∼100% enhancement of cryogenic thermoelectric performance of Bi80Sb20 alloys by incorporation of Fe3O4 nanoparticles

Mengran Chen a, Zhendong Maoab, Heng Liu *cd, Shun Wane, Xugui Xiab, Xuefei Zhangf, Chuanrui Zhanga, Qingfeng Song*b, Shengqiang Baib and Peng-an Zong*a
aCollege of Materials Science and Engineering, Nanjing Tech University, Nanjing 210009, China. E-mail: pazong@njtech.edu.cn
bShanghai Institute of Ceramics, Chinese Academy of Sciences, Shanghai 200050, China. E-mail: qfsong@mail.sic.ac.cn
cState Key Laboratory of Powder Metallurgy, Central South University, Changsha, 410083, China. E-mail: heng.liu.e1@tohoku.ac.jp
dAdvanced Institute for Materials Research (WPI-AIMR), Tohoku University, Sendai 980-8577, Japan
eWuzhen Laboratory, Tongxiang 314500, China
fState Key Laboratory of New Ceramics and Fine Processing, School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China

Received 26th May 2025 , Accepted 15th August 2025

First published on 4th September 2025


Abstract

Bi80Sb20-based alloys are promising candidates for thermoelectric applications below room temperature. However, their inherently low thermoelectric performance, characterized by a limited power factor and high lattice thermal conductivity, severely hinders practical utilization. To address this challenge, we introduce Fe3O4 nanoparticles as a second phase to construct Bi80Sb20/Fe3O4 nanocomposites via a high-efficiency mechanical alloying process, assisted by stearic acid dispersion, followed by spark plasma sintering. The incorporation of nano-Fe3O4 leads to enhanced carrier mobility and intensified phonon scattering at nanoparticle interfaces, resulting in a remarkable threefold increase in the power factor and a 20–40% reduction in lattice thermal conductivity. As a result, the optimized composite with 0.4 vol% Fe3O4 achieves a peak zT of 0.31 at 242 K – representing nearly a 100% improvement over the pristine matrix – and an average zT of 0.29 in the range of 150–300 K. In addition to the thermoelectric enhancements, the Vickers hardness of the composites is also significantly improved. These findings demonstrate the potential of superparamagnetic Fe3O4 nanoparticle incorporation as an effective strategy for the simultaneous enhancement of both thermoelectric and mechanical properties in Bi80Sb20-based alloys, offering a viable pathway for low-temperature power generation and refrigeration applications.


image file: d5ta04222d-p1.tif

Heng Liu

Dr Heng Liu is currently a specially appointed assistant professor at the Advanced Institute for Materials Research (WPIAIMR), Tohoku University, Japan. He earned his PhD degree from the Department of Chemical Engineering, University of Manchester, UK in 2024, and his MS degree in material engineering from Central South University, China, in 2020. His primary research interests focus on first-principles calculations and new computational methodology development for understanding modern energy and environmental technologies.

Introduction

Thermoelectric (TE) refrigeration devices stand out as environmentally friendly solid refrigeration options, eliminating the need for refrigerants or emissions. They offer significant benefits such as compact size, high reliability, and silence, making them highly suitable for cooling precision devices like charge-coupled device cameras, laser diodes, infrared detectors, and blood analyzers, among others.1,2 Besides cooling, TE systems can convert thermal gradients into electricity, enabling self-powered operation for vapor-driven energy harvesting, 3D-printed flexible electronics, and stimuli-responsive sensing, among others. This dual functionality, simultaneous energy conversion and temperature/stress detection, eliminates external power requirements while demonstrating transformative potential for multifunctional integrated devices.3–5 The TE performance of materials is measured by the dimensionless figure of merit (zT), defined as zT = S2σT/κ, where S represents the Seebeck coefficient, σ the electrical conductivity, κ the thermal conductivity, and T the temperature. S2σ is termed the power factor (PF), reflecting the comprehensive electrical performance, while κ comprises both electronic thermal conductivity (κe) and lattice thermal conductivity (κL).6–8 Because these parameters are interdependent, to achieve a high zT value, it is crucial to develop strategies that can decouple these parameters, to simultaneously enhance electronic transport and depress phonon transport.9

Over the past two decades, the TE performance of a wide range of materials, including Bi2Te3,10,11 CoSb3,12,13 PbSe,14 PbTe,15,16 half-Heusler compounds,17,18 Cu2Se,19,20 SiGe21 and others, has been significantly improved via various strategies, such as phonon-glass-electron-crystal and phonon-liquid-electron-crystal concepts, band structure engineering, nanocomposite approaches, etc. These efforts have led to a peak zT exceeding 2.0. However, most of these materials operate effectively only above room temperature (T > 300 K), while the development of low-temperature TE materials (T < 300 K) remains relatively limited. In addition to emerging materials such as YbAl3,22 CeRhSn23 and CsBi4Te6,24 Bi–Sb alloys are among the most promising candidates for cryogenic TE applications.25 Single crystal Bi–Sb alloys have demonstrated a maximum zT of ∼0.6 at 300 K,26 while showing inherent brittleness and cleavage failure. High-energy ball milling27,28 and wet chemical reaction29,30 techniques have been employed to fabricate nanostructured Bi–Sb alloys with improved mechanical integrity. Nonetheless, polycrystalline Bi–Sb alloys typically suffer from reduced electrical transport properties due to random grain orientation compared to their single-crystal counterparts.31 Therefore, achieving fine-grained microstructures that not only suppress cleavage failure but also leverage high-density grain boundaries for effective phonon scattering without significantly degrading carrier transport is critical for advancing cryogenic TE materials.

Besides moderate doping,32,33 the inclusion of second-phase nanoparticles, nanowires, or nanotubes into a crystalline matrix also represents a prominent strategy to improve the TE properties of bulk materials.34 The incorporation of nanoscale second phases inevitably introduces a variety of structural imperfections such as grain boundaries, dislocations, nanopores, and phase interfaces. These defects serve as effective phonon scattering centers, thus reducing lattice thermal conductivity.35 Meanwhile, interfacial band bending and associated energy filtering effects can also enhance the Seebeck coefficient,36 offering a pathway to improve the overall zT value. Recent studies have demonstrated that magnetic nano-inclusions are effective in enhancing the TE performance of TE composites. For example, the incorporation of superparamagnetic Co nanoinclusions into a filled CoSb3 matrix has been shown to improve TE properties by decoupling phonon and electron transport.37 Poudeu et al.38 reported that magnetic interaction between Ti(Ni, Fe)Sn nanoinclusions and the spin of itinerant carriers could significantly enhance the electrical properties while simultaneously suppressing the thermal conductivity of the (Ti, Zr, Hf)Ni(Sn, Sb) alloy. Superparamagnetic Fe3O4 nanoparticles, known for their mixed-valence character, can reduce thermal conductivity through phonon-magnetic moment interactions.39 It has been reported that introduction of Fe3O4 nanoinclusions into BiSbTe and YbAl3 matrices can effectively suppress κL, thereby significantly improving zT.40,41

Based on the above, the application of Fe3O4 magnetic nanoinclusions to enhance the TE properties of n-type Bi–Sb-based alloys below 300 K presents a promising strategy for the commercial development of cryogenic TE materials. To achieve efficient dispersion of Fe3O4 nanoparticles and reduce processing costs, Bi80Sb20/Fe3O4 composites were synthesized via direct ball milling of elemental powders with nano-sized Fe3O4, followed by spark plasma sintering. This ball milling approach offers a higher dispersion efficiency and lower cost compared to the conventional melt-then-mill route, making it more suitable for scalable fabrication. This study presents a detailed analysis of the synthesis procedure, the electrical and thermal transport properties, and the mechanisms responsible for the observed thermoelectric performance.

Experimental section

Synthesis

Bi80Sb20/x vol% Fe3O4 composites were synthesized via dry ball-milling of elemental powders: bismuth (99.9% Sinopharm), antimony (99.99% Alfa Aesar), and nano-sized Fe3O4 powders (Aladdin, 99%) at varying volume fractions (0, 0.2, 0.4, 0.5, and 0.8 vol%). Stearic acid (99.9% Alfa Aesar) was added as a process control agent to prevent powder adhesion to the milling container walls. For this purpose, the weighed powders were transferred into a stainless steel milling container and sealed under an argon atmosphere in a glovebox. Using a ball-to-powder ratio of 6[thin space (1/6-em)]:[thin space (1/6-em)]1 and 1200 rpm, the powders were ball milled for 8 hours. The stearic acid in the Bi–Sb powders was cleaned with acetone before SPS sintering. Specifically, the Bi80Sb20/x vol% Fe3O4 milled alloyed powders were loaded into the graphite SPS die in the glovebox. Then, they were sintered at 513 K through the spark plasma sintering (Sojitz, SPS-725) technique for 10 min with a uniaxial pressure of 65 MPa under vacuum to form a pellet with dimensions of 10 mm diameter and 4 mm thickness, as shown in Fig. 1. For performance comparison, an alternative dispersion approach was investigated. A Bi80Sb20/0.2 vol% Fe3O4 composite was prepared using wet ball milling (in ethanol solvent with 1-hour ultrasonic pre-dispersion), followed by the standard SPS process described above.
image file: d5ta04222d-f1.tif
Fig. 1 Schematic diagram of the preparation process of Bi80Sb20/x vol% Fe3O4 (x = 0, 0.2, 0.4, 0.5, 0.8) composites.

Calculations

Spin-polarized density functional theory computations were conducted via the Vienna ab initio Simulation Package (VASP) code.42 The calculations employed a plane-wave basis set in conjunction with the projector augmented-wave (PAW) method.43 The generalized gradient approximation (GGA),44 incorporating spin–orbit coupling (SOC) effects,45 as parameterized by the Perdew–Burke–Ernzerhof (PBE) functional,46 was employed to describe the electron exchange–correlation interactions. All structural models were constructed using the Atomic Simulation Environment (ASE) Python library.47 The convergence criterion for structural relaxation was set such that the residual force on each atom was less than 0.03 eV Å−1. To account for the strong on-site Coulomb interactions of the localized d electrons in Fe, a Hubbard U correction (DFT + U) was applied.48 The specific U and J values for Fe were 5.3 eV and 0 eV, respectively, based on the recommendations of the Materials Project database.49

Characterization

The phase compositions were analyzed using an X-ray diffractometer with Cu Ka sources (XRD, D8 Advance, Bruker), field emission electron microscopy (FESEM, ZEISS, Supra55) coupled with Energy Dispersive X-ray Spectroscopy (EDS, Oxford, UK) and Field Emission Transmission Electron Microscopy (FETEM, JOEL 2100F) coupled with Energy Dispersive X-ray Spectroscopy (EDS, Oxford, UK). Synchrotron powder diffraction XRD and GIWAXS were performed using a Shanghai Light Source BL17B line station. The Vicker Hardness was measured by using a microhardness tester (HV-1000Z) with a load of 0.2 N and a loading time of 10 s. The electrical conductivity, Seebeck coefficient and thermal conductivity tests were conducted using a comprehensive physical property testing system (PPMS-9, Quantum Design). Specimen preparation followed these protocols: diamond wire cutting of sintered pellets into 5.0 × 2.0 × 1.5 mm3 strips; sequential polishing with abrasive paper; application of colloidal graphite electrodes on terminal faces; vacuum annealing at 110 °C for 15 min. Hall measurements are performed using the five-wire method, with multi-point voltage–current combinations under an externally applied vertical magnetic field. The measurement uncertainties are quantified as follows: electrical conductivity ± 5%, Seebeck coefficient ± 7%, thermal conductivity ± 7%, and Hall coefficient ± 5%.

Results and discussion

Phases and microstructures

During the ball milling process, Bi–Sb powders tended to adhere to the walls of the milling container, which not only reduced the material yield but also increased preparation cost. To mitigate this issue, stearic acid was added in varying concentrations (0.5 wt%, 1 wt%, and 3 wt%) as a processing aid to enhance the powder dispersibility and reduce wall adhesion (Fig. 1). When the stearic acid content was 0.5 wt%, wall sticking was still observed during ball milling. In contrast, no adhesion was observed at concentrations of 1 wt% and 3 wt%. The improvement in powder dispersibility and reduction in wall adhesion could be attributed to the surface-active properties of stearic acid. The stearic acid is a long-chain saturated fatty acid (CH3(CH2)16COOH) containing a polar carboxyl (–COOH) group at one end and a nonpolar hydrocarbon tail. During the ball milling process, the carboxyl group can adsorb onto the surface of metal particles through weak chemical interactions or hydrogen bonding, while the hydrophobic alkyl chain extends outward, forming a steric barrier. This molecular arrangement reduces interparticle cohesion and prevents cold welding between particles, thus enhancing flowability and minimizing their tendency to adhere to the milling vessel walls. However, since stearic acid constitutes an impurity phase, its content should be minimized. Therefore, a stearic acid concentration of 1 wt% was ultimately selected for the synthesis of Bi80Sb20/x vol% Fe3O4 composites. First, due to the principle of “like dissolves like”, the stearic acid was initially removed using acetone, followed by drying and subsequent sintering. In addition, considering that the volatilization temperature of the stearic acid is approximately 423 K,50 any residual stearic acid that may remain after cleaning can still be effectively eliminated during the SPS process at 513 K. Specifically, as shown in Fig. S1, the C[double bond, length as m-dash]O stretching vibration peak at approximately 1700 cm−1 is widely recognized as a key characteristic signal for determining the presence of stearic acid, while the –(CH2)n– peak at 720 cm−1 is a critical identifying peak for long-chain fatty acid compounds. However, comparative spectral analysis revealed that neither of these characteristic peaks was observed in the composite material, indicating that stearic acid had been effectively removed.

The crystalline phases of Bi80Sb20/x vol% Fe3O4 composites were analyzed by XRD. As shown in Fig. 2a, the characteristic diffraction peaks of all the nanocomposites can be well indexed to the rhombohedral phase of Bi (PDF#44-1246). Bi and Sb form a complete solid solution in the rhombohedral A7 crystal structure, owing to their identical and similar lattice parameters (Bi: a = 4.5 Å and c = 11.8 Å; Sb: a = 4.34 Å and c = 11.3 Å).51 No diffraction peaks corresponding to Sb or other impurities are observed in the Bi80Sb20 sample, further confirming the formation of a homogeneous Bi–Sb alloy via a solid solution mechanism. The lattice parameters a and c (Fig. 2b) remained nearly unchanged with varying Fe3O4 content, suggesting that the dispersion of Fe3O4 does not alter the crystal structure of the Bi80Sb20 matrix. Moreover, no characteristic diffraction peaks of Fe3O4 are observed in any of the nanocomposites, which can be attributed to its low volume content, only up to 0.8%, well below the typical XRD detection limit. However, high-sensitivity synchrotron powder XRD enables more refined structural characterization of the composites. As shown in Fig. 2c, synchrotron XRD measurements were conducted on the Bi80Sb20/0.4 vol% Fe3O4 and Bi80Sb20 samples. The results confirm that the diffraction peaks corresponding to the rhombohedral lattice of Bi are consistent with those obtained from the conventional XRD analysis. In addition, weak but distinct peaks attributable to Fe3O4 are observed and match well with the reference data from the standard PDF card (#26-1136).


image file: d5ta04222d-f2.tif
Fig. 2 (a) XRD patterns of the Bi80Sb20/x vol% Fe3O4 (x = 0, 0.2, 0.4, 0.5, 0.8) nanocomposites. (b) Lattice parameters of the composites with different Fe3O4 contents. (c) Synchrotron radiation powder X-ray diffraction (PXRD) of Bi80Sb20 and Bi80Sb20/0.4 vol% Fe3O4 bulks. Grazing Incidence Wide Angle X-ray Scattering (GIWAXS) of Bi80Sb20 (d) and (e) Bi80Sb20/Fe3O4. (f) Surface backscattered electron (BSE) images and energy dispersive X-ray spectroscopy (EDS) elemental mapping for Bi80Sb20/0.4 vol% Fe3O4 bulk. Cross-sectional SEM images of the Bi80Sb20 (g) and (h) Bi80Sb20/0.4 vol% Fe3O4 bulk.

In the GIWAXS 2D scattering patterns shown in Fig. 2d and e, the scattering vector qxy corresponds to specific Bi crystal facets: qxy = 2.76 Å−1 for the (110) plane, 2.65 Å−1 for (104), 1.915 Å−1 for (012), and qxy = 1.58 Å−1 for (003). After the incorporation of 0.4 vol% Fe3O4, the overall diffraction pattern of the Bi80Sb20/Fe3O4 composite remains largely unchanged compared to the pristine Bi80Sb20 sample, suggesting that the small amount of Fe3O4 has minimal impact on the crystal orientation or structural ordering of the matrix. Notably, weak diffraction peaks corresponding to Fe3O4 are observed in the PXRD patterns of the samples, owing to the high sensitivity of this technique. In contrast, no corresponding signals of Fe3O4 are shown in the GIWAXS patterns, likely because it probes only the near-surface region where the Fe3O4 content may be insufficient for detection.

The elemental distribution in Bi80Sb20/x vol% Fe3O4 (x = 0.2, 0.4, 0.5, 0.8) bulk samples was further characterized using backscattered electron (BSE) imaging combined with energy-dispersive X-ray spectroscopy (EDS) elemental mapping. As shown in Fig. 2f and S2, the BSE images show that the Fe3O4 particles are uniformly distributed within the Bi80Sb20 matrix, with no evidence of extraneous impurity elements, confirming the successful formation of the Fe3O4/Bi80Sb20 composite. However, slight inhomogeneities in the Fe distribution are observed in the samples with x = 0.5 and 0.8, likely due to localized agglomeration of the Fe3O4 phase. Even though ball-milling mechanical alloying is a useful and scalable synthesis method, complete homogeneity of a second-phase additive (especially at low content) is hard to guarantee. Some agglomeration is almost unavoidable during the ball milling process unless highly optimized conditions are used. Fig. 2g, h and S3 present scanning electron microscopy (SEM) images of the fractured cross sections of the Bi80Sb20/x vol% Fe3O4 composites. Compared to pristine Bi80Sb20, the composite samples display a mixed microstructure comprising both large, smooth grains and smaller flake-like grains on the order of several hundred nanometers. This multiscale grain structure is beneficial for phonon scattering across a wide range of wavelengths, thereby effectively reducing the lattice thermal conductivity.

The microstructural features of Bi80Sb20/x vol% Fe3O4 composites were further characterized by field emission transmission electron microscopy (FE-TEM) (Fig. 3a). Fast Fourier transform (FFT) analysis of the interface region (Fig. 3b) distinctly delineated the lattice periodicity of the constituent phases: Fe3O4 exhibited an interplanar spacing of 0.244 nm, corresponding to the (311) plane, while Bi80Sb20 displayed an interplanar spacing of 0.324 nm, corresponding to the (012) plane. The findings indicate that the two phases retain their crystal structures despite being tightly bound at the interface. EDS analysis further confirms that Fe3O4 is encapsulated within the Bi80Sb20 main phases (Fig. 3c–f). This interfacial structure could possibly facilitate charge transfer and enhance the electrical properties of the composites.52 In localized TEM-EDS analysis, oxygen signals often appear delocalized due to the high mobility and broad spatial response of light elements. As such, the presence of O outside Fe3O4 particles should be interpreted cautiously, as it does not necessarily indicate incorporation into the Bi80Sb20 matrix.


image file: d5ta04222d-f3.tif
Fig. 3 (a) Field emission transmission electron microscopy (FE-TEM) image of the Bi80Sb20/Fe3O4 sample and (b) magnified view of the local interface. The inset displays the Fast Fourier Transform (FFT) patterns of the two phases and EDS mapping of (c) Fe, (d) O, (e) Bi and (f) Sb within the Bi80Sb20/Fe3O4 in (a).

TE properties of Bi80Sb20/x vol% Fe3O4 nanocomposites

The temperature dependence of the electrical conductivity, σ, of the Bi80Sb20/x vol% Fe3O4 composites in the range of 150–300 K is shown in Fig. 4a. The electrical conductivity, σ, increases with temperature, indicating typical semiconducting behavior. This is because the rise in temperature brings about more excited electrons into the conduction band, which leads to an increase in carrier concentration, n (Fig. 4b). Notably, the σ of the x = 0.4 sample shows a pronounced enhancement, reaching 4.2 × 105 S m−1 at room temperature, almost three times higher than that of pristine Bi80Sb20. This improvement is primarily due to a sharp increase in carrier mobility, μ (Fig. 4c). The observed sharp increase in carrier mobility upon the incorporation of trace amounts of Fe3O4 nanoparticles can be attributed to the dynamic interactions between their superparamagnetic nature and electron transport. Specifically, Fe3O4 nanoparticles exhibit thermally fluctuating magnetic moments, which give rise to localized, time-varying magnetic fields within the Bi80Sb20 matrix. These localized magnetic fields interact with the charge carriers in two key ways. First, they modulate the scattering environment by disrupting conventional scattering mechanisms such as phonon scattering, point defect scattering, and grain boundary scattering. This dynamic magnetic environment effectively perturbs the trajectories of electrons in a manner that reduces their momentum relaxation rate, thereby enhancing mobility. Second, the fluctuating magnetic fields can induce a weak Lorentz-like force on the carriers, subtly redirecting their paths toward lower-scattering channels. This promotes more coherent carrier motion by enabling electrons to bypass energetically unfavorable regions or local scattering centers. Collectively, these effects—scattering suppression and path redirection—synergistically contribute to a significant enhancement in μ (Fig. 4d). This mechanism is supported by prior theoretical and experimental reports on magnetic-field-assisted transport in nanocomposites, particularly those incorporating superparamagnetic inclusions.48–54
image file: d5ta04222d-f4.tif
Fig. 4 Temperature dependence of (a) electrical conductivity, σ; (b) carrier concentration, n; (c) mobility μ for Bi80Sb20/x vol% Fe3O4 (x = 0.2, 0.4, 0.5, 0.8). (d) Mechanisms for enhancement of mobility after the addition of Fe3O4 into Bi80Sb20. (e) Electronic band diagram near the interface of Fe3O4 and Bi80Sb20. (f) Temperature dependence of the Seebeck coefficient, S, for Bi80Sb20/x vol% Fe3O4 (x = 0.2, 0.4, 0.5, 0.8) samples. (g) Room-temperature S as a function of n for all Bi80Sb20/x vol% Fe3O4 samples. The blue solid line represents calculations from the parabolic band model with the effect mass m* = 0.22 me. (h) Temperature-dependence of the power factor, PF, for all Bi80Sb20/x vol% Fe3O4 samples.

As shown in Fig. 4e, when Bi80Sb20 comes into contact with Fe3O4, a Schottky barrier forms at the interface due to the work function difference (ΔΦ ≈ 0.81 eV) between the n-type semiconductor Bi80Sb20 (Φ = 4.98 eV) and the metallic Fe3O4 (Φ = 5.79 eV) and Fig. S4 provides the electronic energy band arrangements of Fe3O4 and Bi80Sb20, which are calculated based on the DFT method. This leads to upward band bending on the Bi80Sb20 side. This creates an energy barrier that carriers must overcome to enter the Fe3O4 phase or traverse the interface. For low-energy electrons, this barrier can act as an obstacle, causing partial carrier reflection or scattering at the interface, thereby reducing the effective carrier mobility. At higher Fe3O4 content over 0.4 vol%, this effect with accumulation would decrease μ. As the temperature increases, the amplitude and frequency of phonons rise, significantly enhancing the probability of carrier-phonon collisions, which leads to a decrease in carrier mobility. Since phonon scattering remains the dominant mechanism in the 150–300 K range, the overall μ continues to decrease, even in the presence of thermionic emission effects across Schottky barriers.

The S of all samples is negative across the measured temperature range, confirming that electrons are the dominant charge carriers (Fig. 4f). Overall, the absolute values of S decrease with increasing temperature, which is inversely related to the trend of n. Notably, Fe3O4 incorporation enhances the S of the composites below 220 K. Above 220 K, the observed reduction in S with temperature for the composites is attributed to the increase of n, as confirmed in Fig. 4b. For the Bi80Sb20 matrix, at lower temperatures (below 180 K), the n increases with temperature (Fig. 4b) as more carriers are thermally excited from impurity states. The S increases concurrently (Fig. 4f) due to the energy-dependent scattering of these activated carriers and the upward shift of the Fermi level towards the conduction band edge. Then, an intermediate plateau occurs until 260 K, indicating the exhaustion of extrinsic carriers before significant intrinsic excitation occurs. The S coefficient reaches its peak value during this regime. Above 260 K, the n rises sharply again (Fig. 4b) due to intrinsic thermal excitation across the narrow band gap. The S subsequently decreases (Fig. 4f) as the detrimental contribution from minority carriers (bipolar diffusion) becomes significant, opposing the thermoelectric voltage generated by majority carriers.

Fig. 4g shows the fitted Pisarenko plot at 300 K on the basis of the single parabolic band (SPB) model under the assumption of acoustic phonon scattering.55,56 The experimental data for all Bi80Sb20/x vol% Fe3O4 samples align well with the fitted curve, indicating that the effective mass, m*, remains nearly unchanged (∼0.22 me) upon Fe3O4 addition. This suggests that the m* of the Bi80Sb20 matrix is largely preserved and that Fe3O4 nanoparticles influence TE performance primarily by tuning the carrier concentration rather than altering the band structure. Finally, based on the measured σ and S values, the PF values were calculated for all samples. Due to the higher σ and S, the x = 0.4 sample shows the highest PF of 34.4 μW cm−1 K−2 at 194 K—nearly three times higher than that of the matrix (Fig. 4h).

The total thermal conductivity, κ, of a material consists of lattice thermal conductivity, κL, electronic thermal conductivity, κe, and bipolar thermal conductivity, κb. κL arises predominantly from phonon transport, whereas κe results from heat transport by carrier transport. The narrow band gaps of Bi–Sb alloys facilitate intrinsic excitations, leading to substantial electron–hole pair diffusion and additional κb.57,58 The temperature dependence of κ for all samples is presented in Fig. 5a. For all the composite samples, κ increases with temperature. However, as the Fe3O4 content increases, the overall thermal conductivity initially increases and then decreases. The κe was calculated (Fig. 5b) according to the Wiedemann–Franz Law:

 
κe = LσT (1)
where L was estimated (Fig. S5) based on the single parabolic band (SPB) model with acoustic phonon scattering. The equations used in the SPB model are shown as follows:
 
image file: d5ta04222d-t1.tif(2)
 
image file: d5ta04222d-t2.tif(3)
 
image file: d5ta04222d-t3.tif(4)
where κB, e, λ, η, F, ξ and j are the Boltzmann constant, elementary charge, scattering factor, reduced chemical potential, Fermi integral function, reduced carrier energy and power of integral variable.


image file: d5ta04222d-f5.tif
Fig. 5 Temperature dependence of (a) thermal conductivity κ; (b) electronic thermal conductivity, κe; (c) lattice thermal conductivity, κL for Bi80Sb20/x vol% Fe3O4 (x = 0.2, 0.4, 0.5, 0.8). (d) Lattice thermal conductivity, κL, as a function of the filling fraction (x) for Bi80Sb20/x vol% Fe3O4 at 300 K (x = 0.2, 0.4). (e) Contribution from various phonon scattering mechanisms to κL in Bi80Sb20, Bi80Sb20/0.4 vol% Fe3O4. U, GB and PD denote the Umklapp phonon–phonon process, grain boundary scattering and point defect scattering, respectively. (f) Schematic representation of different scattering mechanisms within Bi80Sb20/Fe3O4 materials.

By deducting the κe from the total thermal conductivity, the sum of κb and κL is shown in Fig. S6a. By further deducting the estimated κb (Fig. S6b), the κL is obtained (Fig. 5c). It shows that the κL of the Bi80Sb20/Fe3O4 composites at 300 K decreases with increasing Fe3O4 content, reaching a minimum at x = 0.4. Fig. 5d plots room-temperature κL as a function of the Fe3O4 volume fraction. The κL of Bi80Sb20/Fe3O4 composites exhibits a non-monotonic variation with increasing Fe3O4 content. As the Fe3O4 content increases to x = 0.4, κL decreases from ∼0.88 W m−1 K−1 to ∼0.54 W m−1 K−1, representing a ∼39% reduction compared to the matrix. However, it subsequently increases to approximately 1.2 W m−1 K−1 at x = 0.5 vol%, before decreasing again to ∼1.0 W m−1 K−1 at x = 0.8 vol%. At very low loading (x = 0.2 vol%), Fe3O4 nanoparticles primarily act as isolated phonon scattering centers, introducing mass and strain field fluctuations at the interfaces, but their overall density remains limited. The sharp decrease in κL up to x = 0.4 vol% may be attributed to the optimal dispersion of Fe3O4 (Fig. 5c), which maximizes phonon scattering at grain boundaries and heterointerfaces. However, beyond this threshold, particle agglomeration during consolidation likely reduces the effective interfacial area, diminishing phonon scattering efficiency. Simultaneously, the relatively high intrinsic thermal conductivity of Fe3O4 (5.9 ± 0.6 W m−1 K−1)59 contributes to an increased volumetric heat transport, resulting in the observed rise in κL at x = 0.5 vol%. At even higher loading (x = 0.8 vol%), a minor suppression of κL could be due to the increased density of heterointerfaces, which enhances phonon scattering—especially for mid- to long-wavelength phonons. Additionally, localized strain and lattice mismatch at these interfaces may intensify at higher filler content, further impeding phonon transport. Overall, the non-monotonic evolution of κL arises from the competing effects of interface scattering, filler aggregation, and the intrinsic transport characteristics of the Fe3O4 phase.

To further analyze the contributions of different phonon scattering mechanisms to κL, the Callaway model was employed to fit the low-temperature lattice thermal conductivity of both Bi80Sb20 and Bi80Sb20/0.4 vol% Fe3O4. The Callaway model is expressed as

 
image file: d5ta04222d-t4.tif(5)
where x = ħω/kBT, kB is Boltzmann's constant, ħ is the reduced Planck's constant, v is the average sound velocity, θD is the Debye temperature, ω is the phonon frequency, and τ(x) is the total phonon scattering relaxation time. In this study, three primary phonon scattering mechanisms are considered: Umklapp process (U), grain boundary (GB) and point defect (PD). Thus, τ(x) can be expressed as
 
image file: d5ta04222d-t5.tif(6)
with the corresponding terms given by grain boundary scattering, v/G, where G is the average grain size; point defect scattering, 4; Umklapp scattering, 2TeθD/3T. Here, A and B are the fitting parameters for point defect scattering and Umklapp scattering, respectively. The fitting coefficient parameter values are listed in Table S1. Based on these parameters, the relative contributions of different phonon scattering mechanisms to κL are extracted, as shown in Fig. 5e. It is evident that in the low temperature region (T < 50 K), Umklapp and grain boundary scattering are dominant. The Callaway model analysis indicates that point defect scattering becomes the dominant mechanism reducing κL above 50 K, significantly surpassing the contribution from grain boundary scattering. This conclusion is quantitatively supported by the scattering contributions (Table S2) derived from the model fitting. ΔκB represents reduction of κL attributed to grain boundary scattering, while ΔκPD denotes additional reduction attributed to point defect scattering. As shown in Fig. S7, the ratio of ΔκPDκB increases markedly from 3.2 at 50 K to 5.8 at 300 K. This demonstrates that point defect scattering contributes significantly more than grain boundary scattering across this temperature range and its relative dominance increases with rising temperature.

As shown in Fig. 5f, the dominant sources of the scattering captured by the large ΔκPD value in the Callaway model fitting are: (1) intrinsic point defects: the substitutional disorder of Sb atoms on Bi sites) in the Bi80Sb20 solid solution matrix; (2) nano-Fe3O4-induced localized scattering: the atomic-scale disorder and, critically, the strong localized strain fields generated by the embedded nano-Fe3O4 particles. While the interfaces are extended defects, their phonon scattering effect at medium-to-high frequencies (T > 50 K) is effectively modeled within the point defect scattering framework due to its localized, short-wavelength nature. The synergistic combination of these two sources—especially the significant amplification provided by the nanocomposite-induced strain fields—leads to the exceptionally strong suppression of κL observed at higher temperatures. Isolating the exact quantitative contribution solely from the strict point defects versus the interface/strain effects within the aggregate ΔκPD would require sophisticated multiscale modeling beyond standard Callaway analysis, combined with further microstructural characterization.

The dimensionless figure of merit, zT, for all the samples is plotted in Fig. 6a. For every composition, zT increases with temperature and reaches a maximum and then declines at higher temperature. Throughout the entire 150–300 K range, the Bi80Sb20/x vol% Fe3O4 nanocomposites exhibit substantially higher zT values than pristine Bi80Sb20. The highest zT value of 0.31 at 242 K of the x = 0.4% sample represents an improvement of ∼100% over the matrix (Fig. 6b). Therefore, judicious optimization of the nano-Fe3O4 content markedly enhances the TE performance of Bi80Sb20-based composites. The average figure of merit, zTavg (150 to 300 K), for Bi80Sb20/0.4 vol% Fe3O4 and those of typical Bi–Sb based TE materials are presented in Fig. 6c. The zTavg of 0.29 for Bi80Sb20/0.4 vol% Fe3O4 is much higher than that of most Bi–Sb based TE materials and only slightly lower than that of Bi85Sb15/Graphene composites.31,60–65


image file: d5ta04222d-f6.tif
Fig. 6 (a) Temperature dependence of the TE figure of merit (zT) for Bi80Sb20/x vol%Fe3O4 composites. (b) Relative change rate of (ΔzT/zT0) of the zT for Bi80Sb20/x vol%Fe3O4 composites at 300 K, where zT0 is the zT value of Bi80Sb20 and ΔzT represents the zT difference after incorporation of Fe3O4 composites. (c) Comparison of the average zT (zTavg) from 150 to 300 K for Bi80Sb20/0.4 vol% Fe3O4 and other Bi–Sb based TE materials.31,60–65 Indentation for Vickers hardness testing of (d) Bi80Sb20 and (e) Bi80Sb20/0.4 vol% Fe3O4. (f) Comparison of the Vickers hardness values of Bi80Sb20/0.4 vol% iron tetraoxide and several typical TE materials at 0.2N load and 10s loading time.66–72

To validate the advantages of the dry preparation process in enhancing the TE performance, this study also subjected Bi80Sb20 and Bi80Sb20/0.2 vol% Fe3O4 powders to wet ball milling treatment. As shown in Fig. S8, compared to the dry ball milling samples using stearic acid as a dispersant, the wet-prepared composites exhibited a significant overall reduction in TE performance. Therefore, it can be concluded that combining dry ball milling with stearic acid as a dispersant is more advantageous. Quantitative comparison (Table S3) reveals the Bi80Sb20/0.4 vol% Fe3O4 composite achieves a zT of 0.29 at 300 K, which is twice the performance of pristine polycrystalline Bi80Sb20 (zT = 0.14) and surpasses the reported value of polycrystalline Bi85Sb15 (zT ∼0.15).26 This enhancement primarily stems from a striking 107% increase in σ (4220 vs. 1400 S cm−1), yielding a PF of 29.2 μW m−1 K−2 despite a moderate S reduction (83.2 vs. 89.1 μV K−1). While the thermal conductivity increases to 3.1 W m−1 K−1, the composite closes 34% of the zT gap to single-crystal Bi80Sb20 (zT ∼0.58).31 These results demonstrate that magnetic nanoparticle incorporation effectively mitigates the characteristic carrier mobility limitations of polycrystalline Bi–Sb alloys.

Hardness testing of Bi80Sb20/x vol% Fe3O4 nanocomposites

The Vickers hardness test was utilized to accurately evaluate the hardness of the materials (Fig. 6d and e). The measured Vickers hardness is (38.6 ± 0.8) HV for the Bi80Sb20 matrix and (47.5 ± 1.0) HV for the Bi80Sb20/0.4 vol% Fe3O4 composites. The Vickers hardness of Bi80Sb20/x vol% Fe3O4 and several typical TE materials is shown in Fig. 6f. Although the hardness of the composite is lower than that of Ag0.9Sb1.1Te3.1,66 GeTe,67 p-type Mg3Sb2,68 Bi2Te3-based alloys69 and PbSe,70 it is significantly higher than that of Cu2Se71 and PbTe.72 This enhancement in hardness suggests improved resistance to external mechanical stress and vibrations. The Fe3O4 nanoparticles act as reinforcement agents within the Bi80Sb20 matrix, which impede dislocation motion during plastic deformation to increase the hardness. In addition, the applied load applied during indentation can be partially transferred to the stiff Fe3O4 inclusions,73,74 reducing the local stress in the Bi80Sb20 matrix and increasing the overall resistance to deformation. These mechanisms collectively enhance the ability of the composite to resist deformation under mechanical stress.

Conclusions

In summary, a series of n-type Bi80Sb20/x vol% Fe3O4 nanocomposites were successfully fabricated via a combined process of ball milling of elemental Bi and Sb with stearic acid-assisted dispersion and SPS. Their TE properties were systematically investigated. The incorporation of Fe3O4 nanoparticles led to a significant enhancement in TE performance across the entire temperature range of 150–300 K. Among the composites, the sample with 0.4 vol% Fe3O4 exhibited the highest zT of 0.31 at 242 K, which represents an ∼100% improvement compared to that of the Bi80Sb20 matrix. Furthermore, the introduction of nano-Fe3O4 significantly improved the Vickers hardness of the composites. This study demonstrates that the incorporation of superparamagnetic Fe3O4 nanoparticles is an effective strategy to simultaneously enhance both the TE and mechanical performance of n-type Bi80Sb20-based alloys.

Conflicts of interest

The authors declare no competing financial interest.

Data availability

The data that support the findings of this study are available from the corresponding author upon reasonable request.

The supplementary information file contains additional data supporting the main text. See DOI: https://doi.org/10.1039/d5ta04222d.

Acknowledgements

The authors acknowledge financial support from the National Natural Science Foundation of China (NSAF) (No. U2230131 and No. U2141208), JSPS KAKENHI (No. JP24K23069), State Key Laboratory of Powder Metallurgy, Priority Academic Program Development of Jiangsu Higher Education Institutions and Jiangsu Collaborative Innovation Center for Advanced Inorganic Function Composites, Nanjing Tech University. The authors also acknowledge the staff of the BL17B beamline (https://cstr.cn/31129.02.NFPS.BL17B) at the National Facility for Protein Science in Shanghai (NFPS, https://cstr.cn/31129.02.NFPS), Shanghai Advanced Research Institute, Chinese Academy of Sciences, for their technical support in PXRD and GIWAXS data collection and analysis.

References

  1. T. M. Tritt, Science, 1996, 272, 1276,  DOI:10.1126/science.272.5266.1276.
  2. R. T. Littleton, T. M. Tritt, J. W. Kolis and D. R. Ketchum, Phys. Rev. B:Condens. Matter Mater. Phys., 1999, 60, 13453,  DOI:10.1103/PhysRevB.60.13453.
  3. Z. C. Gong, A. Suwardi and J. Cao, Adv. Funct. Mater., 2025, 2423371,  DOI:10.1002/adfm.202423371.
  4. D. W. Zhang, N. N. Y. Qi, S. F. D. Solco, X. W. Li and A. Suwardi, ACS Energy Lett., 2024, 9(5), 2240,  DOI:10.1021/acsenergylett.4c00680.
  5. M. Shu, Z. X. He, J. J. Zhu, Y. R. Ji, X. F. Zhang, C. R. Zhang, M. R. Chen and P. A. Zong, Adv. Funct. Mater., 2025, 35(5), 2414660,  DOI:10.1002/adfm.202414660.
  6. X. Shi, K. Zheng, M. Hong, W. Liu, R. Moshwan, Y. Wang, X. Qu, Z. Chen and J. Zou, Chem. Sci., 2018, 9, 7376,  10.1039/c8sc02397b.
  7. X. Shi, K. Zheng, W. Liu, Y. Wang, Y. Yang, Z. Chen and J. Zou, Adv. Energy Mater., 2018, 8, 1800775,  DOI:10.1002/aenm.201800775.
  8. T. Yang, J. Liu, R. Zhou, Z. Chen, H. Xu, S. Qiao and M. Monteiro, J. Mater. Chem. A, 2014, 2, 18139,  10.1039/c4ta04301d.
  9. X. Shi, L. Chen and C. Uher, Int. Mater. Rev., 2016, 61, 379,  DOI:10.1080/09506608.2016.1183075.
  10. B. Poudel, Q. Hao, Y. Ma, Y. Lan, A. Minnich, B. Yu, X. Yan, D. Wang, A. Muto and D. Vashaee, Science, 2008, 320, 634,  DOI:10.1126/science.1156446.
  11. Y. Saberi and S. A. Sajjadi, J. Alloys Compd., 2022, 904, 163918,  DOI:10.1016/j.jallcom.2022.163918.
  12. X. Shi, J. Yang, J. R. Salvador, M. Chi, J. Y. Cho, H. Wang, S. Q. Bai, J. Yang, W. Q. Zhang and L. D. Chen, J. Am. Chem. Soc., 2011, 133, 7837,  DOI:10.1021/ja111199y.
  13. P. A. Zong, R. Hanus, M. Dylla, Y. S. Tang, J. C. Liao, Q. H. Zhang, G. J. Snyder and L. D. Chen, Energy Environ. Sci., 2017, 10, 183,  10.1039/c6ee02467j.
  14. B. B. Jiang, Y. Yu, J. Cui, X. X. Liu, L. Xie, J. C. Liao, Q. H. Zhang, Y. Huang, S. C. Ning, B. H. Jia, B. Zhu, S. Q. Bai, L. D. Chen, S. J. Pennycook and J. Q. He, Science, 2021, 371, 830,  DOI:10.1126/science.abe1292.
  15. S. W. Tang, M. X. Wu, S. L. Bai, D. M. Luo, J. Y. Zhang and S. B. Yang, J. Alloys Compd., 2022, 907, 164439,  DOI:10.1016/j.jallcom.2022.164439.
  16. J. P. Heremans, V. Jovovic, E. S. Toberer, A. Saramat, K. Kurosaki, A. Charoenphakdee, S. Yamanaka and G. J. Snyder, Science, 2008, 321, 554,  DOI:10.1126/science.1159725.
  17. C. G. Fu, S. Q. Bai, Y. T. Liu, Y. S. Tang, L. D. Chen, X. B. Zhao and T. J. Zhu, Nat. Commun., 2015, 6, 8144,  DOI:10.1038/ncomms9144.
  18. Q. Shen, L. Chen, T. Goto, T. Hirai, J. Yang, G. P. Meisner and C. Uher, Appl. Phys. Lett., 2001, 79, 4165,  DOI:10.1063/1.1425459.
  19. H. Liu, X. Shi, F. Xu, L. Zhang, W. Zhang, L. Chen, Q. Li, C. Uher, T. Day and G. J. Snyder, Nat. Mater., 2012, 11, 422,  DOI:10.1038/NMAT3273.
  20. Y. He, T. Day, T. S. Zhang, H. L. Liu, X. Shi, L. D. Chen and G. J. Snyder, Adv. Mater., 2014, 26, 3974,  DOI:10.1002/adma.201400515.
  21. G. Joshi, H. Lee, Y. C. Lan, X. W. Wang, G. H. Zhu, D. Z. Wang, R. W. Gould, D. C. Cuff, M. Y. Tang, M. S. Dresselhaus, G. Chen and Z. F. Ren, Nano Lett., 2008, 8, 4670,  DOI:10.1021/nl8026795.
  22. J. Liang, D. Fan, P. Jiang, H. Liu and W. Zhao, Intermetallics, 2017, 87, 27,  DOI:10.1016/j.intermet.2017.03.024.
  23. M. S. Kim, T. Sasakawa, Y. Echizen and T. Takabatake, Jpn. J. Appl. Phys., 2003, 42, 6512 CrossRef CAS.
  24. D. Y. Chung, T. Hogan, P. Brazis, M. Rocci-Lane, C. Kannewurf, M. Bastea, C. Uher and M. G. Kanatzidis, Science, 2000, 287, 1024,  DOI:10.1126/science.287.5455.1024.
  25. X. Zhang and L. D. Zhao, J. Mater., 2015, 1, 92,  DOI:10.1016/j.jmat.2015.01.001.
  26. Y. Pan, B. He, X. L. Feng, F. Li, D. Chen, U. Burkhardt and C. Felser, Nat. Mater., 2025, 24, 76–82,  DOI:10.1038/s41563-024-02059-9.
  27. P. K. Nguyen, K. H. Lee, J. Moon, S. I. Kim, K. A. Ahn and L. H. Chen, Nanotechnology, 2012, 23, 415604,  DOI:10.1088/0957-4484/23/41/415604.
  28. B. Poudel, Q. Hao, Y. Ma, Y. Lan, A. Minnich, B. Yu, X. Yan, D. Wang, A. Muto, D. Vashaee, X. Chen, J. Liu, M. S. Dresselhaus, G. Chen and Z. Ren, Science, 2008, 320, 634,  DOI:10.1126/science.1156446.
  29. M. R. Dirmyer, J. Martin, G. S. Nolas, A. Sen and J. V. Badding, Small, 2009, 5, 933,  DOI:10.1002/smll.200801206.
  30. Z. Chen, M. Y. Lin, G. D. Xu, S. Chen, J. H. Zhang and M. M. Wang, J. Alloys Compd., 2014, 588, 384,  DOI:10.1016/j.jallcom.2013.11.065.
  31. M. S. El-Asfoury, M. N. A. Nasr, K. Nakamura and A. Abdel-Moneim, J. Alloys Compd., 2018, 745, 331–340,  DOI:10.1016/j.jallcom.2018.02.040.
  32. R. Huang, C. Song, F. Cai and L. Li, J. Phys. Chem. Solids, 2007, 68, 1400,  DOI:10.1016/j.jpcs.2007.02.047.
  33. C. M. Song, R. J. Huang, M. Zhou, L. H. Gong and L. F. Li, J. Phys. Chem. Solids, 2010, 71, 999,  DOI:10.1016/j.jpcs.2010.04.008.
  34. N. Mingo, D. Hauser, N. P. Kobayashi, M. Plissonnier and A. Shakouri, Nano Lett., 2009, 9, 711,  DOI:10.1021/nl8031982.
  35. S. L. Kim, K. H. Lee, H. A. Mun, H. S. Kim, S. W. Hwang and J. W. Roh, Science, 2015, 348, 109,  DOI:10.1126/science.aaa4166.
  36. J. Zhou, X. Li, G. Chen and R. Yang, Phys. Rev. B:Condens. Matter Mater. Phys., 2010, 82, 115308,  DOI:10.1103/PhysRevB.82.115308.
  37. W. Y. Zhao, Z. Y. Liu, Z. G. Sun, Q. J. Zhang, P. Wei, X. Mu, H. Y. Zhou, C. C. Li, S. F. Ma, D. Q. He, P. X. Ji, W. T. Zhu, X. L. Nie, X. L. Su, X. F. Tang, B. G. Shen, X. L. Dong, J. Yang, Y. Liu and J. Shi, Nature, 2017, 549, 247,  DOI:10.1038/nature23667.
  38. R. Lu, J. S. Lopez, Y. Liu, T. P. Bailey, A. A. Page, S. Wang, C. Uher and P. F. P. Poudeu, J. Mater. Chem. A, 2019, 7, 11095,  10.1039/c9ta01156k.
  39. G. A. Slack, Phys. Rev., 1962, 126, 427,  DOI:10.1103/PhysRev.126.427.
  40. C. C. Li, S. F. Ma, P. Wei, W. T. Zhu, X. L. Nie, X. H. Sang, Z. G. Sun, Q. J. Zhang and W. Y. Zhao, Energy Environ. Sci., 2020, 13, 535,  10.1039/c9ee03446c.
  41. D. He, X. Mu, H. Zhou, C. C. Li, S. F. Ma, P. X. Ji, W. K. Hou, P. Wei, W. T. Zhu, X. L. Nie and W. Y. Zhao, J. Electron. Mater., 2018, 47, 3338,  DOI:10.1007/s11664-017-5842-9.
  42. J. Hafner, J. Comput. Chem., 2008, 29(13), 2044–2078,  DOI:10.1002/jcc.21057.
  43. P. E. Blöchl, Phys. Rev. B:Condens. Matter Mater. Phys., 1994, 50(24), 17953,  DOI:10.1103/PhysRevB.50.17953.
  44. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77(18), 3865,  DOI:10.1103/PhysRevLett.77.3865.
  45. X. Gonze, J. P. Michenaud and J. P. Vigneron, Phys. Scr., 1988, 37(5), 785,  DOI:10.1088/0031-8949/37/5/022.
  46. M. Ernzerhof and G. E. Scuseria, J. Chem. Phys., 1999, 110(11), 5029–5036,  DOI:10.1016/S0009-2614(03)01086-8.
  47. A. H. Larsen, J. J. Mortensen, J. Blomqvist, I. E. Castelli, R. Christensen, M. Dułak, J. Friis, M. N. Groves, B. Hammer, C. Hargus and J. Phys-Condens, Matter, 2017, 29(27), 273002,  DOI:10.1088/1361-648X/aa680e.
  48. S. L. Dudarev, G. A. Botton, S. Y. Savrasov, C. Humphreys and A. P. Sutton, Phys. Rev. B:Condens. Matter Mater. Phys., 1998, 57(3), 1505,  DOI:10.1103/PhysRevB.57.1505.
  49. A. Jain, S. P. Ong, G. Hautier, W. Chen, W. D. Richards, S. Dacek, S. Cholia, D. Gunter, D. Skinner and G. Ceder, APL Mater., 2013, 1, 011002,  DOI:10.1063/1.4812323.
  50. S. Lerdkanchanaporn and D. Dollimore, Thermochim. Acta, 1998, 324, 15–23,  DOI:10.1016/S0040-6031(98)00519-X.
  51. E. Combe, R. Funahashi, T. Takeuchi, T. Barbier and K. Yubuta, J. Alloys Compd., 2017, 692, 563–568,  DOI:10.1016/j.jallcom.2016.09.008.
  52. J. Wang, C. Zhu, F. Luo, J. F. Wang, X. He, Y. Zhang, H. X. Liu and Z. G. Sun, ACS Appl. Mater. Interfaces, 2023, 15(6), 8105–8119,  DOI:10.1021/acsami.2c20778.
  53. A. E. Cohen, J. Phys. Chem. A, 2009, 113, 11084,  DOI:10.1021/jp907113p.
  54. R. Geng, H. M. Luong, M. T. Pham, R. Das, K. S. Repa, J. Robles-Garcia, T. A. Duong, H. T. Pham, T. H. Au, N. D. Lai, G. K. Larsen, M. H. Phan and T. D. Nguyen, Mater. Horiz., 2019, 6, 1913,  10.1039/c9mh00265k.
  55. H. L. Zhuang, J. Pei, B. Cai, J. Dong, H. Hu, F. H. Sun, Y. Pan, G. J. Snyder and J. F. Li, Adv. Funct. Mater., 2021, 31, 2009681,  DOI:10.1002/adfm.202009681.
  56. A. F. May, E. S. Toberer, A. Saramat and G. J. Snyder, Phys. Rev. B:Condens. Matter Mater. Phys., 2009, 80, 125205,  DOI:10.1103/PhysRevB.80.125205.
  57. T. T. Lu, S. Y. Wang, H. Li and X. F. Tang, Intermetallics, 2013, 32, 96–102,  DOI:10.1016/j.intermet.2012.08.007.
  58. X. W. Wu, Z. Fan, H. T. Zhu, T. Y. Wang, M. Liu, J. Li, N. Chen, Q. L. Liu, Z. Lu, G. d. Li, X. Qian, T. H. Liu, R. G. Yang, X. Y. Niu, Q. Zhao, Z. L. Li, S. F. Wang and H. Z. Zhao, Nat. Commun., 2025, 16, 3534,  DOI:10.1038/s41467-025-58491-1.
  59. J. G. Kang, H. Jang, J. Ma, Q. Yang, K. Hattar, Z. Diao, R. L. Yuan, J. M. Zuo, S. Sinha, D. G. Cahill and P. V. Braun, J. Phys. Chem. C, 2021, 125(12), 6897–6908,  DOI:10.1021/acs.jpcc.1c00411.
  60. Z. Chen, Y. M. Han, M. Zhou, C. M. Song, R. J. Huang, Y. Zhou and L. F. Li, J. Phys. Chem. Solids, 2014, 75, 523,  DOI:10.1016/j.jpcs.2013.12.006.
  61. Z. Chen, M. Zhou, R. J. Huang, C. M. Song, Y. Zhou and L. F. Li, J. Alloys Compd., 2012, 511, 85,  DOI:10.1016/j.jallcom.2011.08.076.
  62. Z. Chen, M. Zhou, C. M. Song, Y. Zhou and L. F. Li, J. Electron. Mater., 2012, 41, 1725,  DOI:10.1007/s11664-012-2063-0.
  63. K. C. Lukas, G. Joshi, K. Modic, Z. F. Ren and C. P. Opeil, J. Mater. Sci., 2012, 47, 5729,  DOI:10.1007/s10853-012-6463-6.
  64. M. S. El-Asfoury, S. M. Abdou and A. Nassef, JOM, 2021, 73, 2808,  DOI:10.1007/s11837-021-04699-7.
  65. M. Zhou, Z. Chen, X. Chu and L. Li, J. Electron. Mater., 2012, 41, 1263,  DOI:10.1007/s11664-012-1914-z.
  66. Y. Wu, P. F. Qiu, Y. Yu, Y. F. Xiong, T. T. Deng, O. Cojocaru-Miredin, M. Wuttig, X. Shi and L. D. Chen, J. Mater., 2022, 8, 1095,  DOI:10.1016/j.jmat.2022.07.005.
  67. S. Perumal, S. Roychowdhury and K. Biswas, Inorg. Chem. Front., 2016, 3, 125,  10.1039/c5qi00230c.
  68. J. Li, S. Zhang, F. Jia, S. Q. Zheng, X. L. Shi, D. Q. Jiang, S. Wang, G. W. Lu, L. M. Wu and Z. G. Chen, Mater. Today Phys., 2020, 15, 100269,  DOI:10.1016/j.mtphys.2020.100269.
  69. L. D. Zhao, B. P. Zhang, J. F. Li, M. Zhou, W. S. Liu and J. Liu, J. Alloys Compd., 2008, 455, 259,  DOI:10.1016/j.jallcom.2007.01.015.
  70. M. S. Darrow, W. B. White and R. Roy, J. Mater. Sci., 1969, 4, 313,  DOI:10.1007/BF00550400.
  71. L. Zhao, X. Wang, F. Y. Fei, J. Wang, Z. Cheng, S. Dou, J. Wang and G. J. Snyder, J. Mater. Chem. A, 2015, 3, 9432,  10.1039/c5ta01667c.
  72. Y. Gelbstein, G. Gotesman, Y. Lishzinker, Z. Dashevsky and M. P. Dariel, Scr. Mater., 2008, 58, 251,  DOI:10.1016/j.scriptamat.2007.10.012.
  73. M. Takeda, T. Onishi, S. Nakakubo and S. Fujimoto, Mater. Trans., 2009, 50, 2242,  DOI:10.2320/matertrans.M2009097.
  74. W. Long, X. Zeng and H. Jia, Trans. Indian Inst. Met., 2019, 72, 983,  DOI:10.1007/s12666-018-01559-y.

Footnote

These authors contributed equally.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.