Abstract
GeTe-based materials with superior thermoelectric properties promise great potential for waste heat recovery. However, the lack of appropriate diffusion barrier materials (DBMs) limits not only the energy conversion efficiency but also the service reliability of the thermoelectric devices. Here, we propose a design strategy based on phase equilibria diagrams from first-principles calculations and identify transition metal germanides (e.g., NiGe and FeGe2) as the DBMs. Our validation experiment confirms the excellent chemical and mechanical stabilities of the interfaces between the germanides and GeTe. We also develop a process for scaling up the GeTe production. Combining with module geometry optimization, we fabricate an eight-pair module using mass-produced p-type Ge0.89Cu0.06Sb0.08Te and n-type Yb0.3Co4Sb12 and achieve a record-high efficiency of 12% among all reported single-stage thermoelectric modules. Our work thus paves the way for waste heat recovery based on completely lead-free thermoelectric technology.
High-efficiency GeTe thermoelectric module was realized by computation-guided design of diffusion barrier materials.
INTRODUCTION
Industrial production is one of the top-three energy consuming sectors worldwide and, meanwhile, produces massive waste heat (1). Thermoelectric (TE) generators, which can directly convert such waste heat into electricity through the Seebeck effect, provide a promising strategy to recover low-grade thermal energy (2, 3) and serve as an effective tool for CO2 mitigation. The maximum energy conversion efficiency ηmax of a TE device assembly (or module) can be expressed as (4)
| (1) |
where ΔT = Th − Tc is the difference between the hot-side temperature (Th) and cold-side temperature (Tc) of the TE module and is related to the temperature-dependent Seebeck coefficient (S), electrical conductivity (σ), and lattice and electronic thermal conductivities (κL and κe) of the TE materials through z = S2σ/(κL + κe). While boosting the zT value at a specific temperature is currently the main thrust of the research on TE materials, ηmax is of greater importance when evaluating whether a material is competitive for practical applications.
The temperature range from 500 to 800 K (or the mid-temperature range) is usually considered the most relevant to waste heat recovery (5). Within this range, PbTe has long been perceived to be the best-performed TE material. However, the toxicity of Pb and the evaporation of Pb at high temperature have always been a general concern. To develop lead-free alternatives for PbTe, skutterudites (SKDs) have been intensively studied (6). Unfortunately, p-type SKDs still suffer from inferior performance to their n-type counterparts. In recent years, GeTe stands out for its excellent performance in the mid-temperature range (7–9). By optimizing carrier concentration (10, 11), band structure (12–15), and microstructure (16–18), p-type GeTe has achieved peak zT values above 2.0 and provides an attractive option for integrating with n-type SKDs to fabricate TE modules.
Compared with the extensive studies on material properties, there have been few reports on GeTe-based TE modules. GeTe single legs have demonstrated ηmax up to 14% (19, 20), but such a structure is not practical for TE applications. The p-GeTe/n-Mg3Sb2 module yielded a ηmax higher than 10%, but the Th of ~600 K is not high enough to unleash the full potential of GeTe for mid-temperature applications (21). N-type SKDs were introduced so that the Th can reach ~850 K, but the ηmax of 7.8% is still much lower than its theoretical value (22). Recent work used high-entropy p-(Ge, Pb, Ag, Sb, Bi)Te and n-PbTe, where the single-stage module achieved a ηmax of 10.5%. But this module still highly relies on Pb (23).
Given the zT of a TE material as a function of T, a TE module made of this material has a maximum efficiency ηmax as estimated from Eq. 1 by finite element method (FEM) calculation (see Materials and Methods). Therefore, one can assess the performance of a TE module by defining the completion ratio (rη) between the measured and theoretical values of ηmax. To increase the theoretical ηmax of a TE module, it is necessary to increase the ZT (i.e., the overall zT profile) and ΔT (preferably with the Th higher than the temperature where the zT peaks). Practically, it is usually more challenging to improve rη, i.e., having the measured ηmax approaching the theoretical ηmax. Both the Th and rη are strongly related to the quality of the TE module.
A major obstacle to realizing high-performance GeTe-based modules is the lack of an ideal diffusion barrier material (DBM), which affects not only the energy conversion efficiency through Th and rη but also the high-temperature service reliability of the TE modules (24, 25). However, experimental studies were seldom dedicated to the selection of the DBMs for GeTe-based modules. Although several materials [such as Mo (22), Ti (26), SnTe (21), and Al/Si (27)] have been used as the DBMs in GeTe-based TE modules, the achieved rη as shown in Fig. 1A still cannot compete with the mature TE materials, such as half-Heusler (HH) (28–31), SKDs (32, 33), and Bi2Te3-based modules (34, 35). Thus, designing previously unidentified DBMs has become an urgent need for GeTe-based modules.
Fig. 1. Performance of state-of-the-art thermoelectric (TE) modules.
(A) Ratio between experimental and theoretical ηmax (defined as completion ratio rη) of TE modules plotted versus the average zT of p and n legs for different TE material systems, including HH (28–31), SKD (32, 33), Bi2Te3 (34, 35), Cu2Se (56), Mg3Sb2 (57, 64, 65), PbTe (58), SnSe (66) and GeTe (21, 22)-based modules. The details of these TE modules from these references are listed in Table S1 in the supplementary material (SM). (B) Comparison of ηmax of single-stage TE modules reported in the literature, including GeTe (21–23), SKD (33), PbTe (58), HH (29), Mg3Sb2 (57), and Cu2Se-based (56) modules (HH and SKD stand for half-Heusler and skutterudite, respectively). All the theoretical ηmax values in (A) are recalculated by finite element method (FEM) on the basis of the reported data in the corresponding literature without considering heat loss in the integration process (see Materials and Methods).
Here, by discovering chemically inert germanide DBMs, realizing a scalable GeTe synthesis process, and optimizing module geometry, we develop a full-fledged technique for fabricating high-performance mid-temperature lead-free TE modules ready for practical applications. As both the low Th and low ηmax can be partly attributed to the issues on the DBM, we first carry out screening of previously unidentified DBMs by combining first-principles calculation and experimental validation, which identify transition metal germanides (e.g., NiGe and FeGe2) as the candidates. Next, we develop a process for the mass production of GeTe materials with well-reproducible TE performance. Then, assisted by FEM calculation based on a full-parameter model for module geometry optimization (36), we fabricate a TE module with eight pairs of p-type Ge0.89Cu0.06Sb0.08Te and n-type Yb0.3Co4Sb12 SKD, which achieves a record-high ηmax of 12% under a temperature difference of 545 K. Figure 1B compares our current module with other high-performance single-stage modules reported in the literature. From Fig. 1A, it can also be seen that the rη of our module is above 88%, approaching the most mature TE module based on the HH material. This work demonstrates that GeTe is highly promising for lead-free power generation applications in the mid-temperature range.
RESULTS AND DISCUSSION
Screening of diffusion barrier materials
The DBM layer connects the TE material and the electrode and, therefore, is required to be chemically inert, but mechanically adhesive, to both the TE and electrode materials. The DBMs should also have low thermal and electrical resistivities to minimize energy loss during operation. In addition, matching the coefficient of thermal expansion (CTE) with the TE material is also critical (37, 38). If the selection of DBMs is confined to elemental metals, then there is nearly no satisfactory choice for GeTe. Some commonly used metal elements such as Ti, Mn, Cu, Ag, and Cr are effective heavy dopants in GeTe (9, 15, 18, 39, 40), indicating that they are reactive with GeTe. Other chemically inert metal elements such as W and Mo have too high melting points to form direct bonding at the synthesis temperature of GeTe. In addition, to reduce the overall cost of the module, noble metal elements such as Au and Pt are excluded.
In this work, we started with Fe, Co, and Ni as the candidates because we found that these elements were difficult to be doped into GeTe. Taking a sandwich structure of GeTe/Ni/GeTe, we carried out a bonding experiment. We observed an interfacial reaction layer between Ni and GeTe, as shown in fig. S1. By elemental mapping and line scan under scanning electron microscopy (SEM), it was determined that the reaction layer consists of Ni and Ge, suggesting the formation of a germanide layer. After annealing at 773 K for 30 min, the sample exhibited cracks between the Ni and germanide layer (see fig. S1B), suggesting mismatched CTEs of the two materials. Similar results were also observed when Fe and Co were used in the middle layer of the sandwich structure (see figs. S2 and S3).
The results above, on the one hand, excluded the possibility of using elemental Fe, Co, or Ni as the DBM but, on the other hand, inspired us to consider using germanides as the DBM. The germanides could naturally be inert and adhesive to the GeTe layer. However, there exist a large number of possible germanides of Fe, Co, and Ni. To screen out the best candidate, we resort to the phase equilibria diagram (PED) obtained by first-principles calculations. We considered all possible related compounds in Materials Project (41) and in the Inorganic Crystal Structure Database (42) and calculated their formation energy to obtain the PED (see table S1 for the details of the considered compounds). Figure 2 (A to C) shows the PEDs for the Fe-Ge-Te, Co-Ge-Te, and Ni-Ge-Te systems. The first requirement for a compound to form a thermodynamically stable interface with GeTe is to have a direct line connecting this compound with GeTe on the PED. These lines are highlighted in red color in the figures. It can be seen that FeGe, NiGe, FeTe2, CoTe2, and NiTe2 satisfy this requirement. However, the three tellurides usually have very high resistivity compared with the germanides (43–45) and therefore are not suited to be used as DBM.
Fig. 2. Design of diffusion barrier materials informed by phase equilibrium diagrams.
(A to C) Calculated phase equilibria diagrams (PEDs) at 0 K for the Fe-Ge-Te, Co-Ge-Te, and Ni-Ge-Te systems, respectively. Solid and empty circles represent stable and unstable phases at 0 K. (D to F) Calculated interfacial reaction energy (EIR) between GeTe and potential diffusion barrier materials for the Fe-Ge-Te, Co-Ge-Te, and Ni-Ge-Te systems, respectively.
The PEDs in Fig. 2 (A to C) were obtained by density functional theory calculations at 0 K. For the purpose of DBM screening, it is ideal that the PEDs are calculated at the temperature for synthesizing GeTe. Unfortunately, these calculations are usually too expensive. The missing information from the 0-K PEDs is that the high-temperature stabilized phases might be overlooked. To remedy this issue, we also consider the metastable phases of germanides at 0 K for DBMs, which could be the stable high-temperature phases. These phases are marked by empty circles in Fig. 2 (A to C). We used the interfacial reaction energy (EIR) as the criterion. The details for evaluating EIR are given in the Supplementary Materials. As long as EIR is positive for all possible reactions, the considered compound should be chemically inert to GeTe. Figure 2 (D to F) shows the results for the Fe, Co, and Ni systems (see table S3 for the values). In addition to FeGe and NiGe as found in Fig. 2 (A to C), FeGe2 and NiGe2 also appear to be the candidates.
On the basis of the screening results, we assessed two of the candidates, i.e., NiGe and FeGe2, by the melting-annealing method, as from the Ni-Ge and Fe-Ge binary phase diagrams the two germanides can sustain the soldering temperature (~850 K) for fabricating the GeTe-based module (46, 47). The structure characterizations are given in fig. S4. The measured electrical and thermal conductivity of NiGe and FeGe2 is one order of magnitude higher than that of GeTe (see fig. S5). Also, the CTEs of the two germanides are very close to that of the rhombohedral GeTe (see fig. S6). These properties satisfy the basic requirements for the DBM. We then prepared the sandwich-structured samples of NiGe and FeGe2 to investigate the interfacial properties. As shown in figs. S7 and S8, the elemental line scans show no reaction layers near the interfaces between the germanides and GeTe.
Given the higher electrical and thermal conductivities of NiGe than FeGe2 (fig. S5), we will focus on the GeTe/NiGe interface below. A high-resolution transmission electron microscope was used to further examine the GeTe/NiGe interface. As shown in Fig. 3 (A and B), the element distribution of Ni and Te exhibits a sharp interface, and the elemental line scan in Fig. 3C shows no interdiffusion beyond 2 nm. These results demonstrate that NiGe is chemically inert to GeTe, consistent with our first-principles calculations. The measured interfacial resistivity between the GeTe and NiGe layers is initially ~1 microhm·cm2 and remains below 3 microhm·cm2 after aging at 773 K for 10 days (Fig. 3D and fig. S9). Meanwhile, the interfacial microstructure exhibits no notable changes after the aging test (see fig. S7), demonstrating superior thermal and chemical stability. All the results above support that NiGe is an excellent DBM for GeTe-based modules. To understand the stability of the interface between GeTe and NiGe with no notable interdiffusions observed, we theoretically study the bonding features of this interface. Strong electron accumulation is observed in the middle areas between the atoms around the GeTe/NiGe interface (see fig. S10), suggesting the formation of covalent or metallic-like chemical bonds between these atoms. This bonding feature is also observed in the interface between metal and HH compounds (48) and can account for the firmly bonded interface observed in experiments. We further calculate the work of separation (Wsep) of this interface, which indicates the work required to separate the interface into two free surfaces and can be used to evaluate the bonding strength of the interface (49, 50). The Wsep of the typical interface, i.e., GeTe(001)/NiGe(011), is calculated to be 0.91 J/m2, which is larger than that of the GeTe(001) and GeTe(111) surfaces (0.13 and 0.27 J/m2, respectively). The much larger Wsep of the GeTe/NiGe interface than that of GeTe surfaces means that the bonding of the interface is even stronger than the interlayer interactions in GeTe, thus verifying the strength of the interface in the device.
Fig. 3. Validation of the predicted diffusion barrier materials.
(A) Energy-dispersive spectrometry (EDS) mapping and (C) elemental line scan results of Ni, Te, and Ge under transmission electron microscopy. (B) Atomic structure to show a sharp interface between GeTe and NiGe. (D) Interfacial contact resistivity of NiGe and FeGe2 barrier layers after aging and compared to literature results, such as Mo (22), Ti (26), Al66Si34 (27), and SnTe (21). “Broken” in (D) means that the Fe layer peels off the GeTe matrix, and the GeTe/Fe interface is broken after the aging.
Thermoelectric performance of mass-produced GeTe materials
To scale up TE module fabrication, it is essential to realize the mass production of the GeTe materials that can achieve reproducible high TE performance. In this work, Ge0.89Cu0.06Sb0.08Te (GeTe for short) and Yb0.3Co4Sb12 (SKD for short) were chosen as p-type and n-type TE materials, respectively, for the fabrication of TE modules. Mass production of the SKD materials has been realized in our previous study (36, 51). As for the GeTe materials, although the peak zT values have exceeded 2.0, the preparation of the pellets for device fabrication repeatable with high zT values was still limited to lab scale of ~5 g (21), hindering the development of large GeTe-based modules with high efficiency. Here, we achieved the sintering of cylindrical pellet of ~54 g with uniform material properties as discussed below, which can once support for the integration of three to four modules.
The GeTe samples were cut from the sintered pellet, as illustrated in the inset of Fig. 4A. The x-ray diffraction (XRD) patterns (see fig. S11) show that the main phase of the samples is GeTe in the rhombohedral structure with a small amount of Ge precipitates. SEM and energy-dispersive spectrometry (EDS) analyses (see fig. S12) confirm that Ge, Cu, Sb, and Te are uniformly distributed in the samples. The observation of Ge secondary phases is consistent with the XRD results. The characterizations above confirmed good homogeneity of the large-size GeTe pellet.
Fig. 4. Thermoelectric performance of mass-produced GeTe materials.
Temperature-dependent (A) electrical conductivity, (B) Seebeck coefficient, (C) thermal conductivity, and (D) figure of merit for p-type Ge0.89Cu0.06Sb0.08Te materials. The inset in (A) shows a schematic illustrating the sliced bars and sheets used for measurements of thermoelectric (TE) transport properties. Comparisons are made with lab-scale samples and the record values in the literature (60).
Temperature-dependent electrical and thermal transport properties of three groups of GeTe samples were measured and plotted in Fig. 4. By comparing the mass-produced and lab-scale samples, it can be seen that the TE properties of the GeTe samples are well reproducible (also see fig. S13). The zT values of ~1.7 at 750 K (Fig. 4D) achieved by our mass-produced samples are lower than the lab-scale samples and the record value in the literature (23), which may be attributed to slight oxidation introduced in the hot-pressing process. Nevertheless, the zT of 1.7 manifests that the mass-produced GeTe could yield superior TE performance, compared with p-type SKD for mid-temperature applications.
Geometry optimization of the GeTe/SKD module
A high-performance TE module requires that TE and related physical properties of p- and n-legs should be matched as much as possible. As the TE performance of n-type GeTe is evidently inferior to p-type GeTe (52, 53), we used n-type lead-free SKD to couple with p-type GeTe. Moreover, the CTE of SKD is comparable with the rhombohedral GeTe (54) so that reliable connection through electrodes between the two materials can be made.
Optimization of the module geometry is critical for maximizing performance. FEM-based numerical simulations (see Materials and Methods) were used here to design the optimal geometry of the GeTe/SKD module, as illustrated in Fig. 5A. The Th and Tc were set to 873 and 293 K, respectively, and the total cross-section area, Apn = Ap + An, was kept constant, where Ap and An represent the cross-section areas for a single p- and n-leg, respectively. All these parameters and the measured TE properties (fig. S14) were imported into the full-parameter multi-physical field model that we proposed to optimize the module geometry (36).
Fig. 5. Optimization of the geometry of GeTe/skutterudite (SKD) single-stage module.
(A) Schematic illustration of p-GeTe and n-SKD device and related structural parameters. Simulated (B) power density Pd and (C) maximum conversion efficiency ηmax as a function of the ratio of the cross-sectional areas of the p- to n-legs (Ap/An) and the ratio of the height to the total cross-sectional area (H/Apn) for the GeTe/SKD module. These calculation results are based on the boundary conditions specified by Th = 873 K and Tc = 293 K.
Figure 5A shows a schematic of an eight-pair module. The dependence of open-circuit voltage (Voc), internal resistance (Rin), power density (Pd), and ηmax on the ratio of Ap/An and H/Apn, where H represents the height of the legs, is shown in Fig. 5 (B and C) and fig. S15. As H/Apn increases from 0.1 to 0.5 mm−1, Voc increases by ~25% due to the increased effective temperature difference ΔT (fig. S15A). Meanwhile, Rin becomes several times greater (fig. S15C) leading to a sharp decline of Pd. As H/Apn increases, ηmax increases first but becomes saturated when H/Apn exceeds 0.35 mm−1 because the interfacial contact resistivity with respect to that of the bulk is already low enough. At a given H/Apn, with Ap/An increasing, both Pd and ηmax increase first and then decrease. Our simulation suggests that ηmax and Pd reach the maximum values when Ap/An is about 2.6 and 1.7, respectively.
Fabrication and evaluation of GeTe/SKD module
On the basis of the simulation results, an eight-pair p-GeTe/n-SKD single-stage TE module was fabricated with Ap/An = 2.6 and H/Apn = 0.39 mm−1. A photo of the module is shown in the inset of Fig. 6B. The output performance of the module was tested by a home-built system with Tc maintained at 300 K and Th varied from 473 to 873 K.
Fig. 6. Characterization of the GeTe/skutterudite (SKD) single-stage module.
(A) Output voltage V and output power P, and (B) conversion efficiency η of the module as a function of the current I at different operating temperatures. Comparisons of the measured (C) open-circuit voltage Voc and internal resistance Rin, and (D) maximum conversion efficiency ηmax and maximum output power Pmax with corresponding simulated values. Note that the simulations in (C) and (D) considered the heat loss in the integration process (see Materials and Methods).
Figure 6A plots the dependence of output voltage V and output power P versus the measured current I. As ΔT increases from 173 to 545 K, Voc increases from 0.3 to 1.6 V, while the maximum value of P increases and reaches a peak value of 4.0 W at ΔT = 545 K, which corresponds to power density Pd = 1.0 W cm−2. Figure 6C shows the measured Voc and Rin of the module as a function of ΔT, which are compared with the simulated results. The measured and simulated Voc show excellent consistency in the whole temperature range, suggesting the uniformity of TE materials and a negligible temperature difference loss caused by the contacts. The measured Rin increases before the phase transition temperature of GeTe and at higher temperature reaches a plateau. This behavior is also reflected by the temperature-dependent electrical conductivity curves of GeTe and SKD (fig. S14A).
The measured and simulated Rin curves also show good agreement. In previous studies, the measured Rin of p-GeTe/n-SKD modules is substantially larger than the simulated ones, resulting in relatively deteriorated output performance (22, 55). This result indicates that the NiGe DBM layer and the module fabrication process as developed in this work sufficiently benefit the improvement of TE performance of the GeTe-based modules.
Figure 6B shows the measured energy conversion efficiency η as a function of the current I. A similar current dependence to that of P can be seen. At ΔT = 545 K, η reaches the maximum value of 12.0%. Figure 5D shows the measured Pmax and ηmax of the module as a function of ΔT compared with simulation results. The measured P reasonably agrees with the simulation. As the temperature increases, gradual deviation occurs due to the increased Rin as shown in Fig. 6C.
For comparison, Fig. 1B summarizes the reported ηmax versus ΔT for various TE modules. It can be seen that the present eight-pair p-GeTe/n-SKD single-stage module achieves ηmax of 12% at ΔT = 545 K, which is the highest among all reported single-stage modules (21–23, 29, 56–59) and even close to the values of segmented modules in the same temperature range (29, 36, 58). Note that the peak zT of our mass-produced p-GeTe and n-SKD materials is 1.7 and 1.2, respectively, which are still lower than the record values from lab-scale synthesis methods, suggesting that there is still space to further improve ηmax.
We have conducted a rational interface and structure design of GeTe-based TE modules toward scaling up this technology. A key contribution is a proposal of using transition metal germanides as the DBM. We characterized the interface between NiGe and GeTe and demonstrated superior chemical and mechanical stability in aging experiments. Meanwhile, we explored scale-up production of Ge0.89Cu0.06Sb0.08Te of more than 50 g per pellet and well-reproducible zTmax of 1.7. Combining the mass-produced p-type GeTe and n-type and Yb0.3Co4Sb12 SKD, we fabricated an eight-pair single-stage module and achieved the record-high efficiency of 12% at ΔT = 545 K, reaching 88% of the theoretical limit. The successful use of first-principles calculations for screening DBMs from binary compounds expanded the space of interface design for TE materials.
MATERIALS AND METHODS
Synthesis
High-quality polycrystalline samples of p-type Ge0.89Cu0.06Sb0.08Te were prepared by a melting-annealing method as described in our previous work (60). High-purity raw materials of Ge (pieces, 99.999%), Te (pieces, 99.999%), Cu (shots, 99.999%), and Sb (shots, 99.999%) with stoichiometric composition were used to synthesize GeTe pellets. After annealing, the pellets were hand-grounded into fine powders and well-mixed for high homogeneity.
For the synthesis of transition metal germanides, e.g., NiGe, FeGe2, CoGe, and CoGe2, the raw materials of Ni (slug, 99.99%), Fe (granule, 99.98%), Co (powder, 99.99%), and Ge (pieces, 99.999%) were weighed stoichiometrically and sealed in the evacuated silica tubes. The mixtures were heated up to 1423 K, held for 20 hours, cooled down to room temperature, and then hand-grounded into fine powders.
Characterizations
To characterize the interfacial resistivity between the GeTe layer and DBM layers, sandwich-structured GeTe/DBM/GeTe samples were prepared by spark plasma sintering (SPS; Dr. Sinter: SPS-2040) under a uniaxial pressure of 50 MPa for 10 min at 813 K. Bars of 3 mm by 3 mm by 6 mm were cut from the sintered pellets for long-term aging at 773 K in a vacuum. The measurements of interfacial resistivity were conducted on a home-built four-probe platform (36).
The phase compositions and crystal structures were characterized by powder XRD (D8 Advance, Bruker; Cu Kα: λ = 1.5406 Å) at ambient conditions. The surface morphology and microstructures of the interfaces of the sandwich-structured samples were examined by field emission SEM (ZEISS Supra 55) equipped with EDS. Atomic structures of the GeTe/NiGe interface were investigated by scanning transmission electron microscopy (Hitachi HF5000). The detailed characterization methods for TE properties can be found in our previous work (60).
First-principles calculations
The first-principles calculations were carried out by the Vienna Ab initio Simulation Package program (61). The Perdew-Burke-Ernzerhof functional was used to optimize the crystal structures and calculate the total energies for all the considered elemental substances and compounds (62). The core-valence interaction was described by the projector-augmented wave potentials with a cutoff energy of 400 eV for the plane-wave basis set (63). The k-point meshes satisfy that the products of the lengths of lattice vectors and the numbers of k-points along all directions are about 40 Å. The atoms in the supercell were relaxed until the residual forces on all atoms were smaller than 0.01 eV. The formation energy of the compounds was calculated with respect to their elemental substances. To model the GeTe(001)/NiGe(011) interface, a supercell containing five GeTe(001) and six NiGe(011) atomic layers was constructed which naturally includes two equivalent interfaces. The lattice vector parallel to the interface was fixed as the commensurate value of those of GeTe(001) and NiGe(011) surfaces, while the vector vertical to the interface was allowed to be optimized and the atoms of the supercell were fully relaxed. The work of separation (Wsep) for the interface was then calculated on the basis of , where Etotal is the total energy of the interfacial supercell, and are the energy of the surface slabs of GeTe(001) and NiGe(011), respectively, with the same strain imposed as in the interfacial supercell, and A is the cross-section area of the interface. The Wsep of the GeTe(001), GeTe(111), and NiGe(001) surfaces is essentially equivalent to their surface energy calculated as , where is the energy of the surface slab with all of the lattice vectors and atoms fully relaxed, is the total energy per atom of the bulk material, and N is the number of atoms in the slab.
Finite element method simulations
The FEM simulations of module efficiency were based on the ANSYS-Workbench program. In this study, GeTe and SKD were used as p-type and n-type legs, respectively. The temperature-dependent S, σ, and κ of GeTe and SKD were collected from the experiment as the input to the program. To make the simulations reflect the real status and give effective guidance, the measured σ and κ of the barrier layer were also considered in our simulation. In addition, the interfacial resistivities for the electrode interface in p- and n-legs were also input to our simulation. The interfacial heat transfer coefficients of hot and cold sources, as well as the thermal conductivity of the gap fillers, were obtained from our previous work (36). It is noted that the purpose of Fig. 1 is to compare the intrinsic properties of the materials by ignoring the effects due to device integration; the calculations of theoretical ηmax were only based on the TE properties of various TE material systems without consideration of heat loss caused by the different integration process and different measurement systems.
Module fabrication and measurements
The p-type GeTe legs with NiGe on both sides were fabricated by a one-step sintering process. The NiGe powders and Ge0.89Cu0.06Sb0.08Te powders were sequentially loaded into a graphite die of Φ30 mm and hot-pressed under 50 MPa for 45 min at 873 K. The thickness of each NiGe layer is ~100 μm. The obtained pellets of Φ 30 mm by 12.5 mm were cut into bars with 5.8 × 4 × 12.5 mm3 for TE module integration. The n-type SKD legs were prepared following our previous work (36). The dimensions of the n-type leg were 2.2 × 4 × 12.5 mm3.
The p-GeTe and n-SKD single-stage module was assembled by a soldering process in a vacuum. The p- and n-legs are bridged with Mo-Cu as the hot-side electrode and the direct-bonded copper alumina plate as the cold-side electrode, respectively. The solders used for the hot side and the cold side are Ag-Cu-Zn (soldering temperature, ~850 K) and Sn-based solders (soldering temperature, ~450 K), respectively. The glass fibers were used to fill the gaps between the p- and n-legs to diminish the heat loss caused by convection and radiation. The framework dimensions of the module made of eight pairs are 20 mm by 20 mm by 14.5 mm. The TE conversion efficiency of this module was measured by a home-built testing system when the hot side was heated from 298 to 873 K and the cold side was maintained at room temperature. The conversion efficiency was obtained by
where P is the output power of the module and Qc is the heat flow measured by a flux probe at the cold end. The power density (Pd) was calculated by Pd = P/A, where A is the envelope area of the module.
Acknowledgments
Funding: The work was financially supported by the National Key Research and Development Program of China (grant no. 2019YFE0103500), the National Natural Science Foundation of China (NSFC) (grant nos. U2141208 and 52102330). S.B. acknowledges the support from International Partnership Program of Chinese Academy of Sciences (grant no. 121631KYSB20200012).
Author contributions: S.B. and L.C. designed this work and supervised the experiment. L.X. prepared the GeTe materials and measured the thermoelectric properties. C.M. and Y.-Y.S. performed the calculations. C.Z. and F.X. performed TEM characterization. L.X. and L.W. conducted the numerical modeling. L.X., C.W., and J.L. assembled the modules and measured the performance. L.X., Q.S., C.M., Y.-Y.S., S.B., and L.C. wrote the manuscript. All authors discussed the results and contributed to the data analyses.
Competing interests: The authors declare that they have no competing interests.
Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.
Supplementary Materials
This PDF file includes:
Figs. S1 to S15
Tables S1 to S3
Evaluation of interfacial reaction energy
REFERENCES AND NOTES
- 1.L. Miro, S. Bruckner, L. E. Cabeza, Mapping and discussing industrial waste heat (IWH) potentials for different countries. Renew. Sustain. Energy Rev. 51, 847–855 (2015). [Google Scholar]
- 2.L. E. Bell, Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 321, 1457–1461 (2008). [DOI] [PubMed] [Google Scholar]
- 3.X. Shi, L. D. Chen, Thermoelectric materials step up. Nat. Mater. 15, 691–692 (2016). [DOI] [PubMed] [Google Scholar]
- 4.C. Wood, Materials for thermoelectric energy-conversion. Rep. Prog. Phys. 51, 459–539 (1988). [Google Scholar]
- 5.M. Papapetrou, G. Kosmadakis, Chapter 9 - Resource, environmental, and economic aspects of SGHE, in Salinity Gradient Heat Engines, A. Tamburini, A. Cipollina, G. Micale, Eds. (Woodhead Publishing, 2022), pp. 319–353. [Google Scholar]
- 6.X. Shi, J. Yang, J. R. Salvador, M. Chi, J. Y. Cho, H. Wang, S. Bai, J. Yang, W. Zhang, L. Chen, Multiple-filled skutterudites: High thermoelectric figure of merit through separately optimizing electrical and thermal transports. J. Am. Chem. Soc. 133, 7837–7846 (2011). [DOI] [PubMed] [Google Scholar]
- 7.Z. Zheng, X. Su, R. Deng, C. Stoumpos, H. Xie, W. Liu, Y. Yan, S. Hao, C. Uher, C. Wolverton, M. G. Kanatzidis, X. Tang, Rhombohedral to cubic conversion of GeTe via MnTe alloying leads to ultralow thermal conductivity, electronic band convergence, and high thermoelectric performance. J. Am. Chem. Soc. 140, 2673–2686 (2018). [DOI] [PubMed] [Google Scholar]
- 8.J. Li, X. Zhang, Z. Chen, S. Lin, W. Li, J. Shen, I. T. Witting, A. Faghaninia, Y. Chen, A. Jain, L. Chen, G. J. Snyder, Y. Pei, Low-symmetry rhombohedral GeTe thermoelectrics. Joule 2, 976–987 (2018). [Google Scholar]
- 9.J. Li, X. Y. Zhang, X. Wang, Z. L. Bu, L. T. Zheng, B. Q. Zhou, F. Long, Y. Chen, Y. Z. Pei, High-performance GeTe thermoelectrics in both rhombohedral and cubic phases. J. Am. Chem. Soc. 140, 16190–16197 (2018). [DOI] [PubMed] [Google Scholar]
- 10.S. Perumal, S. Roychowdhury, D. S. Negi, R. Datta, K. Biswas, High thermoelectric performance and enhanced mechanical stability of p-type Ge1–xSbxTe. Chem. Mat. 27, 7171–7178 (2015). [Google Scholar]
- 11.J. Li, Z. Chen, X. Zhang, H. Yu, Z. Wu, H. Xie, Y. Chen, Y. Pei, Simultaneous optimization of carrier concentration and alloy scattering for ultrahigh performance GeTe thermoelectrics. Adv. Sci. 4, 1700341 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Z. Liu, J. Sun, J. Mao, H. Zhu, W. Ren, J. Zhou, Z. Wang, D. J. Singh, J. Sui, C. W. Chu, Z. Ren, Phase-transition temperature suppression to achieve cubic GeTe and high thermoelectric performance by Bi and Mn codoping. Proc. Natl. Acad. Sci. U.S.A. 115, 5332–5337 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.M. Hong, Y. Wang, T. Feng, Q. Sun, S. Xu, S. Matsumura, S. T. Pantelides, J. Zou, Z.-G. Chen, Strong phonon–Phonon interactions securing extraordinary thermoelectric Ge1–xSbxTe with Zn-alloying-induced band alignment. J. Am. Chem. Soc. 141, 1742–1748 (2019). [DOI] [PubMed] [Google Scholar]
- 14.M. Hong, Z. G. Chen, L. Yang, Y. C. Zou, M. S. Dargusch, H. Wang, J. Zou, Realizing zT of 2.3 in Ge1-x-ySbxInyTe via reducing the phase-transition temperature and introducing resonant energy doping. Adv. Mater. 30, 1705942 (2018). [DOI] [PubMed] [Google Scholar]
- 15.M. Li, M. Hong, X. Tang, Q. Sun, W.-Y. Lyu, S.-D. Xu, L.-Z. Kou, M. Dargusch, J. Zou, Z.-G. Chen, Crystal symmetry induced structure and bonding manipulation boosting thermoelectric performance of GeTe. Nano Energy 73, 104740 (2020). [Google Scholar]
- 16.M. Hong, Y. Wang, W. Liu, S. Matsumura, H. Wang, J. Zou, Z.-G. Chen, Arrays of planar vacancies in superior thermoelectric Ge1−x−yCdxBiyTe with band convergence. Adv. Energy Mater. 8, 1801837 (2018). [Google Scholar]
- 17.L. Xie, R. Liu, C. Zhu, Z. Bu, W. Qiu, J. Liu, F. Xu, Y. Pei, S. Bai, L. Chen, Enhanced thermoelectric performance in Ge0.955−xSbxTe/FeGe2 composites enabled by hierarchical defects. Small 17, 2100915 (2021). [DOI] [PubMed] [Google Scholar]
- 18.J. Shuai, Y. Sun, X. Tan, T. Mori, Manipulating the Ge vacancies and Ge precipitates through Cr doping for realizing the high-performance GeTe thermoelectric material. Small 16, 1906921 (2020). [DOI] [PubMed] [Google Scholar]
- 19.S. Perumal, M. Samanta, T. Ghosh, U. S. Shenoy, A. K. Bohra, S. Bhattacharya, A. Singh, U. V. Waghmare, K. Biswas, Realization of high thermoelectric figure of merit in GeTe by complementary co-doping of Bi and In. Joule 3, 2565–2580 (2019). [Google Scholar]
- 20.Z. L. Bu, X. Y. Zhang, B. Shan, J. Tang, H. X. Liu, Z. W. Chen, S. Q. Lin, W. Li, Y. Z. Pei, Realizing a 14% single-leg thermoelectric efficiency in GeTe alloys. Sci. Adv. 7, eabf2738 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Z. Bu, X. Zhang, Y. Hu, Z. Chen, S. Lin, W. Li, Y. Pei, An over 10% module efficiency obtained using non-Bi2Te3 thermoelectric materials for recovering heat of <600 K. Energ. Environ. Sci. 14, 6506–6513 (2021). [Google Scholar]
- 22.T. Xing, Q. Song, P. Qiu, Q. Zhang, M. Gu, X. Xia, J. Liao, X. Shi, L. Chen, High efficiency GeTe-based materials and modules for thermoelectric power generation. Energ. Environ. Sci. 14, 995–1003 (2021). [Google Scholar]
- 23.B. Jiang, W. Wang, S. Liu, Y. Wang, C. Wang, Y. Chen, L. Xie, M. Huang, J. He, High figure-of-merit and power generation in high-entropy GeTe-based thermoelectrics. Science 377, 208–213 (2022). [DOI] [PubMed] [Google Scholar]
- 24.J. Li, S. Zhao, J. Chen, G. Bai, L. Hu, F. Liu, W. Ao, Y. Li, H. Xie, C. Zhang, Enhanced interfacial reliability and mechanical strength of CoSb3-based thermoelectric joints with rationally designed diffusion barrier materials of Ti-based alloys. ACS Appl. Mater. Inter. 12, 44858–44865 (2020). [DOI] [PubMed] [Google Scholar]
- 25.H.-C. Hsieh, C.-H. Wang, T.-W. Lan, T.-H. Lee, Y.-Y. Chen, H.-S. Chu, A. T. Wu, Joint properties enhancement for PbTe thermoelectric materials by addition of diffusion barrier. Mater. Chem. Phys. 246, 122848 (2020). [Google Scholar]
- 26.T. Xing, Q. F. Song, P. F. Qiu, Q. H. Zhang, X. G. Xia, J. C. Liao, R. H. Liu, H. Huang, J. Yang, S. Q. Bai, D. D. Ren, X. Shi, L. D. Chen, Superior performance and high service stability for GeTe-based thermoelectric compounds. Natl. Sci. Rev. 6, 944–954 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.J. Q. Li, S. Y. Zhao, J. L. Chen, C. P. Han, L. P. Hu, F. S. Liu, W. Q. Ao, Y. Li, H. P. Xie, C. H. Zhang, Al-Si alloy as a diffusion barrier for GeTe-based thermoelectric legs with high interfacial reliability and mechanical strength. ACS Appl. Mater. Inter. 12, 18562–18569 (2020). [DOI] [PubMed] [Google Scholar]
- 28.C. G. Fu, S. Q. Bai, Y. T. Liu, Y. S. Tang, L. D. Chen, X. B. Zhao, T. J. Zhu, Realizing high figure of merit in heavy-band p-type half-Heusler thermoelectric materials. Nat. Commun. 6, 8144 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Y. F. Xing, R. H. Liu, J. C. Liao, Q. H. Zhang, X. G. Xia, C. Wang, H. Huang, J. Chu, M. Gu, T. J. Zhu, C. X. Zhu, F. F. Xu, D. X. Yao, Y. P. Zeng, S. Q. Bai, C. Uher, L. D. Chen, High-efficiency half-heusler thermoelectric modules enabled by self-propagating synthesis and topologic structure optimization. Energ. Environ. Sci. 12, 3390–3399 (2019). [Google Scholar]
- 30.J. Yu, Y. Xing, C. Hu, Z. Huang, Q. Qiu, C. Wang, K. Xia, Z. Wang, S. Bai, X. Zhao, L. Chen, T. Zhu, Half-heusler thermoelectric module with high conversion efficiency and high power density. Adv. Energy Mater. 10, 2000888 (2020). [Google Scholar]
- 31.Y. Xing, R. Liu, J. Liao, C. Wang, Q. Zhang, Q. Song, X. Xia, T. Zhu, S. Bai, L. Chen, A device-to-material strategy guiding the “Double-high” thermoelectric module. Joule 4, 2475–2483 (2020). [Google Scholar]
- 32.G. Nie, S. Suzuki, T. Tomida, A. Sumiyoshi, T. Ochi, K. Mukaiyama, M. Kikuchi, J. Q. Guo, A. Yamamoto, H. Obara, Performance of skutterudite-based modules. J. Electron. Mater. 46, 2640–2644 (2017). [Google Scholar]
- 33.J. Chu, J. Huang, R. H. Liu, J. C. Liao, X. G. Xia, Q. H. Zhang, C. Wang, M. Gu, S. Q. Bai, X. Shi, L. D. Chen, Electrode interface optimization advances conversion efficiency and stability of thermoelectric devices. Nat. Commun. 11, 8 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.F. Hao, P. F. Qiu, Y. S. Tang, S. Q. Bai, T. Xing, H. S. Chu, Q. H. Zhang, P. Lu, T. S. Zhang, D. D. Ren, J. K. Chen, X. Shi, L. D. Chen, High efficiency Bi2Te3-based materials and devices for thermoelectric power generation between 100 and 300 °C. Energ. Environ. Sci. 9, 3120–3127 (2016). [Google Scholar]
- 35.X. F. Lu, Q. H. Zhang, J. C. Liao, H. Y. Chen, Y. C. Fan, J. J. Xing, S. J. Gu, J. L. Huang, J. X. Ma, J. C. Wang, L. J. Wang, W. Jiang, High-efficiency thermoelectric power generation enabled by homogeneous incorporation of MXene in (Bi,Sb)2Te3 Matrix. Adv. Energy Mater. 10, 1902986 (2020). [Google Scholar]
- 36.Q. Zhang, J. Liao, Y. Tang, M. Gu, C. Ming, P. Qiu, S. Bai, X. Shi, C. Uher, L. Chen, Realizing a thermoelectric conversion efficiency of 12% in bismuth telluride/skutterudite segmented modules through full-parameter optimization and energy-loss minimized integration. Energ. Environ. Sci. 10, 956–963 (2017). [Google Scholar]
- 37.M. Gu, S. Bai, J. Wu, J. Liao, X. Xia, R. Liu, L. Chen, A high-throughput strategy to screen interfacial diffusion barrier materials for thermoelectric modules. J. Mater. Res. 34, 1179–1187 (2019). [Google Scholar]
- 38.X. Wu, Z. Han, Y. Zhu, B. Deng, K. Zhu, C. Liu, F. Jiang, W. Liu, A general design strategy for thermoelectric interface materials in n-type Mg3Sb1.5Bi0.5 single leg used in TEGs. Acta Mater. 226, 117616 (2022). [Google Scholar]
- 39.Z. Bu, W. Li, J. Li, X. Zhang, J. Mao, Y. Chen, Y. Pei, Dilute Cu2Te-alloying enables extraordinary performance of r-GeTe thermoelectrics. Mater. Today Phys. 9, 100096 (2019). [Google Scholar]
- 40.B. Srinivasan, R. Gautier, F. Gucci, B. Fontaine, J.-F. Halet, F. Cheviré, C. Boussard-Pledel, M. J. Reece, B. Bureau, Impact of coinage metal insertion on the thermoelectric properties of GeTe solid-state solutions. J. Phys. Chem. C 122, 227–235 (2018). [Google Scholar]
- 41.A. Jain, S. P. Ong, G. Hautier, W. Chen, W. D. Richards, S. Dacek, S. Cholia, D. Gunter, D. Skinner, G. Ceder, K. A. Persson, Commentary: The Materials Project: A materials genome approach to accelerating materials innovation. APL Mater. 1, 011002 (2013). [Google Scholar]
- 42.G. Bergerhoff, R. Hundt, R. Sievers, I. D. Brown, The inorganic crystal structure data base. J. Chem. Inf. Comput. Sci. 23, 66–69 (1983). [Google Scholar]
- 43.S. Chen, H. Liu, F. Chen, K. Zhou, Y. Xue, Synthesis, transfer, and properties of layered FeTe2 nanocrystals. ACS Nano 14, 11473–11481 (2020). [DOI] [PubMed] [Google Scholar]
- 44.H. Ma, W. Dang, X. Yang, B. Li, Z. Zhang, P. Chen, Y. Liu, Z. Wan, Q. Qian, J. Luo, K. Zang, X. Duan, X. Duan, Chemical vapor deposition growth of single crystalline CoTe2 nanosheets with tunable thickness and electronic properties. Chem. Mater. 30, 8891–8896 (2018). [Google Scholar]
- 45.B. Zhao, W. Dang, Y. Liu, B. Li, J. Li, J. Luo, Z. Zhang, R. Wu, H. Ma, G. Sun, Y. Huang, X. Duan, X. Duan, Synthetic control of two-dimensional NiTe2 single crystals with highly uniform thickness distributions. J. Am. Chem. Soc. 140, 14217–14223 (2018). [DOI] [PubMed] [Google Scholar]
- 46.A. Nash, P. Nash, The Ge−Ni (germanium-nickel) system. Bull. Alloy Phase Diagrams. 8, 255–264 (1987). [Google Scholar]
- 47.H. Okamoto, Fe-Ge (iron-germanium). J. Phase Equilib. Diffus. 29, 292–292 (2008). [Google Scholar]
- 48.R. Liu, Y. Xing, J. Liao, X. Xia, C. Wang, C. Zhu, F. Xu, Z.-G. Chen, L. Chen, J. Huang, S. Bai, Thermal-inert and ohmic-contact interface for high performance half-Heusler based thermoelectric generator. Nat. Commun. 13, 7738 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.J. Feng, W. Zhang, W. Jiang, Ab initio study of Ag∕Al2O3 and Au∕Al2O3 interfaces. Phys. Rev. B 72, 115423 (2005). [Google Scholar]
- 50.D. Y. Dang, L. Y. Shi, J. L. Fan, H. R. Gong, First-principles study of W–TiC interface cohesion. Surf. Coat. Tech. 276, 602–605 (2015). [Google Scholar]
- 51.F. Chen, R. Liu, Z. Yao, Y. Xing, S. Bai, L. Chen, Scanning laser melting for rapid and massive fabrication of filled skutterudites with high thermoelectric performance. J. Mater. Chem. A 6, 6772–6779 (2018). [Google Scholar]
- 52.M. Samanta, T. Ghosh, R. Arora, U. V. Waghmare, K. Biswas, Realization of both n- and p-type GeTe thermoelectrics: Electronic structure modulation by AgBiSe2 alloying. J. Am. Chem. Soc. 141, 19505–19512 (2019). [DOI] [PubMed] [Google Scholar]
- 53.Z. Liu, N. Sato, Q. Guo, W. Gao, T. Mori, Shaping the role of germanium vacancies in germanium telluride: Metastable cubic structure stabilization, band structure modification, and stable N-type conduction. NPG Asia Mater. 12, 66 (2020). [Google Scholar]
- 54.M. Gu, X. Xia, X. Li, X. Huang, L. Chen, Microstructural evolution of the interfacial layer in the Ti–Al/Yb0.6Co4Sb12 thermoelectric joints at high temperature. J. Alloy. Compd. 610, 665–670 (2014). [Google Scholar]
- 55.G. Bai, Y. Yu, X. Wu, J. Li, Y. Xie, L. Hu, F. Liu, M. Wuttig, O. Cojocaru-Mirédin, C. Zhang, Boron strengthened GeTe-based alloys for robust thermoelectric devices with high output power density. Adv. Energy Mater. 11, 2102012 (2021). [Google Scholar]
- 56.P. Qiu, T. Mao, Z. Huang, X. Xia, J. Liao, M. T. Agne, M. Gu, Q. Zhang, D. Ren, S. Bai, X. Shi, G. J. Snyder, L. Chen, High-efficiency and stable thermoelectric module based on liquid-like materials. Joule 3, 1538–1548 (2019). [Google Scholar]
- 57.Z. H. Liu, N. Sato, W. H. Gao, K. Yubuta, N. Kawamoto, M. Mitome, K. Kurashima, Y. Owada, K. Nagase, C. H. Lee, J. Yi, K. Tsuchiya, T. Mori, Demonstration of ultrahigh thermoelectric efficiency of ∼7.3% in Mg3Sb2/MgAgSb module for low-temperature energy harvesting. Joule 5, 1196–1208 (2021). [Google Scholar]
- 58.X. K. Hu, P. Jood, M. Ohta, M. Kunii, K. Nagase, H. Nishiate, M. G. Kanatzidis, A. Yamamoto, Power generation from nanostructured PbTe-based thermoelectrics: Comprehensive development from materials to modules. Energ. Environ. Sci. 9, 517–529 (2016). [Google Scholar]
- 59.Q. H. Zhang, Z. X. Zhou, M. Dylla, M. T. Agne, Y. Z. Pei, L. J. Wang, Y. S. Tang, J. C. Liao, J. Li, S. Q. Bai, W. Jiang, L. D. Chen, G. J. Snyder, Realizing high-performance thermoelectric power generation through grain boundary engineering of skutterudite-based nanocomposites. Nano Energy 41, 501–510 (2017). [Google Scholar]
- 60.L. Xie, Y. J. Chen, R. H. Liu, E. H. Song, T. Xing, T. T. Deng, Q. F. Song, J. J. Liu, R. K. Zheng, X. Gao, S. Q. Bai, L. D. Chen, Stacking faults modulation for scattering optimization in GeTe-based thermoelectric materials. Nano Energy 68, 104347 (2020). [Google Scholar]
- 61.G. Kresse, J. Furthmüller, Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comp. Mater. Sci. 6, 15–50 (1996). [DOI] [PubMed] [Google Scholar]
- 62.J. P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865–3868 (1996). [DOI] [PubMed] [Google Scholar]
- 63.G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 59, 1758–1775 (1999). [Google Scholar]
- 64.P. J. Ying, R. He, J. Mao, Q. H. Zhang, H. Reith, J. H. Sui, Z. F. Ren, K. Nielsch, G. Schierning, Towards tellurium-free thermoelectric modules for power generation from low-grade heat. Nat. Commun. 12, 1121 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Y. T. Fu, Q. H. Zhang, Z. L. Hu, M. Jiang, A. B. Huang, X. Ai, S. Wan, H. Reith, L. J. Wang, K. Nielsch, W. Jiang, Mg3(Bi,Sb)2-based thermoelectric modules for efficient and reliable waste-heat utilization up to 750 K. Energ. Environ. Sci. 15, 3265–3274 (2022). [Google Scholar]
- 66.B. C. Qin, D. Y. Wang, X. X. Liu, Y. X. Qin, J. F. Dong, J. F. Luo, J. W. Li, W. Liu, G. J. Tan, X. F. Tang, J. F. Li, J. Q. He, L. D. Zhao, Power generation and thermoelectric cooling enabled by momentum and energy multiband alignments. Science 373, 556–561 (2021). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Figs. S1 to S15
Tables S1 to S3
Evaluation of interfacial reaction energy






