Next Article in Journal
Improved Visible Emission from ZnO Nanoparticles Synthesized via the Co-Precipitation Method
Next Article in Special Issue
Innovative Design of Bismuth-Telluride-Based Thermoelectric Transistors
Previous Article in Journal
Enhancing the Electrochemical Stability of LiNi0.8Co0.1Mn0.1O2 Compounds for Lithium-Ion Batteries via Tailoring Precursors Synthesis Temperatures
Previous Article in Special Issue
Preparation and Laser-Induced Thermoelectric Voltage Effect of Bi2Sr2Co2Oy Thin Films Grown on Al2O3 (0001) Substrate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

High-Throughput Screening of High-Performance Thermoelectric Materials with Gibbs Free Energy and Electronegativity

1
Beijing Municipal Key Lab. of Advanced Energy Materials and Technology, School of Materials Science and Engineering, University of Science and Technology Beijing, Beijing 100083, China
2
School of Materials Science and Engineering, Guizhou Minzu University, Guiyang 550025, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(15), 5399; https://doi.org/10.3390/ma16155399
Submission received: 25 June 2023 / Revised: 22 July 2023 / Accepted: 26 July 2023 / Published: 1 August 2023

Abstract

:
Thermoelectric (TE) materials are an important class of energy materials that can directly convert thermal energy into electrical energy. Screening high-performance thermoelectric materials and improving their TE properties are important goals of TE materials research. Based on the objective relationship among the molar Gibbs free energy (Gm), the chemical potential, the Fermi level, the electronegativity (X) and the TE property of a material, a new method for screening TE materials with high throughput is proposed. This method requires no experiments and no first principle or Ab initio calculation. It only needs to find or calculate the molar Gibbs free energy and electronegativity of the material. Here, by calculating a variety of typical and atypical TE materials, it is found that the molar Gibbs free energy of Bi2Te3 and Sb2Te3 from 298 to 600 K (Gm = −130.20~−248.82 kJ/mol) and the electronegativity of Bi2Te3 and Sb2Te3 and PbTe (X = 1.80~2.21) can be used as criteria to judge the potential of materials to become high-performance TE materials. For good TE compounds, Gm and X are required to meet the corresponding standards at the same time. By taking Gm = −130.20~−248.82 kJ/mol and X = 1.80~2.21 as screening criteria for high performance TE materials, it is found that the Gm and X of all 15 typical TE materials and 9 widely studied TE materials meet the requirement very well, except for the X of Mg2Si, and 64 pure substances are screened as potential TE materials from 102 atypical TE materials. In addition, with reference to their electronegativity, 44 pure substances are selected directly from a thermochemical data book as potential high-performance TE materials. A particular finding is that several carbides, such as Be2C, CaC2, BaC2, SmC2, TaC and NbC, may have certain TE properties. Because the Gm and X of pure substances can be easily found in thermochemical data books and calculated using the X of pure elements, respectively, the Gm and X of materials can be used as good high-throughput screening criteria for predicting TE properties.

1. Introduction

Thermoelectric (TE) materials have received widespread attention around the world due to their ability to convert heat and electricity directly to each other. Improving their thermoelectric conversion efficiency or finding materials with high thermoelectric properties is a very important goal of thermoelectric material research. In principle, the thermoelectric figure of merit Z = ɑ2σ/K or ZT = ɑ2σT/K (ɑ, σ, K and T are the Seebeck coefficient, electrical conductivity, thermal conductivity and absolute temperature of the material, respectively) is not only a parameter used to evaluate the performance of a TE material but also a theoretical basis for exploring high-performance TE materials. However, since the ZT of thermoelectric materials usually varies significantly with carrier concentration and temperature, thermoelectric materials with an unknown optimum doping concentration and the maximum figure of merit can only be evaluated by preparing numerous samples with different doping concentrations to measure and analyze the parameters over a wide temperature range. Obviously, the time from material composition design, weighing, synthesis and sintering to thermoelectric property testing and a performance analysis will take as little as three months or as much as one year. Therefore, it is difficult to use Z or ZT values to quickly analyze and judge a large number of unknown materials one by one. With the implementation of genetic engineering projects in recent years, the screening of thermoelectric materials using high-throughput calculations has received widespread attention worldwide. The mainstream high-throughput screening methods are theoretical calculations based on density functional theory and the Boltzmann equation, which establishes the relationship between the lattice structure and the thermoelectric transport coefficient of a material, and uses descriptors such as low thermal conductivity, thermoelectric superiority or the power factor to characterize the thermoelectric properties of the material. For example, elastic properties are used to efficiently evaluate the intensity of anharmonicity and lattice thermal conductivity for the high-throughput and efficient screening of thermoelectric materials with low lattice thermal conductivity [1]. But the high-throughput calculations and screening of high-performance thermoelectric materials also face two important difficulties: (1) precise calculations of the electrical and thermal properties of the materials are difficult and time-consuming, and (2) the existing high-throughput methods for evaluating the electrical and thermal properties of the materials have limitations.
Therefore, there is an urgent need for a simple and effective method to make a preliminary determination of the level of thermoelectric properties of a material or a criterion to determine its potential to become a high-performance thermoelectric material. Earlier, based on the thermoelectric figure of merit (Z) proportional to the previously derived material parameter β (see Equations (1) and (2)), Ioffe proposed a method for finding high-performance TE materials using the β value [2].
ZT = [ ξ ( s + 5 2 ) ] 2 ( β exp ξ ) 1 + ( s + 5 2 )
β = 5.745 × 10 6 μ c K L m * m 3 / 2 T 5 / 2
where ξ, s, μc, KL, m*, and m are the reduced Fermi level, scattering factor, carrier mobility, lattice thermal conductivity, effective mass, and the mass of the free electron, respectively. In Formula (2), the effective mass m*, lattice thermal conductivity KL and carrier mobility μc are generally weakly dependent on the carrier concentration, so the β parameter can be used to initially determine the thermoelectric properties of a material, even for samples that are not optimally doped. Accordingly, Ioffe believes that an effective way to find thermoelectric materials is to first screen the materials for lattice thermal conductivity and then make a further determination by measuring the (m*)3/2μc value of the materials. It is clear that this method avoids the requirement of the optimal doping of the sample, and it is much simpler than the method that uses the original formula of the thermoelectric figure of merit Z or ZT. Nevertheless, the determination of the β parameter involves the measurement of carrier mobility (μc) and effective mass (m*), both of which are more complex to measure than the Seebeck coefficient (ɑ) and electrical conductivity (σ), thus limiting the practical application of this approach. It is also essential to note that, while the variation in the β parameter with carrier concentration is much less pronounced than the thermoelectric figure of merit, it is not a constant.
Based on many years of practice, a number of useful laws, for example, heavy atomic mass [3], a large Fermi surface complexity factor [4], multiple energy valley degeneracies [5,6] or a complex Fermi surface structure [7], the appropriate carrier concentration [8], resonance energy levels [9,10], the energy-filtering effect [11], strong phonon scattering [12], strong anharmonic effects [13] and Phonon Glass–Electron Crystal (PGEC) properties [14], have been gradually obtained that summarize the physical properties of a good thermoelectric material. Among them, the use of materials with a high average relative atomic mass to improve m*/KL and, thus, thermoelectric properties as criteria for selecting thermoelectric materials was first proposed by Goldsmid [15]. This rule was supplemented in 1995 by Slack et al., who noted the relationship of the electronegativity (X) of compounds with mobility, the effective mass and the forbidden band width, and they proposed the use of the electronegativity of compounds as a metric for the first screening of thermoelectric materials. This law can be briefly stated as follows: (1) The greater the sum of the atomic numbers of a compound, the larger the cell size, and the lower the thermal conductivity in general. (2) The smaller the electronegativity of a compound, the larger the product of the effective mass and mobility generally [16]. While the laws summarized by Goldsmid and Slack are useful for a preliminary judgment of element choices for thermoelectric materials, they do not provide sufficient insight into the effects of the elements in the periodic table on the thermoelectric properties of materials. The other parameters listed above suffer from similar problems as β. They all require extensive experimental measurements or calculations to be obtained. They cannot be used as simple, fast and effective criteria to judge whether a compound or alloy has a high thermoelectric performance or thermoelectric figure of merit.
Obviously, the ideal way to find promising thermoelectric materials in a wide variety of materials is to make preliminary judgments based on the periodic table of the elements and the known basic physical properties of the elements. Moreover, it is a fact that there is a wealth of molar Gibbs free energy (Gm) data or thermochemical data of pure substances and convenient calculation methods for molar Gibbs free energy. In this paper, firstly, the rationality of using molar Gibbs free energy to evaluate the thermoelectric properties of materials is described. Then, the molar Gibbs free energy of a series of typical and atypical pure compound thermoelectric semiconductor materials is shown, the electronegativity (X) of the corresponding materials is calculated, and the change rule is analyzed. A new method using the molar Gibbs free energy and electronegativity of pure compounds as a fast and high-throughput preliminary screening method for thermoelectric materials is discussed.

2. Basal Principle

2.1. Fermi Level as a Criterion for High-Performance Thermoelectric Materials

This paper assumes that there is only one type of carrier that obeys the Fermi–Dirac statistical distribution, the isoenergy surface is spherical, the energy band is parabolic, the relaxation time approximation can be used to describe the scattering process in the crystal, and the contribution of the drag effect can be neglected; furthermore, the Fermi level EF is considered an independent variable. The relationship between the reduced Fermi level ξ (ξ = EF/kBT), the Seebeck coefficient (ɑ) and the conductivity (σ) of a material under different degenerate conditions can be seen in Table 1.
As mentioned above, in order to obtain a high thermoelectric figure of merit, Ioffe proposed [15] the use of the β factor to predict the thermoelectric performance of a material. It is believed that the greater the β value, the higher the thermoelectric performance of the material. On this basis, in 1959, Chasmar and Stratton used the Fermi–Dirac statistic to rigorously calculate the dependence of the dimensionless value ZTmax on the reduced Fermi level ξ (ξ = EF/kBT), the β factor and the scattering factor (s). The results are shown in Table 2 [17,18].
It can be found that the ZTmax value increases with an increase in the scattering factor (s) and the β value, but it increases with a decreasing optimal reduced Fermi level (ξopt). In addition, in correspondence to the different scattering mechanisms, the optimal reduced Fermi level (ξopt) also has a certain range of variation. When s = −1/2 and 1 ≤ β ≤ 5, the ZTmax is between 1.8 and 4.6, and the ξopt is between −1.0 and −2.4. An inverse relationship can also be seen between ZTmax and ξopt.
Furthermore, in 1972, Ure [18,19] used a two-band model and disregarded the effects of multi-valley and non-spherical isoenergetic surfaces to avoid excessive complexity. In the study, Ure considered the effects of lattice thermal conductivity and bipolar diffusion and adopted the elastic constant (1.7 × 1011 Nm−2), the deformation potential constant D (7 eV) and the effective mass m* (0.014 m) of silicon. For the case where acoustic phonon scattering predominates, Ure used this method to estimate the actual optimal values of thermoelectric materials. The results indicated that the dimensionless optimal upper limit was ZTmax ≈ 8, corresponding to the optimal reduced Fermi level ξopt ≈ −3.0.
As explained in the previous paragraph, it should be emphasized that, although the formulas in Table 1 are based on the above assumptions, the conclusion that the Fermi level and the scattering factor have a decisive effect on the thermoelectric properties of a material is universal. Based on the research results of the related literature on the effects of two kinds of carriers (including holes and electrons), an asymmetric band structure and a dual-band structure (considering a conduction band and a valence band) on thermoelectric properties, it can be confirmed that the Fermi level and scattering factor have a decisive influence on the thermoelectric properties of materials [20,21,22,23]. In addition, the typical high-performance thermoelectric materials known to us, such as bismuth telluride, antimony telluride and lead telluride, have more energy valleys. Therefore, the conclusion that Fermi levels and scattering factors have a decisive effect on the thermoelectric properties of materials is universal.
Summarizing the research results reported above, it should be possible to give a preliminary judgment on the thermoelectric performance of a material based on the Fermi level (EF) or reduced Fermi level (ξ). Therefore, the relationship between the Gibbs free energy and the Fermi level (EF) of materials is discussed.

2.2. The Relationship between Molar Gibbs Free Energy and Fermi Level

According to thermodynamics theory, the Gibbs free energy G(T, P, N) of a material system has an extensive property, where T, P and N are the absolute temperature, pressure and moles of a pure substance. The G(T, P, N) of the system is equal to the product of the number of moles of the substance (N), and the molar Gibbs free energy Gm(T, P) is equal to the chemical potential (μ) of that substance. This can be expressed as [24,25]
μ = G m ( T , P ) = G ( T , P , N ) N = H TS
where Gm(T, P) (Gm for short), H and S are the Gibbs free energy, enthalpy and entropy per mole of a pure substance.
At 0 K, the Fermi level (EF) of a pure substance is equal to its chemical potential μ [24], namely,
EF = μ
By substituting Formula (3) into Formula (4), we can obtain the following formula:
E F   = μ = G m ( T , P ) = H TS
In Formula (5), it appears that the Fermi level can be solved using the molar Gibbs free energy of the material. However, in addition to the fact that the two are equal at absolute temperatures and not necessarily exactly equal at other temperatures, the ground states calculated for the two are also different. EF has a ground state temperature of 0 K. That is, the Fermi energy is defined as the energy of the topmost filled level in the ground state of the N electron system. The ground state is the state of the N electron system at absolute zero. However, by definition, for a homogeneous crystal with a uniform temperature, S is calculated as [26]
S ( T ) = kln Ω C ( T = 0 ) + 0 T ( C p / T ) dT
where Cp(T) and Ωc are the isobaric heat capacity and the thermodynamic probability of the material, respectively. For a perfect crystal, Ωc(T = 0) = 1; that is, S(T = 0) = 0.
The enthalpy H(T, P) (H for short) of a pure substance is described entirely by independent internal variables T and P. The state function H(T, P) is determined when the pressure P is constant, except for additional uncertain and arbitrarily selected constants. In other words, for any system, the absolute value of enthalpy (H) cannot be determine. For this reason, different books use different conventions for the zero of H, such as the data cited here stating that the standard enthalpy H of all pure elements in its reference phase is zero at P = 1 bar and T = 298.15 K. Therefore, when using enthalpy values from different sources, we must pay attention to the standard state of the reference phase. For pure material enthalpy, its calculation formula can be expressed as [26]
H ( T ) = H ( 298 . 15 ) + 0 T t 1 C p ( T ) dT + Δ H t 1 + T t 1 T t 2 C p ( T ) dT + Δ H t 2 +
where H(298.15) is the enthalpy of the formation of pure matter at 1 bar and 298.15 K, Cp(T) is the temperature function of the heat capacity, and ΔHt1 is the enthalpy of the phase transition at temperature T = Tt1. The corresponding entropy calculation Formula (6) can also be expressed as follows:
S ( T ) = S ( 298 . 15 ) + 0 T t 1 C p ( T ) T dT + Δ H t 1 T t 1 + T t 1 T t 2 C p ( T ) T dT + Δ H t 2 T t 2 +
S ( 298 . 15 ) = S ( T = 0 ) + 0 298 . 15 dH / T
S = kln Ω C ( T = 0 ) + 0 T ( C p / T ) dT
For a complete crystal, at T = 0 K, S = 0, which is Ω = 1.
S ( 298 . 15 ) = 0 298 . 15 dH / T
Then, the molar Gibbs free energy of a pure substance at 1 bar can be calculated as follows:
G m ( T ) = G m ( 298 . 15 ) + 0 T t 1 C p ( T ) dT + Δ H t 1 + T t 1 T t 2 C p ( T ) dT + Δ H t 2 T ( 0 T t 1 C p ( T ) T dT + Δ H t 1 T t 1 + T t 1 T t 2 C p ( T ) T dT + Δ H t 2 T t 2 + )
G m ( 298.15 ) = H ( 298.15 ) TS ( 298.15 )
In Equation (13), H(298.15) and S(298.15) are the standard enthalpy and the standard entropy of the pure substance, respectively. Therefore, G(T) function values are also involved in the H(T) convention. Thus, the molar Gibbs free energy of the reference phase of an element E at 198.15 K and 1 bar is given by using the following formula:
G m ( 298.15 ) = TS ( 298 . 15 )
If there is no phase transition, Equation (12) can be simplified as follows:
G m ( T ) = G m ( 298.15 ) + 0 T C p ( T ) dT T 0 T C p ( T ) T dT
Therefore, the chemical potential (μ) or Fermi level (EF) cannot be calculated simply from the data of the molar Gibbs free energy. However, because of the inevitable relationship between the two, we can still summarize the inevitable relationship between the change law of the molar Gibbs free energy data of thermoelectric materials and the thermoelectric properties of materials and use it as one of the methods for the high-throughput screening of thermoelectric materials.
In addition, as mentioned above, Slack et al. noted the relationship between the electronegativity and mobility, effective mass and band gap width in compounds, and they proposed using the electronegativity of a compound as a metric for a preliminary screening of thermoelectric materials [16]. To this end, following the example of Bulter and Ginley et al., the electronegativity X of the semiconductor compound AnBm is calculated using the geometric mean value of Mulliken electronegativity [27,28]:
X = ( X A n X B m ) 1 / ( n + m )
where XA and XB are the electronegativity of pure elements A and B, respectively.

3. Results and Discussion

Except for special emphasis, the molar Gibbs free energy (Gm) data are selected from Ihsan Barin’s Thermochemical Data of Pure Substances [26] or the Handbook of Inorganic Thermodynamics Data [29].

3.1. Molar Gibbs Free Energy (Gm) of Pure Elements Listed in the Seebeck and Meissner Sequences

The Gms of the substances listed in the Seebeck and Meissner sequences [30] are shown in Table 3. The elements with high Seebeck coefficients listed in both sequences are Bi and Sb. Their Gm values are between −13.572 and −38.325 kJ/mol. Moreover, it is found that their Gm values decrease with an increase in the temperature (the absolute value increases). If Gm = −13.572 and −38.325 kJ/mol are taken as screening criteria, all elements in Table 3, except for element C, may have certain TE properties in the appropriate temperature range, of which the difference is that their optimal working temperature zones are different. At room temperature, only the Gm values of Na, U, Sn, Cd and Au meet the requirements. Although the Gm values of K, Hg, Pb and Cs at 298.15 K are within the above range, their absolute Gm values are larger or comparable to the Gm values of Bi or Sb at higher temperatures, so it is judged that these four elements may have better thermoelectric properties at slightly higher temperatures.

3.2. Molar Gibbs Free Energy and Electronegativity of 15 Typical TE Materials

Table 4 shows the Gm and X of 15 typical TE materials [31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51]. It can be seen that, in the range of 298.15~600 K, the optimal operating temperature range of Bi2Te3 and Sb2Te3, their Gm values are between −130.196 and −248.819 kJ/mol. If these data, or Gm = −130.196~−248.819 kJ/mol, are used as the screening criteria, it can be seen that all the above typical materials have good thermoelectric properties in a certain temperature range, indicating that the Gm of Bi2Te3 and Sb2Te3 at 298.15~600 K is feasible as the basic standard for a preliminary screening of high-thermal-power-factor thermoelectric materials.
Table 5 lists the electronegativity of 15 typical thermoelectric compounds calculated using Equation (16) and the electronegativity data of the elements [52,53,54]. It can be seen that, since the same element has different electronegativity values, the electronegativity of the corresponding compound is not unique. In addition, the electronegativity of Bi2Te3 and Sb2Te3 are very close. Considering that PbTe is a typical medium-temperature thermoelectric material, the electronegativity values of Bi2Te3, Sb2Te3 and PbTe, that is, X = 1.80~2.21, are used as the criteria for screening thermoelectric materials. It can be seen that all other materials, except for Mg2Si, have electronegativity values that meet this requirement, indicating that X = 1.80–2.21 is a suitable criterion.

3.3. Molar Gibbs Free Energy and Electronegativity of 102 Atypical TE Materials

The temperature dependences of the Gm of 102 atypical pure compounds are obtained from References [26,29]. The electronegativity X of the 102 pure atypical compounds are calculated using Equation (16). They are shown in Table 6. If Gm = −130.196~−248.819 kJ/mol is used as the screening criterion for good TE materials, 67 compounds are screened. It can be found that, in addition to Cu2S, Cu2Te, Ag2S, Ag2Se, Ag2Te, SnTe and PbSe, which have been widely investigated as high-performance TE materials [55,56,57,58,59,60,61,62], Bi2S3, Sb2S2, Mn3Si, CoSb2, MoSi2, MnS, MnSe, MnTe2, FeS, FeS2, FeSe0.96, FeTe0.9, FeTe2, CoS0.89, CoS2, NiSe1.05, NiSe1.143, NiSe1.25, NiSe2, NiTe, NiTe1.1, NiS2, NiSe2, CuS, InSb, GeS, GeSe, SnS, PbS, AgP2, AgP3, BeS, Be2C, Ba2C, AlAs, AlP, AlSb, CaC2, CaH2, CaPb, Ca2Pb, CaSi, NaTe, NaTe3, NbC, NbSi2, InSe, CaSi2, Ca2Si, CaSn, CaZn, CaZn2, CrS, CrSi2, GaP, GaSb, GaSe, GaTe, InP, InS, CuO and Cu2O may be good TE materials at suitable temperature ranges. If the Gm of Bi2Te3 at a temperature of 298–800 K, or Gm = −1.61~−3.36 eV, is used as the standard, it can be found that the other 10 compounds, namely, TiS, MoS2, WS2, MnS2, CoP3, CaTe, FeO, NiO, CdO and SnO, may be TE materials.
If X = 1.80~2.21 is used as the screening criterion of high-performance TE materials, 67 compounds are screened out. A comparison of the screening results of the two methods shows that their results are not completely consistent, although most of them are. For compounds that meet the Gm screening criteria, the main difference is reflected in alkali metal and alkaline earth metal compounds. These compounds, such as the alkali metal compounds NaTe and NaTe2 and the alkaline earth metal compounds CaH2, CaPb, Ca2Pb, CaSi, Ca2Si, CaSn, CaTe, CaZn and CaZn2, are less electronegative than the screening criteria. The X value of some transition metal compounds, such as Mn3Si, is also lower than the screening criterion. Transition metal oxides or sulfides, such as FeO, CuO, NiO, CdO, SnO and NiS2, have a larger X value than the screening criteria due to the high electronegativity of O or S. Therefore, although they have TE properties, they are not very good TE materials. So, a bigger X is not better. If both Gm and X are met as screening criteria, a total of 60 pure compounds have the potential to become high-quality TE materials. They are Cu2S, Cu2Te, Ag2S, Ag2Se, Ag2Te, SnTe, Bi2S3, Sb2S2, CoSb2, TiS, MoS2, MoSi2, WS2, MnS, MnSe, MnTe2, FeS, FeS2, FeSe0.96, FeTe0.9, FeTe2, CoS0.89, CoS2, CoP3, NiS, NiSe1.05, NiSe1.143, NiSe1.25, NiSe2, NiTe1.1, NiS2, NiSe2, CuS, InSb, GeS, GeSe, SnS, PbS, AgP2, AgP3, BeS, Be2C, Ba2C, AlAs, AlP, AlSb, CaC2, NbC, NbSi2, InSe, CaSi2, CrS, CrSi2, GaP, GaSe, GaTe, InP, InS and Cu2O.

3.4. Molar Gibbs Free Energy (Gm) and Electronegativity of Some Potential TE Materials

Based on the above Gm and X criteria, 44 possible high-performance thermoelectric compounds are screened directly from the pure substance thermochemical data book [29]. Their electronegativity values are calculated according to Formula (16). The results are presented in Table 7. There are several compounds, such as GeS2, MgB4, Mo3Si, OsSe2, Pd4S, PtBr2, PtI4 and ReS2, whose electronegativity values deviate from the screening criteria.

3.5. The Procedure for the High-Throughput Screening of TE Materials with Gibbs Free Energy and Electronegativity

Because the molar Gibbs free energy of a compound is easily found in the thermochemical data book or calculated, and its electronegativity is easily calculated using the geometric mean value of Mulliken electronegativity, the potential of a material as a high-performance thermoelectric material can be easily and quickly determined. In order to facilitate the screening of TE materials using molar Gibbs free energy (Gm) and electronegativity (X), a schematic diagram of the screening process is shown in Figure 1.
Additionally, one problem should be discussed. From the results of the analysis of the whole paper, the only typical TE compound that cannot meet the requirements of Gm = −130.20~−248.82 kJ/mol and X = 1.80~2.21 at the same time is Mg2Si. That is, its X does not meet the requirements because the electronegativity of the element Mg is too low. But why can Mg2Si become a typical thermoelectric material? The first reason is that the Gm of Mg2Si meets the requirements. The second reason is that X can meet the requirements by changing its composition, which is also the strategy adopted in the research process of Mg2Si TE materials. Therefore, when screening thermoelectric materials, Gm data can be used as the main data, supplemented by X data. It is a reasonable improvement strategy to adjust the composition of a TE material so that its X value meets the requirements.

4. Conclusions

Screening high-performance thermoelectric materials and improving their thermoelectric properties are important goals of thermoelectric materials research. Based on the objective relationship among the molar Gibbs free energy (Gm), the chemical potential, the Fermi level, the electronegativity (X) and the TE property of a material, a new method using molar Gibbs free energy (Gm) and electronegativity (X) for the high-throughput screening of thermoelectric materials is proposed. The molar Gibbs free energy of 15 typical TE materials, 9 widely studied thermoelectric materials and 93 atypical thermoelectric materials were obtained from a thermochemical data book. The electronegativities of the materials above were calculated using the geometric mean value of Mulliken electronegativity. The feasibility of using Gm and X as high-throughput screening thermoelectric materials is discussed in detail. The results are described below.
1. Because it is universal that Fermi levels and scattering factors have a decisive effect on the thermoelectric properties of materials, taking the molar Gibbs free energy Gm and electronegativity X as screening criteria for high-performance TE materials is reasonable.
2. The molar Gibbs free energy Gms of typical TE materials Bi2Te3 and Sb2Te3 range from −130.196 to −248.819 kJ/mol. The electronegativity Xs of Bi2Te3, Sb2Te3 and PbTe range from 1.80 to 2.21. If Gm = −130.20~−248.82 kJ/mol and X = 1.80~2.21 are used as screening criteria for high-performance TE materials, the Gm and X of all of 15 typical TE materials and 9 widely studied thermoelectric materials meet the requirements very well, except for the X of Mg2Si. It is indicated that Gm = −130.20~−248.82 kJ/mol and X = 1.80~2.21 are suitable criteria for screening high-performance TE materials.
3. For TE materials, such as Mg2Si, due to the extremely low electronegativity of the component elements, its X value cannot meet the requirements, but its Gm can meet the requirements very well. Gm data can be used as the main data, supplemented by X data. It is a reasonable improvement strategy to adjust the composition of a TE material so that its X value meets the requirements.
4. For good TE compounds, if Gm and X are required to meet the corresponding standards at the same time, and Gm = −130.196~−248.819 kJ/mol and X = 1.80~2.21 are used as screening criteria, 60 pure substances, including 9 widely studied TE materials, are screened as potential TE materials from 102 atypical TE materials.
5. With reference to their electronegativity, 44 pure substances are selected directly from the thermochemical data book as potential high-performance thermoelectric materials. A particular finding is that several carbides, such as Be2C, CaC2, BaC2 and NbC, may have certain TE properties.
6. Compared with Gm = −130.196~−248.819 kJ/mol, the elemental elements in the Seebeck or Meissner sequence are not good thermoelectric materials. This is consistent with the actual results.
7. The Gm of pure substances can be easily found in thermochemical data books, and the X of compounds can be calculated easily from the X of pure elements, so using Gm and X as high-throughput screening criteria for predicting thermoelectric properties is much more convenient than using the TE figure of merit Z or ZT or the Ab initio calculation method. This method requires no experiments and no first principle or Ab initio calculation.

Author Contributions

Conceptualization, Investigation, Supervision, Methodology, Funding acquisition, Formal analysis, Writing—original draft: G.X.; Data curation, Investigation, Writing—original draft: J.X., H.D., R.S., G.Z. and P.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The National Key Research and Development Program of China (Grant No. 2017YFF0204706); the Fundamental Research Funds for the Central Universities (Grant Nos. FRF-MP-18-005 and FRF-MP-19-005); and the Natural Science Foundation of Guizhou Province, China (Grant No. 20191168).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

All authors agree to publish the paper in this Journal.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jia, T. Binary Chalcogenides as Thermoelectric Materials: High-Throughput Computations and Property Studies. Ph.D. Thesis, University of Science and Technology of China, Hefei, China, June 2020. [Google Scholar]
  2. Ioffe, A.F. Semiconductor Thermoelements and Thermoelectric Cooling; Infosearch: London, UK, 1957. [Google Scholar]
  3. Zeier, W.G.; Zevalkink, A.; Gibbs, Z.M.; Hautier, G.; Kanatzidis, M.G.; Snyder, G.J. Thinking like a chemist: Intuition in thermoelectric materials. Angew. Chem. Int. Ed. 2016, 55, 6826–6841. [Google Scholar] [CrossRef] [PubMed]
  4. Gibbs, Z.M.; Ricci, F.; Li, G.; Zhu, H.; Persson, K.; Ceder, G.; Hautier, G.; Jain, A.; Snyder, G.J. Effective mass and Fermi surface complexity factor from ab initio band structure calculations. NPJ Comput. Mater. 2017, 3, 8. [Google Scholar] [CrossRef] [Green Version]
  5. Xin, J.; Tang, Y.; Liu, Y.; Zhao, X.; Pan, H.; Zhu, T. Valleytronics in thermoelectric materials. NPJ Quantum Mater. 2018, 3, 9. [Google Scholar] [CrossRef]
  6. Nolas, G.; Sharp, J.; Goldsmid, H. Thermoelectrics: Basic Principles and New Materials Developments; Springer: New York, NY, USA, 2001. [Google Scholar]
  7. Reshak, A.H.; Khan, S.A. Thermoelectric properties, electronic structure and optoelectronic properties of anisotropic Ba2Tl2CuO6 single crystal from DFT approach. J. Magn. Magn. Mater. 2014, 354, 216–221. [Google Scholar] [CrossRef]
  8. Shuai, J.; Mao, J.; Song, S.; Zhu, Q.; Sun, J.; Wang, Y.; He, R.; Zhou, J.; Chen, G.; Singh, D.J.; et al. Uning the carrier scattering mechanism to effectively improve the thermoelectric properties. Energy Environ. Sci. 2017, 10, 799–807. [Google Scholar] [CrossRef]
  9. Heremans, J.P.; Wiendlocha, B.; Chamoire, A.M. Resonant levels in bulk thermoelectric semiconductors. Energy Environ. Sci. 2012, 5, 5510–5530. [Google Scholar] [CrossRef]
  10. 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] [Green Version]
  11. Faleev, S.V.; Leonard, F. Theory of enhancement of thermoelectric properties of materials with nanoinclusions. Phys. Rev. B 2008, 77, 214304. [Google Scholar] [CrossRef] [Green Version]
  12. Wu, H.; Carrete, J.; Zhang, Z.; Qu, Y.; Shen, X.; Wang, Z.; Zhao, L.-D.; He, J. Strong enhancement of phonon scattering through nanoscale grains in lead sulfide thermoelectrics. NPG Asia Mater. 2014, 6, el08. [Google Scholar] [CrossRef]
  13. Chang, C.; Zhao, L.-D. Anharmoncity and low thermal conductivity in thermoelectrics. Mater. Today Phys. 2018, 4, 50–57. [Google Scholar] [CrossRef]
  14. Rowe, D.M. Thermoelectrics Handbook Macro to Nano; CRC Press: Boca Raton, FL, USA, 2006. [Google Scholar]
  15. Goldsmid, H.J. Thermoelectric applications of semiconductors. J. Electron. Control. 1955, 1, 218–222. [Google Scholar] [CrossRef]
  16. Slack, G.A. New materials and performance limits for thermoelectric cooling. In CRC Handbook of Thermoelectric; Rowe, D.M., Ed.; CRC Press: London, UK, 1995; Chapter 35. [Google Scholar]
  17. Chasmar, R.P.; Stratton, R.J. The thermoelectric Figure of Merit and its relationship to thermoelectric generator. J. Electron. Control. 1959, 7, 52. [Google Scholar] [CrossRef]
  18. Gao, M.; Zhang, J.S.; Rowe, D.M. Thermoelectric Transformation and Application; Weapon Industry Press: Beijing, China, 1995. [Google Scholar]
  19. Ure, R.W., Jr. Practical limits to the thermoelectrical Figure of merit. Energy Convers. 1972, 45–52. [Google Scholar] [CrossRef]
  20. Mahan, G.D. Figure of merit for thermoelectric. J. Appl. Phys. 1989, 65, 1834–1842. [Google Scholar] [CrossRef]
  21. Seeger, K. Semiconductor Physics; Peoples Education Press: Beijing, China, 1980; p. 336. [Google Scholar]
  22. Simon, R. Thermoelectric figure of merit of two band semiconductor. J. Appl. Phys. 1962, 33, 1830–1841. [Google Scholar] [CrossRef]
  23. Rittner, E.B.; Neumark, G.F. Theorectical bound on the thermoelectric figure of merit of two band semiconductors. J. Appl. Phys. 1963, 34, 2071–2077. [Google Scholar] [CrossRef]
  24. Su, R. Statistical Physics; Fudan University Press: Shanghai, China, 1990. [Google Scholar]
  25. Chen, P.; Yang, C.; Ma, B.; Wang, J.W. University Physics Handbook; Shandong Science and Technology Press: Jinan, China, 1985; p. 153. [Google Scholar]
  26. Barin, I. Thermochemical Data of Pure Substances; Science Press: Beijing, Chian, 2003. [Google Scholar]
  27. Butler, M.A.; Ginley, D.S. Prediction of flat band potentials at semiconductor-electrolyte interface from atomic electronegativities. J. Electrochem Soc. 1978, 125, 228–231. [Google Scholar] [CrossRef]
  28. Chen, J.H.; Feng, Q.M.; Lu, Y.P. Energy band model of electrochemical flotation and its application (I)–Theory and model of energy band of semiconductor-solution interface. Chin. J. Nonferrous Met. 2000, 10, 240–244. [Google Scholar]
  29. Liang, Y.; Che, Y. Handbook of Inorganic Thermodynamics Data; Northestern University Press: Evanstone, IL, USA, 1993. [Google Scholar]
  30. Ioffe, A.F.; Pan, J. Semiconductor Thermocouple; Science Press: Beijing, China, 1958. [Google Scholar]
  31. Sun, G.L.; Li, L.L.; Qin, X.Y.; Li, D.; Zou, T.H.; Xin, H.X.; Ren, B.J.; Zhang, J.; Li, Y.Y.; Li, X.J. Enhanced thermoelectric performance of nanostructured topological insulator Bi2Se3. Appl. Phys. Lett. 2015, 106, 053102. [Google Scholar] [CrossRef]
  32. Lv, T.; Li, Z.M.; Yang, Q.X.; Benton, A.; Zheng, H.T.; Xu, G.Y. Synergistic regulation of electrical-thermal effect leading to an optimized Synergistic regulation of electrical-thermal effect leading to an optimized thermoe lectric performance in Co doping n-type Bi2(Te0.97Se0.03)3. Intermetallics 2020, 118, 106683. [Google Scholar] [CrossRef]
  33. Xu, G.Y.; Niu, S.T.; Wu, X.F. Thermoelectric properties of p-type Bi0.5Sb1.5Te2.7Se0.3 fabricated by high pressure sintering method. J. Appl. Phys. 2012, 112, 073708. [Google Scholar] [CrossRef]
  34. Qin, D.L.; Pan, F.; Zhou, J.; Xu, Z.B.; Deng, Y. High ZT and Performance Controllable Thermoelectric Devices Based on Electrically Gated Bismuth Telluride Thin Films. Nano Energy 2021, 89, 106472. [Google Scholar] [CrossRef]
  35. Liu, W.; Tan, X.; Yin, K.; Liu, H.; Tang, X.; Shi, J. Qingjie Zhang and Ctirad Uher, Convergence of conduction bands as a means of enhancing thermoelectric performance of n-type Mg2Si1-xSnx solids solutions. Phys. Rev. Lett. 2012, 108, 166601. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, Q.; Zhao, X.B.; Yin, H.; Zhu, T.J. Thermoelectric performance of Mg2-xCaxSi compounds. J. Alloys Compd. 2007, 464, 9–12. [Google Scholar] [CrossRef]
  37. Tang, X.F.; Chen, L.D.; Goto, T.; Hirai, T.; Yuan, R.Z. Synthesis and thermoelectric property of filled skutterudite compounds CeyFexCo4-xSb12 by solid state reaction. J. Mater. Sci. 2001, 36, 5435–5439. [Google Scholar] [CrossRef]
  38. Shi, X.; Zhang, W.; Chen, L.; Yang, J.; Uher, C. Thermodynamic analysis of the filling fraction limits for impurities in CoSb3 based on ab initio calculations. Acta Mater. 2008, 56, 1733–1740. [Google Scholar] [CrossRef]
  39. Pei, Y.Z.; Shi, X.; Lalonde, A.; Wang, H.; Chen, L.D.; Snyder, G.J. Convergence of electronic bands for high performance bulk thermoelectrics. Nature 2011, V473, 66–69. [Google Scholar] [CrossRef] [Green Version]
  40. Yang, S.H.; Zhu, T.J.; Sun, T.; Zhang, S.N.; Zhao, X.B.; He, J. Nanostructure in high-performance (GeTe)x(AbSbTe2)100-x thermoelectric materials. Nanotechnology 2008, 19, 245707. [Google Scholar] [CrossRef] [PubMed]
  41. Cui, J.L.; Zhao, X.B. The design and property of pseudobinary alloys (PbTe)1-x-(SnTe)x with gradient composition. Mater. Lett. 2003, 57, 2466–2471. [Google Scholar] [CrossRef]
  42. Madar, N.; Givon, T.; Mogilyanski, D.; Gelbstein, Y. High thermoelectric potential of Bi2Te3 alloyed GeTe- rich phases. J. Appl. Phys. 2016, 120, 035102. [Google Scholar] [CrossRef]
  43. Gelbstein, Y.; Davidow, J.; Leshem, E.; Pinshow, O.; Moisa, S. Significant lattice thermal conductivity reduction following phase separation of the highly efficient GexPb1-xTe thermoelectric alloys. Physica Status Solidi B 2014, 251, 1431–1437. [Google Scholar] [CrossRef]
  44. Ben-Ayoun, D.; Sadia, Y.; Gelbstein, Y. High temperature thermoelectric properties evolution of Pb1-xSnxTe based alloys. J. Alloys Compd. 2017, 722, 33–38. [Google Scholar] [CrossRef]
  45. Meroz, O.; Ben-Ayoun, D.; Beeri, O.; Gelbstein, Y. Development of Bi2Te2.4Se0.6 Alloy for Thermoelectric Power Generation Applications. J. Alloys Compd. 2016, 679, 196–201. [Google Scholar] [CrossRef]
  46. Sadia, Y.; Madar, N.; Kaler, I.; Gelbstein, Y. Thermoelectric properties in the quasi-binary MnSi1.73-FeSi2 system. J. Electron. Mater. 2015, 44, 1637. [Google Scholar] [CrossRef]
  47. Huang, F.; Xu, G.; Ping, Z. Thermoelectric Properties of InxGe1-xTe fabricated by high pressure sintering method. J. Electron. Mater. 2015, 44, 1651–1655. [Google Scholar] [CrossRef]
  48. Liu, H.; Shi, X.; Xu, F.; Zhang, L.; Zhang, W.; Chen, L.; Li, Q.; Uher, C.; Day, T.; Snyder, G.J. Copper ion liquid-like thermoelectrics. Nat. Mater. 2012, 11, 422. [Google Scholar] [CrossRef] [PubMed]
  49. Zhao, L.D.; Lo, S.H.; Zhang, Y.; Sun, H.; Tan, G.; Uher, C.; Wolverton, C.; Dravid, P.V.; Knatzidis, M.G. Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Nature 2014, 508, 373. [Google Scholar] [CrossRef]
  50. Gao, J.; Zhu, H.; Mao, T.; Zhang, L.; Jiaxin, D.; Guiying, X. The effect of Sm doping on the transport and thermoelectric properties of SnSe. Mater. Res. Bulletin. 2017, 93, 366–372. [Google Scholar] [CrossRef]
  51. Zhang, L.-B.; Qi, H.-L.; Gao, J.-L.; Mao, T.; Di, J.-X.; Xu, G.-Y. Thermoelectric Properties of Mn1+x Te-Based Compounds Densified Using High-Pressure Sintering. J. Electron. Mater. 2017, 46, 2894–2899. [Google Scholar] [CrossRef]
  52. Yin, Y.; Gu, Y.; Chen, D.; Luo, X.; Wang, S.; Li, Z.; Liu, Y.; Liu, Z.; Xie, X.; Xi, Z.; et al. University Chemistry Handbook; Shandong Science and Technology Press: Jinan, China, 1985. [Google Scholar]
  53. Gordy, W.; Thomas, W.J.O. Electronegativities of the Elements. J. Chem. Phys. 1956, 24, 439–444. [Google Scholar] [CrossRef]
  54. Cotton, J.A.; Wilknson, G. Advances in Inorganic Chemistry; John Wiley & Sons: Hoboken, NJ, USA, 1972. [Google Scholar]
  55. Qiu, P.; Sh, X.; Chen, L. Cu-based thermoelectric materials. Material 2016, 3, 85–97. [Google Scholar] [CrossRef]
  56. Sridhar, K.; Chattopadhyay, K. Synthesis by mechanical alloying and thermoelectric properties of Cu2Te. J. Alloys Compd. 1998, 264, 293–298. [Google Scholar] [CrossRef]
  57. Wang, T.; Chen, H.-Y.; Qiu, P.-F.; Shi, X.; Chen, L.-D. Thermoelectric properties of Ag2S superionic conductor with intrinsically low lattice thermal conductivity. Acta Phys. Sin. 2019, 68, 090201. [Google Scholar] [CrossRef]
  58. Zhang, Y.; Wu, L.-H.; Zeng, J.-K.; Liu, Y.-F.; Zhang, J.-Y.; Xing, J.-J. Ferhat Marhoun and Nagao Jiro, Thermoelectric and transport properties of β-Ag 2 Se compounds. J. Appl. Phys. 2000, 88, 813. [Google Scholar]
  59. Hu, H.; Xia, K.; Wang, Y.; Fu, C.; Zhu, T.; Zhao, X. Fast synthesis and improved electrical stability in n-type Ag2Te thermoelectric materials. J. Mater. Sci. Technol. 2021, 91, 241–250. [Google Scholar] [CrossRef]
  60. Jun, L. Microstructures and thermoelectric transports in PbSe-MnSe nano-composites. Acta Phys. Sin. 2016, 65, 107201. [Google Scholar]
  61. Yang, Q.; Lyu, T.; Li, Z.; Mi, H.; Dong, Y.; Zheng, H.; Sun, Z.; Feng, W.; Xu, G. Realizing widespread resonance effects to enhance thermoelectric performance of SnTe. J. Alloys Compd. 2021, 852, 156989. [Google Scholar] [CrossRef]
  62. Yang, Q.; Lyu, T.; Dong, Y.; Nan, B.; Tie, J.; Zhou, X.; Zhang, B.; Xu, G. Anion exchanged Cl doping achieving band sharpening and low lattice thermal conductivity for improving thermoelectric performance in SnTe. Inorg. Chem. Front. 2021, 8, 4666–4675. [Google Scholar] [CrossRef]
Figure 1. Flowchart for screening high-performance of TE materials using molar Gibbs free energy (Gm) and electronegativity (X).
Figure 1. Flowchart for screening high-performance of TE materials using molar Gibbs free energy (Gm) and electronegativity (X).
Materials 16 05399 g001
Table 1. The Seebeck coefficients and electrical conductivities under different ξ values.
Table 1. The Seebeck coefficients and electrical conductivities under different ξ values.
ξSeebeck Coefficient (ɑ)Electrical Conductivity (σ)
Any ξ α p , n = ± k B e ( s + 5 2 ) F s + 3 2 ( ξ ) ( s + 3 2 ) F s + 1 2 ( ξ ) ξ σ = 16 π e 2 τ 0 ( 2 m * ) 1 / 2 ( k B T ) 3 / 2 + s 3 h 3 ( s + 3 2 ) F s + 1 / 2 ( ξ )
ξ << 1 α = ± k B e [ ( s + 5 2 ) ξ ] σ = 16 π e 2 τ 0 ( 2 m * ) 1 / 2 ( k B T ) 3 / 2 + s 3 h 3 Γ ( s + 5 2 ) exp ( ξ )
ξ > 4 α = ± π 2 3 k B e ( s + 3 2 ) ξ σ = 16 π e 2 τ 0 ( 2 m * ) 1 / 2 3 h 3 E F s + 3 / 2
Table 2. Correspondence amongst the optimal reduced Fermi level (ξopt), the scattering factor (s), the material parameter (β) and the figure of merit (ZTmax).
Table 2. Correspondence amongst the optimal reduced Fermi level (ξopt), the scattering factor (s), the material parameter (β) and the figure of merit (ZTmax).
Sβ0.10.20.51.02.05.0
−1/2ZTmax0.40.611.82.94.6
ξopt−0.1−0.1−0.8−1.0−1.7−2.4
1/2ZTmax0.81.12.02.83.85.6
ξopt0.80.35−0.4−0.9−1.6−2.3
3/2ZTmax1.42.02.83.85.06.6
ξopt1.20.5−0.3−0.8−1.5−2.2
Table 3. Temperature dependence of the Gm of elementary substances listed in Seebeck and Meissner sequences.
Table 3. Temperature dependence of the Gm of elementary substances listed in Seebeck and Meissner sequences.
Seebeck
Sequence
Meissner
Sequence
α/
(μV/K)
Gm/kJ/mol
298 K300 K400 K500 K600 K700 K800 K900 K1000 K
BiBi−70−16.916−17.021−23.098−29.848−38.325−48.221−58.528−69.185−80.151
CoCo−17.5−8.957−9.012−12.414−16.483−21.119−26.25−31.888−37.932−44.361
NiNi−18−8.907−8.962−12.371−16.496−21.243−26.557−32.322−38.478−44.99
K−12−19.281 −19.401 −26.798 −35.244 −44.309 −53.879 −63.876 −74.247 −84.952
PdPd−6−11.277−11.347−15.541−20.408−25.824−31.704−37.99−47.973−51.611
Na−4.4−15.341 −15.437 −21.246 −28.337 −36.044 −44.251 −52.878 −61.867 −71.178
PtPt−3.3−12.412−12.489−17.061−22.3−28.083−34.326−40.97−38.836−55.299
U −14.994−15.088−20.561−26.783−33.648−41.095−49.086−57.601−66.792
Hg−3.4−22.629 −22.770 −30.795 −39.514 −48.780 −51.584
Al−0.6−8.430 −8.483 −11.699 −15.565 −19.973 −24.852 −30.150 −35.835 −42.645
Mg−0.4−9.743 −9.803 −13.468 −17.790 −22.661 −28.008 −33.780 −39.937 −47.208
Pb−0.1−19.316 −19.436 −26.340 −33.944 −42.122 −51.588 −61.494 −71.776 −82.389
Sn0.1−15.264 −15.359 −20.911 −27.190 −35.389 −44.143 −53.304 −62.823 −72.658
Cs0.2−25.387 −25.544 −35.244 −45.751 −56.887 −68.542 −80.641 −93.127 −99.241
Y2.2−13.248 −13.330 −18.193 −23.742 −29.851 −36.437 −43.442 −50.824 −58.549
RhRh2.5−9.393−9.452−13−17.206−21.958−27.18−32.82−9.393−45.199
ZnZn2.9−12.412−12.489−17.055−22.283−28.061−34.393−42.14−50.28−67.574
C −1.712−1.722−2.449−3.473−4.789−6.386−8.248−10.358−12.7
AgAg1.5/2.4−12.724−12.803−17.471−22.79−28.639−34.939−41.635−48.684−56.057
AuAu1.5/2.7−14.161−14.249−19.398−25.198−31.525−38.297−45.457−52.962−60.779
CuCu2.0/2.6−9.888−9.949−13.654−17.998−22.862−28.17−33.865−39.904−46.255
W2.5/1.5−9.738 −9.798 −13.449 −17.729 −22.519 −27.742 −33.339 −39.271 −45.505
CdCd2.8−15.444−15.54−21.132−27.41−34.322−42.706−51.516−60.698−70.211
Mo5.9−8.525 −8.578 −11.819 −15.692 −20.084 −24.920 −30.142 −35.707 −41.583
FeFe16−8.133−8.184−11.317−15.14−19.559−24.513−29.946−35.891−42.302
As −10.646−10.712−14.674−19.276−24.4−29.968−35.924−42.225−48.837
SbSb35−13.572−13.657−18.609−24.216−30.355−36.949−43.946−51.31−61.128
Se1000−12.599 −12.678 −17.314 −22.738 −29.920 −37.691 −45.966 −54.682 −63.789
Table 4. Temperature dependence of the Gm of typical high-performance thermoelectric materials.
Table 4. Temperature dependence of the Gm of typical high-performance thermoelectric materials.
CompoundsGm/kJ/mol
298 K300 K400 K500 K600 K700 K800 K900 K1000 K
Bi2Se3−211.206 −212.171 −238.210 −267.142 −298.968 −332.723 −368.406 −406.018 −445.559
Bi2Te3−155.271 −155.271 −184.203 −215.064 −248.819 −285.466 −324.043
Sb2Se3−190.954 −190.954 −214.100 −241.103 −270.036 −300.897 −333.687 −369.370 −411.804
Sb2Te3−130.196 −130.196 −157.199 −187.096 −219.886 −255.569 −293.182
GeTe−75.354 −75.521 −85.337 −96.494 −108.745 −121.928 −135.925 −150.652
PbTe−101.263 −101.263 −113.801 −126.338 −140.804 −156.235 −171.666 −188.061 −205.420
MnTe−135.982 −136.947 −146.591 −158.164 −171.666 −185.167 −200.598 −216.029 −231.459
SnSe−115.730 −115.730 −125.374 −135.982 −148.520 −161.057 −175.523 −189.989 −205.420
CoSb3−110.908 −110.908 −129.231 −149.484 −171.666 −196.740 −222.779 −250.747 −279.680
Cu2Se−104.157 −104.157 −118.623 −136.947 −156.235 −177.452 −200.598 −223.744 −247.854
Mg2Si−102.228 −102.228 −110.908 −121.516 −133.089 −146.591 −161.057 −176.488 −193.847
MnSi−91.619 −91.619 −97.406 −104.157 −111.872 −120.552 −130.196 −140.804 −151.413
MnSi1.7−282.573 −281.609 −224.708 −193.847 −174.559 −162.986 −154.306 −149.484 −145.626
Mn5Si3−284.502 −283.537 −228.566 −198.669 −180.345 −168.772 −161.057 −156.235 −152.377
FeSi2−97.406 −97.406 −104.157 −112.836 −122.480 −133.089 −145.626 −158.164
Table 5. The electronegativity of 15 typical thermoelectric materials.
Table 5. The electronegativity of 15 typical thermoelectric materials.
ElementXAXBABXElementXAXBABX
Bi2Se31.672.4232.08Mg2Si1.21.9211.4
Bi2Te31.672.1231.921.21.74211.36
1.82.1231.971.21.8211.37
2.022.1232.071.22.44211.52
Sb2Se31.652.4232.071.322.44211.62
1.82.4232.14MnSi1.61.9111.74
1.822.4232.151.61.74111.67
2.052.4232.251.61.8111.7
Sb2Te31.652.1231.911.62.44111.98
1.82.1231.97MnSi1.71.61.911.71.78
1.822.1231.981.61.7411.71.69
2.052.1232.081.61.811.71.72
GeTe1.82.1111.941.62.4411.72.09
2.012.1112.05Mn5Si31.61.9531.71
2.022.1112.061.61.74531.65
PbTe1.552.1111.81.61.8531.67
1.62.1111.831.62.44531.87
1.82.1111.94FeSi21.641.74121.71
2.332.1112.211.641.8121.75
MnTe1.42.1111.711.641.9121.81
1.552.1111.81.642.44122.14
1.62.1111.831.81.74121.76
SnSe1.72.4112.021.81.8121.8
CoSb31.72.05131.961.81.9121.87
Cu2Se1.82.4211.981.82.44122.2
FeSi21.831.9121.881.831.74121.77
1.832.44122.221.831.8121.81
Table 6. The X and the temperature dependence of the Gm of atypical thermoelectric compounds.
Table 6. The X and the temperature dependence of the Gm of atypical thermoelectric compounds.
CompoundsXGm/kJ/mol
298 K300 K400 K500 K600 K700 K800 K900 K1000 K
Bi2S32.19 −202.527 −203.491 −225.673 −250.747 −277.751 −308.612 −340.438 −374.192 −409.876
Sb2S32.19 −195.776 −196.740 −216.993 −240.139 −265.214 −293.182 −324.043 −359.726 −398.303
Mn3Si1.78 −241.103 −240.139 −193.847 −168.772 −154.306 −145.626 −139.840 −124.409 −135.018
CoSb21.93 −85.833 −85.833 −99.335 −114.765 −131.160 −149.484 −169.737 −189.989 −211.206
Mg3Sb21.49 −737.776 −732.954 −574.790 −484.135 −427.235 −387.694 −360.691 −340.438
MgSe1.70 −1517.022 −1507.378 −1156.331 −948.983 −813.965 −718.488 −649.050 −595.043 −553.573
MgTe1.59 −1126.435 −1119.684 −868.936 −722.345 −625.904 −558.395 −509.210 −472.562 −443.630
TiS2.00 −288.360 −289.324 −295.110 −302.826 −311.505 −321.150 −331.758 −342.367 −353.940
TiS22.15 −430.128 −431.093 −439.772 −450.381 −462.918 −476.420 −490.886 −507.281 −523.676
ZrS22.11 −600.829 −600.829 −609.509 −620.118 −632.655 −646.157 −660.623 −676.054 −692.449
ZrTe21.88 −330.794 −345.260 −378.050 −415.662 −458.096
TaS22.03 −376.121 −376.121 −395.409 −395.409 −407.947 −421.449 −435.915 −451.345 −467.740
MoS22.15 −295.110 −295.110 −301.861 −311.505 −322.114 −333.687 −346.224 −359.726 −374.192
MoSi22.12 −137.911 −137.911 −145.626 −155.271 −165.879 −177.452 −190.954 −205.420 −219.886
WS22.15 −278.715 −278.715 −286.431 −296.075 −306.683 −318.256
MnS1.87 −214.100 −214.100 −216.029 −217.957 −219.886 −220.851 −222.779 −276.787 −278.715
MnS22.15 −253.641 −253.641 −265.214 −277.751 −292.217 −308.612
MnSe1.83 −198.669 −198.669 −208.313 −219.886 −232.424 −244.961 −259.427 −273.893 −289.324
MnTe21.92 −168.772 −168.772 −185.167 −202.527 −221.815 −242.068
FeS2.20 −119.587 −119.587 −126.338 −135.982 −146.591 −158.164 −170.701 −184.203 −198.669
FeS22.17 −187.096 −187.096 −193.847 −201.562 −211.206 −221.815 −233.388 −246.890 −260.392
FeSe0.962.01 −87.762 −87.762 −95.477 −105.121 −115.730 −127.303 −140.804 −155.271 −170.701
FeTe0.91.88 −47.256 −47.256 −55.936 −66.545 −77.153 −89.690 −102.228 −115.730 −130.196
FeTe21.93 −102.228 −102.228 −113.801 −126.338 −141.769 −158.164 −175.523 −193.847 −199.634
CoS0.892.04 −109.943 −109.943 −115.730 −123.445 −131.160 −139.840 −149.484 −160.093 −170.701
CoS22.20 −173.594 −173.594 −182.274 −191.918 −202.527 −215.064 −228.566 −243.997 −259.427
CoP31.99 −309.577 −309.577 −321.150 −335.616 −351.046 −369.370 −388.659 −409.876 −432.057
NiS2.12 −103.192 −104.157 109.943 −116.694 −125.374 −135.018 −146.591 −158.164 −170.701
NiSe1.052.09 −97.406 −97.406 −106.085 −115.730 −126.338 −137.911 −151.413 −164.915 −179.381
NiSe1.1432.10 −102.228 −103.192 −111.872 −121.516 −133.089 −145.626 −159.128 −173.594 −188.061
NiSe1.252.11 −107.050 −107.050 −115.730 −126.338 −137.911 −151.413 −164.915 −180.345 −195.776
NiTe1.94 −59.794 −68.473 −89.690 −115.730
NiTe1.11.95 −82.940 −82.940 −92.584 −103.192 −115.730 −128.267 −141.769 −156.235 −171.666
NiS22.24 −152.377 −153.342 −161.057 −171.666 −183.238 −195.776 −210.242 −225.673 −241.103
NiSe22.18 −139.840 −139.840 −151.413 −164.915 −180.345 −196.740 −215.064 −233.388 −253.641
CuS2.18 −73.295 −73.295 −80.046 −88.726 −98.370 −108.979 −120.552 −132.125 −144.662
Cu2S2.01 −115.730 −115.730 −129.231 −145.626 −162.986 −183.238 −204.456 −226.637 −248.819
Cu2Te1.89 −81.975 −81.975 −97.406 −113.801 −133.089 −154.306 −176.488 −201.562 −226.637
Ag2S2.01 −75.224 −76.189 −91.619 −109.943 −130.196 −151.413 −174.559 −197.705 −222.779
Ag2Se1.98 −82.940 −82.940 −99.335 −120.552 −142.733 −166.843 −191.918 −217.957 −244.961
Ag2Te1.89 −81.975 −81.975 −98.370 −118.623 −141.769 −164.915 −189.989
ZnSb1.84 −43.399 −43.399 −53.043 −63.651 −75.224 −86.797 −100.299
InSb1.91 −55.936 −55.936 −65.580 −76.189 −88.726 −101.263 −114.765 −135.018 −156.235
GeS2.12 −95.477 −95.477 −103.192 −111.872 −121.516 −132.125 −143.698 −155.271 −169.737
GeSe2.08 −92.584 −92.584 −101.263 −110.908 −122.480 −134.053 −146.591 −160.093 −175.523
PbS2.00 −126.338 −126.338 −135.982 −140.804 −159.128 −172.630 −186.132 −200.598 −216.029
PbSe1.90 −130.196 −131.160 −141.769 −154.306 −167.808 −182.274 −196.740 −212.171 −228.566
SnS2.06 −131.160 −131.160 −139.840 −149.484 −160.093 −171.666 −184.203 −197.705 −212.171
SnTe1.89 −91.619 −91.619 −103.192 −114.765 −128.267 −142.733 −157.199 −172.630 −189.025
AgP21.99 −69.438 −69.438 −80.046 −91.619 −105.121 −119.587 −135.018 −152.377
AgP32.02 −101.263 −101.263 −113.801 −128.267 −145.626 −163.950 −185.167 −207.349
BeS1.94 −244.961 −244.961 −249.783 −254.605 −261.356 −268.107 −276.787 −285.466 −294.146
Be2C1.83 −121.516 −121.516 −124.409 −127.303 −132.125 −136.947 −143.698 −150.448 −157.199
BaC22.11 −101.263 −101.263 −110.908 −122.480 −135.982 −149.484 −164.915 −181.310 −198.669
AlAs1.82 −134.053 −134.053 −140.804 −149.484 −158.164 −167.808 −178.416 −189.025 −200.598
AlI32.20 −359.726 −359.726 −380.943 −395.409
AlP1.84 −178.416 −758.993 −184.203 −190.954 −197.705 −205.420 −215.064 −223.744 −233.388
Al2S32.04 −758.993 −758.993 −772.495 −788.890 −808.178 −828.431 −851.577 −875.687 −901.726
Al2Se31.99 −613.367 −613.367 −630.726 −650.979 −674.125 −699.200 −726.203 −755.136 −785.032
AlSb1.82 −69.438 −69.438 −77.153 −85.833 −94.512 −105.121 −115.730 −127.303 −139.840
BaS1.50 −483.171 −484.135 −491.851 −502.459 −513.068 −524.641 −537.178 −550.680 −564.182
CaC21.84 −81.011 −81.011 −88.726 −98.370 −109.943 −122.480 −135.982 −151.413 −166.843
CaH21.64 −189.025 −189.025 −193.847 −200.598 −207.349 −215.064 −223.744 −232.424 −243.032
CaPb1.53 −144.662 −145.626 −154.306 −163.950 −175.523 −188.061 −200.598 −215.064 −229.530
Ca2Pb1.33 −243.997 −244.961 −256.534 −270.036 −284.502 −300.897 −318.256 −336.580 −355.868
CaS1.58 −489.922 −489.922 −496.673 −504.388 −513.068 −522.712 −532.356 −543.929 −555.502
CaSi1.56 −164.915 −164.915 −169.737 −176.488 −183.238 −191.918 −200.598 −210.242 −220.851
CaSe1.55 −387.694 −388.659 −395.409 −404.089 −413.733 −424.342 −435.915 −447.488 −460.025
NaTe1.46 −199.634 −199.634 −208.313 −218.922 −230.495
NaTe31.75 −165.879 −165.879 −181.310 −198.669 −218.922 −240.139
NbC2.06 −149.484 −149.484 −153.342 −159.128 −164.915 −171.666 −179.381 −188.061 −196.740
NbSi22.16 −146.591 −146.591 −154.306 −163.950 −175.523 −188.061 −200.598 −215.064 −230.495
InSe1.90 −142.733 −142.733 −151.413 −162.021 −172.630 −185.167 −198.669 −212.171
CaSi21.81 −165.879 −165.879 −172.630 −180.345 −189.989 −200.598 −212.171 −224.708 −238.210
Ca2Si1.35 −233.388 −233.388 −243.032 −253.641 −266.178 −280.644 −295.110 −311.505 −328.865
CaSn1.40 −180.345 −180.345 −188.061 −196.740 −207.349 −217.957 −230.495 −242.068 −255.569
CaTe1.45 −315.363 −316.328 −324.043 −333.687 −344.295 −354.904 −366.477 −379.014 −391.552
CaZn1.28 −92.584 −93.548 −100.299 −108.979 −119.587 −130.196
CaZn21.40 −124.409 −124.409 −135.982 −149.484 −163.950 −180.345 −197.705 −216.993
CrS1.87 −174.559 −174.559 −182.274 −190.954 −199.634 −210.242 −221.815 −234.352 −246.890
CrSi22.12 −116.694 −116.694 −123.445 −131.160 −140.804 −151.413 −162.986 −175.523 −189.025
GaP1.95 −115.730 −115.730 −121.516 −128.267 −136.947 −145.626 −155.271 −164.915 −176.488
GaSb1.93 −64.616 −64.616 −72.331 −81.011 −89.690 −99.335 −109.943 −119.587
GaSe2.08 −180.345 −180.345 −188.061 −196.740 −207.349 −217.957 −229.530 −242.068 −255.569
GaTe1.95 −148.520 −149.484 −158.164 −168.772 −180.345 −192.883 −206.384 −220.851 −235.317
InP1.93 −106.085 −107.050 −113.801 −121.516 −130.196 −139.840 −150.448 −161.057 −172.630
InS2.11 −154.306 −154.306 −162.021 −171.666 −181.310 −191.918 −203.491 −216.029 −230.495
Ag2O2.25 −67.509 −67.509 −81.011 −95.477
MgFe2O42.48 −1474.588 −1490.018 −1528.595 −1577.780 −1635.645
CaTiO32.33 −1688.687 −1688.687 −1699.296 −1713.762 −1729.193 −1747.517 −1767.769 −1789.951 −1813.097
CaZrO32.30 −1796.702 −1796.702 −1808.275 −1822.741 −1839.136 −1858.424 −1878.677 −1900.858 −1924.004
SrTiO32.33 −1705.082 −1705.082 −1717.620 −1733.050 −1750.410 −1770.663 −1791.880 −1815.026 −1839.136
Mn2O32.49 −992.381 −992.381 −1004.919 −1020.349 −1037.709 −1057.961 −1079.178 −1102.324 −1127.399
FeO2.40 −290.288 −290.288 −297.039 −305.719 −314.399 −325.007 −335.616 −347.189 −359.726
Fe3O42.53 −1162.118 −1162.118 −1179.477 −1200.694 −1226.734 −1255.666 −1287.492 −1323.175 −1360.787
Fe2O32.58 −850.612 −850.612 −861.221 −874.723 −890.153 −909.442 −929.694 −951.876 −976.951
CuO2.47 −168.772 −168.772 −173.594 −180.345 −187.096 −194.811 −203.491 −213.135 −222.779
Cu2O2.20 −198.669 −198.669 −208.313 −220.851 −234.352 −248.819 −264.249 −280.644 −298.004
NiO2.47 −250.747 −250.747 −255.569 −263.285 −268.107 −276.787 −285.466 −295.110 −304.755
ZnO2.29 −363.584 −363.584 −368.406 −375.157 −381.908 −389.623 −398.303 −407.947 −417.591
CdO2.29 −274.858 −275.822 −281.609 −289.324 −297.039 −306.683 −316.328 −326.936 −338.509
SnO2.44 −302.826 −302.826 −309.577 −317.292 −325.972 −335.616 −346.224 −356.833 −368.406
Table 7. The X and the temperature dependence of the Gm of some potential TE compounds.
Table 7. The X and the temperature dependence of the Gm of some potential TE compounds.
CompoundsXGm/kJ/mol
2983004005006007008009001000
GeS22.24−182.972−183.134−192.925−204.453−217.428−231.658−247.008−263.375−280.682
GeSe2.08−92.363−92.509−101.125−111.046−122.039−133.951−146.673−160.125
GeSe22.18−146.525−146.733−159.119−173.371−189.166−206.292−224.597−243.967−264.315
InTe1.94−103.476−103.672−114.999−127.581−141.202−155.719−171.031−187.063
In2Te1.94−125.861−126.148−142.706−161.059−180.918−202.091
IrS22.13−153.634−153.762−161.728−171.474−182.705−195.21−208.836−223.457−239.01
MgB41.62−120.499−120.595−126.939−135.278−145.367−157.024−170.113−184.525−200.173
Mn7C31.83−180.432−180.875−208.533−242.455−281.592−325.235−372.878−424.138−478.719
Mn4N1.81−171.263−171.528−187.866−207.711−230.506
MnP1.83−132.428−132.549−139.843−148.441158.131−168.751−180.179−192.23−205.1
MnTe21.92−168.757−169.026−184.752−202.457−221.747−242.362
Mo2N1.97−100.425−100.524−107.912−117.106−127.873−140.024−153.411−167.92−183.458
MoSi22.12−138.21−138.331−145.876−155.167−165.905−177.882−190.944−204.972−219.876
Mo3Si1.78−133.356−133.554−145.65−160.148−176.604−194.726−214.306−235.189−257.255
NbC0.7021.99−126.67−126.729−130.466−135.13−140.577−146.704−153.431−160.699−168.459
NbC0.8252.02−134.083−134.144−138.012−142.865−148.549−154.954−161.993−169.602−177.729
NbSi22.16−146.353−146.482−154.5−164.221−175.344−187.665−201.037−215.349−230.514
Ni3Sn1.84−132.871−133.115−147.837−165.181−184.701−206.103−229.177−253.766
Ni3Sn21.86−208.67−208.991−228.08−250.01−274.284−300.573−328.641−358.316−389.461
OsP22.07−176.748−176.9−186.225−197.386−210.04−223.964−238.996−255.017−271.933
OsSe22.26−144.406−144.558−153.754−164.629−176.842−190.157−204.405−219.456−235.21
PdI22.04−116.886−117.219−136.408−157.551−180.296−204.414−229.74
PdS22.04−104.438−104.601−114.449−126.078−139.192−153.58−169.089−185.602−203.028
Pd4S2.09−122.901−123.236−143.131−166.052−191.482−219.081−248.607−279.88−312.759
PtBr22.24−116.346−116.445−122.991−131.511−141.651−153.177−165.924
PtI42.24−126.692−127.027−147.125−170.65−197.119−226.212
PtS22.08−132.725−132.864−141.394−151.706−163.501−176.571−190.762−205.958−222.065
ReS22.24−196.745−196.858−203.987−212.896−223.29−234.958−247.748−261.542−276.248
ReSi22.20−112.454−112.591−121.029−131.188−142.771−155.57−169.433−184.245−199.912
Re2Te52.01−169.237−169.704−197.679−230.246−266.6−306.209−348.694−393.774−441.228
RuSe22.01−185.827−185.979−195.29−206.502−219.273−233.369−248.618−264.893−282.093
Ta2Si1.92−156.956−157.151−168.802−182.292−197.3−213.601−231.031−249.465−268.806
TaSi22.16−135.905−136.01−142.703−151.174−161.121−172.329−184.638−197.924−212.09
TiSi2.04−144.299−144.39−150.02−156.89−164.797−173.598−183.187−193.483−204.421
TiSi22.16−152.519−152.632−159.805−168.782−179.278−191.085−204.051−218.057−233.01
USi22.02−154.154−154.306−163.677−175.054−188.12−202.635−218.415−235.315−253.222
VSi22.13−168.213−168.323−175.267−183.976−194.162−205.616−218.183−231.744−246.204
WSi22.12−112.088−112.207−119.641−128.798−139.381−151.181−164.054−177.858−192.528
ZrSi1.91−172.148−172.256−178.783−186.49−195.17−204.68−214.918−225.803−237.276
ZrSi22.07−180.742−180.874−189.059−198.947−210.238−222.727−236.266−250.746−266.078
SmC21.94−122.441−122.604−132.522−144.333−157.717−172.445−188.346−205.292−223.18
TaC2.06−156.734−156.812−161.656−167.544−174.326−181.884−190.128−198.986−208.402
ZnSe2.00−179.949−180.08−187.926−197.097−207.337−218.474−230.388−242.988−256.205
ZnTe1.87−142.447−142.591−151.162−161.036−171.984−183.858−196.552−209.987−224.099
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

Xu, G.; Xin, J.; Deng, H.; Shi, R.; Zhang, G.; Zou, P. High-Throughput Screening of High-Performance Thermoelectric Materials with Gibbs Free Energy and Electronegativity. Materials 2023, 16, 5399. https://doi.org/10.3390/ma16155399

AMA Style

Xu G, Xin J, Deng H, Shi R, Zhang G, Zou P. High-Throughput Screening of High-Performance Thermoelectric Materials with Gibbs Free Energy and Electronegativity. Materials. 2023; 16(15):5399. https://doi.org/10.3390/ma16155399

Chicago/Turabian Style

Xu, Guiying, Jiakai Xin, Hao Deng, Ran Shi, Guangbing Zhang, and Ping Zou. 2023. "High-Throughput Screening of High-Performance Thermoelectric Materials with Gibbs Free Energy and Electronegativity" Materials 16, no. 15: 5399. https://doi.org/10.3390/ma16155399

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