M.; Ghosh, T.; Arora, R.; Waghmare, U.V.; Biswas, K. Realization of both n- and p-type GeTe thermoelectrics: Electronic structure modulation by AgBiSe2 alloying. J. Am. Chem. Soc. 2019, 141, 19505–19512. [Google Scholar] [CrossRef]
  106. Sales, B.C.; Mandrus, D.; Williams, R.K. Filled skutterudite antimonides: A new class of thermoelectric materials. Science 1996, 272, 1325–1328. [Google Scholar] [CrossRef]
  107. Nayak, P.; Srivastava, P.; Gupta, D.C. Theoretical exploration of inherent electronic, structural, mechanical, thermoelectric, and thermophysical response of KRu4Z12 (Z = As12, Sb12) filled skutterudite materials. RSC Adv. 2023, 13, 27873–27886. [Google Scholar] [CrossRef]
  108. Li, H.; Tang, X.; Zhang, Q.; Uher, C. High performance InxCeyCo4Sb12 thermoelectric materials with in situ forming nanostructured InSb phase. Appl. Phys. Lett. 2009, 94, 102114. [Google Scholar] [CrossRef]
  109. Shi, X.; Yang, J.; Salvador, J.R.; Chi, M.; Cho, J.Y.; Wang, H.; Bai, S.; Yang, J.; Zhang, W.; Chen, L. Multiple-filled skutterudites: High thermoelectric figure of merit through separately optimizing electrical and thermal transports. J. Am. Chem. Soc. 2011, 133, 7837–7846. [Google Scholar] [CrossRef]
  110. Sharp, J.W.; Poon, S.J.; Goldsmid, H.J. Boundary scattering and the thermoelectric figure of merit. Phys. Status Solidi A 2001, 187, 507–516. [Google Scholar] [CrossRef]
  111. Yu, C.; Zhu, T.J.; Shi, R.Z.; Zhang, Y.; Zhao, X.B.; He, J. High-performance half-Heusler thermoelectric materials Hf1−x ZrxNiSn1−ySby prepared by levitation melting and spark plasma sintering. Acta Mater. 2009, 57, 2757–2764. [Google Scholar] [CrossRef]
  112. Fu, C.; Zhu, T.; Pei, Y.; Xie, H.; Wang, H.; Snyder, G.J.; Liu, Y.; Liu, Y.; Zhao, X. High band degeneracy contributes to high thermoelectric performance in p-type half-heusler compounds. Adv. Energy Mater. 2014, 4, 1400600. [Google Scholar] [CrossRef]
  113. Pei, Y.; LaLonde, A.D.; Wang, H.; Snyder, G.J. Low effective mass leading to high thermoelectric performance. Energy Environ. Sci. 2012, 5, 7963–7969. [Google Scholar] [CrossRef]
  114. Teunis, M.B.; Jana, A.; Dutta, P.; Johnson, M.A.; Mandal, M.; Muhoberac, B.B.; Sardar, R. Mesoscale growth and assembly of bright luminescent organolead halide perovskite quantum wires. Chem. Mater. 2016, 28, 5043–5054. [Google Scholar] [CrossRef]
  115. Liu, W.; Chang, A.Y.; Schaller, R.D.; Talapin, D.V. Colloidal insb nanocrystals. J. Am. Chem. Soc. 2012, 134, 20258–20261. [Google Scholar] [CrossRef] [PubMed]
  116. Zhao, C.; Wang, M.; Liu, Z. Research Progress on Preparation Methods of Skutterudites. Inorganics 2022, 10, 106. [Google Scholar] [CrossRef]
  117. Dey, T.K. Electrical conductivity, thermoelectric power and figure of merit of doped Bi-Sb tapes produced by melt spinning technique. Pramana 1990, 34, 243–248. [Google Scholar] [CrossRef]
  118. Chen, H.Y.; Zhao, X.B.; Lu, Y.F.; Mueller, E.; Mrotzek, A. Microstructures and thermoelectric properties of Fe0.92Mn0.08Six alloys prepared by rapid solidification and hot pressing. J. Appl. Phys. 2003, 94, 6621–6626. [Google Scholar] [CrossRef]
  119. Purushottam Bhat, C.; Anusha; Ani, A.; Shanubhogue, U.D.; Poornesh, P.; Rao, A.; Chattopadhyay, S. Investigations on Bi Doped Cu2Se Prepared by Solid State Reaction Technique for Thermoelectric Applications. Energies 2023, 16, 3010. [Google Scholar] [CrossRef]
  120. Qin, B.K.; Zhang, L.; Ji, Y.H.; Bai, Z.L.; Zhao, D.; Sun, C. Enhanced thermoelectric properties of FeTe2 by Sb doping prepared by solid-state reaction. Mater. Sci. Semicond. Process. 2024, 174, 108212. [Google Scholar] [CrossRef]
  121. Bai, Z.; Ji, Y.; Qin, B. Microstructure and thermoelectric properties of marcasite-type compounds XTe2 (X = Fe, Co, Ni) prepared by solid-state reaction. J. Mater. Sci. Mater. Electron. 2023, 34, 461. [Google Scholar] [CrossRef]
  122. Jin, M.; Tang, Z.; Jiang, J.; Zhang, R.; Zhou, L.; Zhao, S.; Chen, Y.; Chen, Y.; Wang, X.; Li, R. Growth of SnSe single crystal via vertical vapor deposition method and characterization of its thermoelectric performance. Mater. Res. Bull. 2020, 126, 110819. [Google Scholar] [CrossRef]
  123. Chuai, Y.H.; Wang, Y.F.; Bai, Y. Structural, Optical, Electrical, and Thermoelectric Properties of Bi2Se3 Films Deposited at a High Se/Bi Flow Rate. Nanomaterials 2023, 13, 2785. [Google Scholar] [CrossRef] [PubMed]
  124. Son, J.S.; Choo, S.; Lee, J.; Sisik, B.; Jung, S.J.; Kim, K.; Yang, S.E.; Jo, S.; Nam, C.; Ahn, S.; et al. Geometric design of Cu2Se-based thermoelectric device for enhancing power generation. Res. Sq. 2023. [Google Scholar] [CrossRef]
  125. Bharti, M.; Singh, A.; Samanta, S.; Aswal, D. Conductive polymers for thermoelectric power generation. Prog. Mater. Sci. 2018, 93, 270–310. [Google Scholar] [CrossRef]
  126. Singh, R.; Dogra, S.; Dixit, S.; Vatin, N.I.; Bhardwaj, R.; Sundramoorthy, A.K.; Perera, H.; Patole, S.P.; Mishra, R.K.; Arya, S. Advancements in thermoelectric materials for efficient waste heat recovery and renewable energy generation. Hybrid Adv. 2024, 5, 100176. [Google Scholar] [CrossRef]
  127. Palaporn, D.; Tanusilp, S.-A.; Sun, Y.; Pinitsoontorn, S.; Kurosaki, K. Thermoelectric materials for space explorations. Mater. Adv. 2024, 5, 5351–5364. [Google Scholar] [CrossRef]
  128. Summerer, L.; Roux, J.P.; Pustovalov, A.; Gusev, V.; Rybkin, N. Technology-based design and scaling for RTGs for space exploration in the 100 W range. Acta Astronaut. 2011, 68, 873–882. [Google Scholar] [CrossRef]
  129. Dai, Y.; Wang, H.; Qi, K.; Ma, X.; Wang, M.; Ma, Z.; Wang, Z.; Yang, Y.; Ramakrishna, S.; Ou, K. Electrode-dependent thermoelectric effect in ionic hydrogel fiber for self-powered sensing and low-grade heat harvesting. Chem. Eng. J. 2024, 497, 154970. [Google Scholar] [CrossRef]
  130. Sajid, M.; Hassan, I.; Rahman, A. An overview of cooling of thermoelectric devices. Renew. Sustain. Energy Rev. 2017, 78, 15–22. [Google Scholar] [CrossRef]
  131. Liu, Y.; Wang, X.; Hou, S.; Wu, Z.; Wang, J.; Mao, J.; Zhang, Q.; Liu, Z.; Cao, F. Scalable-produced 3D elastic thermoelectric network for body heat harvesting. Nat. Commun. 2023, 14, 1–8. [Google Scholar] [CrossRef] [PubMed]
  132. He, Y.; Li, R.; Fan, Y.; Zheng, Y.; Chen, G. Study on the performance of a solid-state thermoelectric refrigeration system equipped with ionic wind fans for ultra-quiet operation. Int. J. Refrig. 2021, 130, 441–451. [Google Scholar] [CrossRef]
  133. Qamar, A.; Kanwal, A.; Amjad, M.; Farooq, M.; Munir, A.; Ahmad, S.; Abdollahian, M. Advancing sustainable cooling: Performance analysis of a solar-driven thermoelectric refrigeration system for eco-friendly solutions. Case Stud. Therm. Eng. 2024, 60, 104781. [Google Scholar] [CrossRef]
  134. Twaha, S.; Zhu, J.; Yan, Y.; Li, B. A comprehensive review of thermoelectric technology: Materials, applications, modelling and performance improvement. Renew. Sustain. Energy Rev. 2016, 65, 698–726. [Google Scholar] [CrossRef]
  135. 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]
  136. Hendricks, T.; Caillat, T.; Mori, T. Keynote Review of Latest Advances in Thermoelectric Generation Materials, Devices, and Technologies 2022. Energies 2022, 15, 7307. [Google Scholar] [CrossRef]
  137. Yazawa, K.; Bahk, J.-H.; Shakouri, A. Thermoelectric Energy Conversion Devices and Systems; World Scientific Pub Co Pte Ltd.: Singapore, 2021; ISBN 9789811218262. [Google Scholar]
  138. Zheng, L.J.; Lim, S.; Kim, N.K.; Kang, D.H.; Youn, Y.J.; Lee, W.; Kang, H.W. Experimental study of a thin water-film evaporative cooling system to enhance the energy conversion efficiency of a thermoelectric device. Energy 2020, 211, 119040. [Google Scholar] [CrossRef]
  139. Singh, B.S.B. Thermoelectric Generators: Design, Operation, and Applications. In New Materials and Devices for Thermoelectric Power Generation; IntechOpen: London, UK, 2023. [Google Scholar]
  140. Mastsuura, K.; Rowe, D.M.; Tsvyoshi, A.; Min, G. Large scale thermoelectric generator’s of low grade heat. In Proceedings of the 10th International Conference on Thermoelectric, Cardiff, UK, 10–12 September 1991; pp. 9–12. [Google Scholar]
  141. Meng, F.; Chen, L.; Xie, Z.; Ge, Y. Thermoelectric generator with air-cooling heat recovery device from wastewater. Therm. Sci. Eng. Prog. 2017, 4, 106–112. [Google Scholar] [CrossRef]
  142. Telkes, M. Solar thermoelectric generators. J. Appl. Phys. 1954, 25, 765–777. [Google Scholar] [CrossRef]
  143. Guo, J.; Huai, X. Maximizing Electric Power through Spectral-Splitting Photovoltaic-Thermoelectric Hybrid System Integrated with Radiative Cooling. Adv. Sci. 2023, 10, 2206575. [Google Scholar] [CrossRef]
  144. Lamba, R.; Kaushik, S. Modeling and performance analysis of a concentrated photovoltaic–thermoelectric hybrid power generation system. Energy Convers. Manag. 2016, 115, 288–298. [Google Scholar] [CrossRef]
  145. Ge, M.; Zhao, Y.; Li, Y.; He, W.; Xie, L.; Zhao, Y. Structural optimization of thermoelectric modules in a concentration photovoltaic–thermoelectric hybrid system. Energy 2022, 244 Pt B, 123202. [Google Scholar] [CrossRef]
  146. Ying, P.; He, R.; Mao, J.; Zhang, Q.; Reith, H.; Sui, J.; Ren, Z.; Nielsch, K.; Schierning, G. Towards tellurium-free thermoelectric modules for power generation from low-grade heat. Nat. Commun. 2021, 12, 1121. [Google Scholar] [CrossRef] [PubMed]
  147. Huang, B.; Shen, Z.G. Performance assessment of annular thermoelectric generators for automobile exhaust waste heat recovery. Energy 2022, 246, 123375. [Google Scholar] [CrossRef]
  148. Siddique, A.R.M.; Bozorgi, M.; Venkateshwar, K.; Tasnim, S.; Mahmud, S. Phase change material-enhanced solid-state thermoelectric cooling technology for food refrigeration and storage applications. J. Energy Storage 2023, 60, 106569. [Google Scholar] [CrossRef]
  149. Ab Rahman, R.; Mohamad, M.A.H.; Kaamin, M.; Batcha, M.F.M.; Mazlan, M.D.A.; Rosli, M.L.; Aziz, M.A.A.C. Experimental study of peltier-based thermoelectric cooling box system. J. Adv. Res. Appl. Mech. 2022, 94, 1–6. [Google Scholar] [CrossRef]
  150. Wu, X.; Wang, Z.; Liu, Y.; Sun, X.; Xu, Y.; Tian, Y.; Wang, B.; Sang, X.; Shi, J.; Xiong, R. Enhanced performance of Bi2Te3-based thermoelectric materials by incorporating Bi2Fe4O9 magnetic nanoparticles. J. Alloys Compd. 2022, 904, 163933. [Google Scholar] [CrossRef]
  151. Liu, C.; Liang, D.; Chen, T.; Ye, X.; He, D.; Zhu, W.; Nie, X.; Wei, P.; Zhao, W.; Zhang, Q. Optimizing Room-Temperature Thermoelectric and Magnetocaloric Performance via Constructing Multi-Scale Interfacial Phases in LaFeSi/BiSbTe Thermo-Electro-Magnetic Refrigeration Materials. Adv. Funct. Mater. 2024, 35, 2415368. [Google Scholar] [CrossRef]
Figure 1. (a) Seebeck effect; (b) Peltier effect.
Figure 1. (a) Seebeck effect; (b) Peltier effect.
Energies 18 02122 g001
Figure 2. Variation in S, σ, K, and ZT with carrier concentration (n) [1].
Figure 2. Variation in S, σ, K, and ZT with carrier concentration (n) [1].
Energies 18 02122 g002
Figure 3. (a,b) Temperature dependence of carrier concentration and carrier mobility of Bi2-xGexO2Se [30]; (c,d) room-temperature carrier concentration and mobility for Zn1−xCuxSb and PF curves dependent on carrier concentration (n) using the SPB model compared with the experimental data [28]; (e,f) temperature-dependent nH of CaMg2-xCdxBi2 samples and room-temperature nH and μH for Ca1−yNayMg1.85Cd0.15Bi2 samples [31].
Figure 3. (a,b) Temperature dependence of carrier concentration and carrier mobility of Bi2-xGexO2Se [30]; (c,d) room-temperature carrier concentration and mobility for Zn1−xCuxSb and PF curves dependent on carrier concentration (n) using the SPB model compared with the experimental data [28]; (e,f) temperature-dependent nH of CaMg2-xCdxBi2 samples and room-temperature nH and μH for Ca1−yNayMg1.85Cd0.15Bi2 samples [31].
Energies 18 02122 g003
Figure 4. Energy band structures through DFT calculations: (a) Sn27Te27, (b) Sn25BiCdTe26Se, and (c) schematic diagram of band structure changes in SnTe with CdSe alloying [39]. Band structures of (d) Mg24Sb16, (e) Mg24Sb12Bi4, and (f) Mg24Sb15Te1 [40].
Figure 4. Energy band structures through DFT calculations: (a) Sn27Te27, (b) Sn25BiCdTe26Se, and (c) schematic diagram of band structure changes in SnTe with CdSe alloying [39]. Band structures of (d) Mg24Sb16, (e) Mg24Sb12Bi4, and (f) Mg24Sb15Te1 [40].
Energies 18 02122 g004aEnergies 18 02122 g004b
Figure 5. (a) Schematic diagram of the mechanism of enhanced phonon scattering by nanoprecipitates, nanopores, and grain boundaries; (b) SEM images of SnTe, showing the grain boundary; (c) SEM images of Sn0.97Zn0.03Te, showing the grain boundary and nanoprecipitates; (d) enlarged SEM images of Sn0.97Zn0.03Te, clearly showing nanoprecipitates; (e) BSE image of the as-sintered Sn0.97Zn0.03Te [48].
Figure 5. (a) Schematic diagram of the mechanism of enhanced phonon scattering by nanoprecipitates, nanopores, and grain boundaries; (b) SEM images of SnTe, showing the grain boundary; (c) SEM images of Sn0.97Zn0.03Te, showing the grain boundary and nanoprecipitates; (d) enlarged SEM images of Sn0.97Zn0.03Te, clearly showing nanoprecipitates; (e) BSE image of the as-sintered Sn0.97Zn0.03Te [48].
Energies 18 02122 g005
Figure 6. Peak ZT of Bi2Te3-based thermoelectric materials [57,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94].
Figure 6. Peak ZT of Bi2Te3-based thermoelectric materials [57,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94].
Energies 18 02122 g006
Figure 7. (a) Radioisotope thermoelectric generator; (b) PbTe thermoelectric modules [99].
Figure 7. (a) Radioisotope thermoelectric generator; (b) PbTe thermoelectric modules [99].
Energies 18 02122 g007
Figure 8. Conversion efficiency measurement of the Bi2Te3-based module. (a,b) Optical images of the fabricated 2-pair module and the power generation setup. Current dependence of the (c) output power (P), and (d) conversion efficiency [131].
Figure 8. Conversion efficiency measurement of the Bi2Te3-based module. (a,b) Optical images of the fabricated 2-pair module and the power generation setup. Current dependence of the (c) output power (P), and (d) conversion efficiency [131].
Energies 18 02122 g008
Figure 10. Commercial thermoelectric module CP-031 (a) front view; (b) back view; (c) TEC module attached to the container [148]; and (d) power supply, heatsinks, Peltier module and CPU fan preparation, (e) fastening process for the Peltier module between small and big heatsinks [149].
Figure 10. Commercial thermoelectric module CP-031 (a) front view; (b) back view; (c) TEC module attached to the container [148]; and (d) power supply, heatsinks, Peltier module and CPU fan preparation, (e) fastening process for the Peltier module between small and big heatsinks [149].
Energies 18 02122 g010
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

Wang, J.; Yin, Y.; Che, C.; Cui, M. Research Progress of Thermoelectric Materials—A Review. Energies 2025, 18, 2122. https://doi.org/10.3390/en18082122

AMA Style

Wang J, Yin Y, Che C, Cui M. Research Progress of Thermoelectric Materials—A Review. Energies. 2025; 18(8):2122. https://doi.org/10.3390/en18082122

Chicago/Turabian Style

Wang, Jun, Yonggao Yin, Chunwen Che, and Mengying Cui. 2025. "Research Progress of Thermoelectric Materials—A Review" Energies 18, no. 8: 2122. https://doi.org/10.3390/en18082122

APA Style

Wang, J., Yin, Y., Che, C., & Cui, M. (2025). Research Progress of Thermoelectric Materials—A Review. Energies, 18(8), 2122. https://doi.org/10.3390/en18082122

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