Abstract
Thermoelectric technology offers the opportunity of direct conversion between heat and electricity, and new and exciting materials that can enable this technology to deliver higher efficiencies have been developed in recent years. This mini-review covers the most promising advances in thermoelectric materials as they pertain to their potential in being implemented in devices and modules with an emphasis on thermoelectric power generation. Classified into three groups in terms of their operating temperature, the thermoelectric materials that are most likely to be used in future devices are briefly discussed. We summarize the state-of-the-art thermoelectric modules/devices, among which nanostructured PbTe modules are particularly highlighted. At the end, key issues and the possible strategies that can help thermoelectric power generation technology move forward are considered.
This article is part of a discussion meeting issue ‘Energy materials for a low carbon future’.
Keywords: thermoelectrics, power generation, materials, modules, devices
1. Introduction
Thermoelectric materials have drawn tremendous attention in the past two decades because they can enable devices that can harvest waste heat and convert it to electrical power thereby promising to improve the efficiency of fuel utilization [1]. The efficiency of a thermoelectric material is defined by the dimensionless figure of merit ZT = S2σT/κ, where S is the Seebeck coefficient, σ is the electrical conductivity, T is the operation temperature and κ is thermal conductivity that is composed of carrier contribution (κcar) and lattice contribution (κlat). The efficiency of a thermoelectric device (η), however, depends not on ZT maximum at a given temperature but on the average ZT (ZTave) value over a wide temperature range as η = [(TH − TC)/TH] [(1 + ZTave)1/2 − 1]/[(1 + ZTave)1/2 + TC/TH], where TH and TC are the hot side and cold side temperature, respectively. Thus, for practical device applications, the ZTave is the critical quantity that must be maximized over the entire operation temperature range, rather than maximum ZT (ZTmax). In addition to requiring high ZTs and low cost, the materials must have enough mechanical strength for robust device fabrication and high thermal stability for long-term service [2].
There have been several important breakthroughs in increasing the ZT of thermoelectric materials since 2000, as summarized in table 1 [3–26]. The ZT enhancement originates either from increase of power factors (PF = S2σ) [17,27,28] or decrease of κlat through microstructure engineering [6,9,29–31]. Several materials have now been reported to display high ZT values over 2 [9,19,25,26,30,32–37]. However, only a few of them have been demonstrated in thermoelectric devices or modules [38].
Table 1.
Record ZTs of some typical thermoelectric materials since the year 2000.
| material | ZTmax at T | ZTave (TC − TH) | p/n | reference |
|---|---|---|---|---|
| Bi83.5Sb16.5 | 0.52 (80 K) | 0.47 (50–100 K) | n | [3] |
| CsBi4Te6 | 0.8 (225 K) | 0.44 (100–250 K) | p | [4] |
| Bi2Te2.79Se0.21 | 1.2 (357 K) | 1.02 (300–450 K) | n | [5] |
| Bi0.5Sb1.5Te3 + x%Te | 1.86 (320 K) | 1.42 (300–450 K) | p | [6] |
| MgAg0.97Sb0.99 | 1.3 (533 K) | 1.1 (300–550 K) | p | [7] |
| PbTe0.998I0.002-3%Sb | 1.8 (773 K) | 0.97 (300–773 K) | n | [8] |
| Pb0.98Na0.02Te-8%SrTe | 2.5 (923 K) | 1.67 (300–900 K) | p | [9] |
| PbSe0.998Br0.002-2%Cu2Se | 1.8 (723 K) | 1.1 (300–823 K) | n | [10] |
| Pb0.98Na0.02Se-2%HgSe | 1.7 (973 K) | 0.63 (400–900 K) | p | [11] |
| PbS-4.4%Ag | 1.7 (850 K) | 0.93 (300–850 K) | n | [12] |
| Pb0.975Na0.025S-3%CdS | 1.3 (923 K) | 0.55 (300–923 K) | p | [13] |
| β-Zn4Sb3 | 1.4 (750 K) | 0.81 (300–750 K) | p | [14] |
| Ba0.08La0.05Yb0.04Co4Sb12 | 1.7 (850 K) | 1.06 (300–850 K) | n | [15] |
| DD0.7Fe2.7Co1.3Sb11.8Sn0.2 | 1.45 (823 K) | 0.82 (423–823 K) | p | [16] |
| Mg2.15Sb0.006Si0.28Sn0.71 | 1.3 (700 K) | 0.93 (300–773 K) | n | [17] |
| Mg2Li0.025Si0.4Sn0.6 | 0.7 (650 K) | 0.41 (300–750 K) | p | [18] |
| SnSe1−xBrx | 2.8 (773 K) | 1.29 (300–773 K) | n | [19] |
| SnSe | 2.6 (923 K) | 1.37 (600–923 K) | p | [20] |
| (Si95Ge5)0.65(Si70Ge30P3)35 | 1.3 (1173 K) | 0.7 (300–1173 K) | n | [21] |
| Si80Ge20Bx | 0.6 (1100 K) | 0.55 (373–1300 K) | p | [22] |
| Zr0.2Hf0.8NiSn0.985Sb0.015 | 1.1 (1000 K) | 0.66 (300–1100 K) | n | [23] |
| Nb0.88Hf0.12FeSb | 1.5 (1200 K) | 1.0 (500–1200 K) | p | [24] |
| Cu2Se-x%CNT | 2.4 (1000 K) | 1.5 (450–1000 K) | p | [25] |
| Cu1.97S | 1.7 (1000 K) | 0.77 (300–1000 K) | p | [26] |
This article focuses on the potential of the new thermoelectric materials with extraordinary performance in being implemented in thermoelectric modules specifically for power generation (organic materials and their devices fall out of scope because efficiencies are still too low [39]). Unless the promise to build robust thermoelectric modules is ultimately realized, thermoelectric materials will remain in the basic research domain of this important energy conversion technology. The article is organized as follows: first, we briefly discuss the most important recent developments in high-performance thermoelectric materials, particularly those with good stability, low cost and scalable potential; second, we discuss the novel device fabrication technologies made of these materials; finally, we consider future challenges and possible strategies for further enhancing ZTs and device performance. We note that this article is not a comprehensive review of the literature but is a selected snapshot of the module-related field reflecting the authors' individual interests in developing new materials for modules.
2. Thermoelectric materials
(a). Low-temperature materials (300–500 K)
Since its discovery in the 1950s, Bi2Te3 and its alloys with ZT values of approximately 0.9–1.0 around room temperature have been widely commercialized but mainly for cooling applications [40]. Thermoelectric modules made of conventional Bi2Te3-based materials have been commercialized for decades. Recent researches mainly focus on their soldering and welding technologies [41,42] as well as fabrication of flexible Bi2Te3 thermoelectric devices on organic substrates [43]. However, because these modules are based on conventional material we are not going to give a detailed survey of Bi2Te3 thermoelectric devices in this article.
The concept of nanosizing was applied in the 2000s to this system and marginally boosted ZTs to just over 1 [29,44–49]. The pioneering work was done by the Chen and Tang groups in late 2000s, where ball milling [29] and melt spinning techniques [47] were used to create nanostructuring inside the bulk materials. These nanosized grains were found to be effective in scattering heat-carrying acoustic phonons and reducing thermal conductivity. Hot deformation [5,49] developed by the Zhao group was also shown to lower thermal conductivity of bismuth telluride-based materials. In addition to reducing κlat, researchers also tried to enhance ZT values of Bi2Te3-based alloys by increasing S2σ through band structure engineering. For instance, Sn was found to form resonant impurity states in the valence band of Bi2Te3, thereby largely enhancing S of the host material at cryogenic temperatures through resonant scattering [50,51]. The well-accepted best ZT values reported for p- and n-type Bi2Te3-based thermoelectric alloys are approximately 1.5 [29,47] and 1.2 [5] (both at approx. 400 K), respectively. It is worth mentioning that Kim et al. [6] in 2015 reported a peak ZT of approximately 1.86 at 320 K in p-type Bi2Te3, which was a purported record high. However, a very recent study clearly suggests that such an abnormal ‘too good to be true’ ZT might originate from an incorrect combination of values: (i) thermal conductivity measured along the pressing direction of anisotropy, (ii) charge transport measured in the direction perpendicular to the anisotropic direction [52].
Although improved thermoelectric performance was reported in Bi2Te3 alloys in a laboratory scale [5,6,44,45,47,49], the ability to scale up of these materials, a key factor for industry production, remains an open question. Recently, Zheng et al. prepared large p-type Bi2Te3-based pellets of 30, 40 and 60 mm in diameter by melt spinning followed by plasma-activated sintering techniques (PAS), and examined their uniformity, thermoelectric properties as well as mechanical strength [53]. It was shown that, influenced by an ellipsoidal-shape-distributed temperature field during the sintering process, the as-prepared large pellet is inhomogeneous because of compositional fluctuation, figure 1a. However, by appropriately elongating the sintering time (figure 1b–d) or designing graphite dies (figure 1e–g), the temperature distribution homogeneity is largely improved during the sintering process, and the pressed pellets' uniformity is therefore greatly enhanced. The large Bi2Te3-based pellets show a peak ZT of 1.15 at 373 K, which is approximately 20% improvement over the commercial zone-melted Bi2Te3 ingots (figure 1h). Moreover, the compressive and bending strengths of the pressed pellets are significantly improved compared to zone-melted ingots, figure 1i. This work highlighted the possibility of fabricating large mechanically robust new generation of Bi2Te3-based materials with high thermoelectric performance at temperatures less than 500 K.
Figure 1.
Sample homogeneity demonstration in Bi2Te3 bulk thermoelectric materials. The distribution of Seebeck coefficients over the cross section of pressed pellets with the diameter of 30 mm and height of 12 mm sintered for the time of (a) 5 min, (b) 10 min, (c) 20 min, (d) 30 min and for the graphite dies with an outer diameter of (e) 60 mm, (f) 90 mm, (g) 120 mm. (h) A comparison of ZT values measured on different samples. Square symbols are for samples prepared by melt spinning and spark plasma sintering (MS-SPS); [53] solid curves are for samples prepared by conventional melting followed by spark plasma sintering (M-SPS); the dashed line is for zone-melted (ZM) sample. and represent the directions perpendicular and parallel to the sintering pressure, respectively. (i) The compressive and bending strength of Φ30 mm diameter MS-SPS samples with sintering duration of 5 min and 20 min, respectively. The data for the ZM sample are also included for comparison. Adapted from Zheng [53]. (Online version in colour.)
α-MgAgSb is an interesting new entry in the field of thermoelectrics with a narrow band gap (approx. 0.1 eV at 300 K), which may be a candidate for making device modules despite some undesirable phase transitions above room temperature [54,55]. It has been reported that ball milling combined with hot pressing technology can be employed to synthesize this material and [56] allows the precise control and fine-tuning of chemical stoichiometry. MgAgSb features complex crystal structures [54] and low sound velocities (1276 m s−1) [57], which are supposed to contribute to its intrinsically low thermal conductivity (less than 0.7 Wm−1 K−1 at T > 300 K) [56]. Moreover, it has a high valence band degeneracy of 8 and a large density-of-states effective masses (m* = 0.6 me, me is the inertial mass of free electrons), [58] enabling high power factors over 20 µW cm−1 K−2 above 300 K [56]. Altogether, a peak ZT of approximately 1.2–1.4 at 550 K was reported for this material with p-type conduction [56,59,60]. More importantly, a realistic conversion efficiency of approximately 8.5% for a unileg under a temperature difference of 225 K has been experimentally demonstrated [61]. Even so, the main problem with the MgAgSb-based thermoelectric materials is the presence of silver metal which may make it cost prohibitive to develop an affordable technology.
(b). Mid-temperature materials (500–900 K)
In terms of ZT lead chalcogenides are the highest performing thermoelectric materials operating in the mid-temperature range and have a long history of powering several spacecraft launched by NASA [8–11,13,27,30–34,37,38]. They crystalize in a simple cubic rock-salt structure and possess intrinsically low thermal conductivity, which is ascribed to many factors including the unique ‘off-centring’ behaviour of Pb atoms [62]. Among them PbTe is the most studied and advanced in thermoelectric figure of merit with ZTs exceeding 2 for p-type PbTe [9,30,32,34] and 1.5 for n-type systems [33,63,64]. A maximum conversion efficiency of approximately 8.8% at a temperature difference of 570 K has already been demonstrated in the nanostructured PbTe-based module to be described below [38]. When segmented with Bi2Te3, the efficiency improved to approximately 11% under a temperature difference of 590 K, a record high value for PbTe-based systems [38]. We will discuss more on this PbTe device later in this article.
In consideration of the high cost of Te, there have been incentives to develop high performance in PbSe or PbS as well since they too have low thermal conductivity (1.7 Wm−1 K−1 and 2.6 Wm−1 K−1 at 300 K, respectively) and similar crystal and electronic structures [65]. Lee et al. [66] reported that Sb-doped PbSe shows a ZT of approximately 1.5 at 800 K, while Bi-doped material has a ZT of only approximately 0.9 at 920 K [66]. Very recently, Luo et al. [67] reported n-type GeSe-alloyed PbSe which attains a peak ZT approximately 1.54 at 773 K. This extraordinary ZT matches that of PbTe materials and stems from its ultralow lattice thermal conductivity of approximately 0.36 Wm−1 K−1 at 573 K, which approaches the theoretical limit of amorphous PbSe. First-principles calculations unravel that the alloyed Ge atoms prefer to stay at off-centre lattice positions (discordant atoms), inducing low-frequency modes and contributing to low κlat [67].
There have been great progresses in p-type PbSe materials as well. For example, CdSe or HgSe alloying has been used to converge the two valence bands of PbSe for higher Seebeck coefficients [11,68]. MSe (M = Ca, Sr and Ba) as second phases enhance the ZT values of PbSe by reducing the thermal conductivity without harming the electrical conductivity too much [69]. The best ZT reported so far for p-type PbSe is approximately 1.7 around 900 K [11]. As both high-performance n- and p-type PbSe materials can now be made, they might be on their way to becoming competitive with PbTe for mid-temperature thermoelectric power generation, and if they succeed this will be a great development considering their lower cost and higher melting point [70].
Figure 2a,b shows the ZT values and ZTave of the best performing lead chalcogenides [8–13] reported so far. Undoubtedly, both p- and n-type PbTe have the highest performance over the entire temperature range of interest. All n-type lead chalcogenides show ZTave values close to or greater than unity within their respective working temperature range, which suggests a great promise for real application. Although p-type PbSe and PbTe display ZTmax > 1.5 at elevated temperatures, their lower ZTave respect to the n-type counterparts needs to be improved in the future studies.
Figure 2.
Thermoelectric figure of merit for the best performing p-type and n-type lead chalcogenides [8–13]: (a) ZT and (b) ZTave. (Online version in colour.)
Although lead chalcogenides hold the greatest promise for use in the temperature range of 600–900 K, notable materials in this range are also β-Zn4Sb3, skutterudites (also known as CoSb3), magnesium silicides and single crystal SnSe [14,15,17–20,35,71–73].
β-Zn4Sb3 has long be viewed as a high-performance material with an outstanding ZT > 1.2, the material, however, has been problematic [74]. The main issue that prevents its use in devices is its thermal instability [71]. A recent study by Lin et al. [14] revealed that Zn atoms migrate from crystalline sites to interstitial positions after 425 K and gradually decompose the compound into Zn and ZnSb. However, above 565 K the material recovers its stability and any damage caused by the phase decomposition is repaired. They also demonstrated an excellent ZT of 1.4 at 750 K where β-Zn4Sb3 is thermodynamically stable [14]. Despite this fascinating behaviour, it seems that β-Zn4Sb3 has poor prospects for being implemented in thermoelectric devices.
Skutterudites have been some of the most promising materials for devices for over two decades, especially for n-type systems because of their larger power factors and good mechanical properties [75,76]. The cubic structure of skutterudite has void cages that can be filled by guest atoms which act point defects and rattlers that cause strong phonon scattering thereby significantly reducing κlat [72,77]. By carefully combining multiple species of filling guest atoms, one can introduce two or more localized rattling modes with different frequencies into the skutterudite void to scatter a wide range of phonons to reach very low thermal conductivities [78]. For example, Ba-La-Yb triple-filled skutterudites show a much-reduced κlat to their amorphous limit and outstanding thermoelectric performance (ZT = 1.7 at 850 K) [15]. High efficiency was also demonstrated in thermoelectric modules made of skutterudite compounds [79,80], which will be detailed later in this article.
Magnesium silicides/stannides made of light and cheap elements also show excellent thermoelectric performance [17,18]. It has been reported that n-type Mg2Si-Mg2Sn solid solutions prepared by a two-step solid-state reaction method display the greatest ZT of approximately 1.3 (700 K) at 70 mol.% Mg2Sn, as a result of composition induced conduction band convergence [17]. However, it is very difficult to scale up the preparation of magnesium silicides by high-temperature melting or solid-state reaction technique, because of the large difference in melting temperature between Mg and Si. Moreover, due to the high vapour pressure, Mg can easily evaporate during high-temperature reactions, which leads to phase instability and degraded thermoelectric performance of Mg2Si in service. Recently, ball milling as an easy-to-scale-up technology has been successfully employed to synthesize single phase magnesium silicides [81,82]. This synthesis route also avoids the Mg evaporation problem. Other than materials, a lot of device work has been done on the basis of magnesium silicides as well [83–87]. For example, Sakamoto et al. fabricated Sb-doped Mg2Si thermoelectric device with Ni as electrodes using a PAS technique, which has no notable deterioration even after ageing for 1000 h [83]. Even so, the progress of Mg2Si-based thermoelectric device has been stagnant because of the low performance of its p-type counterpart (ZT approx. 0.7) [18].
SnSe as an emerging compound has received increasing attention in thermoelectric community since 2014 [20]. It has many merits for advanced thermoelectric materials, for example, strong anharmonic chemical bonding for low thermal conductivity, [20] multiple electronic bands for high Seebeck coefficients [35] and ‘3D charge and 2D phonon transport’ behaviour in the out-of-plane direction of its single crystal ingot for high n-type electrical conductivity [19]. These give peak ZTs of approximately 2 and approximately 2.8 (both at 773 K) for single crystals of p- and n-type SnSe, respectively [19,35]. Despite very high thermoelectric performance found in SnSe single crystals, their easy-to-cleave nature limits their processability during cutting and machining. Researchers have been focusing intensely on polycrystalline SnSe, which is more mechanically robust. However, polycrystalline SnSe prepared and handled in air displays significantly higher thermal conductivity and therefore much lower ZTs than their single crystalline form [88–90]. This is surprising because usually, polycrystals have more grains and defects than single crystals, which should contribute to lower thermal conductivities. This was attributed to surface oxidation forming SnO2 which has a very high thermal conductivity [91]. A recent study by Chung and colleagues confirms that the underperformance of polycrystalline SnSe is mainly due to the presence of tin oxides on the surface of the sample [92]. They used an H2 reduction process to remove the majority of tin oxide films that ubiquitously cover SnSe grains and eventually unveiled the ultralow thermal conductivity intrinsic in this material. This study resolves the thermal conductivity discrepancy between single and poly-crystalline SnSe, and attains a ZT approximately 2.5 for polycrystalline SnSe which nearly matches the record-high value of the single crystals. We believe SnSe might revolutionize thermoelectric power generation technology if its device can be demonstrated, but to date such reports are few and will be discussed later [93,94].
(c). High-temperature materials (greater than 900 K)
SiGe alloys were some of the prototypical high-temperature thermoelectric materials to be studied for more than half a century [21,22]. The n-type SiGe shows a promising ZT of 1.1 at approximately 1200 K [21]. However, the relatively poor thermoelectric performance of the p-type counterpart (peak ZT = approx. 0.6) and the high cost of Ge limit the large-scale commercial application of SiGe alloys [22]. The old Si/Ge-based devices are a legacy system and they are outside the scope of this article.
Half-Heusler (HH) alloys are another group of materials which have gained popularity as new candidates for high-temperature thermoelectric power generation because of their excellent electrical properties [24,95–99], very good mechanical robustness [100] and high thermal stability [101]. They belong to a large family of ternary intermetallics with a general formula of ABX, where A is usually the most electropositive rare earth or early transition metal, B is a less electropositive transition metal and X is a main group element [102]. Their physical properties are largely determined by the valence electron count (VEC) [103]. High thermoelectric performance is generally achieved in the semiconducting HH phases with VEC = 18 [23,97,99,104–112], while VEC > 18 usually leads to metallic conduction behaviour [113].
The HH alloys have seen great advances in ZT of both n- and p-type samples, figure 3a,b. ZrNiSn and ZrCoSb are the two most important n-type compositions [95,104,107,109–111,114,117,118,123–125]. ZrNiSn, for example, exhibits a high power factor [23,110,117,126] which originates from a large Seebeck coefficient because of a large density-of-states conduction band effective mass (2.8 me at 300 K) [110]. Sb doping at Sn sites is often used to tune the carrier density to an optimal level (nopt = approx. 3–4 × 1020 cm−3), leading to a ZT of approximately 0.8 at 873 K [127]. The main shortcoming of HH materials is their high lattice thermal conductivity and enormous effort has gone into finding ways to decrease it [23]. Grain size refinement for example by ball milling or using in situ second phase nanostructuring similar to the PbTe strategies mentioned above have been used to decrease κlat of ZrNiSn alloys [104,106,119]. However, this reduction in κlat is often offset by the loss of carrier mobility because of the comparable mean free paths of electrons and phonons (approx. 1 nm) in the ZrNiSn system, making the net enhancement of ZT marginal [131]. Recent studies suggest that isoelectric substitution of Zr by heavier Hf effectively may reduce κlat of ZrNiSn compounds via strong point defect scattering, without obviously deteriorating their electrical properties [95,123,126]. The highest ZT of approximately 1.1 at 1000 K and a ZTave of 0.58 (300–973 K) has been obtained for Zr0.2Hf0.8NiSn0.985Sb0.015 [95].
Figure 3.
Temperature-dependent ZT values for typical (a) n-type [23,96,104,105,110,113–116] and (b) p-type HH alloys [24,99,108,117–122]. (c) Evolution of the peak ZT values for some typical HH alloys in 20 years [24,95,97–99,104,106–112,114,115,117,123–129]. Adapted from Yu [130]. (d) Maximum power output and conversion efficiency as a function of hot side temperature for the thermoelectric device made of n-type ZrNiSn and p-type FeNbSb-based HH materials. The dash line is theoretically calculated conversion efficiency of the module by assuming that no electrical or thermal contact resistances exist. Adapted from Fu [24]. (Online version in colour.)
The thermoelectric properties of p-type HH alloys have attracted significant effort in the last 5 years because ZT in these systems is even lower than in the n-type systems (figure 3c). For instance, the peak ZT of NbFeSb has been reported to increase rapidly from only approximately 0.4 in 2011 [108] to a record value of approximately 1.5 in 2015 [24] by alloying with Ti [112]. However, limited by the low solubility of Ti, the optimal carrier concentration (nopt) could not be reached [98]. Two approaches were attempted to solve this problem: (i) since nopt is proportional to m*, substituting Nb partially by V could lower m* (due to the less spatially localized 4d orbital of Nb than the 3d orbital of V) and therefore decrease nopt [97]; (ii) looking for a more efficient dopant than Ti, for example Hf. This is not only good for electrical properties enhancement, but provides more opportunities for thermal conductivity reduction because the heavier Hf doping results in stronger mass fluctuation and reinforced phonon scattering [24,107,110,123,126]. As a result, an outstanding ZT of approximately 1.5 at 1200 K together with a promising ZTave of approximately 1 (500–1200 K) is achieved in the composition Nb0.88Hf0.12FeSb [24]. More importantly, an eight-couple prototype thermoelectric module made of n-type ZrNiSn and p-type NbFeSb with a high conversion efficiency of 6.2% has been demonstrated experimentally at a temperature difference of approximately 650 K [24]. We will discuss this in detail later. In addition, some facile and cost-efficient fabrication methods (self-propagating high-temperature synthesis, thermal combustion, laser) have been developed to prepare HH alloys [132–134], which point to a reasonable path to scalability [24,110,123,127]. All these advances clearly demonstrate the great potential of HH alloys for high-temperature thermoelectric power generation.
There are other thermoelectric materials that have been explored in the past several years for relevant to high temperatures, for example copper chalcogenides [25,26]. These materials have rigid anion sub-lattices but contain too much copper whose state in the lattice is ‘liquid-like’. This leads to extremely low thermal conductivities but also to destructive electroplating of copper metal prolonged under bias [135]. Although high ZTs over 1.5 have been seen in copper chalcogenides [25,26,135], their stability to high electrical currents or large temperature gradients is poor because of the highly mobile copper ions [136,137]. Recently, efforts have been described aimed to address this issue. By constructing a series of electronically conducting, but ion-blocking barriers to reset the chemical potential of such mixed ionic–electronic conductors seems to help keep the system below the threshold (approx. 0.11 V) for decomposition [138]. Even so, there is still a long way to go before the copper-rich chalcogenides are brought into real use.
Despite the above-mentioned extensive efforts to develop high-ZT materials, and the dramatic progress achieved [1,139–148], there have been few efforts to move the field to the next level and closer to commercialization, which is module development employing the new higher performance materials. Module fabrication using newly developed high-ZT materials is not a trivial pursuit but is needed for achieving practical thermoelectric power generation and to justify continued research efforts in the field of thermoelectrics.
3. Thermoelectric devices
(a). Design of thermoelectric devices and modules
Below we describe the most important developments and efforts in using these very new materials described above in fabrication of functional power generation devices and modules. The maximum conversion efficiency of a thermoelectric device for power generation (ηmax) theoretically defined using two terms, Carnot efficiency (Th − Tc)/Th and the average (device) ZT of the temperature drop (ZTave) [149–151]. The actual efficiency obtained is normally lower than the calculated value because of parasitic ohmic losses at the electrode interfaces and heat losses in various parts in a module.
Figure 4a shows a schematic of a thermoelectric module made from thermocouple of p- and n-type thermoelectric legs. The thermocouples are alternately joined by interconnect electrodes in electrical series but in a thermally parallel connection (figure 4b). The interconnect metallic electrodes need to have very low electrical contact resistance to extract effectively electric power from legs and high thermal conductivity to supply effectively thermal power to legs. An effective way to enhance the conversion efficiency is a stacking of different materials/legs (segment type) or stacking of different modules (cascade type). A segmented module is fabricated by stacking materials/legs with high ZT at high temperature on top of materials/legs with high ZT at low temperature (figure 4c). In a cascaded module, at least two different modules are used. One is a module with high efficiency at higher temperature and is stacked on top of a second module with high efficiency at lower temperature (figure 4d).
Figure 4.
Schematic of (a) a thermoelectric module made from thermocouple of p- and n-type thermoelectric legs. Thermocouple configuration: (b) single stage, (c) segment type and (d) cascade type. (Online version in colour.)
In fabricating good thermoelectric modules, the most important factor is to realize stable low resistance electrical contacts (especially on the hot side) between the interconnect electrodes and the thermoelectric materials [140,152,153]. A module must be exposed to high temperature, which can result in chemical reaction and/or atomic diffusion at interface between the interconnect electrodes and the thermoelectric materials to instability and degradation. These phenomena lead to the gradual development of high-resistance secondary phases at the interface and/or degeneration of the thermoelectric materials, reducing conversion efficiency of a module. Therefore, the formation of excellent diffusion barriers between interconnect electrodes and the thermoelectric materials is needed to prevent the chemical reaction and/or atomic diffusion on the sides of the modules (figure 4b)). This is a challenging task because a fundamental understanding of how to achieve such barriers while at the same time achieving minimal contact resistance is not yet available and basic research on the subject is lagging. Part of the reason for scarce basic research activity on this subject in the past decades, we believe, is the broad perception that this is essentially a device engineering problem and can be solved empirically in a straightforward fashion, an attitude which up to now has clearly been proven incorrect.
The high-temperature operation imposes additional challenges in the joining methods that can be used between materials, diffusion barriers and interconnect electrodes. Good match in the coefficient of thermal expansion between these materials is also needed to reduce thermal mechanical stress [140,152,153]. Therefore, to achieve full benefit from the high-performance materials now available, we must learn how to create low electrical contacts while taking into account the nature of the diffusion barrier, the joining method and the coefficient of thermal expansion. This is a real fundamental materials research problem with many unanswered questions, which needs the support of funding agencies engaged in supporting basic research.
A reduction in heat loses is critical for the development of efficient modules operating at high temperatures (especially greater than 573 K). Heat leaks through parasitic radiation from the hot side to the cold side in the open space between thermoelectric legs, radiation losses from the side walls of legs, poor thermal contacts between the hot source and the hot side of the modules all reduce conversion efficiency.
Table 2 lists ηmax values for the state-of-the-art modules fabricated with newly developed high-ZT materials. ηmax > 7% has been achieved in single-stage modules of nanostructured PbTe [38,154], Sb2Te3/Bi2Te3 [155], skutterudite [79,80,157,163] and silicides [155]. For example, ηmax of approximately 7% was achieved in skutterudite-based modules with Mo diffusion barrier for Th = 872 K and Tc = 314 K [157]. The ηmax was enhanced in segmented and cascaded modules [38,154,155,159–161]. For example, ηmax of silicide-module increases from approximately 8% in single stage Th = 823 K and Tc = 303 K to approximately 12.1% in cascaded module with Sb2Te3/Bi2Te3 for Th = 853 K and Tc = 303 K [155]. Recently, ‘potential’ ηmax have been reported in unileg only composed of p-type leg or n-type leg, including HH, MgAgSb, Mg3Sb2 and Cu26Nb2Ge6S32 [61,164–167]. For MgAgSb, the p-type unileg with Ag contact pad barrier shows ηmax of approximately 8.5% for Th = 518 K and Tc = 293 K. In this case, we do not have real devices or modules but a value predicted from single leg measurements.
Table 2.
Maximum conversion efficiency (ηmax) of the state-of-the-art module with newly developed high-performance materials.
| Th (K) | Tc (K) | module configuration | p-type | n-type | diffusion barrier | ηmax (%) | reference |
|---|---|---|---|---|---|---|---|
| 873 | 303 | single stage | nanostructured PbTe | PbTe | p- and n-type: Co–Fe | 8.8 | [38] |
| 873 | 283 | single stage | nanostructured PbTe | PbTe | p-type: Fe | 8.5 | [154] |
| n-type: Co–Fe | |||||||
| 553 | 303 | single stage | Sb2Te3 | Bi2Te3 | 7.8 | [155] | |
| 873 | 323 | single stage | skutterudite | skutterudite | p- and n-type: Co–Fe–Ni | 8 | [156] |
| 773 | 313 | single stage | skutterudite | skutterudite | p- and n-type: Mo | 7 | [157] |
| 873 | 296 | single stage | skutterudite | skutterudite | 8.4 | [79] | |
| 872 | 314 | single stage | skutterudite | skutterudite | 9.3 | [80] | |
| 820 | 293 | single stage | half-Heusler | half-Heusler | *power density: approximately 3.2 W cm−2 | [158] | |
| 991 | 336 | single stage | half-Heusler | half-Heusler | 6.2 | [24] | |
| 823 | 303 | single stage | higher manganese silicide | Mg2Si | 8 | [155] | |
| 673 | 303 | single stage | clathrate | clathrate | 4.8 | [159] | |
| 674 | 368 | segment type | Sb2Te3/Ag0.9Pb9Sn9Sb0.6Te20 (LASTT) | Bi2Te3/Ag0.86Pb19+xSbTe20 (LAST) | 6.56 | [160] | |
| 873 | 283 | segment type | Bi2Te3/nanostructured PbTe | Bi2Te3/PbTe | p-nanostructured PbTe and n-PbTe: Co–Fe | 11 | [38] |
| 873 | 283 | cascade type | Bi2Te3/nanostructured PbTe | Bi2Te3/PbTe | p-nanostructured PbTe: Fe | 12 | [154] |
| n-PbTe: Co–Fe | |||||||
| 863 | 363 | segment type | Sb2Te3/Zn4Sb3/skutterudite | Bi2Te3/skutterudite | 10 | [161] | |
| 849 | 308 | segment type | Sb2Te3/skutterudite | Bi2Te3/skutterudite | p- and n-skutterudite: Ti-Al | 12 | [162] |
| 753 K | 303 | segment type | Sb2Te3/Zn4Sb3 | Bi2Te3/skutterudite | 7.5 | [159] | |
| 853 | 303 | cascade type | Sb2Te3/higher manganese silicide | Bi2Te3-Bi2Se3/Mg2Si | 12.1 | [155] | |
| 673 | 303 | segment type | Sb2Te3/clathrate | Bi2Te3/clathrate | 7.5 | [159] | |
| 823 | 343 | single stage (unicouple) | skutterudite | skutterudite | p-type: Co2Si | 9.1 | [163] |
| n-type: CoSi2 | |||||||
| 823 | 323 | single stage (unileg) | half-Heusler | — | 9 | [164] | |
| 973 | 317 | single stage (unileg) | half-Heusler | — | 11.4 | [165] | |
| 518 | 293 | single stage (unileg) | MgAgSb | — | *Ag contact layer | 8.5 | [61] |
| 773 | 373 | single stage (unileg) | Mg3Sb2 | — | Fe | 10.6 | [166] |
| 570 | 297 | single stage (unileg) | Cu26Nb2Ge6S32 | Au | 3.3 | [167] |
(b). Modules based on nanostructured PbTe materials
The nanostructured PbTe-based materials have begun to make the transition from mainly materials development to module fabrication and device assembly, and simulation [38,154]. As listed in table 2, high ηmax was demonstrated in nanostructured PbTe-based modules. Here, we review recent progress to bridge the technology gap between materials development and module fabrication in nanostructured PbTe.
PbTe itself is a traditional thermoelectric material, and shows high ZT of approximately 1.1 in the temperature range of 600 K to 900 K. PbTe-based modules have been used decades ago in radioisotope generators for space missions [168]. Using a radioisotope heat source the nominal efficiency of these old modules was approximately 5.1% for Th = 783 K and Tc = 366 K [169].
Because the ZT of PbTe-based materials has now been dramatically boosted through the new strategies discussed above [28,30,170–174], recently we achieved higher ηmax in nanostrucuted Pb-based modules than the old PbTe modules [38,154]. The new modules were constructed from nanostructured Pb0.940Mg0.020Na0.040Te or Pb0.953Ge0.007Na0.040Te as p-type legs, and non-nanostructured PbTe0.9960I0.0040 or PbTe0.9964I0.0036 as n-type legs. As shown in figure 5, ZT values of approximately 1.8 at 810 K (approx. 1.9 at 805 K) and approximately 1.4 at 750 K (approx. 1.3 at 780 K) are easily achieved for scaled-up nanostructured p- and non-nanostructured n-type materials, respectively.
Figure 5.

Temperature dependence of the thermoelectric figure of merit ZT in nanostructured Pb0.940Mg0.020Na0.040Te (p-type) [38], non-nanostructured PbTe0.9960I0.0040 (n-type) [38], nanostructured Pb0.953Ge0.007Na0.040Te (p-type) [154] and non-nanostructured PbTe0.9960I0.0040 (n-type) [154]. (Online version in colour.)
The fabrication process used for the nanostructured PbTe-based module is shown in figure 6 [38,154]. First, p-type and n-type pucks were produced by pressure-assisted sintering of PbTe and diffusion barriers (figure 6a). For the new PbTe-based modules, Fe-, Ni and Nb-based alloys were investigated as diffusion barriers [38,175–179]. In newly developed high-performance module, Fe and 80% Co-20% Fe were used as a diffusion barrier for p-type Pb0.953Ge0.007Na0.040Te and n-type PbTe0.9964I0.0036, respectively [154]. Then, the two pucks were polished together to uniformize these heights (figure 6b) and diced to into rectangular legs (figure 6c). The p- and n-type legs were alternately positioned onto the insulated metal substrate (cold side), where Cu patterns were printed onto the polymer film. In hot side, the legs were interconnected by the Cu electrodes. For the module of p-type Pb0.953Ge0.007Na0.040Te and n-type PbTe0.9964I0.0036, Ag-based nanopaste and Pb–Sn-based solder were used to join the legs and Cu interconnecting electrodes at hot and cold sides, respectively. A photograph of single-stage thermoelectric module of p-type Pb0.953Ge0.007Na0.040Te/n-type PbTe0.9964I0.0036 is shown in figure 6d.
Figure 6.
(a) Pucks of Pb0.953Na0.040Ge0.007Te (p-type) and PbTe0.9964I0.0036 (n-type). The modules were prepared by (b) polishing, (c) cutting and assembling the components onto the insulated metal substrate with the dimension of 18 mm × 15 mm × 1 mm. (d) Single-stage nanostructured PbTe module (p-type: Pb0.953Ge0.007Na0.040Te; n-type: PbTe0.9964I0.0036), (e) segmented Bi2Te3/nanostructured PbTe module (p-type: Pb0.940Mg0.020Na0.040Te; n-type: PbTe0.9960I0.0040), and (f) cascaded Bi2Te3/nanostructured PbTe module (p-type: Pb0.953Ge0.007Na0.040Te; n-type: PbTe0.9964I0.0036). (d,f) PbTe-based legs: 2.0 mm × 2.0 mm × approximately 4.3 mm. (e) PbTe-based legs: 2.0 mm × 2.0 mm × approximately 2.8 mm, and (e,f) Bi2Te3-based legs: 2.0 mm × 2.0 mm × 2.0 mm. (g) Home-built testing system for thermoelectric power generation module. Adapted from Jood [154] and Hu [38]. (Online version in colour.)
The power generation characteristics of the nanostructured PbTe-based module were measured for Th of 873 K–573 K and for Tc of 283 K in vacuum (10−2–10−3 Pa), figure 7g [38,154,159]. In this system, the conversion efficiency η was calculated from P and Qout:
| 3.1 |
The terminal voltage (V), electrical power output (P), heat dissipated from the cold side of the module (Qout) and conversion efficiency (η) of the PbTe-based module comprising eight p-n couples are shown as a function of electrical current I in figure 7. The measured maximum power output (Pmax), open-circuit heat flow (Qoc) and ηmax are listed in table 3. The ηmax approximately 8.5% for Th = 873 K and Tc = 283 K was obtained in single-stage p-type Pb0.953Ge0.007Na0.040Te/n-type PbTe0.9964I0.0036 (figure 6d).
Figure 7.
Power generation characteristics of the nanostructured PbTe-based module (p-type: Pb0.953Ge0.007Na0.040Te; n-type: PbTe0.9964I0.0036). Measured (a) terminal voltage (V), (b) electrical power output (P), (c) heat dissipation from the cold side of the module (Qout) and (d) conversion efficiency (η) as functions of electrical current (I). The hot side temperature (Th) was changed from 873 K to 573 K, while the cold side temperature (Tc) was maintained at 283 K. Adapted from Jood [154]. (Online version in colour.)
Table 3.
Measured and simulated maximum power output (Pmax), open-circuit heat flow (Qoc) and maximum conversion efficiency () in single-stage nanostructured PbTe module (p-type: Pb0.953Ge0.007Na0.040Te; n-type: PbTe0.9964I0.0036) [154], segmented Bi2Te3/nanostructured PbTe module (p-type: Pb0.940Mg0.020Na0.040Te; n-type: PbTe0.9960I0.0040) [38], and cascaded Bi2Te3/nanostructured PbTe module (p-type: Pb0.953Ge0.007Na0.040Te; n-type: PbTe0.9964I0.0036) [154] for the hot side temperature (Th) of 873 K and the cold side temperature (Tc) of 283 K.
| module configuration | Th (K) | Tc (K) | Pmax (W) | Qoc (W) | (%) | |
|---|---|---|---|---|---|---|
| single stage | measured | 873 | 283 | 2.23 | 24.0 | 8.5 |
| simulated | 2.58 | 20.6 | 11.1 | |||
| segment type | measured | 873 | 283 | 2.34 | 15.8 | 11 |
| simulated | 2.55 | 10.8 | 15.6 | |||
| cascade type | measured | 873 | 283 | 1.77 | 13.2 | 12 |
To leverage a wider temperature range of high ZT and particularly to increase the efficiency at the lower temperatures, segmented and cascaded modules can be fabricated using good low-temperature materials such as the Bi2Te3 alloys. In the segmented module, the nanostructured PbTe legs were stacked on top of low-temperature materials Bi2Te3 legs (figure 6e). For the cascaded module, the nanostructured PbTe module was stacked on top of another Bi2Te3 module (figure 6f). This led to a significant enhancement in ηmax to approximately 11% for the segmented module (p-type: Pb0.940Mg0.020Na0.040Te; n-type: PbTe0.9960I0.0040) and approximately 12% for the cascaded module (p-type: Pb0.953Ge0.007Na0.040Te; n-type: PbTe0.9964I0.0036), respectively, for Th = 873 K and Tc = 283 K (table 3). These ηmax values are among the highest achieved in modules using the new state-of-the-art materials (table 2).
A 24 h stability test was performed using various temperature gradients in the single-stage module of p-type Pb0.953Ge0.007Na0.040Te and n-type PbTe0.9964I0.0036 (figure 8) [154]. The module shows good durability in ηmax over 24 h with the exception of at the largest temperature gradient (Th = 873 K and Tc = 283 K). A slight decrease in ηmax was observed after testing with largest temperature gradient of 590 K. This is due to a small increase in internal resistance, which probably was caused by the deterioration of the interfaces between the legs and electrodes and/or the degassing of Te as a result of the large temperature gradients applied over 24 h.
Figure 8.

A 24 h stability test of ηmax as a function of time (hours) with Th = 873, 773, 673 and 573 K and Tc = 283 K in single-stage nanostructured PbTe module (p-type: Pb0.953Ge0.007Na0.040Te; n-type: PbTe0.9964I0.0036). Adapted from Jood [154]. (Online version in colour.)
We modelled the power generation characteristics of the nanostructured PbTe-based module using the commercial three-dimensional finite-element software (COMSOL Multiphysics with Heat Transfer Module) using a measured Seebeck coefficient, electrical resistivity and thermal conductivity of the materials. The simulated values of Pmax, Qoc and ηmax for single-stage nanostructured PbTe module and segmented Bi2Te3/nanostructured PbTe module for Th = 873 K and Tc = 283 K are listed in table 3 along with the experimentally measured values for comparison. The simulated ηmax are 11.1% for single-stage module and 15.6% for cascaded module, which are approximately 30% and approximately 40% higher than the measured values. The reason for the lower measured values of ηmax is the lower Pmax and higher Qoc values of the fabricated module. In both modules, the measured values of Pmax (approx. 2.23 W for single stage and approx. 2.34 W for cascade type) were lower than the theoretically simulated ones (2.58 W for single stage and 2.55 W for cascade type). This difference arises from the parasitic electrical resistances at the material interfaces. The parasitic electrical resistances result in ohmic losses, leading to the lower power output. Further improvement of the electrical contact at material interfaces such as thermoelectric materials and interconnect electrodes will enable higher Pmax and higher ηmax. Moreover, in both modules, simulated values of Qoc (approx. 20.6 W for single stage and approx. 10.8 W for cascade type) are lower than measured ones (approx. 24.0 W for single stage and approx. 15.8 W for cascade type). The deviation can be explained by radiative heat exchange from hot side to cold side in the module as a loss mechanism. The open area between the thermoelectric legs acts as a path for thermal radiation transfer. Further optimization of geometrical configuration of the module will reduce these radiation heat losses and enable lower Qoc and higher ηmax.
(c). Skutterudite-based module
Skutterudites have also demonstrated attractive performance in temperature range of 600–900 K. The fabrication process used for the segmented Bi2Te3/skutterudite (p-type: CeFe3.85Mn0.15Sb12; n-type: Yb0.3Co4Sb12) module is shown in figure 9a [162]. The p- and n-type samples were hot-pressed with diffusion barrier (Ti-Al) and contact layer (Ni). After polishing and cutting, interconnect electrodes (Mo-Cu) were joined at both hot and cold sides. The Ti-Al-based diffusion barrier yields a low specific electrical contact resistance of approximately 10 µΩ m2 at the interface between Ti-Al and skutterudite. For the hot side, the electrode and Ni contact layer on top of the leg were joined by brazing with Cu-Ag-Zn. An Sn-based solder was used for joining Bi2Te3 and skutterudite. Glass fibres were filled into the gaps between legs to reduce thermal convection and radiation losses.
Figure 9.
(a) Fabrication process used for the segmented Bi2Te3/skutterudite module (p-type: CeFe3.85Mn0.15Sb12; n-type: Yb0.3Co4Sb12) and (b) the conversion efficiency (η) as functions of electrical current (I) [162]. Adapted from Zhang [162]. (Online version in colour.)
Salvador et al. have achieved approximately 7% conversion efficiency in a skutterudite-based module for Th = 773 K and Tc = 313 K (table 2) [157,180]. In this module, Mm0.30Fe1.46Co2.54Sb12.05 (Mm stands for Misch metal which is an alloy of La, Ce, Pr and Nd) was used as a p-type material and Yb0.09Ba0.05La0.05Co4Sb12 was used as an n-type material (figure 10a). Mo diffusion barriers were formed to prevent the Sb in the skutterudites from reacting with the braze and interconnect electrode for both the p- and n-type legs. Recently, ηmax of a skutterudite-based module was improved through grain-boundary engineering. For example, ηmax approximately 8.4% for Th = 873 K and Tc = 296 K was measured in modules based on grain-boundary engineered skutterudite with graphene [79] and ηmax approximately 9.3% for Th = 872 K and Tc = 314 K was found in that of grain-boundary engineered skutterudite with carbon nanotube (table 2) [80].
Figure 10.
Skutterudite-based module and interface between skutterudites and diffusion barrier. (a) Single-stage module of p-0Fe1.46Co2.54Sb12.05 (Mm: Misch metal) and n-type Yb0.09Ba0.05La0.05Co4Sb12 with Mo diffusion barrier [157,180]. (b) Single-stage module of p-type La0.7Ba0.1Ga0.1Ti0.1Fe3CoSb12 and n-type Yb0.3Ca0.1Al0.1Ga0.1In0.1Co3.75Fe0.25Sb12 with Co–Fe–Ni diffusion barrier [156,181]. (a) Adapted from Salvador [157,180], (b) Adapted from Guo [156]. (Online version in colour.)
Like the PbTe-module above, the ηmax was dramatically enhanced by segmenting the skutterudite-based device with a Bi2Te3-based device on the cold side (figure 9b). This led to a ηmax approximately 12% for Th = 849 K and Tc = 363 K, figure 9b and table 2. In this module, CeFe3.85Mn0.15Sb12, Yb0.3Co4Sb12 and Ti-Al were used as p-type skutterudite, n-type skutterudite and diffusion barriers, respectively [162].
Long-term stability test has been reported on p-type CeyFeCo3Sb12 and n-type Yb0.3Co4Sb12 modules [152]. The ηmax of this module was approximately 7.2% for Th = 843 K and Tc = 338 K. The power output of the module was steady for 840 h and more for Th of 773 K and Tc of 298 K. The module also shows good stability in the power output for the heat cycle test over 2400 times under Th of 373 K– 823 K and Tc of 298 K. Moreover, the 8000 h stability test under Th = 873 K and Tc = 353 K was performed on a module of p-type La0.7Ba0.1Ga0.1Ti0.1Fe3CoSb12 and n-type Yb0.3Ca0.1Al0.1Ga0.1In0.1Co3.75Fe0.25Sb12 with a Co–Fe–Ni diffusion barrier (figure 10b) [156,181]. The ηmax of this module was approximately 8% for Th = 843 K and Tc = 338 K. A power output initially decreased by approximately 6% in 300 h then remained constant. This module also shows good stability in the power output for heat cycle test over 500 times under Th of 473 K– 873 K and Tc of 313 K. These life tests exhibit promising durability of skutterudite-based modules for practical use.
(d). Half-Heusler-based module
As mentioned above in §2c, ηmax of approximately 6.2% has been demonstrated experimentally in an HH-based module of p-type FeNbSb and n-type ZrNiSn-based alloys for Th = 991 K and Tc = 336 K (figure 3d and table 2) [24]. As shown in figure 3d, ηmax of 11.3% is estimated for Th = 991 K and Tc = 336 K, assuming no electrical and thermal contact resistances in the module. The difference between measured and estimated values in ηmax is due to insufficient electrical contact and the large heat radiation and convection. In particular, the contact resistance was estimated to contribute to about 3.2% efficiency loss.
Long-term stability of HH-based module of p-type Zr0.5Hf0.5CoSb0.8Sn0.2 and n-type Zr0.4Hf0.6NiSn0.98Sb0.02 was investigated through heat cycle and heat shock tests [158]. Power output of the module with substrate size of 16 mm × 16 mm reached approximately 2.8 W (power density: approx. 3.2 W cm−2) for Th = 820 K and Tc = 293 K (table 2). The power output was unchanged under several heat cycle tests (such as 130 cycles of hot side temperature change from 523 K to 823 K) and six shocks of the module by increasing the hot side temperature from 373 K to 873 K with a heating rate of 100 K min−1.
Recently, the ZT value in p-type HH materials was improved and corresponding efficiency was demonstrated in these p-type unilegs. Figure 11 shows ZT value in p-type HH materials and ηmax [164,165]. High ZT values of approximately 1.42 and approximately 1.52 at 973 K were found in ZrCoBi0Sb0.15Sn0.20 and Ta0.74V0.1Ti0.16FeSb (figure 11a), respectively, leading to higher conversion efficiency in unilegs. For example, ηmax values of approximately 9% and approximately 11.4% have been reported in unileg of ZrCoBi0.65Sb0.15Sn0.20 for Th = 823 K and Tc = 323 K and Ta0.74V0.1Ti0.16FeSb for Th = 973 K and Tc = 317 K, respectively (figure 11b and table 2). The results indicate the p-type TaFeSb-based HHs are promising for thermoelectric power generation. For real application, development high-performance n-type HH legs is required.
Figure 11.
(a) ZT value in p-type half-Heuslers and (b) conversion efficiency ηmax for these unilegs [24,109,118,165,182]. The efficiency was estimated for Ta0.74V0.1Ti0.16FeSb (Theoretical prediction). Adapted from Zhu [165]. (Online version in colour.)
(e). Other modules
Zintl compounds are an important class of thermoelectrics [183,184]. One example is Yb14MnSb11, which shows high ZT approximately 1.0 at high temperature (1223 K) [184]. The module based on Yb14MnSb11 had been developed for radioisotope thermoelectric generators in space missions [185]. The Yb14MnSb11 leg demonstrated good durability in thermoelectric properties over the six month test at 1273 K and 1323 K under vacuum. The sublimation of Yb14MnSb11 was suppressed through formation sublimation barrier made of porous alumina layer. Stable contact with low resistance was achieved between the Mo and Yb14MnSb11.
As mentioned above (§2b), SnSe has recently emerged as a new class of thermoelectric material. The fabrication of a module based on polycrystalline SnSe has just begun. Ag/Ni bilayer and Ag/Co/Ti multi-layer were developed as diffusion barriers [93,186]. For example, no secondary phases and no cracks were observed between the Ag/Co/Ti multi-layer and the SnSe interface [186]. Moreover, the specific contact resistance between the Ag/Co/Ti multi-layer and the SnSe was observed to be approximately 1.53 mΩ cm2 at room temperature. However, after heat treatment at 723 K for 20 h under Ar atmosphere the specific contact resistance increased to approximately 2.27 mΩ cm2 at room temperature because of diffusion of Sn and Se into Ag layer. It is necessary to develop stable diffusion barriers for the fabrication of SnSe module.
4. Summary and future prospects
There has been tremendous progress in the science of thermoelectric materials, but more progress needs to be made when it comes to transitioning from materials to devices. This is crucial for commercialization and practical use. On the material side, many outstanding ZTs > 1.5 or even over 2 have been reported; however, one should keep in mind that the conversion efficiency is more related to a material's average ZT over the whole operating temperature. Moreover, in addition to thermoelectric performance, we also need to look into a material's mechanical strength and thermal stability, which are crucial in real applications but have been long ignored in most studies. Going forward, researchers should put more effort into fabricating novel thermoelectric devices with high efficiency from the new materials, including the design of new device structures and evaluation protocols of device performance/stability.
Acknowledgements
M.O. thanks Dr Priyanka Jood, Dr Raju Chetty and Mr Atsushi Yamamoto of AIST for lending their expertise on thermoelectric modules.
Data accessibility
This article has no additional data.
Competing interests
We declare we have no competing interests.
Funding
G.T. would like to acknowledge the recent financial support from the Natural Science Foundation of China (grant no. 11804261) and ‘the Fundamental Research Funds for the Central Universities (WUT: 2019IVA068 and 2019IVB049)’. The work at AIST was partially supported as part of the Development of Thermal Management Materials and Technology funded by the New Energy and Industrial Technology Development Organization (NEDO) and the International Joint Research Program for Innovative Energy Technology funded by the Ministry of Economy, Trade and Industry (METI). At Northwestern University, the work was supported primarily by the Department of Energy, Office of Science, Basic Energy Sciences under grant no. DE-SC0014520 (M.G.K.)
References
- 1.Tan G, Zhao L-D, Kanatzidis MG. 2016. Rationally designing high-performance bulk thermoelectric materials. Chem. Rev. 116, 12 123–12 149. ( 10.1021/acs.chemrev.6b00255) [DOI] [PubMed] [Google Scholar]
- 2.Snyder GJ, Toberer ES. 2008. Complex thermoelectric materials. Nat. Mater. 7, 105–114. ( 10.1038/nmat2090) [DOI] [PubMed] [Google Scholar]
- 3.Lenoir B, Dauscher A, Cassart M, Ravich YI, Scherrer H. 1998. Effect of antimony content on the thermoelectric figure of merit of Bi1−xSbx alloys. J. Phys. Chem. Solids 59, 129–134. ( 10.1016/S0022-3697(97)00187-X) [DOI] [Google Scholar]
- 4.Chung D-Y, Hogan T, Brazis P, Rocci-Lane M, Kannewurf C, Bastea M, Uher C, Kanatzidis MG. 2000. CsBi4Te6: a high-performance thermoelectric material for low-temperature applications. Science 287, 1024–1027. ( 10.1126/science.287.5455.1024) [DOI] [PubMed] [Google Scholar]
- 5.Hu L, Wu H, Zhu T, Fu C, He J, Ying P, Zhao X. 2015. Tuning multiscale microstructures to enhance thermoelectric performance of n-type bismuth-telluride-based solid solutions. Adv. Energy Mater. 5, 1500411 ( 10.1002/aenm.201500411) [DOI] [Google Scholar]
- 6.Kim SI, et al. 2015. Dense dislocation arrays embedded in grain boundaries for high-performance bulk thermoelectrics. Science 348, 109–114. ( 10.1126/science.aaa4166) [DOI] [PubMed] [Google Scholar]
- 7.Liu Z, Geng H, Mao J, Shuai J, He R, Wang C, Cai W, Sui J, Ren Z. 2016. Understanding and manipulating the intrinsic point defect in α-MgAgSb for higher thermoelectric performance. J. Mater. Chem. A 4, 16834–16 840. ( 10.1039/C6TA06832D) [DOI] [Google Scholar]
- 8.Fu L, Yin M, Wu D, Li W, Feng D, Huang L, He J. 2017. Large enhancement of thermoelectric properties in n-type PbTe via dual-site point defects. Energy Environ. Sci. 10, 2030–2040. ( 10.1039/C7EE01871A) [DOI] [Google Scholar]
- 9.Tan G, et al. 2016. Non-equilibrium processing leads to record high thermoelectric figure of merit in PbTe–SrTe. Nat. Commun. 7, 12167 ( 10.1038/ncomms12167) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Zhou C, et al. 2018. High-performance n-type PbSe–Cu2Se thermoelectrics through conduction band engineering and phonon softening. J. Am. Chem. Soc. 140, 15 535–15 545. ( 10.1021/jacs.8b10448) [DOI] [PubMed] [Google Scholar]
- 11.Hodges JM, et al. 2018. Chemical insights into PbSe–x%HgSe: high power factor and improved thermoelectric performance by alloying with discordant atoms. J. Am. Chem. Soc. 140, 18 115–18 123. ( 10.1021/jacs.8b11050) [DOI] [PubMed] [Google Scholar]
- 12.Ibáñez M, et al. 2016. High-performance thermoelectric nanocomposites from nanocrystal building blocks. Nat. Commun. 7, 10766 ( 10.1038/ncomms10766) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Zhao L-D, He J, Hao S, Wu C-I, Hogan TP, Wolverton C, Dravid VP, Kanatzidis MG. 2012. Raising the thermoelectric performance of p-type PbS with endotaxial nanostructuring and valence-band offset engineering using CdS and ZnS. J. Am. Chem. Soc. 134, 16 327–16 336. ( 10.1021/ja306527n) [DOI] [PubMed] [Google Scholar]
- 14.Lin J, et al. 2014. Unexpected high-temperature stability of β-Zn4Sb3 opens the door to enhanced thermoelectric performance. J. Am. Chem. Soc. 136, 1497–1504. ( 10.1021/ja410605f) [DOI] [PubMed] [Google Scholar]
- 15.Shi X, et al. 2011. Multiple-filled skutterudites: high thermoelectric figure of merit through separately optimizing electrical and thermal transports. J. Am. Chem. Soc. 133, 7837–7846. ( 10.1021/ja111199y) [DOI] [PubMed] [Google Scholar]
- 16.Rogl G, et al. 2015. New bulk p-type skutterudites DD0.7Fe2.7Co1.3Sb12−xXx (X = Ge, Sn) reaching ZT > 1.3. Acta Mater. 91, 227–238. ( 10.1016/j.actamat.2015.03.008) [DOI] [Google Scholar]
- 17.Liu W, Tan X, Yin K, Liu H, Tang X, Shi J, Zhang Q, Uher C. 2012. Convergence of conduction bands as a means of enhancing thermoelectric performance of n-type Mg2Si1−xSnx solid solutions. Phys. Rev. Lett. 108, 166601 ( 10.1103/PhysRevLett.108.166601) [DOI] [PubMed] [Google Scholar]
- 18.Gao P, Davis JD, Poltavets VV, Hogan TP. 2016. The p-type Mg2LixSi0.4Sn0.6 thermoelectric materials synthesized by a B2O3 encapsulation method using Li2CO3 as the doping agent. J. Mater. Chem. C 4, 929–934. ( 10.1039/C5TC03692E) [DOI] [Google Scholar]
- 19.Chang C, et al. 2018. 3D charge and 2D phonon transports leading to high out-of-plane ZT in n-type SnSe crystals. Science 360, 778–783. ( 10.1126/science.aaq1479) [DOI] [PubMed] [Google Scholar]
- 20.Zhao L-D, Lo S-H, Zhang Y, Sun H, Tan G, Uher C, Wolverton C, Dravid VP, Kanatzidis MG. 2014. Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Nature 508, 373–377. ( 10.1038/nature13184) [DOI] [PubMed] [Google Scholar]
- 21.Yu B, et al. 2012. Enhancement of thermoelectric properties by modulation-doping in silicon germanium alloy nanocomposites. Nano Lett. 12, 2077–2082. ( 10.1021/nl3003045) [DOI] [PubMed] [Google Scholar]
- 22.Vining CB, Laskow W, Hanson JO, Beck RRVD, Gorsuch PD. 1991. Thermoelectric properties of pressure-sintered Si0.8Ge0.2 thermoelectric alloys. J. Appl. Phys. 69, 4333–4340. ( 10.1063/1.348408) [DOI] [Google Scholar]
- 23.Liu Y, Xie H, Fu C, Snyder GJ, Zhao X, Zhu T. 2015. Demonstration of a phonon-glass electron-crystal strategy in (Hf,Zr)NiSn half-Heusler thermoelectric materials by alloying. J. Mater. Chem. A 3, 22 716–22 722. ( 10.1039/C5TA04418A) [DOI] [Google Scholar]
- 24.Fu C, Bai S, Liu Y, Tang Y, Chen L, Zhao X, Zhu T. 2015. Realizing high figure of merit in heavy-band p-type half-Heusler thermoelectric materials. Nat. Commun. 6, 8144 ( 10.1038/ncomms9144) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Nunna R, et al. 2017. Ultrahigh thermoelectric performance in Cu2Se-based hybrid materials with highly dispersed molecular CNTs. Energy Environ. Sci. 10, 1928–1935. ( 10.1039/C7EE01737E) [DOI] [Google Scholar]
- 26.He Y, Day T, Zhang T, Liu H, Shi X, Chen L, Snyder GJ. 2014. High thermoelectric performance in non-toxic earth-abundant copper sulfide. Adv. Mater. 26, 3974–3978. ( 10.1002/adma.201400515) [DOI] [PubMed] [Google Scholar]
- 27.Heremans JP, Jovovic V, Toberer ES, Saramat A, Kurosaki K, Charoenphakdee A, Yamanaka S, Snyder GJ. 2008. Enhancement of thermoelectric efficiency in PbTe by distortion of the electronic density of states. Science 321, 554–557. ( 10.1126/science.1159725) [DOI] [PubMed] [Google Scholar]
- 28.Pei Y, Shi X, LaLonde A, Wang H, Chen L, Snyder GJ. 2011. Convergence of electronic bands for high performance bulk thermoelectrics. Nature 473, 66 ( 10.1038/nature09996) [DOI] [PubMed] [Google Scholar]
- 29.Poudel B, et al. 2008. High-thermoelectric performance of nanostructured bismuth antimony telluride bulk alloys. Science 320, 634–638. ( 10.1126/science.1156446) [DOI] [PubMed] [Google Scholar]
- 30.Biswas K, He J, Blum ID, Wu C-I, Hogan TP, Seidman DN, Dravid VP, Kanatzidis MG. 2012. High-performance bulk thermoelectrics with all-scale hierarchical architectures. Nature 489, 414 ( 10.1038/nature11439) [DOI] [PubMed] [Google Scholar]
- 31.Biswas K, He J, Zhang Q, Wang G, Uher C, Dravid VP, Kanatzidis MG. 2011. Strained endotaxial nanostructures with high thermoelectric figure of merit. Nat. Chem. 3, 160 ( 10.1038/nchem.955) [DOI] [PubMed] [Google Scholar]
- 32.Korkosz RJ, et al. 2014. High ZT in p-type (PbTe)1–2x(PbSe)x(PbS)x thermoelectric materials. J. Am. Chem. Soc. 136, 3225–3237. ( 10.1021/ja4121583) [DOI] [PubMed] [Google Scholar]
- 33.Sarkar S, Zhang X, Hao S, Hua X, Bailey TP, Uher C, Wolverton C, Dravid VP, Kanatzidis MG. 2018. Dual alloying strategy to achieve a high thermoelectric figure of merit and lattice hardening in p-type nanostructured PbTe. ACS Energy Lett. 3, 2593–2601. ( 10.1021/acsenergylett.8b01684) [DOI] [Google Scholar]
- 34.Wu HJ, Zhao LD, Zheng FS, Wu D, Pei YL, Tong X, Kanatzidis MG, He JQ. 2014. Broad temperature plateau for thermoelectric figure of merit ZT > 2 in phase-separated PbTe0.7S0.3. Nat. Commun. 5, 4515 ( 10.1038/ncomms5515) [DOI] [PubMed] [Google Scholar]
- 35.Zhao L-D, et al. 2016. Ultrahigh power factor and thermoelectric performance in hole-doped single-crystal SnSe. Science 351, 141–144. ( 10.1126/science.aad3749) [DOI] [PubMed] [Google Scholar]
- 36.Zhang X, Li J, Wang X, Chen Z, Mao J, Chen Y, Pei Y. 2018. Vacancy manipulation for thermoelectric enhancements in GeTe alloys. J. Am. Chem. Soc. 140, 15 883–15 888. ( 10.1021/jacs.8b09375) [DOI] [PubMed] [Google Scholar]
- 37.Chen Z, et al. 2017. Lattice dislocations enhancing thermoelectric PbTe in addition to band convergence. Adv. Mater. 29, 1606768 ( 10.1002/adma.201606768) [DOI] [PubMed] [Google Scholar]
- 38.Hu X, Jood P, Ohta M, Kunii M, Nagase K, Nishiate H, Kanatzidis MG, Yamamoto A. 2016. Power generation from nanostructured PbTe-based thermoelectrics: comprehensive development from materials to modules. Energy Environ. Sci. 9, 517–529. ( 10.1039/C5EE02979A) [DOI] [Google Scholar]
- 39.Wei Q, Mukaida M, Kirihara K, Naitoh Y, Ishida T. 2015. Recent progress on PEDOT-based thermoelectric materials. Materials 8, 732–750. ( 10.3390/ma8020732) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Goldsmid HJ, Douglas RW. 1954. The use of semiconductors in thermoelectric refrigeration. Br. J. Appl. Phys. 5, 386–390. ( 10.1088/0508-3443/5/11/303) [DOI] [Google Scholar]
- 41.Lin W-C, Li Y-S, Wu AT. 2018. Study of diffusion barrier for solder/n-type Bi2Te3 and bonding strength for p- and n-type thermoelectric modules. J. Electron. Mater. 47, 148–154. ( 10.1007/s11664-017-5906-x) [DOI] [Google Scholar]
- 42.Zhou H, Mu X, Zhao W, Tang D, Wei P, Zhu W, Nie X, Zhang Q. 2017. Low interface resistance and excellent anti-oxidation of Al/Cu/Ni multilayer thin-film electrodes for Bi2Te3-based modules. Nano Energy 40, 274–281. ( 10.1016/j.nanoen.2017.08.034) [DOI] [Google Scholar]
- 43.Kato K, Hatasako Y, Kashiwagi M, Hagino H, Adachi C, Miyazaki K. 2014. Fabrication of a flexible bismuth telluride power generation module using microporous polyimide films as substrates. J. Electron. Mater. 43, 1733–1739. ( 10.1007/s11664-013-2852-0) [DOI] [Google Scholar]
- 44.Xie W, Tang X, Yan Y, Zhang Q, Tritt TM. 2009. Unique nanostructures and enhanced thermoelectric performance of melt-spun BiSbTe alloys. Appl. Phys. Lett. 94, 102111 ( 10.1063/1.3097026) [DOI] [Google Scholar]
- 45.Tang X, Xie W, Li H, Zhao W, Zhang Q, Niino M. 2007. Preparation and thermoelectric transport properties of high-performance p-type Bi2Te3 with layered nanostructure. Appl. Phys. Lett. 90, 012102 ( 10.1063/1.2425007) [DOI] [Google Scholar]
- 46.Ma Y, Hao Q, Poudel B, Lan Y, Yu B, Wang D, Chen G, Ren Z. 2008. Enhanced thermoelectric figure-of-merit in p-type nanostructured bismuth antimony tellurium alloys made from elemental chunks. Nano Lett. 8, 2580–2584. ( 10.1021/nl8009928) [DOI] [PubMed] [Google Scholar]
- 47.Xie W, et al. 2010. Identifying the specific nanostructures responsible for the high thermoelectric performance of (Bi,Sb)2Te3 nanocomposites. Nano Lett. 10, 3283–3289. ( 10.1021/nl100804a) [DOI] [PubMed] [Google Scholar]
- 48.Shen J-J, Zhu T-J, Zhao X-B, Zhang S-N, Yang S-H, Yin Z-Z. 2010. Recrystallization induced in situ nanostructures in bulk bismuth antimony tellurides: a simple top down route and improved thermoelectric properties. Energy Environ. Sci. 3, 1519–1523. ( 10.1039/C0EE00012D) [DOI] [Google Scholar]
- 49.Hu L, Gao H, Liu X, Xie H, Shen J, Zhu T, Zhao X. 2012. Enhancement in thermoelectric performance of bismuth telluride based alloys by multi-scale microstructural effects. J. Mater. Chem. 22, 16 484–16 490. ( 10.1039/C2JM32916F) [DOI] [Google Scholar]
- 50.Jaworski CM, Kulbachinskii V, Heremans JP. 2009. Resonant level formed by tin in Bi2Te3 and the enhancement of room-temperature thermoelectric power. Phys. Rev. B 80, 233201 ( 10.1103/PhysRevB.80.233201) [DOI] [Google Scholar]
- 51.Zhitinskaya MK, Nemov SA, Svechnikova TE, Reinshaus P, Müller E. 2000. Influence of Sn resonance states on the electrical homogeneity of Bi2Te3 single crystals. Semiconductors 34, 1363–1364. ( 10.1134/1.1331791) [DOI] [Google Scholar]
- 52.Deng R, et al. 2018. Thermal conductivity in Bi0.5Sb1.5Te3+x and the role of dense dislocation arrays at grain boundaries. Sci. Adv. 4, eaar5606 ( 10.1126/sciadv.aar5606) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Zheng G, et al. 2016. Toward high-thermoelectric-performance large-size nanostructured BiSbTe alloys via optimization of sintering-temperature distribution. Adv. Energy Mater. 6, 1600595 ( 10.1002/aenm.201600595) [DOI] [Google Scholar]
- 54.Kirkham MJ, dos Santos AM, Rawn CJ, Lara-Curzio E, Sharp JW, Thompson AJ. 2012. Abinitio determination of crystal structures of the thermoelectric material MgAgSb. Phys. Rev. B 85, 144120 ( 10.1103/PhysRevB.85.144120) [DOI] [Google Scholar]
- 55.Ying P, Liu X, Fu C, Yue X, Xie H, Zhao X, Zhang W, Zhu T. 2015. High performance α-MgAgSb thermoelectric materials for low temperature power generation. Chem. Mater. 27, 909–913. ( 10.1021/cm5041826) [DOI] [Google Scholar]
- 56.Zhao H, et al. 2014. High thermoelectric performance of MgAgSb-based materials. Nano Energy 7, 97–103. ( 10.1016/j.nanoen.2014.04.012) [DOI] [Google Scholar]
- 57.Li D, et al. 2015. Atomic disorders induced by silver and magnesium ion migrations favor high thermoelectric performance in α-MgAgSb-based materials. Adv. Funct. Mater. 25, 6478–6488. ( 10.1002/adfm.201503022) [DOI] [Google Scholar]
- 58.Tan X, Wang L, Shao H, Yue S, Xu J, Liu G, Jiang H, Jiang J. 2017. Improving thermoelectric performance of α-MgAgSb by theoretical band engineering design. Adv. Energy Mater. 7, 1700076 ( 10.1002/aenm.201700076) [DOI] [Google Scholar]
- 59.Liu Z, Mao J, Sui J, Ren Z. 2018. High thermoelectric performance of α-MgAgSb for power generation. Energy Environ. Sci. 11, 23–44. ( 10.1039/C7EE02504A) [DOI] [Google Scholar]
- 60.Liu Z, Wang Y, Gao W, Mao J, Geng H, Shuai J, Cai W, Sui J, Ren Z. 2017. The influence of doping sites on achieving higher thermoelectric performance for nanostructured α-MgAgSb. Nano Energy 31, 194–200. ( 10.1016/j.nanoen.2016.11.010) [DOI] [Google Scholar]
- 61.Kraemer D, Sui J, McEnaney K, Zhao H, Jie Q, Ren ZF, Chen G. 2015. High thermoelectric conversion efficiency of MgAgSb-based material with hot-pressed contacts. Energy Environ. Sci. 8, 1299–1308. ( 10.1039/C4EE02813A) [DOI] [Google Scholar]
- 62.Božin ES, Malliakas CD, Souvatzis P, Proffen T, Spaldin NA, Kanatzidis MG, Billinge SJ. L. 2010. Entropically stabilized local dipole formation in lead chalcogenides. Science 330, 1660–1663. ( 10.1126/science.1192759) [DOI] [PubMed] [Google Scholar]
- 63.Xiao Y, et al. 2018. Realizing high performance n-type PbTe by synergistically optimizing effective mass and carrier mobility and suppressing bipolar thermal conductivity. Energy Environ. Sci. 11, 2486–2495. ( 10.1039/C8EE01151F) [DOI] [Google Scholar]
- 64.Zhang J, Wu D, He D, Feng D, Yin M, Qin X, He J. 2017. Extraordinary thermoelectric performance realized in n-type PbTe through multiphase nanostructure engineering. Adv. Mater. 29, 1703148 ( 10.1002/adma.201703148) [DOI] [PubMed] [Google Scholar]
- 65.Pei Y-L, Liu Y. 2012. Electrical and thermal transport properties of Pb-based chalcogenides: PbTe, PbSe, and PbS. J. Alloys Compd. 514, 40–44. ( 10.1016/j.jallcom.2011.10.036) [DOI] [Google Scholar]
- 66.Lee Y, Lo S-H, Chen C, Sun H, Chung D-Y, Chasapis TC, Uher C, Dravid VP, Kanatzidis MG. 2014. Contrasting role of antimony and bismuth dopants on the thermoelectric performance of lead selenide. Nat. Commun. 5, 3640 ( 10.1038/ncomms4640) [DOI] [PubMed] [Google Scholar]
- 67.Luo Z-Z, et al. 2018. Soft phonon modes from off-center Ge atoms lead to ultralow thermal conductivity and superior thermoelectric performance in n-type PbSe–GeSe. Energy Environ. Sci. 11, 3220–3230. ( 10.1039/C8EE01755G) [DOI] [Google Scholar]
- 68.Zhao L-D, et al. 2013. High thermoelectric performance via hierarchical compositionally alloyed nanostructures. J. Am. Chem. Soc. 135, 7364–7370. ( 10.1021/ja403134b) [DOI] [PubMed] [Google Scholar]
- 69.Lee Y, Lo S-H, Androulakis J, Wu C-I, Zhao L-D, Chung D-Y, Hogan TP, Dravid VP, Kanatzidis MG. 2013. High-performance tellurium-free thermoelectrics: all-scale hierarchical structuring of p-type PbSe–MSe systems (M = Ca, Sr, Ba). J. Am. Chem. Soc. 135, 5152–5160. ( 10.1021/ja400069s) [DOI] [PubMed] [Google Scholar]
- 70.Wang H, Pei Y, LaLonde AD, Snyder GJ. 2011. Heavily doped p-type PbSe with high thermoelectric performance: an alternative for PbTe. Adv. Mater. 23, 1366–1370. ( 10.1002/adma.201004200) [DOI] [PubMed] [Google Scholar]
- 71.Snyder GJ, Christensen M, Nishibori E, Caillat T, Iversen BB. 2004. Disordered zinc in Zn4Sb3 with phonon-glass and electron-crystal thermoelectric properties. Nat. Mater. 3, 458 ( 10.1038/nmat1154) [DOI] [PubMed] [Google Scholar]
- 72.Sales BC, Mandrus D, Williams RK. 1996. Filled skutterudite antimonides: a new class of thermoelectric materials. Science 272, 1325–1328. ( 10.1126/science.272.5266.1325) [DOI] [PubMed] [Google Scholar]
- 73.Duong AT, et al. 2016. Achieving ZT = 2.2 with Bi-doped n-type SnSe single crystals. Nat. Commun. 7, 13713 ( 10.1038/ncomms13713) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Toberer ES, Rauwel P, Gariel S, Taftø J, Jeffrey Snyder G. 2010. Composition and the thermoelectric performance of β-Zn4Sb3. J. Mater. Chem. 20, 9877–9885. ( 10.1039/C0JM02011G) [DOI] [Google Scholar]
- 75.Caillat T, Borshchevsky A, Fleurial JP. 1994. Preparation and thermoelectric properties of p- and n-type CoSb3. AIP Conf. Proc. 316, 58–61. ( 10.1063/1.46835) [DOI] [Google Scholar]
- 76.Caillat T, Borshchevsky A, Fleurial JP. 1994. Existence and some properties of new ternary skutterudite phases. AIP Conf. Proc. 316, 209–211. ( 10.1063/1.46796) [DOI] [Google Scholar]
- 77.Sales BC, Mandrus D, Chakoumakos BC, Keppens V, Thompson JR. 1997. Filled skutterudite antimonides: electron crystals and phonon glasses. Phys. Rev. B 56, 15 081–15 089. ( 10.1103/PhysRevB.56.15081) [DOI] [Google Scholar]
- 78.Xi L, Yang J, Lu C, Mei Z, Zhang W, Chen L. 2010. Systematic study of the multiple-element filling in caged skutterudite CoSb3. Chem. Mater. 22, 2384–2394. ( 10.1021/cm903717w) [DOI] [Google Scholar]
- 79.Zong P-A, Hanus R, Dylla M, Tang Y, Liao J, Zhang Q, Snyder GJ, Chen L. 2017. Skutterudite with graphene-modified grain-boundary complexion enhances zT enabling high-efficiency thermoelectric device. Energy Environ. Sci. 10, 183–191. ( 10.1039/C6EE02467J) [DOI] [Google Scholar]
- 80.Zhang Q, et al. 2017. Realizing high-performance thermoelectric power generation through grain boundary engineering of skutterudite-based nanocomposites. Nano Energy 41, 501–510. ( 10.1016/j.nanoen.2017.10.003) [DOI] [Google Scholar]
- 81.Bux SK, Yeung MT, Toberer ES, Snyder GJ, Kaner RB, Fleurial J-P. 2011. Mechanochemical synthesis and thermoelectric properties of high quality magnesium silicide. J. Mater. Chem. 21, 12 259–12 266. ( 10.1039/C1JM10827A) [DOI] [Google Scholar]
- 82.Kato D, Iwasaki K, Yoshino M, Yamada T, Nagasaki T. 2018. Control of Mg content and carrier concentration via post annealing under different Mg partial pressures for Sb-doped Mg2Si thermoelectric material. J. Solid State Chem. 258, 93–98. ( 10.1016/j.jssc.2017.10.003) [DOI] [Google Scholar]
- 83.Sakamoto T, Iida T, Kurosaki S, Yano K, Taguchi H, Nishio K, Takanashi Y. 2011. Thermoelectric behavior of Sb- and Al-doped n-type Mg2Si device under large temperature differences. J. Electron. Mater. 40, 629–634. ( 10.1007/s11664-010-1489-5) [DOI] [Google Scholar]
- 84.Inoue H, Yoneda S, Kato M, Ohsugi IJ, Kobayashi T. 2018. Examination of oxidation resistance of Mg2Si thermoelectric modules at practical operating temperature. J. Alloys Compd. 735, 828–832. ( 10.1016/j.jallcom.2017.11.202) [DOI] [Google Scholar]
- 85.de Boor J, Gloanec C, Kolb H, Sottong R, Ziolkowski P, Müller E.. 2015. Fabrication and characterization of nickel contacts for magnesium silicide based thermoelectric generators. J. Alloys Compd. 632, 348–353. ( 10.1016/j.jallcom.2015.01.149) [DOI] [Google Scholar]
- 86.Thimont Y, Lognoné Q, Goupil C, Gascoin F, Guilmeau E. 2014. Design of apparatus for Ni/Mg2Si and Ni/MnSi1.75 contact resistance determination for thermoelectric legs. J. Electron. Mater. 43, 2023–2028. ( 10.1007/s11664-013-2940-1) [DOI] [Google Scholar]
- 87.Park SH, Kim Y, Yoo C-Y. 2016. Oxidation suppression characteristics of the YSZ coating on Mg2Si thermoelectric legs. Ceram. Int. 42, 10 279–10 288. ( 10.1016/j.ceramint.2016.03.161) [DOI] [Google Scholar]
- 88.Wei T-R, Tan G, Zhang X, Wu C-F, Li J-F, Dravid VP, Snyder GJ, Kanatzidis MG. 2016. Distinct impact of alkali-ion doping on electrical transport properties of thermoelectric p-type polycrystalline SnSe. J. Am. Chem. Soc. 138, 8875–8882. ( 10.1021/jacs.6b04181) [DOI] [PubMed] [Google Scholar]
- 89.Chere EK, Zhang Q, Dahal K, Cao F, Mao J, Ren Z. 2016. Studies on thermoelectric figure of merit of Na-doped p-type polycrystalline SnSe. J. Mater. Chem. A 4, 1848–1854. ( 10.1039/C5TA08847J) [DOI] [Google Scholar]
- 90.Fu Y, et al. 2016. Enhanced thermoelectric performance in p-type polycrystalline SnSe benefiting from texture modulation. J. Mater. Chem. C 4, 1201–1207. ( 10.1039/C5TC03652F) [DOI] [Google Scholar]
- 91.Chang C, Tan G, He J, Kanatzidis MG, Zhao L-D. 2018. The thermoelectric properties of SnSe continue to surprise: extraordinary electron and phonon transport. Chem. Mater. 30, 7355–7367. ( 10.1021/acs.chemmater.8b03732) [DOI] [Google Scholar]
- 92.Lee YK, Luo Z, Cho SP, Kanatzidis MG, Chung I. 2019. Surface oxide removal for polycrystalline SnSe reveals near-single-crystal thermoelectric performance. Joule 3, 719–731. ( 10.1016/j.joule.2019.01.001) [DOI] [Google Scholar]
- 93.Park SH, Jin Y, Ahn K, Chung I, Yoo C-Y. 2017. Ag/Ni metallization bilayer: a functional layer for highly efficient polycrystalline SnSe thermoelectric modules. J. Electron. Mater. 46, 848–855. ( 10.1007/s11664-016-4972-9) [DOI] [Google Scholar]
- 94.Kim Y, Yoon G, Cho BJ, Park SH. 2017. Multi-layer metallization structure development for highly efficient polycrystalline SnSe thermoelectric devices. Appl. Sci. 7, 1116 ( 10.3390/app7111116) [DOI] [Google Scholar]
- 95.Chen S, Lukas KC, Liu W, Opeil CP, Chen G, Ren Z. 2013. Effect of Hf concentration on thermoelectric properties of nanostructured n-type half-Heusler materials HfxZr1–xNiSn0.99Sb0.01. Adv. Energy Mater. 3, 1210–1214. ( 10.1002/aenm.201300336) [DOI] [Google Scholar]
- 96.Fu C, Liu Y, Xie H, Liu X, Zhao X, Snyder GJ, Xie J, Zhu T. 2013. Electron and phonon transport in Co-doped FeV0.6Nb0.4Sb half-Heusler thermoelectric materials. J. Appl. Phys. 114, 134 905 ( 10.1063/1.4823859) [DOI] [Google Scholar]
- 97.Fu C, Zhu T, Liu Y, Xie H, Zhao X. 2015. Band engineering of high performance p-type FeNbSb based half-Heusler thermoelectric materials for figure of merit zT > 1. Energy Environ. Sci. 8, 216–220. ( 10.1039/C4EE03042G) [DOI] [Google Scholar]
- 98.Fu C, Zhu T, Pei Y, Xie H, Wang H, Snyder GJ, Liu Y, Liu Y, Zhao X. 2014. High band degeneracy contributes to high thermoelectric performance in p-type half-Heusler compounds. Adv. Energy Mater. 4, 1400600 ( 10.1002/aenm.201400600) [DOI] [Google Scholar]
- 99.Yu J, Fu C, Liu Y, Xia K, Aydemir U, Chasapis TC, Snyder GJ, Zhao X, Zhu T. 2018. Unique role of refractory Ta alloying in enhancing the figure of merit of NbFeSb thermoelectric materials. Adv. Energy Mater. 8, 1701313 ( 10.1002/aenm.201701313) [DOI] [Google Scholar]
- 100.Rogl G, et al. 2016. Mechanical properties of half-Heusler alloys. Acta Mater. 107, 178–195. ( 10.1016/j.actamat.2016.01.031) [DOI] [Google Scholar]
- 101.Zhu T, Fu C, Xie H, Liu Y, Zhao X. 2015. High efficiency half-Heusler thermoelectric materials for energy harvesting. Adv. Energy Mater. 5, 1500588 ( 10.1002/aenm.201500588) [DOI] [Google Scholar]
- 102.Ouardi S, et al. 2010. Electronic transport properties of electron- and hole-doped semiconducting C1b Heusler compounds: NiTi1−xMxSn (M = Sc, V). Phys. Rev. B 82, 085108 ( 10.1103/PhysRevB.82.085108) [DOI] [Google Scholar]
- 103.Graf T, Felser C, Parkin SS. P. 2011. Simple rules for the understanding of Heusler compounds. Prog. Solid State Chem. 39, 1–50. ( 10.1016/j.progsolidstchem.2011.02.001) [DOI] [Google Scholar]
- 104.He R, Huang L, Wang Y, Samsonidze G, Kozinsky B, Zhang Q, Ren Z. 2016. Enhanced thermoelectric properties of n-type NbCoSn half-Heusler by improving phase purity. APL Mater. 4, 104804 ( 10.1063/1.4952994) [DOI] [Google Scholar]
- 105.Mao J, Zhou J, Zhu H, Liu Z, Zhang H, He R, Chen G, Ren Z. 2017. Thermoelectric properties of n-type ZrNiPb-based half-Heuslers. Chem. Mater. 29, 867–872. ( 10.1021/acs.chemmater.6b04898) [DOI] [Google Scholar]
- 106.Minmin Z, Jing-Feng L, Peijun G, Takuji K. 2010. Synthesis and thermoelectric properties of fine-grained FeVSb system half-Heusler compound polycrystals with high phase purity. J. Phys. D: Appl. Phys. 43, 415403 ( 10.1088/0022-3727/43/41/415403) [DOI] [Google Scholar]
- 107.Xie W, Jin Q, Tang X. 2008. The preparation and thermoelectric properties of Ti0.5Zr0.25Hf0.25Co1−xNixSb half-Heusler compounds. J. Appl. Phys. 103, 043711 ( 10.1063/1.2885113) [DOI] [Google Scholar]
- 108.Yan X, et al. 2011. Enhanced thermoelectric figure of merit of p-type half-Heuslers. Nano Lett. 11, 556–560. ( 10.1021/nl104138t) [DOI] [PubMed] [Google Scholar]
- 109.Yan X, et al. 2012. Stronger phonon scattering by larger differences in atomic mass and size in p-type half-Heuslers Hf1−xTixCoSb0.8Sn0.2. Energy Environ. Sci. 5, 7543–7548. ( 10.1039/C2EE21554C) [DOI] [Google Scholar]
- 110.Yu C, Zhu T-J, Shi R-Z, Zhang Y, Zhao X-B, He J. 2009. High-performance half-Heusler thermoelectric materials Hf1−xZrxNiSn1−ySby prepared by levitation melting and spark plasma sintering. Acta Mater. 57, 2757–2764. ( 10.1016/j.actamat.2009.02.026) [DOI] [Google Scholar]
- 111.Zhou M, Chen L, Feng C, Wang D, Li J-F. 2007. Moderate-temperature thermoelectric properties of TiCoSb-based half-Heusler compounds Ti1−xTaxCoSb. J. Appl. Phys. 101, 113714 ( 10.1063/1.2738460) [DOI] [Google Scholar]
- 112.Zou M, Li J-F, Kita T. 2013. Thermoelectric properties of fine-grained FeVSb half-Heusler alloys tuned to p-type by substituting vanadium with titanium. J. Solid State Chem. 198, 125–130. ( 10.1016/j.jssc.2012.09.043) [DOI] [Google Scholar]
- 113.Zhang H, Wang Y, Huang L, Chen S, Dahal H, Wang D, Ren Z. 2016. Synthesis and thermoelectric properties of n-type half-Heusler compound VCoSb with valence electron count of 19. J. Alloys Compd. 654, 321–326. ( 10.1016/j.jallcom.2015.09.082) [DOI] [Google Scholar]
- 114.He R, Zhu H, Sun J, Mao J, Reith H, Chen S, Schierning G, Nielsch K, Ren Z. 2017. Improved thermoelectric performance of n-type half-Heusler MCo1−xNixSb (M = Hf, Zr). Materials Today Physics 1, 24–30. ( 10.1016/j.mtphys.2017.05.002) [DOI] [Google Scholar]
- 115.Schwall M, Balke B. 2013. Phase separation as a key to a thermoelectric high efficiency. Phys. Chem. Chem. Phys. 15, 1868–1872. ( 10.1039/C2CP43946H) [DOI] [PubMed] [Google Scholar]
- 116.Huang L, He R, Chen S, Zhang H, Dahal K, Zhou H, Wang H, Zhang Q, Ren Z. 2015. A new n-type half-Heusler thermoelectric material NbCoSb. Mater. Res. Bull. 70, 773–778. ( 10.1016/j.materresbull.2015.06.022) [DOI] [Google Scholar]
- 117.Kimura Y, Tanoguchi T, Kita T. 2010. Vacancy site occupation by Co and Ir in half-Heusler ZrNiSn and conversion of the thermoelectric properties from n-type to p-type. Acta Mater. 58, 4354–4361. ( 10.1016/j.actamat.2010.04.028) [DOI] [Google Scholar]
- 118.Yan X, Liu W, Chen S, Wang H, Zhang Q, Chen G, Ren Z. 2013. Thermoelectric property study of nanostructured p-type half-Heuslers (Hf, Zr, Ti)CoSb0.8Sn0.2. Adv. Energy Mater. 3, 1195–1200. ( 10.1002/aenm.201200973) [DOI] [Google Scholar]
- 119.Rausch E, Balke B, Stahlhofen JM, Ouardi S, Burkhardt U, Felser C. 2015. Fine tuning of thermoelectric performance in phase-separated half-Heusler compounds. J. Mater. Chem. C 3, 10409–10 414. ( 10.1039/C5TC01196E) [DOI] [Google Scholar]
- 120.Guanghe L, Ken K, Yuji O, Hiroaki M, Shinsuke Y. 2013. High temperature thermoelectric properties of half-Heusler compound PtYSb. Jpn. J. Appl. Phys. 52, 041804 ( 10.7567/JJAP.52.041804) [DOI] [Google Scholar]
- 121.Kawano K, Kurosaki K, Muta H, Yamanaka S. 2008. Substitution effect on the thermoelectric properties of p-type half-Heusler compounds: ErNi1−xPdxSb. J. Appl. Phys. 104, 013714 ( 10.1063/1.2956699) [DOI] [Google Scholar]
- 122.Kimura Y, Zama A. 2006. Thermoelectric properties of p-type half-Heusler compound HfPtSn and improvement for high-performance by Ir and Co additions. Appl. Phys. Lett. 89, 172110 ( 10.1063/1.2364721) [DOI] [Google Scholar]
- 123.Culp SR, Simonson JW, Poon SJ, Ponnambalam V, Edwards J, Tritt TM. 2008. (Zr,Hf)Co(Sb,Sn) half-Heusler phases as high-temperature (>700°C) p-type thermoelectric materials. Appl. Phys. Lett. 93, 022105 ( 10.1063/1.2959103) [DOI] [Google Scholar]
- 124.Qiu P, Huang X, Chen X, Chen L. 2009. Enhanced thermoelectric performance by the combination of alloying and doping in TiCoSb-based half-Heusler compounds. J. Appl. Phys. 106, 103703 ( 10.1063/1.3238363) [DOI] [Google Scholar]
- 125.Takeyuki S, Ken K, Hiroaki M, Shinsuke Y. 2007. High-thermoelectric figure of merit realized in p-type half-Heusler compounds: ZrCoSnxSb1−x. Jan. J. Appl. Phys. 46, L673 ( 10.1143/JJAP.46.L673) [DOI] [Google Scholar]
- 126.Gürth M, Rogl G, Romaka VV, Grytsiv A, Bauer E, Rogl P. 2016. Thermoelectric high ZT half-Heusler alloys Ti1−x−yZrxHfyNiSn (0 ≤ x ≤ 1; 0 ≤ y ≤ 1). Acta Mater. 104, 210–222. ( 10.1016/j.actamat.2015.11.022) [DOI] [Google Scholar]
- 127.Culp SR, Poon SJ, Hickman N, Tritt TM, Blumm J. 2006. Effect of substitutions on the thermoelectric figure of merit of half-Heusler phases at 800 °C. Appl. Phys. Lett. 88, 042106 ( 10.1063/1.2168019) [DOI] [Google Scholar]
- 128.Joshi G, et al. 2014. NbFeSb-based p-type half-Heuslers for power generation applications. Energy Environ. Sci. 7, 4070–4076. ( 10.1039/C4EE02180K) [DOI] [Google Scholar]
- 129.Kim SW, Kimura Y, Mishima Y. 2004. Enhancement of high temperature thermoelectric properties of intermetallic compounds based on a skutterudite IrSb3 and a half-Heusler TiNiSb. Sci. Technol. Adv. Mater. 5, 485–489. ( 10.1016/j.stam.2004.02.006) [DOI] [Google Scholar]
- 130.Junjie Y, Kaiyang X, Xinbing Z, Tiejun Z. 2018. High performance p-type half-Heusler thermoelectric materials. J. Phys. D: Appl. Phys. 51, 113001 ( 10.1088/1361-6463/aaaa58) [DOI] [Google Scholar]
- 131.Bhattacharya S, Skove MJ, Russell M, Tritt TM, Xia Y, Ponnambalam V, Poon SJ, Thadhani N. 2008. Effect of boundary scattering on the thermal conductivity of TiNiSn-based half-Heusler alloys. Phys. Rev. B 77, 184203 ( 10.1103/PhysRevB.77.184203) [DOI] [Google Scholar]
- 132.Xing Y, Liu R, Sun Y-Y, Chen F, Zhao K, Zhu T, Bai S, Chen L. 2018. Self-propagation high-temperature synthesis of half-Heusler thermoelectric materials: reaction mechanism and applicability. J. Mater. Chem. A 6, 19 470–19 478. ( 10.1039/C8TA07411A) [DOI] [Google Scholar]
- 133.Hu T, Yang D, Su X, Yan Y, You Y, Liu W, Uher C, Tang X. 2018. Interpreting the combustion process for high-performance ZrNiSn thermoelectric materials. ACS Appl. Mater. Interfaces 10, 864–872. ( 10.1021/acsami.7b15273) [DOI] [PubMed] [Google Scholar]
- 134.Yan Y, Geng W, Qiu J, Ke H, Luo C, Yang J, Uher C, Tang X. 2018. Thermoelectric properties of n-type ZrNiSn prepared by rapid non-equilibrium laser processing. RSC Adv. 8, 15 796–15 803. ( 10.1039/C8RA00992A) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Liu H, et al. 2012. Copper ion liquid-like thermoelectrics. Nat. Mater. 11, 422 ( 10.1038/nmat3273) [DOI] [PubMed] [Google Scholar]
- 136.Brown DR, Day T, Caillat T, Snyder GJ. 2013. Chemical stability of (Ag,Cu)2Se: a historical overview. J. Electron. Mater. 42, 2014–2019. ( 10.1007/s11664-013-2506-2) [DOI] [Google Scholar]
- 137.Dennler G, Chmielowski R, Jacob S, Capet F, Roussel P, Zastrow S, Nielsch K, Opahle I, Madsen GK-H. 2014. Are binary copper sulfides/selenides really new and promising thermoelectric materials? Adv. Energy Mater. 4, 1301581 ( 10.1002/aenm.201301581) [DOI] [Google Scholar]
- 138.Qiu P, et al. 2018. Suppression of atom motion and metal deposition in mixed ionic electronic conductors. Nat. Commun. 9, 2910 ( 10.1038/s41467-018-05248-8) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139.Kanatzidis MG. 2010. Nanostructured thermoelectrics: the new paradigm? Chem. Mater. 22, 648–659. ( 10.1021/cm902195j) [DOI] [Google Scholar]
- 140.Zebarjadi M, Esfarjani K, Dresselhaus MS, Ren ZF, Chen G. 2012. Perspectives on thermoelectrics: from fundamentals to device applications. Energy Environ. Sci. 5, 5147–5162. ( 10.1039/C1EE02497C) [DOI] [Google Scholar]
- 141.Sootsman JR, Chung DY, Kanatzidis MG. 2009. New and old concepts in thermoelectric materials. Angew. Chem. Int. Ed. 48, 8616–8639. ( 10.1002/anie.200900598) [DOI] [PubMed] [Google Scholar]
- 142.Vineis CJ, Shakouri A, Majumdar A, Kanatzidis MG. 2010. Nanostructured thermoelectrics: big efficiency gains from small features. Adv. Mater. 22, 3970–3980. ( 10.1002/adma.201000839) [DOI] [PubMed] [Google Scholar]
- 143.Shi X, Chen L, Uher C. 2016. Recent advances in high-performance bulk thermoelectric materials. Int. Mater. Rev. 61, 379–415. ( 10.1080/09506608.2016.1183075) [DOI] [Google Scholar]
- 144.Zeier WG, Zevalkink A, Gibbs ZM, Hautier G, Kanatzidis MG, Snyder GJ. 2016. Thinking like a chemist: intuition in thermoelectric materials. Angew. Chem. Int. Ed. 55, 6826–6841. ( 10.1002/anie.201508381) [DOI] [PubMed] [Google Scholar]
- 145.Jood P, Ohta M. 2015. Hierarchical architecturing for layered thermoelectric sulfides and chalcogenides. Materials 8, 1124 ( 10.3390/ma8031124) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.He J, Tritt TM. 2017. Advances in thermoelectric materials research: looking back and moving forward. Science 357, eaak9997 ( 10.1126/science.aak9997) [DOI] [PubMed] [Google Scholar]
- 147.Yang L, Chen Z-G, Dargusch MS, Zou J. 2018. High performance thermoelectric materials: progress and their applications. Adv. Energy Mater. 8, 1 701 797 ( 10.1002/aenm.201701797) [DOI] [Google Scholar]
- 148.Suekuni K, Takabatake T. 2016. Research update: Cu–S based synthetic minerals as efficient thermoelectric materials at medium temperatures. APL Mater. 4, 104503 ( 10.1063/1.4955398) [DOI] [Google Scholar]
- 149.Rowe DM. 1995. Conversion efficiency and figure-of-merit. In CRC handbook of thermoelectrics (ed. Rowe DM.), pp. 19–25. Boca Raton: CRC press. [Google Scholar]
- 150.Kim HS, Liu W, Ren Z. 2017. The bridge between the materials and devices of thermoelectric power generators. Energy Environ. Sci. 10, 69–85. ( 10.1039/C6EE02488B) [DOI] [Google Scholar]
- 151.Snyder GJ, Snyder AH. 2017. Figure of merit ZT of a thermoelectric device defined from materials properties. Energy Environ. Sci. 10, 2280–2283. ( 10.1039/C7EE02007D) [DOI] [Google Scholar]
- 152.Zhang QH, Huang XY, Bai SQ, Shi X, Uher C, Chen LD. 2016. Thermoelectric devices for power generation: recent progress and future challenges. Adv. Enger Mater. 18, 194–213. ( 10.1002/adem.201500333) [DOI] [Google Scholar]
- 153.He R, Schierning G, Nielsch K. 2018. Thermoelectric devices: a review of devices, architectures, and contact optimization. Adv. Mater. Technol. 3, 1700256 ( 10.1002/admt.201700256) [DOI] [Google Scholar]
- 154.Jood P, Ohta M, Yamamoto A, Kanatzidis MG. 2018. Excessively doped PbTe with Ge-induced nanostructures enables high-efficiency thermoelectric modules. Joule 2, 1339–1355. ( 10.1016/j.joule.2018.04.025) [DOI] [Google Scholar]
- 155.Kaibe H, et al. 2005. Development of thermoelectric generating stacked modules aiming for 15% of conversion efficiency. In ICT 2005. 24th International Conference on Thermoelectrics, 242–247. ( 10.1109/ICT.2005.1519929) [DOI]
- 156.Guo JQ, Geng HY, Ochi T, Suzuki S, Kikuchi M, Yamaguchi Y, Ito S. 2012. Development of skutterudite thermoelectric materials and modules. J. Electron. Mater. 41, 1036–1042. ( 10.1007/s11664-012-1958-0) [DOI] [Google Scholar]
- 157.Salvador JR, et al. 2014. Conversion efficiency of skutterudite-based thermoelectric modules. Phys. Chem. Chem. Phys. 16, 12 510–12 520. ( 10.1039/C4CP01582G) [DOI] [PubMed] [Google Scholar]
- 158.Bartholomé K, Balke B, Zuckermann D, Köhne M, Müller M, Tarantik K, König J. 2014. Thermoelectric modules based on half-Heusler materials produced in large quantities. J. Electron. Mater. 43, 1775–1781. ( 10.1007/s11664-013-2863-x) [DOI] [Google Scholar]
- 159.Ohta M, Jood P, Murata M, Lee C-H, Yamamoto A, Obara H. In press. An integrated approach to thermoelectrics: combining phonon dynamics, nanoengineering, novel materials development, module fabrication, and metrology. Adv. Energy Mater. 1801304 ( 10.1002/aenm.201801304) [DOI] [Google Scholar]
- 160.D'Angelo J, Case ED, Matchanov N, Wu C-I, Hogan TP, Barnard J, Cauchy C, Hendricks T, Kanatzidis MG. 2011. Electrical, thermal, and mechanical characterization of novel segmented-leg thermoelectric modules. J. Electron. Mater. 40, 2051–2062. ( 10.1007/s11664-011-1717-7) [DOI] [Google Scholar]
- 161.Caillat T, Fleurial J, Snyder GJ, Borshchevsky A. 2001. Development of high efficiency segmented thermoelectric unicouples. In Proc. ICT2001 20 Int. Conf. on Thermoelectrics, pp. 282–285. ( 10.1109/ICT.2001.979888) [DOI]
- 162.Zhang Q, et al. 2017. Realizing a thermoelectric conversion efficiency of 12% in bismuth telluride/skutterudite segmented modules through full-parameter optimization and energy-loss minimized integration. Energy Environ. Sci. 10, 956–963. ( 10.1039/C7EE00447H) [DOI] [Google Scholar]
- 163.Muto A, Yang J, Poudel B, Ren Z, Chen G. 2013. Skutterudite unicouple characterization for energy harvesting applications. Adv. Energy Mater. 3, 245–251. ( 10.1002/aenm.201200503) [DOI] [Google Scholar]
- 164.Zhu H, et al. 2018. Discovery of ZrCoBi based half Heuslers with high thermoelectric conversion efficiency. Nat. Commun. 9, 2497 ( 10.1038/s41467-018-04958-3) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Zhu H, et al. 2019. Discovery of TaFeSb-based half-Heuslers with high thermoelectric performance. Nat. Commun. 10, 270 ( 10.1038/s41467-018-08223-5) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Zhu Q, Song S, Zhu H, Ren Z. 2019. Realizing high conversion efficiency of Mg3Sb2-based thermoelectric materials. J. Power Sources 414, 393–400. ( 10.1016/j.jpowsour.2019.01.022) [DOI] [Google Scholar]
- 167.Chetty R, Kikuchi Y, Bouyrie Y, Jood P, Yamamoto A, Suekuni K, Ohta M. 2019. Power generation from the Cu26Nb2Ge6S32-based single thermoelectric element with Au diffusion barrier. J. Mater. Chem. C 7, 5184–5192. ( 10.1039/C9TC00868C) [DOI] [Google Scholar]
- 168.LaLonde AD, Pei Y, Wang H, Jeffrey Snyder G. 2011. Lead telluride alloy thermoelectrics. Mater. Today 14, 526–532. ( 10.1016/S1369-7021(11)70278-4) [DOI] [Google Scholar]
- 169.Abelson RD. 2005. Space missions and applications. In Thermoelectrics handbook: macro to nano (ed. DM Rowe), pp. 56-1–56-29 Boca Raton, FL: CRC press. [Google Scholar]
- 170.Ohta M, Biswas K, Lo S-H, He J, Chung DY, Dravid VP, Kanatzidis MG. 2012. Enhancement of thermoelectric figure of merit by the insertion of MgTe nanostructures in p-type PbTe doped with Na2Te. Adv. Energy Mater. 2, 1117–1123. ( 10.1002/aenm.201100756) [DOI] [Google Scholar]
- 171.Zhao L-D, Dravid VP, Kanatzidis MG. 2014. The panoscopic approach to high performance thermoelectrics. Energy Environ. Sci. 7, 251–268. ( 10.1039/C3EE43099E) [DOI] [Google Scholar]
- 172.Pei Y, Wang H, Snyder GJ. 2012. Band engineering of thermoelectric materials. Adv. Mater. 24, 6125–6135. ( 10.1002/adma.201202919) [DOI] [PubMed] [Google Scholar]
- 173.Jood P, Ohta M, Kunii M, Hu X, Nishiate H, Yamamoto A, Kanatzidis MG. 2015. Enhanced average thermoelectric figure of merit of n-type PbTe1−xIx–MgTe. J. Mater. Chem. C 3, 10 401–10 408. ( 10.1039/C5TC01652E) [DOI] [Google Scholar]
- 174.Eaksuwanchai P, Tanusilp S-A, Jood P, Ohta M, Kurosaki K. 2018. Increased Seebeck coefficient and decreased lattice thermal conductivity in grain-size-controlled p-type PbTe–MgTe system. ACS Appl. Energy Mater. 1, 6586–6592. ( 10.1021/acsaem.8b01491) [DOI] [Google Scholar]
- 175.Elsner NB. 2011. Review of lead-telluride bonding concepts. MRS Proceedings 234, 167–177. ( 10.1557/PROC-234-167) [DOI] [Google Scholar]
- 176.Hori Y, Ito T. 2006. Fabrication of 500°C Class Thermoelectric Module and Evaluation of its High Temperature Stability. In 2006 25th international conference on thermoelectrics. pp. 642–645. ( 10.1109/ICT.2006.331223) [DOI]
- 177.Singh A, Bhattacharya S, Thinaharan C, Aswal DK, Gupta SK, Yakhmi JV, Bhanumurthy K. 2008. Development of low resistance electrical contacts for thermoelectric devices based on n-type PbTe and p-type TAGS-85 ((AgSbTe2)0.15(GeTe)0.85). J. Phys. D: Appl. Phys. 42, 015502 ( 10.1088/0022-3727/42/1/015502) [DOI] [Google Scholar]
- 178.Xia H, Drymiotis F, Chen C-L, Wu A, Snyder GJ. 2014. Bonding and interfacial reaction between Ni foil and n-type PbTe thermoelectric materials for thermoelectric module applications. J. Mater. Sci. 49, 1716–1723. ( 10.1007/s10853-013-7857-9) [DOI] [Google Scholar]
- 179.Xia H, Chen C-L, Drymiotis F, Wu A, Chen Y-Y, Snyder GJ. 2014. Interfacial reaction between Nb foil and n-type PbTe thermoelectric materials during thermoelectric contact fabrication. J. Electron. Mater. 43, 4064–4069. ( 10.1007/s11664-014-3350-8) [DOI] [Google Scholar]
- 180.Salvador JR, et al. 2013. Thermal to electrical energy conversion of skutterudite-based thermoelectric modules. J. Electron. Mater. 42, 1389–1399. ( 10.1007/s11664-012-2261-9) [DOI] [Google Scholar]
- 181.Ochi T, Nie G, Suzuki S, Kikuchi M, Ito S, Guo JQ. 2014. Power-generation performance and durability of a skutterudite thermoelectric generator. J. Electron. Mater. 43, 2344–2347. ( 10.1007/s11664-014-3060-2) [DOI] [Google Scholar]
- 182.He R, et al. 2016. Achieving high power factor and output power density in p-type half-Heuslers Nb1−xTixFeSb. Proc. Natl Acad. Sci. USA 113, 13 576–13 581. ( 10.1073/pnas.1617663113) [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Kauzlarich SM, Brown SR, Jeffrey Snyder G. 2007. Zintl phases for thermoelectric devices. Dalton Trans. 2099–2107. ( 10.1039/B702266B) [DOI] [PubMed] [Google Scholar]
- 184.Brown SR, Kauzlarich SM, Gascoin F, Snyder GJ. 2006. Yb14MnSb11: new high efficiency thermoelectric material for power generation. Chem. Mater. 18, 1873–1877. ( 10.1021/cm060261t) [DOI] [Google Scholar]
- 185.Paik J-A, Brandon E, Caillat T, Ewell R, Fleurial J-P. 2011. Life testing of Yb14MnSb11 for high performance thermoelectric couples. Proc. Nucl. Emerg. Technol. Space, 616–622. [Google Scholar]
- 186.Kim Y, Jin Y, Yoon G, Chung I, Yoon H, Yoo C-Y, Park SH. 2019. Electrical characteristics and detailed interfacial structures of Ag/Ni metallization on polycrystalline thermoelectric SnSe. J. Mater. Sci. Technol. 35, 711–718. ( 10.1016/j.jmst.2018.11.020) [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
This article has no additional data.









