Next Article in Journal
Fabrication of Au–Ag Bimetallic Nanoparticles Using Pulsed Laser Ablation for Medical Applications: A Review
Previous Article in Journal
Donor-Acceptor Copolymers with 9-(2-Ethylhexyl)carbazole or Dibenzothiophene-5,5-dioxide Donor Units and 5,6-Difluorobenzo[c][1,2,5]thiadiazole Acceptor Units for Photonics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Lattice Thermal Conductivity of Mg3(Bi,Sb)2 Nanocomposites: A First-Principles Study

1
School of Science, Harbin Institute of Technology, Shenzhen 518055, China
2
The State Key Laboratory of Nonlinear Mechanics, Institute of Mechanics, Chinese Academy of Sciences, Beijing 100190, China
3
Guangdong Aerospace Research Academy, Guangzhou 511458, China
4
Department of Modern Mechanics, University of Science and Technology of China, Hefei 230026, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2023, 13(22), 2938; https://doi.org/10.3390/nano13222938
Submission received: 3 October 2023 / Revised: 7 November 2023 / Accepted: 9 November 2023 / Published: 13 November 2023

Abstract

:
Mg3(BixSb1−x)2 (0 ≤ x ≤ 1) nanocomposites are a highly appealing class of thermoelectric materials that hold great potential for solid-state cooling applications. Tuning of the lattice thermal conductivity is crucial for improving the thermoelectric properties of these materials. Hereby, we investigated the lattice thermal conductivity of Mg3(BixSb1−x)2 nanocomposites with varying Bi content (x = 0.0, 0.25, 0.5, 0.75, and 1.0) using first-principles calculations. This study reveals that the lattice thermal conductivity follows a classical inverse temperature-dependent relationship. There is a significant decrease in the lattice thermal conductivity when the Bi content increases from 0 to 0.25 or decreases from 1.0 to 0.75 at 300 K. In contrast, when the Bi content increases from 0.25 to 0.75, the lattice thermal conductivity experiences a gradual decrease and reaches a plateau. For the nanohybrids (x = 0.25, 0.5, and 0.75), the distribution patterns of the phonon group velocity and phonon lifetime are similar, with consistent distribution intervals. Consequently, the change in lattice thermal conductivity is not pronounced. However, the phonon group speed and phonon lifetime are generally lower compared to those of the pristine components with x = 0 and x = 1.0. Our results suggest that the lattice thermal conductivity is sensitive to impurities but not to concentrations. This research provides valuable theoretical insights for adjusting the lattice thermal conductivity of Mg3(BixSb1−x)2 nanocomposites.

1. Introduction

Thermoelectric (TE) materials can convert thermal energy into electricity, making them a potential solution to the current energy crisis [1,2,3,4,5,6,7,8,9]. Among these materials, Mg3(BixSb1−x)2 is the most promising candidate for TE applications near room temperature (RT) due to its lower cost than the commercially available Bi2Te3−xSex [10,11,12]. As a result, it has received considerable attention [13,14,15,16,17,18,19,20,21,22,23,24,25,26]. Researchers are constantly working to improve the thermoelectric performance of these materials. The thermoelectric figure of merit, ZT, is typically used to evaluate performance, which is defined as z T = S 2 σ T / ( κ l + κ e ) . The parameters T, S, σ , κ l , and κ e represent the Kelvin temperature, Seebeck coefficient, electrical conductivity, lattice thermal conductivity, and electronic thermal conductivity, respectively [27]. Achieving a high thermoelectric figure of merit requires a high Seebeck coefficient (S), high electrical conductivity ( σ ), and low thermal conductivity (κ) [28]. Yet, achieving this is challenging due to the complicated interplay among these parameters. It is believed that reducing the lattice thermal conductivity is crucial for achieving high TE performance.
There are currently many methods available to reduce the thermal conductivity of thermoelectric materials. Biswas et al. achieved the maximum reduction in the lattice thermal conductivity of PbTe thermoelectric materials via considering scattering sources on all relevant length scales in a hierarchical manner, ranging from atomic-scale lattice disorder and nanoscale endotaxial precipitates to mesoscale grain boundaries [29]. Nakamura et al. overcame the amorphous limit of thermal conductivity in Si thermoelectric materials through constructing a Si nanostructure with “well-controlled nanoscale-shaped interfaces” and oriented nanocrystals (NCs), resulting in a significant reduction in thermal conductivity [30]. Zhang et al. achieved a reduction in thermal conductivity and improved thermoelectric performance through doping Sb in Mg2(Si,Sn) compounds [31]. Mo et al. demonstrated that doping a small amount of Se into Mg3Bi1.4Sb0.6 can effectively reduce the thermal conductivity of the system [32]. Knura et al. found that alloying PbTe with SnTe leads to a significant reduction in lattice thermal conductivity to values below approximately 1.0 W m−1 K−1 across a wide range of compositions (from x = 0.25 to x = 0.80) [33]. Here, we focus on the effect of alloying Mg3Bi2 with Mg3Sb2 on the lattice thermal conductivity of the Mg3(Sb,Bi)2 system.
The lattice thermal conductivity of Mg3X2 (X = Bi, Sb) has been thoroughly researched. Mg3X2 (X = Bi and Sb) has an uncommonly low lattice thermal conductivity that can be attributed to the notable softening and flattening of low-energy transverse acoustic phonons [15,34,35]. Zhang et al. offered new insights into the low lattice thermal conductivity of Mg3Bi2 through relating the pronounced phonon anharmonicity to the asymmetric nature of Bismuth’s 6s lone-pair electrons [15]. Contrary to the anticipated T−1 temperature-dependent κ l at high temperatures, Zhu et al. discovered that Mg3Sb2 exhibits a feeble temperature dependence of κ l following a power law of T−0.48 from theory and T−0.57 from experimental measurements. This weak dependence can be traced back to the stiffening of low-lying phonons and diminished anharmonicity at high temperatures, as indicated by the authors [36]. Mg3Sb2 and Mg3Bi2 can be combined in different stoichiometric ratios [37,38,39,40]. The combination of these compounds has been found to lower lattice thermal conductivity [41]. Nonetheless, it is still a major challenge to synthesize ternary Mg3(Bi,Sb)2 with a controllable Bi/Sb ratio [42].
Computational simulations are crucial for studying the thermal properties of Mg3(Bi,Sb)2. Various first-principles-based software, including phono3py [43] and ShengBTE [44], have been used to explore phonon transport phenomena. The lattice thermal conductivity of Mg3(Bi,Sb)2 has been studied to a limited extent using first-principles calculations, unlike that of Mg3X2 (X = Bi, Sb), due to its computational cost. Additionally, machine learning potentials have emerged as an alternative approach to expedite research on heat transport [45,46,47,48]. hiPhive, developed by Eriksson et al., efficiently extracts high-order force constants from density functional theory calculations [49,50]. Yang et al. utilized a machine learning potential function based on dual adaptive sampling to investigate the lattice thermal conductivity of Mg3Sb2, and obtained results consistent with the experimental data [51]. Ouyang and colleagues conducted a molecular dynamics (MD) exploration of the thermal properties of Mg3(BixSb1−x)2 using a moment tensor potential (MTP) model [52,53]. They developed an MTP model, which was based on machine learning (ML). Their predictions of changes in thermal conductivity with varying solution concentrations were generated using molecular dynamics simulations based on ML-interatomic potential (MLIAP) [54].
In this research, we analyze the lattice thermal conductivity of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1) utilizing first-principles calculations. Initially, we assess the crystal structure, thermodynamic stability, and dynamic stability of Mg3(BixSb1−x)2 at different Bi contents (x = 0.0, 0.25, 0.5, 0.75, and 1.0). Subsequently, we examine the heat capacity, phonon group velocity, and phonon lifetime of Mg3(BixSb1−x)2 to identify the factors that affect the changes in lattice thermal conductivity in the presence of various Bi contents.

2. Methods

The unit cell structure of Mg3Bi2 and Mg3Sb2 comprises five atoms. To investigate alternative structures, we expanded the Mg3Bi2 unit cell to create a 1 × 1 × 2 supercell. In this supercell, Sb atoms substitute Bi atoms, yielding three nanohybrid structures: Mg3Bi1.5Sb0.5, Mg3BiSb, and Mg3Bi0.5Sb1.5. We used first-principles calculations with VASP [55,56] to determine the lowest-energy structures for each composition. To compute the phonon spectrum, we utilized Phonopy [57,58,59] via the finite-displacement supercell approach. In the finite-displacement supercell approach, the first-principles calculation is utilized to acquire atomic forces in the supercell crystal structure model. Force constants were computed from an adequate number of supercells with distinct sets of displacements and corresponding forces were obtained using the first-principles calculation. For Mg3Bi2 and Mg3Sb2, 3 × 3 × 2 supercells were utilized, while 3 × 3 × 1 supercells were employed for Mg3Bi1.5Sb0.5, Mg3BiSb, and Mg3Bi0.5Sb1.5.
The thermal conductivity κ l is closely related to heat capacity, phonon group velocity, and phonon lifetime, as shown in Equation (1):
κ l = 1 N V 0 λ C λ v λ 2 τ λ
where λ represents the phonon mode, N is the total number of q points used to sample the Brillouin zone, V 0 stands for the volume of a unit cell, and C λ , v λ , and τ λ indicate the specific heat capacity, group velocity, and phonon lifetime, respectively. The C λ values were determined via analyzing the phonon density of states using the Bose–Einstein distribution. On the other hand, the v λ values were obtained via calculating the gradient of the phonon dispersion relation. To account for various phonon scattering mechanisms, the phonon relaxation time τ λ can be divided into three contributions: phonon–phonon scattering, phonon–grain boundary scattering, and phonon–defect scattering. This division is performed using Matthiessen’s rule [60]. In the current study, our main focus is on phonon–phonon scattering. We calculated the phonon relaxation time for this type of scattering using the harmonic and third-order anharmonic force constants obtained from density functional theory (DFT) calculations. To solve the phonon Boltzmann transport equation (BTE) [61] and calculate the lattice thermal conductivity, we employed the phono3py framework within the DFT framework. Phono3py is a tool that allows us to compute properties related to phonon–phonon interactions. At each step, the users invoked phono3py with at least the unit cell and the supercell matrix. In the first step, we generated supercells and sets of atomic displacements. These sets of displacements are referred to as “displacement datasets”. We constructed supercells using these displacements. Next, we calculated the forces of the supercell models using the VASP software, which we refer to as “force sets”. In the second step, we computed the second- and third-order force constants (fc2 and fc3) from the displacement datasets and force sets obtained in the first step. Finally, in the third step, we utilized the force constants obtained in the second step to calculate various properties, such as the lattice thermal conductivity, specific heat capacity, phonon group velocity, phonon lifetime, and mode-Grüneisen parameters. Due to the computationally intensive nature of lattice thermal conductivity calculations, particularly for larger unit cells, we used relatively smaller supercells in our computations. We utilized a 2 × 2 × 1 supercell for Mg3Bi1.5Sb0.5, Mg3BiSb, and Mg3Bi0.5Sb1.5. Considering that the original structure of Mg3(BixSb1−x)2 (x = 0.5, 1.0, and 1.5) is based on the 1 × 1 × 2 supercell of Mg3Bi2, the lattice thermal conductivity was calculated based on the 2 × 2 × 1 supercell. Therefore, for Mg3Sb2 and Mg3Bi2, we used a 2 × 2 × 2 supercell when calculating the lattice thermal conductivity. In this way, the calculation of the lattice thermal conductivity of Mg3(BixSb1−x)2(0 ≤ x ≤ 1) can be kept in the same dimension, and the number of atoms in the supercell is 40. For the purpose of comparing the impact of cell size on the calculation of lattice thermal conductivity, we utilized Mg3Sb2 as an example and calculated the lattice thermal conductivity of a larger 4 × 4 × 3 supercell using the hiPhive software in combination with phono3py.
The VASP calculation parameters were chosen as follows. The exchange and correlation functions were approximated using the general gradient approximation with Perdew–Burke–Ernzerhof parameterization (GGA-PBE) [62]. The energy cutoff for the plane-wave basis expansion was set to 450 eV. k-mesh were produced using VASPKIT [63] in the Gamma scheme at a density of 2 π × 0.03 Å−1. For geometry optimization, a convergence criterion of 1 × 10−5 eV in energy and 1 × 10−4 eV/Å in force was applied. For the calculations of the phonon spectrum and lattice thermal conductivity, a convergence criterion of 1 × 10−8 eV in energy was utilized.

3. Results and Discussions

The formation energies and crystal structures of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1) are presented in Figure 1. The formation energies of Mg3(BixSb1−x)2 were calculated using Equation (2):
E f o r m = E u n i t c e l l M g 3 ( B i x S b 1 x ) 2 3 E f o r m M g 2 x E f o r m B i 2 ( 1 x ) E f o r m S b / 5
In this equation, E u n i t c e l l M g 3 ( B i x S b 1 x ) 2 represents the total energy of a single chemical formula Mg3(BixSb1−x)2, and E f o r m M g , E f o r m B i , E f o r m S b denote the single atomic energies of Mg, Bi, and Sb pure constituents, respectively. The structures of Mg, Bi, and Sb were retrieved from the Material Project database [64] and have space groups of P63/mmc, R 3 ¯ m , and R 3 ¯ m , respectively. The single atomic energies of the pure constituents Mg, Bi, and Sb are −1.508 eV/atom, −3.9 eV/atom, and −4.13 eV/atom, respectively. As demonstrated in Figure 1, the formation energies of Mg3(BixSb1−x)2 are negative, indicating thermodynamic favorability.
Additionally, the formation energy increases linearly as the Bi content increases. Figure 2 illustrates the phonon spectra of Mg3(BixSb1−x)2. With the exception of the negligible imaginary frequency observed for Mg3BiSb at the L2 point, the remaining structures exhibit no imaginary frequencies, indicating dynamic stability. Moreover, the absence of a phonon bandgap facilitates a plethora of phonon–phonon scattering processes.
The Zintl compounds Mg3Sb2 and Mg3Bi2 share a trigonal CaAl2Si2-type structure and belong to the space group P 3 ¯ m 1 [65]. The polyanions (Mg2X2)2− (X = Bi, Sb), in which Mg is tetrahedrally coordinated, form covalently bonded layers. These layers are ionically bonded with the octahedrally coordinated Mg2+ cation layers, forming the overall framework. Substituting Bi atoms with Sb results in a decrease in system symmetry. As the concentration of Bi increases, the gradual substitution of Bi atoms by Sb atoms occurs in a layer-by-layer manner, resulting in an expansion of the unit cell volume. Figure S1 in the Supplementary Information depicts the average lengths of Mg-X bonds in Mg3(BixSb1−x)2 (0 ≤ x ≤ 1). In the (Mg2X2)2− network, the vertical Mg-X bonds are longer than the three symmetry-equivalent tilted Mg-X bonds. In addition, the ionic M-X bonds are longer than the covalent M-X bonds. As the Bi content increases, so does the average bond length between M and X, suggesting a diminished bonding ability between Mg-X and the increased Bi content.
Figure 3a presents the average lattice thermal conductivities of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1) obtained from the BTE utilizing second- and third-order force constants calculated using DFT. These conductivities demonstrate a classical temperature-dependent relationship of T−1 instead of a weak temperature-dependent κ l . It is observed that the order of lattice thermal conductivity is as follows: Mg3Sb2 (x = 0) > Mg3Bi2 (x = 1) > Mg3(BixSb1−x)2 (x= 0.25, 0.5, 0.75). Alloying Mg3Bi2 with Mg3Sb2 significantly reduces the lattice thermal conductivity. We also analyzed the lattice thermal conductivity at 300 K for various Bi contents and compared it with the previous literature [41,54]. Notably, there were marked reductions in the lattice thermal conductivity as the Bi content increased from 0 to 0.25 or decreased from 1.0 to 0.75. However, as the Bi content increased from 0.25 to 0.75, the lattice thermal conductivity showed a slight decrease, indicative of a plateau. Thus, adjusting the Bi content within the range of x = 0.25 to 0.75 may have a negligible effect on the lattice thermal conductivity. Ouyang and colleagues employed ML-IAP to compute the lattice thermal conductivity of the Mg3(BixSb1−x)2 alloy at different alloying concentrations for MCMD structures at 300 K [54]. Our findings are in qualitative concurrence with the experiment [41] and calculation by Ouyang et al., albeit our computed lattice thermal conductivities are comparatively lower. This difference could be due to the reduced size of the supercell employed in the phono3py computations. Due to the significant computational resources needed for larger supercells when calculating the lattice thermal conductivity, we selected Mg3Sb2 as an illustrative example. We assessed the lattice thermal conductivity in a 4 × 4 × 3 supercell using the hiPhive software and the results, demonstrated in Figure 4, generally indicate that the thermal conductivity acquired from the 4 × 4 × 3 supercell using hiPhive exceeds that of the 2 × 2 × 2 supercell calculated using phono3py. Furthermore, the outcomes generated using hiPhive coincide with the calculations carried out by Ouyang et al. [54], even though both sets of calculations produce marginally lower values in comparison to the experimental data [66].
Figure S2 in the Supplementary Information presents the heat capacity of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1) at 300 K. The difference in thermal conductivity mainly arises from the phonon group velocity and phonon lifetime rather than the heat capacity, which exhibits only a small discrepancy. Figure 5 and Figure 6 display the phonon group velocity and phonon lifetime, respectively, of Mg3(BixSb1−x)2. The phonon group velocity distribution range remains similar when x equals 0.25, 0.5, and 0.75. However, the overall group velocity is lower when compared to x equals 0 and 1.0, with the exception of a few higher group velocities in the low-frequency region. As shown in Figure 5, for x equals 0.25, 0.5, and 0.75, the proportion of phonon group speeds below 100 m/s increases significantly. Although the range of phonon group velocity distributions in Mg3Sb2 and Mg3Bi2 is consistent, the high-frequency region reveals that Mg3Bi2 has a significantly lower phonon lifetime than Mg3Sb2. Consequently, the lattice thermal conductivity of Mg3Bi2 is lower than that of Mg3Sb2. For x = 0.25, 0.5, and 0.75, the lifetime distribution pattern of phonons in Mg3(BixSb1−x)2 is similar, with the distribution interval falling within the same range. Nonetheless, the overall distribution is lower in comparison to x = 0 and x = 1.0. According to Figure 6, there is a significant increase in the proportion of phonon lifetimes below 1.0 ps at x = 0.25, 0.5, and 0.75. We further examined the mode-Grüneisen parameters of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1), as illustrated in Figure 7. We found that compared to Mg3Sb2 and Mg3Bi2, there is a significant increase in the proportion of regions in the low-frequency range of Mg3(BixSb1−x)2 with absolute values greater than 10 when x = 0.25, 0.5, and 0.75. This indicates that Mg3(BixSb1−x)2 (x = 0.25, 0.5, and 0.75) exhibits higher anharmonicity. Therefore, as the Bi content increases from 0 to 0.25 or decreases from 1 to 0.75, the lattice thermal conductivity of Mg3(BixSb1−x)2 decreases considerably. However, when the concentration changes from 0.25 to 0.75, the change in lattice thermal conductivity reaches a plateau.

4. Conclusions

We conducted a comprehensive investigation into the lattice thermal conductivity of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1) using DFT calculations. The average lengths of the Mg-X ionic bonds (d1), vertical Mg-X covalent bonds (d2), and three symmetry-equivalent tilted covalent Mg-X bonds (d3) all increase as the Bi content increases. The formation energies of Mg3(BixSb1−x)2 are all negative, indicating that their crystal structures are thermodynamically stable. The dynamic stability of these structures is confirmed through the absence of imaginary frequencies in their phonon spectra.
The results from the phono3py calculations reveal that the lattice thermal conductivity of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1) with different Bi contents follows a descending order as Mg3Sb2 > Mg3Bi2 > Mg3(BixSb1−x)2 (x = 0.25, 0.5, 0.75). At 300 K, the lattice thermal conductivity of Mg3(BixSb1−x)2 is higher at both the upper and lower bounds of the composition range. The effect of Bi content on the lattice thermal conductivity is negligible for x in the range between 0.25 and 0.75. This trend is directly related to the phonon group velocity and phonon lifetime. The phonon lifetime of Mg3Sb2 is higher than that of Mg3Bi2. When x = 0 and x = 1.0, the phonon group velocity and phonon lifetime of Mg3(BixSb1−x)2 are significantly higher compared to the values when x = 0.25, 0.5, and 0.75. When the Bi content increases from 0.25 to 0.75, there is little change in the distribution patterns for the phonon group velocity and phonon lifetime. Despite the increase in Bi content, the distribution range remains consistent, leading to negligible changes in the lattice thermal conductivity.
It is worth noting that the current calculation of lattice thermal conductivity for Mg3(BixSb1−x)2 is based on a relatively small supercell with just 40 atoms. While a qualitative agreement between the calculated results and the experimental results is observed, some excluded phonons with shorter wavelengths lead to lower results due to the simulation box’s limitation. Our research contributes not only to the comprehension of thermal transport in Mg3(BixSb1−x)2 but also to the optimization of its thermal conductivity through experimentation.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/nano13222938/s1, Figure S1. (a) Crystal structure model of Mg3X2 (X = Sb, Bi); (b) The average lengths of Mg-X ionic bonds ( d 1 ¯ ), vertical Mg-X covalent bonds ( d 2 ¯ ), and three symmetry-equivalent tilted covalent Mg-X bonds ( d 3 ¯ ); Figure S2. The heat capacity of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1) at 300 K.

Author Contributions

Q.P. and X.-J.C. conceived the project; X.Y. and S.Z. performed the first-principles-based calculation and analysis. All authors analyzed the data, discussed the results, wrote the draft, and reviewed and edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Shenzhen Science and Technology Program with grant number KQTD20200820113045081, the National Natural Science Foundation of China with Grant No. 12272378, the High-level Innovation Research Institute Program of Guangdong Province (Grant No. 2020B0909010003), and the Institute of Mechanics, Chinese Academy of Sciences with grant number E1Z1011001.

Data Availability Statement

Data available on request from the authors upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. He, J.; Tritt, T.M. Advances in thermoelectric materials research: Looking back and moving forward. Science 2017, 357, eaak9997. [Google Scholar] [CrossRef] [PubMed]
  2. 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]
  3. Yang, L.; Chen, Z.-G.; Dargusch, M.S.; Zou, J. High performance thermoelectric materials: Progress and their applications. Adv. Energy Mater. 2018, 8, 1701797. [Google Scholar] [CrossRef]
  4. Tritt, T.M. Thermoelectric phenomena, materials, and applications. Annu. Rev. Mater. Sci. 2011, 41, 433–448. [Google Scholar] [CrossRef]
  5. Twaha, S.; Zhu, J.; Yan, Y.; Li, B. A comprehensive review of thermoelectric technology: Materials, applications, modelling and performance improvement. Renew. Sust. Energy Rev. 2016, 65, 698–726. [Google Scholar] [CrossRef]
  6. Shi, X.; Chen, L.; Uher, C. Recent advances in high-performance bulk thermoelectric materials. Int. Mater. Rev. 2016, 61, 379–415. [Google Scholar] [CrossRef]
  7. Shuai, J.; Mao, J.; Song, S.; Zhang, Q.; Chen, G.; Ren, Z. Recent progress and future challenges on thermoelectric Zintl materials. Mater. Today Phys. 2017, 1, 74–95. [Google Scholar] [CrossRef]
  8. Zhang, Q.H.; Huang, X.Y.; Bai, S.Q.; Shi, X.; Uher, C.; Chen, L.D. Thermoelectric devices for power generation: Recent progress and future challenges. Adv. Eng. Mater. 2016, 18, 194–213. [Google Scholar] [CrossRef]
  9. Wei, J.; Yang, L.; Ma, Z.; Song, P.; Zhang, M.; Ma, J.; Yang, F.; Wang, X. Review of current high-ZT thermoelectric materials. J. Mater. Sci. 2020, 55, 12642–12704. [Google Scholar] [CrossRef]
  10. Hu, L.; Wu, H.; Zhu, T.; Fu, C.; He, J.; Ying, P.; Zhao, X. Tuning multiscale microstructures to enhance thermoelectric performance of n-type Bismuth-Telluride-based solid solutions. Adv. Energy Mater. 2015, 5, 1500411. [Google Scholar] [CrossRef]
  11. Liu, W.; Wang, H.; Wang, L.; Wang, X.; Joshi, G.; Chen, G.; Ren, Z. Understanding of the contact of nanostructured thermoelectric n-type Bi2Te2.7Se0.3 legs for power generation applications. J. Mater. Chem. A 2013, 1, 13093–13100. [Google Scholar] [CrossRef]
  12. Mao, J.; Zhu, H.; Ding, Z.; Liu, Z.; Gamage, G.A.; Chen, G.; Ren, Z. High thermoelectric cooling performance of n-type Mg3Bi2-based materials. Science 2019, 365, 495–498. [Google Scholar] [CrossRef] [PubMed]
  13. Tamaki, H.; Sato, H.K.; Kanno, T. Isotropic conduction network and defect chemistry in Mg3+ δSb2-based layered zintl compounds with high thermoelectric performance. Adv. Mater. 2016, 28, 10182–10187. [Google Scholar] [CrossRef]
  14. Zhang, J.; Song, L.; Pedersen, S.H.; Yin, H.; Hung, L.T.; Iversen, B.B. Discovery of high-performance low-cost n-type Mg3Sb2-based thermoelectric materials with multivalley conduction bands. Nat. Commun. 2017, 8, 13901. [Google Scholar] [CrossRef]
  15. Zhang, J.; Song, L.; Iversen, B.B. Insights into the design of thermoelectric Mg3Sb2 and its analogs by combining theory and experiment. NPJ Comput. Mater. 2019, 5, 76. [Google Scholar] [CrossRef]
  16. Shi, X.; Wang, X.; Li, W.; Pei, Y. Advances in thermoelectric Mg3Sb2 and its derivatives. Small Methods 2018, 2, 1800022. [Google Scholar] [CrossRef]
  17. Zhou, Z.; Han, G.; Lu, X.; Wang, G.; Zhou, X. High-performance magnesium-based thermoelectric materials: Progress and challenges. J. Magnes. Alloys 2022, 10, 1719–1736. [Google Scholar] [CrossRef]
  18. Shang, H.; Liang, Z.; Xu, C.; Mao, J.; Gu, H.; Ding, F.; Ren, Z. N-Type Mg3Sb2-xBix alloys as promising thermoelectric materials. Research 2020, 2020, 1219461. [Google Scholar] [CrossRef]
  19. Chen, Y.; Wang, C.; Ma, Z.; Li, L.; Li, S.; Wang, J. Improved thermoelectric performance of n-type Mg3Sb2–Mg3Bi2 alloy with Co element doping. Curr. Appl. Phys. 2021, 21, 25–30. [Google Scholar] [CrossRef]
  20. Jiang, F.; Feng, T.; Zhu, Y.; Han, Z.; Shu, R.; Chen, C.; Zhang, Y.; Xia, C.; Wu, X.; Yu, H.; et al. Extraordinary thermoelectric performance, thermal stability and mechanical properties of n-type Mg3Sb1.5Bi0.5 through multi-dopants at interstitial site. Mater. Today Phys. 2022, 27, 100835. [Google Scholar] [CrossRef]
  21. Shang, H.; Liang, Z.; Xu, C.; Song, S.; Huang, D.; Gu, H.; Mao, J.; Ren, Z.; Ding, F. Ntype Mg3Sb2-xBix with improved thermal stability for thermoelectric power generation. Acta Mater. 2020, 201, 572–579. [Google Scholar] [CrossRef]
  22. Yang, J.; Li, G.; Zhu, H.; Chen, N.; Lu, T.; Gao, J.; Guo, L.; Xiang, J.; Sun, P.; Yao, Y.; et al. Next-generation thermoelectric cooling modules based on high-performance Mg3(Bi, Sb)2 material. Joule 2022, 6, 193–204. [Google Scholar] [CrossRef]
  23. Peng, Q.; Yuan, X.; Zhao, S.; Zhou, Y.; Wen, X.; Chen, X.-j. Active-learning search for unitcell structures: A case study on Mg3Bi2-xSbx. Comput. Mater. Sci. 2023, 226, 112260. [Google Scholar] [CrossRef]
  24. Peng, Q.; Zhao, S.; Yuan, X.; Chen, X.-J. Elasticity of Mg3Bi2-xSbx. Materials 2022, 15, 7161. [Google Scholar] [CrossRef]
  25. Huang, B.; Luo, P.; Li, Z.; Liu, X.; Zhang, Y.; Tang, Y.; Xing, J.; Zhang, J.; Guo, K.; Dong, Z.; et al. Improving Thermoelectric Performance of n-Type Mg3Bi2-Based Materials by Introducing a Spatially Confined Magnetic Ordered Structure. ACS Appl. Energy Mater. 2023, 6, 8032–8041. [Google Scholar] [CrossRef]
  26. Zhang, Y.-b.; Liang, J.-S.; Liu, C.; Zhou, Q.; Xu, Z.; Chen, H.-b.; Li, F.-c.; Peng, Y.; Miao, L. Enhancing thermoelectric performance in P-Type Mg3Sb2-based zintls through optimization of band gap structure and nanostructuring. J. Mater. Sci. Technol. 2024, 170, 25–32. [Google Scholar] [CrossRef]
  27. Witkoske, E.; Wang, X.; Maassen, J.; Lundstrom, M. Universal behavior of the thermoelectric figure of merit, zT, vs. quality factor. Mater. Today Phys. 2019, 8, 43–48. [Google Scholar] [CrossRef]
  28. Chen, K.-X.; Li, M.-S.; Mo, D.-C.; Lyu, S.-S. Nanostructural thermoelectric materials and their performance. Front. Energy 2018, 12, 97–108. [Google Scholar] [CrossRef]
  29. Biswas, K.; He, J.; Blum, I.D.; Wu, C.-I.; Hogan, T.P.; Seidman, D.N.; Dravid, V.P.; Kanatzidis, M.G. High-performance bulk thermoelectrics with all-scale hierarchical architectures. Nature 2012, 489, 414–418. [Google Scholar] [CrossRef]
  30. Nakamura, Y.; Isogawa, M.; Ueda, T.; Yamasaka, S.; Matsui, H.; Kikkawa, J.; Ikeuchi, S.; Oyake, T.; Hori, T.; Shiomi, J.; et al. Anomalous reduction of thermal conductivity in coherent nanocrystal architecture for silicon thermoelectric material. Nano Energy 2015, 12, 845–851. [Google Scholar] [CrossRef]
  31. Zhang, Q.; He, J.; Zhu, T.; Zhang, S.; Zhao, X.; Tritt, T.M. High figures of merit and natural nanostructures in Mg2Si0.4Sn0.6 based thermoelectric materials. Appl. Phys. Lett. 2008, 93, 102109. [Google Scholar] [CrossRef]
  32. Mo, X.; Liao, J.; Yuan, G.; Zhu, S.; Lei, X.; Huang, L.; Zhang, Q.; Wang, C.; Ren, Z. High thermoelectric performance at room temperature of n-type Mg3Bi2-based materials by Se doping. J. Magnes. Alloys 2022, 10, 1024–1032. [Google Scholar] [CrossRef]
  33. Knura, R.; Parashchuk, T.; Yoshiasa, A.; Wojciechowski, K.T. Origins of low lattice thermal conductivity of Pb1-xSnx Te alloys for thermoelectric applications. Dalton Trans. 2021, 50, 4323–4334. [Google Scholar] [CrossRef] [PubMed]
  34. Ding, J.; Lanigan-Atkins, T.; Calderón-Cueva, M.; Banerjee, A.; Abernathy, D.L.; Said, A.; Zevalkink, A.; Delaire, O. Soft anharmonic phonons and ultralow thermal conductivity in Mg3(Sb,Bi)2 thermoelectrics. Sci. Adv. 2021, 7, eabg1449. [Google Scholar] [CrossRef] [PubMed]
  35. Peng, W.; Petretto, G.; Rignanese, G.-M.; Hautier, G.; Zevalkink, A. An unlikely route to low lattice thermal conductivity: Small atoms in a simple layered structure. Joule 2018, 2, 1879–1893. [Google Scholar] [CrossRef]
  36. Zhu, Y.; Xia, Y.; Wang, Y.; Sheng, Y.; Yang, J.; Fu, C.; Li, A.; Zhu, T.; Luo, J.; Wolverton, C.; et al. Violation of the T- 1 relationship in the lattice thermal conductivity of Mg3Sb2 with locally asymmetric vibrations. Research 2020, 2020, 4589786. [Google Scholar] [CrossRef]
  37. Han, Z.; Gui, Z.; Zhu, Y.; Qin, P.; Zhang, B.-P.; Zhang, W.; Huang, L.; Liu, W. The electronic transport channel protection and tuning in real space to boost the thermoelectric performance of Mg3+ δSb2-yBiy near room temperature. Research 2020, 2020, 1672051. [Google Scholar] [CrossRef]
  38. Imasato, K.; Kang, S.D.; Snyder, G.J. Exceptional thermoelectric performance in Mg3Sb0.6Bi1.4 for low-grade waste heat recovery. Energy Environ. Sci. 2019, 12, 965–971. [Google Scholar] [CrossRef]
  39. Shi, X.; Zhang, X.; Ganose, A.; Park, J.; Sun, C.; Chen, Z.; Lin, S.; Li, W.; Jain, A.; Pei, Y. Compromise between band structure and phonon scattering in efficient nMg3SbBi2-xBix thermoelectrics. Mater. Today Phys. 2021, 18, 100362. [Google Scholar] [CrossRef]
  40. Xu, C.; Liang, Z.; Shang, H.; Wang, D.; Wang, H.; Ding, F.; Mao, J.; Ren, Z. Scalable synthesis of n-type Mg3SbBi2-xBix for thermoelectric applications. Mater. Today Phys. 2021, 17, 100336. [Google Scholar] [CrossRef]
  41. Imasato, K.; Wood, M.; Anand, S.; Kuo, J.J.; Snyder, G.J. Understanding the high thermoelectric performance of Mg3Sb2-Mg3Bi2 alloys. Adv. Energy Sustain. Res. 2022, 3, 2100208. [Google Scholar] [CrossRef]
  42. Pan, Y.; Yao, M.; Hong, X.; Zhu, Y.; Fan, F.; Imasato, K.; He, Y.; Hess, C.; Fink, J.; Yang, J.; et al. Mg3(Bi, Sb)2 single crystals towards high thermoelectric performance. Energy Environ. Sci. 2020, 13, 1717–1724. [Google Scholar] [CrossRef]
  43. Togo, A.; Chaput, L.; Tanaka, I. Distributions of phonon lifetimes in Brillouin zones. Phys. Rev. B 2015, 91, 094306. [Google Scholar] [CrossRef]
  44. Li, W.; Carrete, J.; Katcho, N.A.; Mingo, N. ShengBTE: A solver of the Boltzmann transport equation for phonons. Comput. Phys. Commun. 2014, 185, 1747–1758. [Google Scholar] [CrossRef]
  45. Korotaev, P.; Shapeev, A. Lattice dynamics of YbxCo4Sb12 skutterudite by machine learning interatomic potentials: Effect of filler concentration and disorder. Phys. Rev. B 2020, 102, 184305. [Google Scholar] [CrossRef]
  46. Verdi, C.; Karsai, F.; Liu, P.; Jinnouchi, R.; Kresse, G. Thermal transport and phase transitions of zirconia by on-the-fly machine-learned interatomic potentials. NPJ Comput. Mater. 2021, 7, 156. [Google Scholar] [CrossRef]
  47. Li, R.; Liu, Z.; Rohskopf, A.; Gordiz, K.; Henry, A.; Lee, E.; Luo, T. A deep neural network interatomic potential for studying thermal conductivity of β-Ga2O3. Appl. Phys. Lett. 2020, 117, 152102. [Google Scholar] [CrossRef]
  48. George, J.; Hautier, G.; Bartók, A.P.; Csányi, G.; Deringer, V.L. Combining phonon accuracy with high transferability in Gaussian approximation potential models. J. Chem. Phys. 2020, 153, 044104. [Google Scholar] [CrossRef]
  49. Eriksson, F.; Fransson, E.; Erhart, P. The Hiphive Package for the extraction of high order force constants by machine learning. Adv. Theory Simul. 2019, 2, 1800184. [Google Scholar] [CrossRef]
  50. Fransson, E.; Eriksson, F.; Erhart, P. Efficient construction of linear models in materials modeling and applications to force constant expansions. NPJ Comput. Mater. 2020, 6, 135. [Google Scholar] [CrossRef]
  51. Yang, H.; Zhu, Y.; Dong, E.; Wu, Y.; Yang, J.; Zhang, W. Dual adaptive sampling and machine learning interatomic potentials for modeling materials with chemical bond hierarchy. Phys. Rev. B 2021, 104, 094310. [Google Scholar] [CrossRef]
  52. Zuo, Y.; Chen, C.; Li, X.; Deng, Z.; Chen, Y.; Behler, J.; Csányi, G.; Shapeev, A.V.; Thompson, A.P.; Wood, M.A.; et al. Performance and cost assessment of machine learning interatomic potentials. J. Phys. Chem. A 2020, 124, 731–745. [Google Scholar] [CrossRef]
  53. Shapeev, A.V. Moment tensor potentials: A class of systematically improvable interatomic potentials. Multiscale Model. Simul. 2016, 14, 1153–1173. [Google Scholar] [CrossRef]
  54. Ouyang, P.; Yuan, M.-h.; Tang, P.; Zhang, Q.; Liu, S.; Shuai, J.; Li, X.-G. Atomic Local Ordering and Alloying Effects on the Mg3(Sb1–xBix)2 Thermoelectric Material. ACS Appl. Mater. Interfaces 2023, 15, 37554–37562. [Google Scholar] [CrossRef] [PubMed]
  55. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  56. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  57. Wu, X.; Vanderbilt, D.; Hamann, D. Systematic treatment of displacements, strains, and electric fields in density-functional perturbation theory. Phys. Rev. B 2005, 72, 035105. [Google Scholar] [CrossRef]
  58. Huang, L.-F.; Zeng, Z. Roles of mass, structure, and bond strength in the phonon properties and lattice anharmonicity of single-layer Mo and W dichalcogenides. J. Phys. Chem. C 2015, 119, 18779–18789. [Google Scholar] [CrossRef]
  59. Togo, A.; Tanaka, I. First principles phonon calculations in materials science. Scr. Mater. 2015, 108, 1–5. [Google Scholar] [CrossRef]
  60. Ziman, J.M. Electrons and Phonons: The Theory of Transport Phenomena in Solids; Oxford University Press: Oxford, UK, 2001. [Google Scholar]
  61. Omini, M.; Sparavigna, A. An iterative approach to the phonon Boltzmann equation in the theory of thermal conductivity. Phys. B Condens. Matter 1995, 212, 101–112. [Google Scholar] [CrossRef]
  62. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, V.; Xu, N.; Liu, J.-C.; Tang, G.; Geng, W.-T. VASPKIT: A user-friendly interface facilitating high-throughput computing and analysis using VASP code. Comput. Phys. Commun. 2021, 267, 108033. [Google Scholar] [CrossRef]
  64. Jain, A.; Ong, S.P.; Hautier, G.; Chen, W.; Richards, W.D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.; Ceder, G.; et al. Commentary: The Materials Project: A materials genome approach to accelerating materials innovation. APL Mater. 2013, 1, 011002. [Google Scholar] [CrossRef]
  65. Gladishevskij, E.; Kripjakevic, P.; Bodak, O. The crystal structures of the compound CaAl2Si2 and its analogues. Ukr. Fiz. Zh. 1967, 12, 447–453. [Google Scholar]
  66. Shuai, J.; Wang, Y.; Kim, H.S.; Liu, Z.; Sun, J.; Chen, S.; Sui, J.; Ren, Z. Thermoelectric properties of Na-doped Zintl compound: Mg3-xNaxSb2. Acta Mater. 2015, 93, 187–193. [Google Scholar] [CrossRef]
Figure 1. The formation energies and crystal structures of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1).
Figure 1. The formation energies and crystal structures of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1).
Nanomaterials 13 02938 g001
Figure 2. The phonon spectra of Mg3(BixSb1−x)2 for (a) x = 0.0; (b) x = 0.25; (c) x = 0.5; (d) x = 0.75; (e) x = 1.0.
Figure 2. The phonon spectra of Mg3(BixSb1−x)2 for (a) x = 0.0; (b) x = 0.25; (c) x = 0.5; (d) x = 0.75; (e) x = 1.0.
Nanomaterials 13 02938 g002
Figure 3. (a) Average lattice thermal conductivity of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1) at various temperatures; (b) Average lattice thermal conductivity of Mg3(BixSb1−x)2 at 300 K for different Bi concentration compared with the experiment [41] and calculation by Ouyang et al. [54].
Figure 3. (a) Average lattice thermal conductivity of Mg3(BixSb1−x)2 (0 ≤ x ≤ 1) at various temperatures; (b) Average lattice thermal conductivity of Mg3(BixSb1−x)2 at 300 K for different Bi concentration compared with the experiment [41] and calculation by Ouyang et al. [54].
Nanomaterials 13 02938 g003
Figure 4. Average lattice thermal conductivity of Mg3Sb2 at various temperatures in 2 × 2 × 2 supercells compared with the hiPhive calculation in 4 × 4 × 3 supercells, experiment [66], and the MD calculation by Ouyang et al. in 2 × 2 × 30 supercells [54].
Figure 4. Average lattice thermal conductivity of Mg3Sb2 at various temperatures in 2 × 2 × 2 supercells compared with the hiPhive calculation in 4 × 4 × 3 supercells, experiment [66], and the MD calculation by Ouyang et al. in 2 × 2 × 30 supercells [54].
Nanomaterials 13 02938 g004
Figure 5. The phonon group velocity of Mg3(BixSb1−x)2 for (a) x = 0.0; (b) x = 0.25; (c) x = 0.5; (d) x = 0.75; (e) x = 1.0.
Figure 5. The phonon group velocity of Mg3(BixSb1−x)2 for (a) x = 0.0; (b) x = 0.25; (c) x = 0.5; (d) x = 0.75; (e) x = 1.0.
Nanomaterials 13 02938 g005
Figure 6. The phonon lifetime of Mg3(BixSb1−x)2 at 300 K for (a) x = 0.0; (b) x = 0.25; (c) x = 0.5; (d) x = 0.75; (e) x = 1.0.
Figure 6. The phonon lifetime of Mg3(BixSb1−x)2 at 300 K for (a) x = 0.0; (b) x = 0.25; (c) x = 0.5; (d) x = 0.75; (e) x = 1.0.
Nanomaterials 13 02938 g006
Figure 7. The mode-Grüneisen parameters of Mg3(BixSb1−x)2 for (a) x = 0.0; (b) x = 0.25; (c) x = 0.5; (d) x = 0.75; (e) x = 1.0.
Figure 7. The mode-Grüneisen parameters of Mg3(BixSb1−x)2 for (a) x = 0.0; (b) x = 0.25; (c) x = 0.5; (d) x = 0.75; (e) x = 1.0.
Nanomaterials 13 02938 g007
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

Peng, Q.; Yuan, X.; Zhao, S.; Chen, X.-J. Lattice Thermal Conductivity of Mg3(Bi,Sb)2 Nanocomposites: A First-Principles Study. Nanomaterials 2023, 13, 2938. https://doi.org/10.3390/nano13222938

AMA Style

Peng Q, Yuan X, Zhao S, Chen X-J. Lattice Thermal Conductivity of Mg3(Bi,Sb)2 Nanocomposites: A First-Principles Study. Nanomaterials. 2023; 13(22):2938. https://doi.org/10.3390/nano13222938

Chicago/Turabian Style

Peng, Qing, Xiaoze Yuan, Shuai Zhao, and Xiao-Jia Chen. 2023. "Lattice Thermal Conductivity of Mg3(Bi,Sb)2 Nanocomposites: A First-Principles Study" Nanomaterials 13, no. 22: 2938. https://doi.org/10.3390/nano13222938

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