Next Article in Journal
Combined Effect of a Microporous Layer and Type I Collagen Coating on a Biphasic Calcium Phosphate Scaffold for Bone Tissue Engineering
Next Article in Special Issue
Chemical Potential Tuning and Enhancement of Thermoelectric Properties in Indium Selenides
Previous Article in Journal
Corrosion Behavior of Detonation Gun Sprayed Fe-Al Type Intermetallic Coating
Previous Article in Special Issue
Structure and Transport Properties of the BiCuSeO-BiCuSO Solid Solution
 
 
Correction published on 21 September 2015, see Materials 2015, 8(9), 6482-6483.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hierarchical Architecturing for Layered Thermoelectric Sulfides and Chalcogenides

Energy Technology Research Institute, National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba, Ibaraki 305-8568, Japan
*
Author to whom correspondence should be addressed.
Materials 2015, 8(3), 1124-1149; https://doi.org/10.3390/ma8031124
Submission received: 4 February 2015 / Revised: 25 February 2015 / Accepted: 26 February 2015 / Published: 16 March 2015
(This article belongs to the Special Issue Low-Dimensional Anisotropic Thermoelectrics)

Abstract

:
Sulfides are promising candidates for environment-friendly and cost-effective thermoelectric materials. In this article, we review the recent progress in all-length-scale hierarchical architecturing for sulfides and chalcogenides, highlighting the key strategies used to enhance their thermoelectric performance. We primarily focus on TiS2-based layered sulfides, misfit layered sulfides, homologous chalcogenides, accordion-like layered Sn chalcogenides, and thermoelectric minerals. CS2 sulfurization is an appropriate method for preparing sulfide thermoelectric materials. At the atomic scale, the intercalation of guest atoms/layers into host crystal layers, crystal-structural evolution enabled by the homologous series, and low-energy atomic vibration effectively scatter phonons, resulting in a reduced lattice thermal conductivity. At the nanoscale, stacking faults further reduce the lattice thermal conductivity. At the microscale, the highly oriented microtexture allows high carrier mobility in the in-plane direction, leading to a high thermoelectric power factor.

Graphical Abstract

1. Introduction

The growing demand for energy throughout the world is causing an energy crisis and aggravating environmental burden. However, more than 50% of the primary energy supplied is not utilized and is wasted in the form of heat. Solid-state devices based on thermoelectrics can directly convert the waste heat generated from various sources into useful electrical energy and can thus provide a new approach to improving energy management and sustainability while reducing greenhouse-gas emissions [1,2,3,4,5]. Currently, thermoelectric devices are found only in niche applications because thermoelectric materials show low efficiency and mainly consist of toxic and scarce (expensive) elements. Therefore, extensive efforts have been devoted to develop high-efficiency, environment-friendly, and cost-effective thermoelectric materials for large-scale practical applications, such as waste-heat recovery in vehicles and industries.
The thermoelectric figure of merit (ZT) expresses the efficiency of thermoelectric materials and is defined as ZT = (S2/ρκtotal)T, where S, ρ, κtotal, and T are the Seebeck coefficient, electrical resistivity, total thermal conductivity, and absolute temperature, respectively. An increase in the quantity S2/ρ, which is known as the thermoelectric power factor, leads to an increase in the electrical performance. The S2/ρ is typically optimized by tuning the carrier concentration of materials. The κtotal is the sum of two parts: the charge carriers transporting heat (i.e., electronic thermal conductivity, κel) and phonons traveling through the lattice (i.e., lattice thermal conductivity, κlat). Therefore, κtotal = κel + κlat. The former is directly related to ρ and can be estimated using the Wiedemann-Franz law: κel = LT/ρ, where L is the Lorenz number. Therefore, one way to reduce the κtotal is to minimize κlat, which is a carrier-independent parameter. In order to increase S2/ρ and reduce κlat, thermoelectric materials have been strategically developed with the phonon glass-electron crystal (PGEC) concept proposed by Slack [6]. A PGEC system consists of an electron crystal region and a phonon glass region. The electron crystal region allows carrier transmission, resulting in high carrier mobility and high S2/ρ. The phonon glass region effectively scatters phonons, resulting in low κlat. The PGEC concept has been realized using three methods: nanoblock integration, a nano- and meso-structuring/panoscopic approach, and rattling/low-energy atomic vibration.
Layered cobaltite oxides such as NaxCoO and Ca–Co–O have been demonstrated to be promising high-temperature thermoelectric materials [7,8,9,10,11,12,13]. An important advantage of thermoelectric oxides is their high chemical stability in air over a wide temperature range of 300–1200 K. The high ZT of these oxides originates from the integration of physical properties of the metallic electron-crystal nanoblock layer and disordered phonon-glass nanoblock layer. While the metallic Co–O nanoblock layers provide a high S2/ρ, the disordered Na and Ca–Co–O nanoblock layers scatter phonons and reduce κlat [10,11,12,13]. The impact of layered oxides has generated considerable interest in layered compounds, particularly layered sulfides and chalcogenides. The covalency generally increases when moving from oxides to sulfides and chalcogenides; therefore, the layered sulfides and chalcogenides show lower ρ and higher S2/ρ [14].
PbTe-based materials represent one of the most successful examples of nano- and meso-structuring [15,16,17,18,19,20,21]. The insertion of endotaxial nanostructures in PbTe bulk materials causes the effective scattering of short-mean-free-path phonons without affecting the carrier mobility. Moreover, long-mean-free-path phonons are scattered at grain boundaries by controlling and fine-tuning the mesoscale architecture of nanostructured materials. As a result, the all-length-scale hierarchical architecture (panoscopic approach) in PbTe bulk thermoelectric materials causes a significant reduction in κlat and dramatically increases ZT to an exceptionally high value of 2.2 at 915 K [17].
Skutterudites and clathrates are highly promising for practical thermoelectric devices operating in the intermediate temperature range of 573–973 K [22,23,24,25,26,27,28]. These systems contain large cages of host atoms in their crystal structures. Filling the large cages with guest atoms reduces κlat because the guest atoms rattle at low energy inside the cages and effectively scatter phonons. A very high ZT of 1.9 at 835 K was achieved in the skutterudite Sr0.09Ba0.11Yb0.05Co4Sb12 [27], and a conversion efficiency of 7% was demonstrated in a skutterudite-based thermoelectric module [28].
This review discusses recent progress in an intriguing new class of thermoelectric materials, layered chalcogenides. Among the layered chalcogenides, we primarily focus on the sulfide systems because sulfur is an environment-friendly and cost-effective element. Figure 1 shows the crystal structures of the thermoelectric sulfides and chalcogenides addressed in this article, including layered sulfide TiS2, misfit layered sulfide [LaS]1.14NbS2, a member of cannizzarite homologous series Pb5Bi6Se14, accordion-like layered selenide SnSe, and thermoelectric mineral Cu12Sb4S13. The nanoblock integration, nano- and meso-structuring/panoscopic approach, and rattling/low-energy atomic vibration to layered sulfides and chalcogenides are important guidance for ZT enhancement through all-length-scale hierarchical architecture.
Figure 1. Crystal structures of the thermoelectric sulfides and chalcogenides addressed in this article: (a) TiS2 [29,30]; (b) [LaS]1.14NbS2 [31,32]; (c) Pb5Bi6Se14 [33,34]; (d) SnSe [35]; and (e) Cu12Sb4S13 [36,37]. Sizes of atoms in this figure are arbitrary.
Figure 1. Crystal structures of the thermoelectric sulfides and chalcogenides addressed in this article: (a) TiS2 [29,30]; (b) [LaS]1.14NbS2 [31,32]; (c) Pb5Bi6Se14 [33,34]; (d) SnSe [35]; and (e) Cu12Sb4S13 [36,37]. Sizes of atoms in this figure are arbitrary.
Materials 08 01124 g001

2. CS2 Sulfurization and Pressurized Sintering

We first discuss the synthesis and processing of thermoelectric sulfides, which drastically affects their properties. In the case of TiS2-based layered sulfides, a single crystal can be grown using the chemical-vapor-transport method with I2 as the transport agent [38,39,40,41]. Polycrystalline TiS2-based layered sulfides have been commonly prepared by melting stoichiometric amounts of the constituent elements in evacuated and sealed quartz tubes at 905–1273 K, followed by pressurized sintering at 873–1173 K [30,42,43,44,45,46,47]. However, these methods are difficult to apply because of the large difference in vapor pressure between metals and sulfur; consequently, there is less motivation for researchers in the thermoelectric community to investigate thermoelectric sulfides further. In this review, we focused on CS2 sulfurization, which is a low-temperature preparation method for thermoelectric sulfides [29,48,49,50,51,52,53].
The temperature dependence of the standard free-energy change (ΔG°) for the sulfurization of TiO2, La2O3, CeO2, and Pr6O11 with CS2 and H2S gases, calculated from thermodynamic data [54,55], is shown in Figure 2 [29,51,52]. The value of ΔG° for CS2 sulfurization is lower than that for H2S sulfurization in the temperature range of 600–1300 K. For example, the values of ΔG° at 1100 K for the sulfurization of La2O3 with CS2 and H2S are −126 and −26 kJ·mol−1, respectively, which implies that CS2 gas is a powerful sulfurizing agent for oxides, allowing the low-temperature formation of sulfides.
Figure 2. Temperature dependence of the standard free-energy change (ΔG°) for the sulfurization of TiO2 [29], La2O3, CeO2 [51], and Pr6O11 [52] with CS2 and H2S.
Figure 2. Temperature dependence of the standard free-energy change (ΔG°) for the sulfurization of TiO2 [29], La2O3, CeO2 [51], and Pr6O11 [52] with CS2 and H2S.
Materials 08 01124 g002
A schematic diagram of the experimental apparatus used for CS2 sulfurization is shown in Figure 3. The apparatus was referenced previously in a report by Hirai et al. [51]. The starting oxide powders were placed in quartz boats and set in a quartz reaction tube. The quartz tube was evacuated and purged with Ar gas. The powders were heated to 573 K under Ar gas flow. A mixture of CS2 and Ar gases was introduced into the quartz tube as soon as the powders reached 573 K. The CS2 gas was obtained by bubbling Ar carrier gas through CS2 liquid. Single phases of TiS2, [LaS]1.20CrS2, and [LaS]1.14NbS2 were successfully prepared through CS2 sulfurization at 1073 K for 4–12 h [29,32].
Powders of layered thermoelectric sulfides and chalcogenides were sintered under uniaxial pressure to obtain high-density compacts. Figure 4 shows scanning electron microscopy (SEM) micrographs of the fractured section of sintered compacts of Ti1.008S2 [29], [LaS]1.20CrS2 [32], and Pb5Bi6Se14 [34] that were fractured parallel to the pressure direction. The micrographs reveal a dense structure and well-organized microtexture, indicating that the pressurized sintering promotes densification and grain orientation. The XRD patterns show that the crystalline c-axis is preferentially oriented along the pressing (out-of-plane) direction [29,32,34]. A centrifugal heating technique also provides the dense samples with highly oriented microtexture [56]. The highly oriented microtextures allow high carrier mobility in the in-plane direction, leading to a high thermoelectric power factor.
Figure 3. Schematic diagram of the experimental apparatus used for CS2 sulfurization. The apparatus was referenced previously in a report by Hirai et al. [51].
Figure 3. Schematic diagram of the experimental apparatus used for CS2 sulfurization. The apparatus was referenced previously in a report by Hirai et al. [51].
Materials 08 01124 g003
Figure 4. Scanning electron micrographs of the fractured section of sintered compacts of (a) Ti1.008S2 [29]; (b) [LaS]1.20CrS2 [32]; and (c) Pb5Bi6Se14 [34].
Figure 4. Scanning electron micrographs of the fractured section of sintered compacts of (a) Ti1.008S2 [29]; (b) [LaS]1.20CrS2 [32]; and (c) Pb5Bi6Se14 [34].
Materials 08 01124 g004

3. TiS2-Based Layered Sulfides

Because TiS2-based layered sulfides are primarily composed of earth-abundant, low-toxicity, and light elements, they will pave the way for environment-friendly, cost-effective, and lightweight thermoelectric devices [29,30,39,40,41,42,43,44,45,46,47,56,57,58,59,60,61,62,63,64,65,66]. The two-dimensional crystal structure of these systems is a great example in which the thermoelectric performance can be improved through nanoblock integration and hierarchical architecturing. The recently developed TiS2-based sulfides with high thermoelectric figure of merit (ZT) are summarized in Figure 5. In 2001, Imai et al. [40] demonstrated a high power factor (S2/ρ) of ~3710 μW·K−2·m−1 at 300 K in a single crystal of near-stoichiometric TiS2, which is comparable to that of commercially available thermoelectric materials such as Bi2Te3. However, the single crystal shows a high lattice thermal conductivity κlat of ~6.35 W·K−1·m−1, which limits the ZT to ~0.16 at 300 K (Table 1). Recently, the intercalation of guest atoms and guest layers into TiS2 layers was found to reduce the κlat of these systems, similar to layered cobaltite oxides [7,8,9,10,11,12,13]. For example, Cu intercalation reduces κlat and enhances ZT to ~0.5 at 823 K [44].
Figure 5. Current state-of-the-art TiS2-based layered sulfides. Temperature dependence of the thermoelectric figure of merit (ZT) in the in-plane (ab-plane) direction for single-crystal TiS2 [40] and polycrystalline Ti1.008S2 [29], Ti1.025S2 [30], Cu0.1TiS2 [44], TiS1.5Se0.5 [45], Ti0.95Ta0.05S2 [46], [BiS]1.2[TiS2]2 [42], [PbS]1.18[TiS2]2 [42], [SnS]1.2[TiS2]2 [43], and [BiS]1.2[Ti0.95Cr0.05S2]2 [66].
Figure 5. Current state-of-the-art TiS2-based layered sulfides. Temperature dependence of the thermoelectric figure of merit (ZT) in the in-plane (ab-plane) direction for single-crystal TiS2 [40] and polycrystalline Ti1.008S2 [29], Ti1.025S2 [30], Cu0.1TiS2 [44], TiS1.5Se0.5 [45], Ti0.95Ta0.05S2 [46], [BiS]1.2[TiS2]2 [42], [PbS]1.18[TiS2]2 [42], [SnS]1.2[TiS2]2 [43], and [BiS]1.2[Ti0.95Cr0.05S2]2 [66].
Materials 08 01124 g005
Table 1. Seebeck coefficient (S), electrical resistivity (ρ), carrier mobility (μ), power factor (S2/ρ), lattice thermal conductivity (κlat), and thermoelectric figure of merit (ZT) at room temperature in the in-plane (ab-plane) and out-of-plane (c-axis) directions for a single crystal of nearly stoichiometric TiS2 [40] and polycrystalline Ti1.008S2 [29].
Table 1. Seebeck coefficient (S), electrical resistivity (ρ), carrier mobility (μ), power factor (S2/ρ), lattice thermal conductivity (κlat), and thermoelectric figure of merit (ZT) at room temperature in the in-plane (ab-plane) and out-of-plane (c-axis) directions for a single crystal of nearly stoichiometric TiS2 [40] and polycrystalline Ti1.008S2 [29].
Sample DirectionS (μV·K1)ρ (μΩ·m)μ (cm2·V1·s1)S2/ρ (μW·K−2·m−1)κlat (W·K−1·m−1)ZT
Single crystalIn-plane−251171537106.350.16
Single crystalOut-of-plane-13,0000.017-4.21-
PolycrystallineIn-plane−806.22.310301.50.12
PolycrystallineOut-of-plane−84111.26301.30.10
The crystal structure of TiS2 consists of a stacked CdI2-type layer along the crystalline c-axis and is shown in Figure 1a. The TiS2 phase has a wide range of chemical compositions ranging from stoichiometric to Ti-rich: Ti1+xS2 with 0 ≤ x < 0.1 [67,68,69]. Because the TiS2 layer is weakly stacked through van der Waals forces, the excess Ti atoms occupy the space between TiS2 layers. The lattice parameter a monotonically increases from 0.3405 nm for Ti1.001S2 to 0.3412 nm for Ti1.093S2, and c increases from 0.5691 nm for Ti1.001S2 to 0.5711 nm for Ti1.093S2 [70].
To interpret the thermoelectric properties of TiS2, we used the following formulas for approximations of electrical resistivity ρ, carrier mobility μ, κlat, and Seebeck coefficient S [71,72,73]:
1 ρ = n e μ
μ = e τ e m *
κ lat = 1 3 C v v a l p
S = 8 π 2 k B 2 3 e h 2 m * T ( π 3 n ) 2 / 3
where n, e, τe, m*, Cv, va, lp, kB, and h are the carrier concentration, electronic charge, relaxation time of charge carriers, effective mass, heat capacity at constant pressure, average sound velocity, mean free path of phonons, Boltzmann constant, and Planck’s constant, respectively. These relationships are derived from the parabolic band model and energy-independent scattering time for metals and degenerate semiconductors.
The layered crystal structure results in highly anisotropic electrical and thermal transport properties [29,40,43]. Table 1 lists S, ρ, μ, S2/ρ, κlat, and ZT at room temperature in the in-plane (ab-plane) and out-of-plane (c-axis) directions for a single crystal of nearly stoichiometric TiS2 [40] and polycrystalline Ti1.008S2 [29]. The polycrystalline sample possesses a well-organized microtexture (Figure 4a) with the crystalline ab-plane preferentially oriented along the in-plane direction. The room-temperature n of the single crystal and polycrystalline samples are ~2.8 × 1020 cm3 and ~4.5 × 1021 cm3, respectively. For single crystal, the out-of-plane ρ (~13,000 μΩ·m) is nearly 750 times higher than the in-plane ρ (~17 μΩ·m), and the out-of-plane μ (~0.017 cm2·V1·s1) is significantly lower than the in-plane μ (~15 cm2·V1·s1) at 300 K. Moreover, the out-of-plane κlat (~4.21 W K−1·m−1) is slightly lower than the in-plane κlat (~6.35 W·K−1·m−1) at 300 K. The highly oriented polycrystalline sample also exhibits lower ρ, higher μ, and higher κlat in the in-plane direction. The high value of the out-of-plane ρ and low value of the out-of-plane κlat are principally due to the reduced relaxation time (reduced mean free path) resulting from the significant electron and phonon scatterings, respectively, at the interfaces between the TiS2 layers (see Equations (1)–(3)). The sign of S is negative for both samples, confirming n-type carrier transport. Unlike ρ and κlat, S was found to be insensitive to the crystal orientation in the sintered compacts; as there is no relationship between S and τ in Equation (1), S is nearly isotropic. The in-plane S (~−80 μV·K1) is nearly the same as the out-of-plane S (~−84 μV·K1) at 300 K. The higher S2/ρ (~1030 μW·K−2·m−1) and higher ZT (~0.12) in the in-plane direction at 300 K for polycrystalline Ti1.008S2 is due to the lower in-plane ρ and insensitivity of S to the crystal orientation.
The intercalated Ti provides n-type carriers to the system; therefore, n can be tuned to optimize S2/ρ by self-intercalation. Figure 6 shows the room-temperature S, ρ, and S2/ρ plotted against n. The solid line in Figure 6a represents the values calculated using Equation (4) with m*/m0 (m0 is the free electron mass) equal to 2.88 when n is greater than 5 × 1020 cm3 [74,75,76]. The measured values fall on this calculated line. An increase in Ti content results in an increase in n and, hence, a decrease in S and ρ. The S2/ρ peaks around n ~ 2.8 × 1021 cm3 in Ti1+xS2 and reaches a value of ~3700 μW·K−2·m−1.
Figure 6. (a) Seebeck coefficient (S); (b) electrical resistivity (ρ); and (c) power factor (S2/ρ) plotted against the carrier concentration (n) at 300 K in the in-plane (ab-plane) direction for Ti1+xS2 [29,30,41,43,44,75,76,77]. The solid line in (a) represents the values calculated using Equation (4) with m*/m0 (m0 is the free electron mass) equal to 2.88 when n is greater than 5 × 1020 cm3 [74,75,76].
Figure 6. (a) Seebeck coefficient (S); (b) electrical resistivity (ρ); and (c) power factor (S2/ρ) plotted against the carrier concentration (n) at 300 K in the in-plane (ab-plane) direction for Ti1+xS2 [29,30,41,43,44,75,76,77]. The solid line in (a) represents the values calculated using Equation (4) with m*/m0 (m0 is the free electron mass) equal to 2.88 when n is greater than 5 × 1020 cm3 [74,75,76].
Materials 08 01124 g006
One way to reduce κlat is the intercalation of guest atoms and guest layers into the host TiS2 layer [42,43,44,65,66]. For example, Cu intercalation reduces κlat from ~2.0 W·K−1·m−1 for TiS2 to ~1.7 W·K−1·m−1 for Cu0.1TiS2, and SnS intercalation dramatically reduces κlat to ~1.0 W·K−1·m−1 for [SnS]1.2[TiS2]2 at 300 K (Figure 7a). Substitution is another way to reduce κlat [45,46,47]; for example, the substitution of Ti by Ta slightly reduces κlat to ~1.8 W·K−1·m−1 for Ti0.95Ta0.05S2 at 300 K (Figure 7b).
Figure 7. Temperature dependence of the lattice thermal conductivity (κlat) in the in-plane (ab-plane) direction for Ti1.008S2 [29]: (a) intercalated systems [BiS]1.2[TiS2]2 [42], [PbS]1.18[TiS2]2 [42], [SnS]1.2[TiS2]2 [43], and Cu0.1TiS2 [44]; (b) substituted systems TiS1.5Se0.5 [45] and Ti0.95Ta0.05S2 [46].
Figure 7. Temperature dependence of the lattice thermal conductivity (κlat) in the in-plane (ab-plane) direction for Ti1.008S2 [29]: (a) intercalated systems [BiS]1.2[TiS2]2 [42], [PbS]1.18[TiS2]2 [42], [SnS]1.2[TiS2]2 [43], and Cu0.1TiS2 [44]; (b) substituted systems TiS1.5Se0.5 [45] and Ti0.95Ta0.05S2 [46].
Materials 08 01124 g007
The reduction in κlat through intercalation and substitution is limited; therefore, ZT is below 0.5. Future work should aim to develop new methods for a dramatic reduction in κlat to achieve ZT enhancement in TiS2-based layered sulfides. Recently, Koumoto et al. [64,66] have demonstrated that the stacking faults in the layers can be tuned to reduce the thermal conductivity. We believe that well-organized stacking faults can ideally reduce κlat and enhance ZT beyond 0.5.

4. Misfit Layered Sulfides [MS]1+m[TS2]n

TiS2-based layered sulfides are a part of a large family of misfit layered sulfides with a general formula [MS]1+m[TS2]n, where M = Sn, Pb, Sb, Bi, and rare-earth metals; T = Ti, V, Cr, Nb, and Ta; and n = 1, 2, 3 [78]. These compounds consist of alternating MS and TS2 layers, which belong to separate subgroups; therefore, these compounds lack three-dimensional periodicity. The crystal-structure analysis of rare-earth misfit compound (MS)1+mNbS2 has been performed using different subgroups including the (3+1)-dimensional superspace group [31,79,80], and different structural variants depending upon the intercalation layer were proposed. Miyazaki et al. [81] have reported a significant variation in the c-axis depending on the ionic radius of the rare-earth element. Previous results have shown that NbS2 [32,53,81,82], TaS2 [83,84], and TiS2 [42,43,65,66] sandwich structures show metallic conductivity, whereas CrS2 [32,85] and VS2 [78,85] sandwich structures show semiconducting properties. The crystal structure and physical properties of this misfit family were extensively studied in the 1990s [78]; however, not many reports on their thermoelectric properties exist to date. In recent years, the misfit layered sulfides have gained interest in the thermoelectric community for their PGEC behavior, where the intercalated NaCl-type MS layer is primarily responsible for the disorder (reduction in κlat) and the CdI2-type TS2 layer behaves as a charge-carrier pathway [86,87]. The entire crystal structure is stabilized through charge transfer from the MS layer to the TS2 layer [78,88]. Through the concept of bond valences, it is established that the interlayer La–S bond valences are much larger than the interlayer M-S bond valences with M = Sn or Bi [88,89,90], which makes the LnS-based misfit layered compounds (Ln = rare-earth metals) much more mechanically hard (due to strong interlayer bonding) than the SnS-, PbS-, BiS-, and SbS-based systems [88].
Thus far, only [MS]1+mNbS2 with M = rare-earth elements [32,53,81]; [LaS]1.2CrS2 [32]; and [MS]1+mTiS2 with M = Sn, Pb, or Bi [42,43,64,65,66] (discussed in the previous section) have been studied in depth for their thermoelectric properties. The thermoelectric properties of [MS]1+mTS2 (M = La, Yb; T = Cr, Nb) in the in-plane (ab-plane) and out-of-plane (c-axis) directions are summarized in Table 2. The NbS2 systems show p-type S, whereas CrS2 systems show n-type S. Among all compounds in the LnS-based series, [Yb2S2]0.62NbS2 shows the highest ZT of ~0.10 at 300 K because it shows the highest S2/ρ [81]. The S of the Yb sample is ~35%–60% higher than those of samples consisting of the other elements in the lanthanide series at 300 K. This is due to its slightly lower n (~5 × 1020 cm−3), the reason for which is not very well understood and requires further study. At 300 K, a very low κlat ~0.41 W·K−1·m−1 is achievable for [Yb2S2]0.62NbS2, which reduces even further (~0.18 W·K−1·m−1) with the removal of ~5% Yb and S from the [Yb2S2] layer, slightly enhancing the room-temperature ZT to ~0.11. This is due to the high phonon scattering originating from the highly modulated structure and atomic deficiency. Further improvements in ZT are expected at higher temperatures.
Table 2. Seebeck coefficient (S); electrical resistivity (ρ); total thermal conductivity (κtotal); lattice thermal conductivity (κlat); power factor (S2/ρ); and thermoelectric figure of merit (ZT) in the in-plane (ab-plane) and out-of-plane (c-axis) directions of state-of-the-art misfit layered sulfides: [MS]1+mTS2 (M = La, Yb; T = Cr, Nb) [32,81].
Table 2. Seebeck coefficient (S); electrical resistivity (ρ); total thermal conductivity (κtotal); lattice thermal conductivity (κlat); power factor (S2/ρ); and thermoelectric figure of merit (ZT) in the in-plane (ab-plane) and out-of-plane (c-axis) directions of state-of-the-art misfit layered sulfides: [MS]1+mTS2 (M = La, Yb; T = Cr, Nb) [32,81].
SampleDirectionT (K)ρ (μΩ·m)S (μV·K−1)κtotal (W·K−1·m−1)κlat (W·K−1·m−1)S2/ρ (μW·K−2·m−1)ZTReference
(Yb2S2)0.62NbS2In-plane30019.0600.800.412000.1[81]
(La2S2)0.62NbS2In-plane30011.522--50-[81]
(LaS)1.14NbS2 aIn-plane3007.6372.501.501770.02[32]
95022.0832.000.933160.15
Out-of-plane30013.3252.041.48490.01
95032.1721.620.881620.09
(LaS)1.14NbS2 bIn-plane3005.2354.883.452330.02[32]
95016.9833.251.864050.15
Out-of-plane3009.3251.560.75700.01
95028.5561.340.521110.09
(LaS)1.2CrS2 aIn-plane950207−1721.161.041430.12[32]
Out-of-plane950223−1741.020.911370.13
(LaS)1.2CrS2 bIn-plane950171−1721.251.111740.13[32]
Out-of-plane950278−1540.920.84840.08
a Small grains (~1 μm), weak/random orientation of grains; b Large grains (>20 μm), strong orientation of grains perpendicular to the pressing direction.
As seen from Table 2, misfit compounds are highly sensitive to microstructural variations because of the anisotropic nature of their atomic bonds. Therefore, one strategy to enhance ZT is tuning the microstructure. Microstructure control was achieved in (LaS)1+mTS2 (T = Nb, Cr) [32] by varying the sulfurization duration from 6 to 12 h, followed by pressure-assisted sintering to obtain samples with randomly and highly oriented textures. Extended sulfurization (for 12 h) resulted in higher carbon content in the samples, which impeded grain growth [91]. On the other hand, the grains in the sample sulfurized for 6 h grew to >20 μm and self-arranged into a layered structure after sintering (Figure 4b). Large anisotropy was observed in thermoelectric properties for both the systems (Table 2). Both systems show low total thermal conductivity κtotal, especially in the out-of-plane direction, and the lowest κtotal was observed for highly oriented (LaS)1.2CrS2 (~0.92 W·K−1·m−1) and (LaS)1.14NbS2 (~1.34 W·K−1·m−1), which mainly results from the very low κlat ~0.84 W·K−1·m−1 and ~0.52 W·K−1·m−1, respectively, at 950 K. The low κlat is attributed to the fact that the interfaces between the host TS2 and guest MS layers act as effective phonon scatterers. Similar to ρ and κlat, S was found to be highly anisotropic for the highly oriented NbS2 misfit compounds. For instance, the in-plane and out-of-plane S values were ~83 μV·K1 and ~56 μV·K1, respectively, for (LaS)1.14NbS2 at 950 K. The anisotropy in S may have originated from the anisotropic band structure. Band-structure calculations showed that strong intralayer energy dispersions (~1.0 eV) occur near the Fermi level, while smaller dispersions of 0.01–0.05 eV occur in the interlayer direction [79,88].
Defects such as the presence of extra bright contrast planes are observed in the HRTEM image (indicated by white arrows) and ED patterns (streaks along the c-axis) of the CrS2 system (Figure 8), which exhibits coherent stacking-fault-induced intergrowth of CrS2 layers [32]. With proper manipulation, these crystal defects can help improve the thermoelectric properties. These nanoscale defects, along with the phonon scattering at the atomic scale (layer interfaces, dopants, and point defects) and micro scale (microstructure control) make these compounds good candidates for ZT enhancement through all-length-scale hierarchical architecturing [17,18,20].
Figure 8. High-resolution transmission electron microscopy (HRTEM) image of (LaS)1.2CrS2 and corresponding electron diffraction pattern. The structural model superimposed on the enlarged HRTEM image is given as inset in the bottom-left corner of the figure. La, Cr, and S are represented by blue, red, and yellow, respectively. White arrows indicate stacking faults/defects (taken with permission from reference [32]).
Figure 8. High-resolution transmission electron microscopy (HRTEM) image of (LaS)1.2CrS2 and corresponding electron diffraction pattern. The structural model superimposed on the enlarged HRTEM image is given as inset in the bottom-left corner of the figure. La, Cr, and S are represented by blue, red, and yellow, respectively. White arrows indicate stacking faults/defects (taken with permission from reference [32]).
Materials 08 01124 g008
The highly oriented texture provided the highest ZT ~0.14 at 950 K in (LaS)1.2CrS2, while the sample with weakly oriented texture had the highest ZT ~0.15 at 950 K among the (LaS)1.14NbS2 samples [32]. Carrier mobility (μ) is quite low in these misfit compounds with μ ~ 0.5–1.6 cm2·V1·s1 for the CrS2 system and μ ~ 7–11 cm2·V1·s1 for the NbS2 system at 300 K. The carrier concentrations at 300 K are n ~ 3 × 1020 cm3 for the CrS2 system and n ~ 1 × 1021 cm3 for the NbS2 system; these values need to be optimized further for enhancing S2/ρ. The ZT can, therefore, be further improved through appropriate doping and substitutions not only to optimize n but also to tune the stacking faults, as reported for the TiS2 misfit system [64,66]. These misfit layered sulfides are promising candidates for high-temperature thermoelectric applications as they are environmentally benign, nontoxic, and stable at high temperature; further, they provide tremendous opportunities for enhancing ZT.

5. Homologous Chalcogenides

A structure of the homologous series is built on the same structural principle with certain module(s) expanding in various dimension(s) by regular increments [92,93,94,95,96]. For example, the crystal structure of the cannizzarite homologous series consists of alternating infinite PbSe- and Bi2Se3-type layers [33,95,97], as shown in Figure 1c and Figure 9. The two layers are stacked alternately along the c-axis, resulting in a three-dimensional structure with varying thicknesses that form different members in the homology. The thermoelectric properties of homologous compounds can be tuned by modifying the size and shape of the structural module(s); therefore, the homologous compounds are good platforms for developing new thermoelectric materials [34,94,98,99,100,101,102,103,104,105,106,107,108,109].
Figure 9. Structural evolution in the cannizzarite homologous series [(PbSe)5]m[(Bi2Se3)3]n: (a) Pb5Bi6Se14 (m = 1, n = 1); (b) Pb5Bi12Se23 (m = 1, n = 2); (c) Pb5Bi18Se32 (m = 1, n = 3); and (d) Pb5Bi24Se41 (m = 1, n = 4) [33,95,97]. The sizes of atoms in this figure are arbitrary.
Figure 9. Structural evolution in the cannizzarite homologous series [(PbSe)5]m[(Bi2Se3)3]n: (a) Pb5Bi6Se14 (m = 1, n = 1); (b) Pb5Bi12Se23 (m = 1, n = 2); (c) Pb5Bi18Se32 (m = 1, n = 3); and (d) Pb5Bi24Se41 (m = 1, n = 4) [33,95,97]. The sizes of atoms in this figure are arbitrary.
Materials 08 01124 g009
Figure 10 shows the κlat, ZT, and crystal structure of CsBi4Te6, a member of the Cs4[Bi2n+4Te3n+6] homologous series [99,102]; SnBi2Te4, a member of the [SnTe]n[Bi2Te3]m homologous series [106,107]; Pb5Bi6Se14, a member of the cannizzarite homologous series [34]; PbBi2S4, a member of the galenobismuthite homologous series [34]; and Pb7Bi4Se13, a member of lillianite homologous series [109]. CsBi4Te6 and SnBi2Te4 show p-type behavior, while Pb5Bi6Se14, PbBi2S4, and Pb7Bi4Se13 show n-type behavior. The complex crystal structures yield low κlat in the range of 0.15 to 1.15 W·K−1·m−1 over the temperature range of 100 K to 763 K, leading to high ZT. A record-high ZT of ~0.8 at 225 K and a high ZT of ~0.9 at 775 K were found in CsBi4Te6 and Pb7Bi4Se13, respectively.
As shown in Figure 4c, the scanning electron microscope images of the fractured section of sintered compacts of Pb5Bi6Se14 reveal that the crystal grains were preferentially grown in the direction perpendicular to the pressure applied during sintering, forming needle-like grains with a mean length of ~30 μm. The XRD patterns show that the crystalline c-axis is preferentially oriented along the pressing (out-of-plane) direction [34]. Figure 11 shows the temperature dependence of the S, ρ, κtotal, κlat, and ZT for sintered compacts of the cannizzarite homologous series member Pb5Bi6Se14 in the in-plane and out-of-plane directions. The room-temperature carrier concentration of the system is ~4.8 × 1019 cm−3. As in TiS2, S was observed to be insensitive to the crystal orientation (Figure 11a); for example, at 705 K, the in-plane S value (~210 μV·K−1) is in rough agreement with the out-of-plane S value (~230 μV·K−1). On the other hand, the interfaces between the PbSe layer and the Bi2Se3 layer effectively scatter the charge carriers and heat-carrying phonons, leading to highly anisotropic ρ and κlat. For example, at 705 K, the in-plane ρ value (~130 μΩ·m) is 60% lower than the out-of-plane ρ value (~320 μΩ·m) (Figure 11b), and the in-plane κlat value (~0.39 W·K−1·m−1) is slightly higher than the out-of-plane κlat value (~0.31 W·K−1·m−1) (Figure 11c). The lower ρ in the in-plane direction and the insensitivity of S to crystal orientation results in a higher S2/ρ of 320 μW·K−2·m−1 and higher ZT of 0.46 in the in-plane direction at 705 K (Figure 11d).
Figure 10. Temperature dependence of the (a) lattice thermal conductivity (κlat); (b) thermoelectric figure of merit (ZT); and (c) crystal structure of CsBi4Te6, a member of the Cs4[Bi2n+4Te3n+6] homologous series [99,102]; SnBi2Te4, a member of the [SnTe]n[Bi2Te3]m homologous series [106,107]; Pb5Bi6Se14, a member of the cannizzarite homologous series [34]; PbBi2S4, a member of the galenobismuthite homologous series [34]; and Pb7Bi4Se13, a member of the lillianite homologous series [109].
Figure 10. Temperature dependence of the (a) lattice thermal conductivity (κlat); (b) thermoelectric figure of merit (ZT); and (c) crystal structure of CsBi4Te6, a member of the Cs4[Bi2n+4Te3n+6] homologous series [99,102]; SnBi2Te4, a member of the [SnTe]n[Bi2Te3]m homologous series [106,107]; Pb5Bi6Se14, a member of the cannizzarite homologous series [34]; PbBi2S4, a member of the galenobismuthite homologous series [34]; and Pb7Bi4Se13, a member of the lillianite homologous series [109].
Materials 08 01124 g010
Figure 11. Temperature dependence of the (a) Seebeck coefficient (S), (b) electrical resistivity (ρ), (c) total thermal conductivity (κtotal) and lattice thermal conductivity (κlat), and (d) thermoelectric figure of merit (ZT) for sintered compacts of the cannizzarite homologous member Pb5Bi6Se14 in the in-plane (ab-plane) and out-of-plane (c-axis) directions [34].
Figure 11. Temperature dependence of the (a) Seebeck coefficient (S), (b) electrical resistivity (ρ), (c) total thermal conductivity (κtotal) and lattice thermal conductivity (κlat), and (d) thermoelectric figure of merit (ZT) for sintered compacts of the cannizzarite homologous member Pb5Bi6Se14 in the in-plane (ab-plane) and out-of-plane (c-axis) directions [34].
Materials 08 01124 g011aMaterials 08 01124 g011b
One way to enhance the ZT of homologous compounds is the tuning of carrier concentration n through doping for optimizing S2/ρ. Another way is to reduce κlat further by increasing the complexity of the crystal structure using crystal-structural evolution enabled by the homologous series (Figure 9). Table 3 lists the room-temperature thermoelectric properties of three members of the cannizzarite homologous series, Pb5Bi6Se14, Pb5Bi12Se23, and Pb5Bi18Se32 [103]. An important result here is that κlat systematically reduces from 0.72 W·K−1·m−1 for Pb5Bi6Se14 (m = 1, n = 1) to 0.49 W·K−1·m−1 for Pb5Bi18Se32 (m = 1, n = 3), demonstrating the reduction in κlat through crystal-structural evolution at the atomic scale.
Table 3. Seebeck coefficient (S), electrical resistivity (ρ), carrier concentration (n), carrier mobility (μ), power factor (S2/ρ), lattice thermal conductivity (κlat), and thermoelectric figure of merit (ZT) at room temperature in three members of the cannizzarite homologous series, Pb5Bi6Se14, Pb5Bi12Se23, and Pb5Bi18Se32 [103].
Table 3. Seebeck coefficient (S), electrical resistivity (ρ), carrier concentration (n), carrier mobility (μ), power factor (S2/ρ), lattice thermal conductivity (κlat), and thermoelectric figure of merit (ZT) at room temperature in three members of the cannizzarite homologous series, Pb5Bi6Se14, Pb5Bi12Se23, and Pb5Bi18Se32 [103].
SampleS (μV·K1)ρ (μΩ·m)n (cm3)μ (cm2·V1·s1)S2/ρ (μW·K−2·m−1)κlat (W·K−1·m−1)ZT
Pb5Bi6Se14−28220.86 × 102033360.720.01
Pb5Bi12Se23−27271.15 × 102020270.590.01
Pb5Bi18Se32−52391.19 × 102014690.490.03

6. Accordion-Like Layered SnQ (Q = S, Se)

The major strategies for achieving good thermoelectric performance have mostly involved the use of heavy elements (often toxic elements such as Pb), nanostructuring (which requires precise control of synthesis procedures) [15,16,17,18,19,20,21], and complex unit cells such as skutterudites [22,23,24,25,26,27] and zintl phases [110]. SnSe, however, possesses a very simple crystal structure with light, earth-abundant elements (Figure 1d) and exhibits an intrinsically ultra-low thermal conductivity [35]. SnSe has a layered orthorhombic crystal structure with zigzag (accordion-like) atomic chains (space group Pnma) below ~750 K [111]; at ~750 K, it undergoes a phase transition towards its higher-symmetry phase (space group Cmcm) [112,113]. SnSe, which has been ignored by the thermoelectric community in the past, has attracted tremendous interest after the very recent report by Kanatzidis et al. [35], which demonstrates a record-high ZT of ~2.62 at 923 K along the b-axis, high ZT of ~2.3 along the c-axis, and moderate ZT of ~0.8 along the a-axis in single-crystal SnSe. The lighter analogous sulfide SnS has also received attention in the thermoelectric community [114,115,116].
Figure 12 shows a comparison of the thermoelectric properties of single-crystal [35], polycrystalline [117], and Ag-alloyed SnSe [118]. There are no significant differences observed S between the a-, b-, and c-axis for single-crystal SnSe. S decreases at the transition temperature of ~750 K but retains a high value (Figure 12a). For example, the b-axis S for single crystals decreases from ~570 μV·K−1 at 573 K to ~340 μV·K−1 at 823 K. Unlike S, a large anisotropy in ρ can be seen for the single crystal, where ρ along the b- and c-axis is much smaller than that along the a-axis (Figure 12b). This is because the carrier mobilities μ ~ 250 cm2·V1·s1 in the b-axis direction and μ ~ 130 cm2·V1·s1 in the c-axis direction at 300 K are higher within the plane of the slab than in the inter-slab direction. The in-plane and out-of-plane ρ of the polycrystalline samples yielded similar observations. The value of ρ dramatically decreases at ~750 K and consequently increases S2/ρ. In the case of single crystals, the b-axis ρ decreases from ~2000 μΩ·m at 573 K to ~120 μΩ·m at 823 K. With Ag alloying of polycrystalline SnSe, the room-temperature carrier concentration increases from 2 × 1017 cm−3 for a pure system to 9 × 1018 cm3 for a 7% Ag system [118]. The 1% Ag-alloyed SnSe system produces the highest ZT of ~0.6 at 750 K, in contrast to the pure SnSe system, which has a ZT of ~0.3 at the same temperature and ~0.5 at 820 K (Figure 12d).
Figure 12. Temperature dependence of the (a) Seebeck coefficient (S); (b) electrical resistivity (ρ); (c) total thermal conductivity (κtotal); and (d) thermoelectric figure of merit (ZT) for single-crystal (SC) SnSe measured along the a-, b-, and c-axis [35], polycrystalline (PC) SnSe in the in-plane (crystalline bc-plane) and out-of-plane (crystalline a-axis) directions [117], and 1% Ag-alloyed polycrystalline SnSe in the in-plane direction [118].
Figure 12. Temperature dependence of the (a) Seebeck coefficient (S); (b) electrical resistivity (ρ); (c) total thermal conductivity (κtotal); and (d) thermoelectric figure of merit (ZT) for single-crystal (SC) SnSe measured along the a-, b-, and c-axis [35], polycrystalline (PC) SnSe in the in-plane (crystalline bc-plane) and out-of-plane (crystalline a-axis) directions [117], and 1% Ag-alloyed polycrystalline SnSe in the in-plane direction [118].
Materials 08 01124 g012
The performance of SnSe surpasses that of other materials, especially above the transition temperature of ~750 K, where the single-crystal κlat less than 0.25 W·K−1·m−1 is achieved along all crystallographic axes (Figure 12c) [35]. This value is significantly lower than the κlat obtained for all-length-scale hierarchically architectured PbTe (PbTe-4 mol% SrTe-2 mol% Na has κlat ~ 0.5 W·K−1·m−1 and ZT ~ 2.2 at 900 K) [17] and skutterudites exhibiting rattling (for example, LaFe3CoSb12 has κlat ~ 1 W·K−1·m−1 and ZT ~ 0.9 at 750 K) [22].
One would expect the κtotallat) of polycrystalline samples to be less than that of single-crystal samples because of increased phonon scattering from grain boundaries in polycrystalline samples; however, the opposite is observed for SnSe (Figure 12c). For instance, in-plane κtotal values of ~0.6 W·K−1·m−1 (at 850 K) and ~0.7 W·K−1·m−1 (at 750 K) were obtained for pure polycrystalline and Ag-alloyed SnSe, respectively. However, very low κtotal values of ~0.23 W·K−1·m−1 (a-axis), ~0.34 W·K−1·m−1 (b-axis), and ~0.29 W·K−1·m−1 (c-axis) were obtained for the single-crystal sample at 973 K. Ab initio studies suggest a strong anisotropy between all the crystallographic axes with the conductivity in the b-axis direction (κlat ~ 1.4 W·K−1·m−1) higher than that in the c-axis direction (κlat ~ 0.7 W·K−1·m−1) at 300 K [119]. This theoretical study relates quite well with the polycrystalline samples but differs markedly from the values for single-crystal samples, in which κlat values along b and c axis are reported to be similar (~0.7 W·K−1·m−1 at 300 K). As data for only two directions (i.e., parallel and perpendicular to the pressing direction) are available for polycrystalline samples, it is difficult to draw conclusions. Furthermore, the room-temperature κtotal of the single crystal differs markedly from the earlier reported work on SnSe system [120] and therefore, further verification is warranted. Very recent investigations on Te-substituted SnSe [121] and non-stoichiometric SnSe [118] have indicated that n-type conductivity in SnSe is feasible.
The lighter analog SnS is also gaining increased interest, backed up by theoretical work [122,123] supporting the good thermoelectric properties of this compound. SnS also has an accordion-like layered structure and is environmentally compatible and cost-effective. Among the few reports on the thermoelectric properties of SnS [114,115,116], the Ag-doped system demonstrates the highest ZT of ~0.6 at 973 K [116]. There is a big scope for improvement in ZT through improved doping and an all-length-scale hierarchical architecture approach.

7. Thermoelectric Minerals

The thermoelectric minerals tetrahedrites (Cu12Sb4S13) and colusites (Cu26V2M6S32; M = Ge, Sn) include no layers in their crystal structures. However, the p-type ZT demonstrated in these mineral-based sulfide systems is among the highest achieved in Pb-free thermoelectric sulfides (Figure 13) and is therefore worth mentioning [36,37,124,125,126].
The terms “tetrahedrite” and “colusite” are derived from naturally occurring minerals (Cu,Fe,Ag,Zn)12Sb4S13 and Cu24–26V2(As,Sn,Sb)6S32, respectively, which are mainly composed of the earth-abundant and low-toxicity elements Cu and S. The tetrahedrites possess a cubic structure of I 4 ¯ 3 m symmetry with SbS3 pyramids, CuS4 tetrahedra, and CuS3 triangles. The colusites crystallize in a cubic structure of P 4 ¯ 3 n symmetry with CuS4 tetrahedra and VS4 tetrahedra.
Figure 13. Temperature dependence of (a) thermoelectric figure of merit (ZT) (p-type); and (b) lattice thermal conductivity (κlat) for the tetrahedrites Cu10.5Ni1.5Sb4S13 [36], Cu11.5Zn0.5Sb4S13 [124], Cu11.6Mn0.4Sb4S13 [125], and Cu10.5Ni1.0Zn0.5Sb4S13 [37] and the colusite Cu26V2Ge6S32 [126].
Figure 13. Temperature dependence of (a) thermoelectric figure of merit (ZT) (p-type); and (b) lattice thermal conductivity (κlat) for the tetrahedrites Cu10.5Ni1.5Sb4S13 [36], Cu11.5Zn0.5Sb4S13 [124], Cu11.6Mn0.4Sb4S13 [125], and Cu10.5Ni1.0Zn0.5Sb4S13 [37] and the colusite Cu26V2Ge6S32 [126].
Materials 08 01124 g013
In 2012, Suekuni et al. [127] first reported a notable S2/ρ and low κlat at room temperature in tetrahedrite systems. Since then, many efforts have been devoted to enhance the ZT of tetrahedrite systems [36,37,124,125] and to find new mineral-based thermoelectric materials [126,128]. As shown in Figure 13, mineral-based sulfide systems show extremely low κlat, resulting in high ZT. In the case of the tetrahedrites, the low κlat is due to the low-energy vibration of the Cu atom out of the [CuS3] trigonal planar unit, which appears similar to the rattling in skutterudites and clathrates (Figure 1e) [36]. The κlat of both systems ranges from ~0.1 W·K−1·m−1 to ~0.7 W·K−1·m−1 over the temperature range of 300 to 730 K (Figure 13b). The value of S2/ρ can be increased by tuning the carrier concentration by substituting Tr = Mn [125,127], Fe [124,127], Co [127], Ni [36,37,127], and Zn [37,124,127] at the Cu sites in Cu12−xTrxSb4S13. For example, the value of S2/ρ at 660 K was slightly boosted through Ni doping from 1170 μW·K−2·m−1 for Cu12Sb4S13 to 1210 μW·K−2·m−1 for Cu11.5Ni0.5Sb4S13. A ZT of 1.03 at 723 K was achieved for Cu10.5Ni1.0Zn0.5Sb4S13 (Figure 13a) [37]. Recently, the structural stability and purity of synthetic Cu12Sb4S13 and Cu10.4Ni1.6Sb4S13 were also reported [129]. These high-ZT sulfide minerals may help realize feasible methods for environment-friendly and cost-effective thermoelectric waste heat recovery.

8. Conclusions and Insights

The decoupling of the interdependent thermoelectric properties of a material has always been a challenge in the development of high-efficiency thermoelectric materials. A layered structure not only provides vast opportunities for enhancing the thermoelectric figure of merit ZT through appropriate optimizations at all length scales but also provides methods for individual tuning of thermal and electron transports. Ultra-low lattice thermal conductivities below 0.5 W·K−1·m−1 have been obtained through intercalation, crystal-structural evolution, or the formation of stacking faults in layered sulfides. Future work must consider scattering centers on all relevant length scales in a hierarchical manner, i.e., from atomic-scale lattice disorders (layer interfaces, substitution sites, and structural evolution) to nano-scale disorders (stacking faults) and microscale grain boundaries (through microtexturing), to achieve the maximum reduction in lattice thermal conductivity. Higher ZT is expected with further work to optimize the carrier concentrations in this new category of materials.

Acknowledgments

The authors thank Atsushi Yamamoto (AIST), Shinji Hirai (Muroran IT), Toshihiro Kuzuya (Muroran IT), Oleg I. Lebedev (CRISMAT), David Berthebaud (CRISMAT), Duck Young Chung (ANL), Mercouri G. Kanatzidis (Northwestern University and ANL), Mikio Koyano (JAIST), Koichiro Suekuni (Hiroshima University), and Toshiro Takabatake (Hiroshima University) for helpful discussions and fruitful collaborations. Financial support from the Japan-U.S. Cooperation Project for Research and Standardization of Clean Energy Technologies funded by the Ministry of Economy, Trade and Industry (METI) and JSPS KAKENHI Grant Number 25420699 is gratefully acknowledged.

Author Contributions

Michihiro Ohta designed the work. Priyanka Jood and Michihiro Ohta wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kajikawa, T. Thermoelectric power generation system recovering industrial waste heat. In Thermoelectrics Handbook: Macro to Nano; Rowe, D.M., Ed.; CRC Press: Boca Raton, FL, USA, 2005; pp. 50:1–50:28. [Google Scholar]
  2. Matsubara, K.; Matsuura, M. A Thermoelectric application to vehicles. In Thermoelectrics Handbook: Macro to Nano; Rowe, D.M., Ed.; CRC Press: Boca Raton, FL, USA, 2005; pp. 52:1–52:11. [Google Scholar]
  3. Bell, L.E. Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 2008, 321, 1457–1461. [Google Scholar] [CrossRef] [PubMed]
  4. Crane, D.T.; LaGrandeur, J.W. Progress report on BSST-led US department of energy automotive waste heat recovery program. J. Electron. Mater. 2010, 39, 2142–2148. [Google Scholar] [CrossRef]
  5. Yang, J.; Stabler, F.R. Automotive applications of thermoelectric materials. In Thermoelectrics and Its Energy Harvesting: Modules, Systems, and Applications in Thermoelectrics; Rowe, D.M., Ed.; CRC Press: Boca Raton, FL, USA, 2012; pp. 25:1–25:15. [Google Scholar]
  6. Slack, G.A. New materials and performance limits for thermoelectric cooling. In CRC Handbook of Thermoelectrics; Rowe, D.M., Ed.; CRC Press: Boca Raton, FL, USA, 1995; pp. 407–440. [Google Scholar]
  7. Terasaki, I.; Sasago, Y.; Uchinokura, K. Large thermoelectric power in NaCo2O4 single crystals. Phys. Rev. B 1997, 56, R12685–R12687. [Google Scholar] [CrossRef]
  8. Masset, A.C.; Michel, C.; Maignan, A.; Hervieu, M.; Toulemonde, O.; Studer, F.; Raveau, B.; Hejtmanek, J. Misfit-layered cobaltite with an anisotropic giant magnetoresistance: Ca3Co4O9. Phys. Rev. B 2000, 62, 166–175. [Google Scholar] [CrossRef]
  9. Funahashi, R.; Matsubara, I.; Ikuta, H.; Takeuchi, T.; Mizutani, U.; Sodeoka, S. An oxide single crystal with high thermoelectric performance in air. Jpn. J. Appl. Phys. 2000, 39, L1127–L1129. [Google Scholar] [CrossRef]
  10. Koumoto, K.; Terasaki, I.; Kajitani, T.; Ohtaki, M.; Funahashi, R. Oxide thermoelectrics. In Thermoelectrics Handbook: Macro to Nano; Rowe, D.M., Ed.; CRC Press: Boca Raton, FL, USA, 2005; pp. 35:1–35:15. [Google Scholar]
  11. Koumoto, K.; Terasaki, I.; Funahashi, R. Complex oxide materials for potential thermoelectric applications. MRS Bull. 2006, 31, 206–210. [Google Scholar] [CrossRef]
  12. Koumoto, K.; Wang, Y.F.; Zhang, R.Z.; Kosuga, A.; Funahashi, R. Oxide thermoelectric materials: A nanostructuring approach. Annu. Rev. Mater. Res. 2010, 40, 363–394. [Google Scholar] [CrossRef]
  13. Koumoto, K.; Funahashi, R.; Guilmeau, E.; Miyazaki, Y.; Weidenkaff, A.; Wang, Y.F.; Wan, C.L. Thermoelectric ceramics for energy harvesting. J. Am. Ceram. Soc. 2013, 96, 1–23. [Google Scholar] [CrossRef]
  14. Hébert, S.; Kobayashi, W.; Muguerra, H.; Bréard, Y.; Raghavendra, N.; Gascoin, F.; Guilmeau, E.; Maignan, A. From oxides to selenides and sulfides: The richness of the CdI2 type crystallographic structure for thermoelectric properties. Phys. Status Solidi A 2013, 210, 69–81. [Google Scholar] [CrossRef]
  15. Biswas, K.; He, J.Q.; Zhang, Q.C.; Wang, G.Y.; Uher, C.; Dravid, V.P.; Kanatzidis, M.G. Strained endotaxial nanostructures with high thermoelectric figure of merit. Nat. Chem. 2011, 3, 160–166. [Google Scholar] [CrossRef] [PubMed]
  16. Ohta, M.; Biswas, K.; Lo, S.H.; He, J.Q.; Chung, D.Y.; Dravid, V.P.; Kanatzidis, M.G. Enhancement of thermoelectric figure of merit by the insertion of MgTe nanostructures in p-type PbTe doped with Na2Te. Adv. Energy Mater. 2012, 2, 1117–1123. [Google Scholar] [CrossRef]
  17. Biswas, K.; He, J.Q.; Blum, I.D.; Wu, C.I.; Hogan, T.P.; Seidman, D.N.; Dravid, V.P.; Kanatzidis, M.G. High-performance bulk thermoelectrics with all-scale hierarchical architectures. Nature 2012, 489, 414–418. [Google Scholar] [CrossRef] [PubMed]
  18. He, J.Q.; Kanatzidis, M.G.; Dravid, V.P. High performance bulk thermoelectrics via a panoscopic approach. Mater. Today 2013, 5, 166–176. [Google Scholar] [CrossRef]
  19. Zhao, L.D.; Wu, H.J.; Hao, S.Q.; Wu, C.I.; Zhou, X.Y.; Biswas, K.; He, J.Q.; Hogan, T.P.; Uher, C.; Wolverton, C.; et al. All-scale hierarchical thermoelectrics: MgTe in PbTe facilitates valence band convergence and suppresses bipolar thermal transport for high performance. Energy Environ. Sci. 2013, 6, 3346–3355. [Google Scholar] [CrossRef]
  20. Zhao, L.D.; Dravid, V.P.; Kanatzidis, M.G. The panoscopic approach to high performance thermoelectrics. Energy Environ. Sci. 2014, 7, 251–268. [Google Scholar] [CrossRef]
  21. Wu, H.J.; Zhao, L.D.; Zheng, F.S.; Wu, D.; Pei, Y.L.; Tong, X.; Kanatzidis, M.G.; He, J.Q. Broad temperature plateau for thermoelectric figure of merit ZT > 2 in phase-separated PbTe0.7S0.3. Nat. Commun. 2014, 5, 4515:1–4515:9. [Google Scholar]
  22. Sales, B.C.; Mandrus, D.; Williams, R.K. Filled skutterudite antimonides: A new class of thermoelectric materials. Science 1996, 272, 1325–1328. [Google Scholar] [CrossRef] [PubMed]
  23. Nolas, G.S.; Cohn, J.L.; Slack, G.A.; Schujman, S.B. Semiconducting Ge clathrates: Promising candidates for thermoelectric applications. Appl. Phys. Lett. 1998, 73, 178–180. [Google Scholar] [CrossRef]
  24. Nolas, G.S.; Morelli, D.T.; Tritt, T.M. Skutterudites: A phonon-glass-electron crystal approach to advanced thermoelectric energy conversion applications. Annu. Rev. Mater. Sci. 1999, 29, 89–116. [Google Scholar] [CrossRef]
  25. Guo, J.Q.; Geng, H.Y.; Ochi, T.; Suzuki, S.; Kikuchi, M.; Yamaguchi, Y.; Ito, S. Development of skutterudite thermoelectric materials and modules. J. Electron. Mater. 2012, 41, 1036–1042. [Google Scholar] [CrossRef]
  26. Takabatake, T. Nano-cage structured materials: Clathrates. In Thermoelectric Nanomaterials: Materials Design and Applications; Koumoto, K., Mori, T., Eds.; Springer: Berlin, Germany, 2013; pp. 33–49. [Google Scholar]
  27. Rogl, G.; Grytsiv, A.; Rogl, P.; Peranio, N.; Bauer, E.; Zehetbauer, M.; Eibl, O. n-type skutterudites (R,Ba,Yb)yCo4Sb12 (R = Sr, La, Mm, DD, SrMm, SrDD) approaching ZT ≈ 2.0. Acta Mater. 2014, 63, 30–43. [Google Scholar] [CrossRef]
  28. Salvador, J.R.; Cho, J.Y.; Ye, Z.; Moczygemba, J.E.; Thompson, A.J.; Sharp, J.W.; Koenig, J.D.; Maloney, R.; Thompson, T.; Sakamoto, J.; et al. Conversion efficiency of skutterudite-based thermoelectric modules. Phys. Chem. Chem. Phys. 2014, 16, 12510–12520. [Google Scholar] [CrossRef] [PubMed]
  29. Ohta, M.; Satoh, S.; Kuzuya, T.; Hirai, S.; Kunii, M.; Yamamoto, A. Thermoelectric properties of Ti1+xS2 prepared by CS2 sulfurization. Acta Mater. 2012, 60, 7232–7240. [Google Scholar] [CrossRef]
  30. Beaumale, M.; Barbier, T.; Bréard, Y.; Guelou, G.; Powell, A.V.; Vaqueiro, P.; Guilmeau, E. Electron doping and phonon scattering in Ti1+xS2 thermoelectric compounds. Acta Mater. 2014, 78, 86–92. [Google Scholar] [CrossRef]
  31. Meerschaut, A.; Rabu, P.; Rouxel, J. Preparation and characterization of new mixed sandwiched layered compounds Ln32Nb28S88 (Ln = La, Ce). J. Solid State Chem. 1989, 78, 35–45. [Google Scholar] [CrossRef]
  32. Jood, P.; Ohta, M.; Nishiate, H.; Yamamoto, A.; Lebedev, O.I.; Berthebaud, D.; Suekuni, K.; Kunii, M. Microstructural control and thermoelectric properties of misfit layered sulfides (LaS)1+mTS2 (T = Cr, Nb): The natural superlattice systems. Chem. Mater. 2014, 26, 2684–2692. [Google Scholar] [CrossRef]
  33. Zhang, Y.G.; Wilkinson, A.P.; Lee, P.L.; Shastri, S.D.; Shu, D.; Chung, D.Y.; Kanatzidis, M.G. Determining metal ion distributions using resonant scattering at very high-energy K-edges: Bi/Pb in Pb5Bi6Se14. J. Appl. Crystallogr. 2005, 38, 433–441. [Google Scholar] [CrossRef]
  34. Ohta, M.; Chung, D.Y.; Kunii, M.; Kanatzidis, M.G. Low lattice thermal conductivity in Pb5Bi6Se14, Pb3Bi2S6, and PbBi2S4: Promising thermoelectric materials in the cannizzarite, lillianite, and galenobismuthite homologous series. J. Mater. Chem. A 2014, 2, 20048–20058. [Google Scholar] [CrossRef]
  35. Zhao, L.D.; Lo, S.H.; Zhang, Y.S.; Sun, H.; Tan, G.J.; Uher, C.; Wolverton, C.; Dravid, V.P.; Kanatzidis, M.G. Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Nature 2014, 508, 373–377. [Google Scholar] [CrossRef] [PubMed]
  36. Suekuni, K.; Tsuruta, K.; Kunii, M.; Nishiate, H.; Nishibori, E.; Maki, S.; Ohta, M.; Yamamoto, A.; Koyano, M. High-performance thermoelectric mineral Cu12−xNixSb4S13 tetrahedrite. J. Appl. Phys. 2013, 113, 043712:1–043712:5. [Google Scholar] [CrossRef]
  37. Lu, X.; Morelli, D.T.; Xia, Y.; Ozolins, V. Increasing the thermoelectric figure of merit of tetrahedrites by Co-doping with nickel and zinc. Chem. Mater. 2015, 27, 408–413. [Google Scholar] [CrossRef]
  38. Rimmington, H.P.B.; Balchin, A.A. The growth by iodine vapour transport techniques and the crystal structures of layer compounds in the series TiSxSe2−x, TiSxTe2−x, TiSexTe2−x. J. Cryst. Growth. 1974, 21, 171–181. [Google Scholar] [CrossRef]
  39. Han, S.H.; Cook, B.A. An experimental search for high ZT semiconductors: A survey of the preparation and properties of several alloy systems. AIP Conf. Proc. 1995, 316, 66–70. [Google Scholar]
  40. Imai, H.; Shimakawa, Y.; Kubo, Y. Large thermoelectric power factor in TiS2 crystal with nearly stoichiometric composition. Phys. Rev. B 2001, 64, 241104(R):1–241104(R):4. [Google Scholar] [CrossRef]
  41. Abbott, E.E.; Kolis, J.W.; Lowhorn, N.D.; Sams, W.; Rao, A.; Tritt, T.M. Thermoelectric properties of doped titanium disulfides. Appl. Phys. Lett. 2006, 88, 262106:1–262106:3. [Google Scholar]
  42. Wan, C.L.; Wang, Y.F.; Wang, N.; Koumoto, K. Low-thermal-conductivity (MS)1+x(TiS2)2 (M = Pb, Bi, Sn) misfit layer compounds for bulk thermoelectric materials. Materials 2010, 3, 2606–2617. [Google Scholar] [CrossRef]
  43. Wan, C.L.; Wang, Y.F.; Wang, N.; Norimatsu, W.; Kusunoki, M.; Koumoto, K. Intercalation: Building a natural superlattice for better thermoelectric performance in layered chalcogenides. J. Electron. Mater. 2011, 40, 1271–1280. [Google Scholar] [CrossRef]
  44. Guilmeau, E.; Bréard, Y.; Maignan, A. Transport and thermoelectric properties in Copper intercalated TiS2 chalcogenide. Appl. Phys. Lett. 2011, 99, 052107:1–052107:3. [Google Scholar] [CrossRef]
  45. Gascoin, F.; Raghavendra, N.; Guilmeau, E.; Bréard, Y. CdI2 structure type as potential thermoelectric materials: Synthesis and high temperature thermoelectric properties of the solid solution TiSxSe2−x. J. Alloy. Compd. 2012, 521, 121–125. [Google Scholar] [CrossRef]
  46. Beaumale, M.; Barbier, T.; Bréard, Y.; Hébert, S.; Kinemuchi, Y.; Guilmeau, E. Thermoelectric properties in the series Ti1−xTaxS2. J. Appl. Phys. 2014, 115, 043704:1–043704:7. [Google Scholar] [CrossRef]
  47. Beaumale, M.; Barbier, T.; Bréard, Y.; Raveau, B.; Kinemuchi, Y.; Funahashi, R.; Guilmeau, E. Mass fluctuation effect in Ti1−xNbxS2 bulk compounds. J. Electron. Mater. 2014, 43, 1590–1596. [Google Scholar] [CrossRef]
  48. Henderson, J.R.; Muramoto, M.; Loh, E.; Gruber, J.B. Electronic structure of rare-earth sesquisulfide crystals. J. Chem. Phys. 1967, 47, 3347–3356. [Google Scholar] [CrossRef]
  49. Toide, T.; Utsunomiya, T.; Sato, M.; Hoshino, Y.; Hatano, T.; Akimoto, Y. Preparation of lanthanum sulfides using carbon disulfide as sulfurization agent and the change of these sulfides on heating in air. Bull. Tokyo Inst. Technol. 1973, 117, 41–48. [Google Scholar]
  50. Guittard, M.; Flahaut, J. Preparation of rare earth sulfides and selenides. In Synthesis of Lanthanide and Actinide Compounds (Topics in f-Element Chemistry, Volume 2); Meyer, G., Morss, L.R., Eds.; Kluwer Academic: Dordrecht, The Netherlands, 1991; pp. 321–352. [Google Scholar]
  51. Hirai, S.; Shimakage, K.; Saitou, Y.; Nishimura, T.; Uemura, Y.; Mitomo, M.; Brewer, L. Synthesis and sintering of cerium(III) sulfide powders. J. Am. Ceram. Soc. 1998, 81, 145–151. [Google Scholar] [CrossRef]
  52. Hirai, S.; Suzuki, K.; Shimakage, K.; Nishimura, T.; Uemura, Y.; Mitomo, M. Preparations of γ-Pr2S3 and γ-Nd2S3 powders by sulfurization of Pr6O11 and Nd2O3 powders using CS2 gas, and their sintering. J. Jpn. Inst. Metals 2003, 67, 15–21. [Google Scholar]
  53. Miyazaki, Y.; Ogawa, H.; Kajitani, T. Preparation and thermoelectric properties of misfit layered sulfide [Yb1.90S2]0.62NbS2. Jpn. J. Appl. Phys. 2004, 43, L1202–L1204. [Google Scholar] [CrossRef]
  54. Barin, I.; Knacke, O. Thermochemical Properties of Inorganic Substances; Springer-Verlag: Berlin, Germany, 1973. [Google Scholar]
  55. Barin, I.; Knacke, O.; Kubaschewski, O. Thermochemical Properties of Inorganic Substances: Supplement; Springer-Verlag: Berlin, Germany, 1977. [Google Scholar]
  56. Aoki, T.; Wan, C.L.; Ishiguro, H.; Morimitsu, H.; Koumoto, K. Evaluation of layered TiS2-based thermoelectric elements fabricated by a centrifugal heating technique. J. Ceram. Soc. Jpn. 2011, 119, 382–385. [Google Scholar] [CrossRef]
  57. Li, D.; Qin, X.Y.; Zhang, J.; Wang, L.; Li, H.J. Enhanced thermoelectric properties of bismuth intercalated compounds BixTiS2. Solid State Commun. 2005, 135, 237–240. [Google Scholar] [CrossRef]
  58. Li, D.; Qin, X.Y.; Zhang, J.; Li, H.J. Enhanced thermoelectric properties of neodymium intercalated compounds NdxTiS2. Phys. Lett. A 2006, 348, 379–385. [Google Scholar] [CrossRef]
  59. Li, D.; Qin, X.Y.; Zhang, J. Improved thermoelectric properties of gadolinium intercalated compounds GdxTiS2 at the temperatures from 5 to 310 K. J. Mater. Res. 2006, 21, 480–483. [Google Scholar] [CrossRef]
  60. Zhang, J.; Qin, X.Y.; Li, D.; Dong, H.Z. The electrical and thermal conductivity and thermopower of nickel doped compounds (NixTi1−x)1+yS2 at low temperatures. J. Phys. D Appl. Phys. 2006, 39, 1230–1236. [Google Scholar] [CrossRef]
  61. Qin, X.Y.; Zhang, J.; Li, D.; Dong, H.Z.; Wang, L. The effect of Mg substitution for Ti on transport and thermoelectric properties of TiS2. J. Appl. Phys. 2007, 102, 073703:1–073703:7. [Google Scholar]
  62. Zhang, J.; Qin, X.Y.; Li, D.; Xin, H.X.; Pan, L.; Zhang, K.X. The transport and thermoelectric properties of Cd doped compounds (CdxTi1−x)1+yS2. J. Alloy. Compd. 2009, 479, 816–820. [Google Scholar] [CrossRef]
  63. Zhang, J.; Qin, X.Y.; Xin, H.X.; Li, D.; Song, C.J. Thermoelectric properties of Co-doped TiS2. J. Electron. Mater. 2011, 40, 980–986. [Google Scholar] [CrossRef]
  64. Wan, C.L.; Wang, Y.F.; Norimatsu, W.; Kusunoki, M.; Koumoto, K. Nanoscale stacking faults induced low thermal conductivity in thermoelectric layered metal sulfides. Appl. Phys. Lett. 2012, 100, 101913:1–101913:4. [Google Scholar]
  65. Putri, Y.E.; Wan, C.L.; Wang, Y.F.; Norimatsu, W.; Kusunoki, M.; Koumoto, K. Effects of alkaline earth doping on the thermoelectric properties of misfit layer sulfides. Scr. Mater. 2012, 66, 895–898. [Google Scholar] [CrossRef]
  66. Putri, Y.E.; Wan, C.L.; Dang, F.; Mori, T.; Ozawa, Y.; Norimatsu, W.; Kusunoki, M.; Koumoto, K. Effects of transition metal substitution on the thermoelectric properties of metallic (BiS)1.2(TiS2)2 misfit layer sulfide. J. Electron. Mater. 2014, 43, 1870–1874. [Google Scholar] [CrossRef]
  67. Takeuchi, S.; Katsuta, H. Characteristics of nonstoichiometry and lattice defects of the TiS2 phase. J. Jpn. Inst. Metals 1970, 34, 758–763. [Google Scholar]
  68. Thompson, A.H.; Gamble, F.R.; Symon, C.R. The verification of the existence of TiS2. Mater. Res. Bull. 1975, 10, 915–919. [Google Scholar] [CrossRef]
  69. Murray, J.L. The S-Ti (Sulfur-Titanium) system. J. Phase Equilib. 1986, 7, 156–163. [Google Scholar]
  70. Kobayashi, H.; Sakashita, K.; Sato, M.; Nozue, T.; Suzuki, T.; Kamimura, T. Electronic specific heat of Ti1+xS2 (0 < x < 0.1). Physica B 1997, 237–238, 169–171. [Google Scholar] [CrossRef]
  71. Cutler, M.; Leavy, J.F.; Fitzpatrick, R.L. Electronic transport in semimetallic cerium sulfide. Phys. Rev. 1964, 133, A1143–A1152. [Google Scholar] [CrossRef]
  72. Snyder, G.J.; Tobere, E.S. Complex thermoelectric materials. Nat. Mater. 2008, 7, 105–114. [Google Scholar] [CrossRef] [PubMed]
  73. May, A.F.; Fleurial, J.-P.; Snyder, G.J. Thermoelectric performance of lanthanum telluride produced via mechanical alloying. Phys. Rev. B 2008, 78, 125205:1–125205:12. [Google Scholar] [CrossRef]
  74. Logothetis, E.M.; Kaiser, W.J.; Kukkonen, C.A.; Faile, S.P.; Colella, R.; Gambold, J. Transport properties and the semiconducting nature of TiS2. Physica B+C 1980, 99, 193–198. [Google Scholar] [CrossRef]
  75. Kukkonen, C.A.; Kaiser, W.J.; Logothetis, E.M.; Blumenstock, B.J.; Schroeder, P.A.; Faile, S.P.; Colella, R.; Gambold, J. Transport and optical properties of Ti1+xS2. Phys. Rev. B 1981, 24, 1691–1709. [Google Scholar] [CrossRef]
  76. Koyano, M.; Negishi, H.; Ueda, Y.; Sasaki, M.; Inoue, M. Electrical resistivity and thermoelectric power of intercalation compounds MxTiS2 (M = Mn, Fe, Co, and Ni). Phys. Status Solidi B 1986, 138, 357–363. [Google Scholar] [CrossRef]
  77. Amara, A.; Frongillo, Y.; Aubin, M.J.; Jandl, S.; Lopez-Castillo, J.M.; Jay-Gerin, J.-P. Thermoelectric power of TiS2. Phys. Rev. B 1987, 36, 6415–6419. [Google Scholar] [CrossRef]
  78. Wiegers, G.A. Misfit layer compounds: Structures and physical properties. Prog. Solid State Chem. 1996, 24, 1–139. [Google Scholar] [CrossRef]
  79. Van Smaalen, S. Superspace-group approach to the modulated structure of the inorganic misfit layer compound (LaS)1.14NbS2. J. Phys. Condens. Matter 1991, 3, 1247–1263. [Google Scholar] [CrossRef]
  80. Wiegers, G.A.; Meetsma, A.; Haange, R.J.; van Smaalen, S.; de Boer, J.L.; Meerschaut, A.; Rabu, P.; Rouxel, J. The incommensurate misfit layer structure of (PbS)1.14NbS2, “PbNbS3” and (LaS)1.14NbS2, “LaNbS3”: An X-ray diffraction study. Acta Crystallogr. B 1990, 46, 324–332. [Google Scholar] [CrossRef]
  81. Miyazaki, Y.; Ogawa, H.; Nakajo, T.; Kikuchii, Y.; Hayashi, K. Crystal structure and thermoelectric properties of misfit-layered sulfides [Ln2S2]pNbS2 (Ln = Lanthanides). J. Electron. Mater. 2013, 42, 1335–1339. [Google Scholar] [CrossRef]
  82. Terashima, T.; Kojima, N. Electrical transport properties of incommensurate layer compounds (RES)xNbS2 (RE = rare-earth metals; x = 1.2, 0.6). J. Phys. Soc. Jpn. 1994, 63, 658–673. [Google Scholar] [CrossRef]
  83. Suzuki, K.; Enoki, T.; Imaeda, K. Synthesis, characterization and physical properties of incommensurate layered compounds/(RES)xTaS2 (RE = rare earth metal). Solid State Commun. 1991, 78, 73–77. [Google Scholar] [CrossRef]
  84. Suzuki, K.; Enoki, T.; Bandow, S. Electronic properties and valence state of Sm in (SmS)1.19TaS2. Phys. Rev. B 1993, 48, 11077–11085. [Google Scholar] [CrossRef]
  85. Cho, N.; Kikkawa, S.; Kanamaru, F.; Takeda, Y.; Yamamoto, O.; Kido, H.; Hoshikawa, T. Crystal structural, electric and magnetic studies on the misfit layer compounds “LnMS3” (Ln = rare-earth metal; M = Ti, V, Cr). Solid State Ion. 1993, 63–65, 696–701. [Google Scholar] [CrossRef]
  86. Kato, K.; Kawada, I.; Takahashi, T. Die Kristallstruktur von LaCrS3. Acta Crystallogr. B 1977, 33, 3437–3443. [Google Scholar] [CrossRef]
  87. Wiegers, G.A.; Meerschaut, A. Structures of misfit layer compounds (MS)nTS2 (M = Sn, Pb, Bi, rare earth metals; T = Nb, Ta, Ti, V, Cr; 1.08 < n < 1.23). J. Alloy. Compd. 1992, 178, 351–368. [Google Scholar] [CrossRef]
  88. Fang, C.M.; van Smaalen, S.; Wiegers, G.A.; Haas, C.; de Groot, R.A. Electronic structure of the misfit layer compound (LaS)1.14NbS2: Band-structure calculations and photoelectron spectra. J. Phys. Condens. Matter 1996, 8, 5367–5382. [Google Scholar] [CrossRef]
  89. Fang, C.M.; Ettema, A.R.H.; Haas, C.; Wiegers, G.A.; van Leuken, H.; de Groot, R.A. Electronic structure of the misfit-layer compound (SnS)1.17NbS2 deduced from band-structure calculations and photoelectron spectra. Phys. Rev. B 1995, 52, 2336–2347. [Google Scholar] [CrossRef]
  90. Fang, C.M.; de Groot, R.A.; Wiegers, G.A.; Haas, C. Electronic structure of the misfit layer compound (SnS)1.20TiS2: Band structure calculations and photoelectron spectra. J. Phys. Condens. Matter 1996, 8, 1663–1676. [Google Scholar] [CrossRef]
  91. Ohta, M.; Hirai, S.; Ma, Z.; Nishimura, T.; Uemura, Y.; Shimakage, K. Phase transformation and microstructures of Ln2S3 (Ln = La, Sm) with different impurities content of oxygen and carbon. J. Alloy. Compd. 2006, 408–412, 551–555. [Google Scholar] [CrossRef]
  92. Makovicky, E. The building principles and classification of bismuth-lead sulphosalts and related compounds. Fortschr. Mineral. 1981, 59, 137–190. [Google Scholar]
  93. Mrotzek, A.; Kanatzidis, M.G. “Design” in solid-state chemistry based on phase homologies. The concept of structural evolution and the new megaseries Am[M1+lSe2+l]2m[M2l+nSe2+3l+n]. Acc. Chem. Res. 2003, 36, 111–119. [Google Scholar] [CrossRef] [PubMed]
  94. Kanatzidis, M.G. New bulk materials for thermoelectric applications: Synthetic strategies based on phase homologies. In Chemistry, Physics, and Materials Science of Thermoelectric Materials; Kanatzidis, M.G., Mahanti, S.D., Hogan, T.P., Eds.; Kluwer Acadmeic/Plenum Publishers: New York, NY, USA, 2003; pp. 35–54. [Google Scholar]
  95. Kanatzidis, M.G. Structural evolution and phase homologies for “design” and prediction of solid-state compounds. Acc. Chem. Res. 2005, 38, 361–370. [Google Scholar] [CrossRef]
  96. Makovicky, E. Crystal structures of sulfides and other chalcogenides. Rev. Mineral. Geochem. 2006, 61, 7–125. [Google Scholar] [CrossRef]
  97. Segawa, K.; Taskin, A.A.; Ando, Y. Pb5Bi24Se41: A new member of the homologous series forming topological insulator heterostructures. J. Solid State Chem. 2015, 221, 196–201. [Google Scholar] [CrossRef]
  98. Chung, D.-Y.; Choi, K.-S.; Iordanidis, L.; Schindler, J.L.; Brazis, P.W.; Kannewurf, C.R.; Chen, B.; Hu, S.; Uher, C.; Kanatzidis, M.G. High thermopower and low thermal conductivity in semiconducting ternary K-Bi-Se compounds. Synthesis and properties of β-K2Bi8Se13 and K2.5Bi8.5Se14 and their Sb analogues. Chem. Mater. 1997, 9, 3060–3071. [Google Scholar] [CrossRef]
  99. Chung, D.Y.; Hogan, T.; Brazis, P.; Rocci-Lane, M.; Kannewurf, C.; Bastea, M.; Uher, C.; Kanatzidis, M.G. CsBi4Te6: A high-performance thermoelectric material for low-temperature applications. Science 2000, 287, 1024–1027. [Google Scholar] [CrossRef] [PubMed]
  100. Kuznetsova, L.A.; Kuznetsov, V.L.; Rowe, D.M. Thermoelectric properties and crystal structure of ternary compounds in the Ge(Sn,Pb)Te–Bi2Te3 systems. J. Phys. Chem. Solids 2000, 61, 1269–1274. [Google Scholar] [CrossRef]
  101. Caillat, T.; Huang, C.K.; Fleurial, J.-P.; Snyder, G.J.; Borshchevsky, A. Synthesis and thermoelectric properties of some materials with the PbBi4Te7 crystal structure. In Proceedings of the 19th International Conference on Thermoelectrics, Cardiff, UK, 20–24 August 2000; Rowe, D.M., Ed.; Babrow Press: Wales, UK, 2000; pp. 151–154. [Google Scholar]
  102. Chung, D.Y.; Hogan, T.P.; Rocci-Lane, M.; Brazis, P.; Ireland, J.R.; Kannewurf, C.R.; Bastea, M.; Uher, C.; Kanatzidis, M.G. A new thermoelectric material: CsBi4Te6. J. Am. Chem. Soc. 2004, 126, 6414–6428. [Google Scholar] [CrossRef] [PubMed]
  103. Shelimova, L.E.; Karpinskii, O.G.; Konstantinov, P.P.; Avilov, E.S.; Kretova, M.A.; Lubman, G.U.; Nikhezina, I.Yu.; Zemskov, V.S. Composition and properties of compounds in the PbSe-Bi2Se3 system. Inorg. Mater. 2010, 46, 120–126. [Google Scholar] [CrossRef]
  104. Zemskov, V.S.; Shelimova, L.E.; Konstantinov, P.P.; Avilov, E.S.; Kretova, M.A.; Nikhezina, I.Y. Thermoelectric materials with low heat conductivity based on PbSe-Bi2Se3 compounds. Inorg. Mater. Appl. Res. 2011, 2, 405–413. [Google Scholar] [CrossRef]
  105. Zemskov, V.S.; Shelimova, L.E.; Konstantinov, P.P.; Avilov, E.S.; Kretova, M.A.; Nikhezina, I.Yu. Thermoelectric materials based on layered chalcogenides of bismuth and lead. Inorg. Mater. Appl. Res. 2012, 3, 61–68. [Google Scholar] [CrossRef]
  106. Kuropatwa, B.A.; Kleinke, H. Thermoelectric properties of stoichiometric compounds in the (SnTe)x(Bi2Te3)y system. Z. Anorg. Allg. Chem. 2012, 638, 2640–2647. [Google Scholar] [CrossRef]
  107. Kuropatwa, B.A.; Assoud, A.; Kleinke, H. Effects of cation site substitutions on the thermoelectric performance of layered SnBi2Te4 utilizing the triel elements Ga, In, and Tl. Z. Anorg. Allg. Chem. 2013, 639, 2411–2420. [Google Scholar] [CrossRef]
  108. Medlin, D.L.; Snyder, G.J. Atomic-scale interfacial structure in rock salt and tetradymite chalcogenide thermoelectric materials. JOM 2013, 65, 390–400. [Google Scholar] [CrossRef]
  109. Olvera, A.; Shi, G.G.; Djieutedjeu, H.; Page, A.; Uher, C.; Kioupakis, E.; Poudeu, P.F.P. Pb7Bi4Se13: A lillianite homologue with promising thermoelectric properties. Inorg. Chem. 2015, 54, 746–755. [Google Scholar] [CrossRef] [PubMed]
  110. Kauzlarich, S.M.; Brown, S.R.; Snyder, G.J. Zintl phases for thermoelectric devices. Dalton Trans. 2007, 2099–2107. [Google Scholar] [CrossRef]
  111. Lefebvre, I.; Szymanski, M.A.; Olivier-Fourcade, J.; Jumas, J.C. Electronic structure of tin monochalcogenides from SnO to SnTe. Phys. Rev. B 1998, 58, 1896–1906. [Google Scholar] [CrossRef]
  112. Chattopadhyay, T.; Pannetier, J.; von Schnering, H.G. Neutron diffraction study of the structural phase transition in SnS and SnSe. J. Phys. Chem. Solids 1986, 47, 879–885. [Google Scholar] [CrossRef]
  113. Peters, M.J.; McNeil, L.E. High-pressure Mössbauer study of SnSe. Phys. Rev. B 1990, 41, 5893–5897. [Google Scholar] [CrossRef]
  114. Fadavieslam, M.R.; Shahtahmasebi, N.; Rezaee-Roknabadi, M.; Bagheri-Mohagheghi, M.M. A study of the photoconductivity and thermoelectric properties of SnxSy optical semiconductor thin films deposited by the spray pyrolysis technique. Phys. Scr. 2011, 84, 035705:1–035705:8. [Google Scholar] [CrossRef]
  115. Tan, Q.; Li, J.F. Thermoelectric properties of Sn-S bulk materials prepared by mechanical alloying and spark plasma sintering. J. Electron. Mater. 2014, 43, 2435–2439. [Google Scholar] [CrossRef]
  116. Tan, Q.; Zhao, L.D.; Li, J.F.; Wu, C.F.; Wei, T.R.; Xing, Z.B.; Kanatzidis, M.G. Thermoelectrics with earth abundant elements: Low thermal conductivity and high thermopower in doped SnS. J. Mater. Chem. A 2014, 2, 17302–17306. [Google Scholar] [CrossRef]
  117. Sassi, S.; Candolfi, C.; Vaney, J.-B.; Ohorodniichuk, V.; Masschelein, P.; Dauscher, A.; Lenoir, B. Assessment of the thermoelectric performance of polycrystalline p-type SnSe. Appl. Phys. Lett. 2014, 104, 212105:1–212105:4. [Google Scholar] [CrossRef]
  118. Chen, C.L.; Wang, H.; Chen, Y.Y.; Day, T.; Snyder, G.J. Thermoelectric properties of p-type polycrystalline SnSe doped with Ag. J. Mater. Chem. A 2014, 2, 11171–11176. [Google Scholar] [CrossRef]
  119. Carrete, J.; Mingo, N.; Curtarolo, S. Low thermal conductivity and triaxial phononic anisotropy of SnSe. Appl. Phys. Lett. 2014, 105, 101907:1–101907:4. [Google Scholar] [CrossRef]
  120. Wasscher, J.D.; Albers, W.; Haas, C. Simple evaluation of the maximum thermoelectric figure of merit, with application to mixed crystals SnS1-xSex. Solid-State Electron. 1963, 6, 261–264. [Google Scholar] [CrossRef]
  121. Chen, S.; Cai, K.; Zhao, W. The effect of Te doping on the electronic structure and thermoelectric properties of SnSe. Physica B 2012, 407, 4154–4159. [Google Scholar] [CrossRef]
  122. Parker, D.; Singh, D.J. First Principles investigations of the thermoelectric behavior of tin sulfide. J. Appl. Phys. 2010, 108, 083712:1–083712:3. [Google Scholar] [CrossRef]
  123. Bera, C.; Jacob, S.; Opahle, I.; Gunda, N.S.H.; Chmielowski, R.; Dennler, G.; Madsen, G.K.H. Integrated computational materials discovery of silver doped tin sulfide as a thermoelectric material. Phys. Chem. Chem. Phys. 2014, 16, 19894–19899. [Google Scholar] [CrossRef] [PubMed]
  124. Lu, X.; Morelli, D.T.; Xia, Y.; Zhou, F.; Ozolins, V.; Chi, H.; Zhou, X.Y.; Uher, C. High performance thermoelectricity in earth-abundant compounds based on natural mineral tetrahedrites. Adv. Energy Mater. 2013, 3, 342–348. [Google Scholar] [CrossRef]
  125. Chetty, R.; Kumar, D.S.P.; Rogl, G.; Rogl, P.; Bauer, E.; Michor, H.; Suwas, S.; Puchegger, S.; Giester, G.; Mallik, R.C.; et al. Thermoelectric properties of a Mn substituted synthetic tetrahedrite. Phys. Chem. Chem. Phys. 2015, 17, 1716–1727. [Google Scholar] [CrossRef] [PubMed]
  126. Suekuni, K.; Kim, F.S.; Nishiate, H.; Ohta, M.; Tanaka, H.I.; Takabatake, T. High-performance thermoelectric minerals: Colusites Cu26V2M6S32 (M = Ge, Sn). Appl. Phys. Lett. 2014, 105, 132107:1–132107:4. [Google Scholar] [CrossRef]
  127. Suekuni, K.; Tsuruta, K.; Ariga, T.; Koyano, M. Thermoelectric properties of mineral tetrahedrites Cu10Tr2Sb4S13 with low thermal conductivity. Appl. Phys. Express 2012, 5, 051201:1–051201:3. [Google Scholar] [CrossRef]
  128. Tsujii, N.; Mori, T. Stability and thermoelectric property of Cu9Fe9S16: Sulfide mineral as a promising thermoelectric material. MRS Proc. 2014, 1680. [Google Scholar] [CrossRef]
  129. Barbier, T.; Lemoine, P.; Gascoin, S.; Lebedev, O.I.; Kaltzoglou, A.; Vaqueiro, P.; Powell, A.V.; Smith, R.I.; Guilmeau, E. Structural stability of the synthetic thermoelectric ternary and nickel-substituted tetrahedrite phases. J. Alloy. Compd. 2015, 634, 253–262. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Jood, P.; Ohta, M. Hierarchical Architecturing for Layered Thermoelectric Sulfides and Chalcogenides. Materials 2015, 8, 1124-1149. https://doi.org/10.3390/ma8031124

AMA Style

Jood P, Ohta M. Hierarchical Architecturing for Layered Thermoelectric Sulfides and Chalcogenides. Materials. 2015; 8(3):1124-1149. https://doi.org/10.3390/ma8031124

Chicago/Turabian Style

Jood, Priyanka, and Michihiro Ohta. 2015. "Hierarchical Architecturing for Layered Thermoelectric Sulfides and Chalcogenides" Materials 8, no. 3: 1124-1149. https://doi.org/10.3390/ma8031124

Article Metrics

Back to TopTop