Thermoelectrics

Today, most energy resources are discharged as waste heat into the environment without being applied. Such exhaust heat reaches approximately 2/3 of the primary energy. Hence, thermoelectric energy conversion technology has attracted great attention for converting waste heat into electricity1,2. The principle of thermoelectric energy conversion was first discovered by T.J. Seebeck in 18213. He found that a voltage is generated between two ends of a metal bar by introducing a temperature difference. Thus, when electric loads are connected at both ends of a metal bar, an electric current can be obtained. This phenomenon is called the Seebeck effect. Conversely, in 1834, J.C.A. Peltier discovered that heating or cooling of the junctions can occur during electric current application to a heterogeneous metal circuit. This phenomenon is called the Peltier effect, and it has been commercially applied in electronic refrigerators, among other applications.

Generally, the performance of thermoelectric materials is evaluated in terms of a dimensionless figure of merit, ZT = S2·σ·T·κ−1, where Z is a figure of merit, T is the absolute temperature, S is the thermopower (≡ Seebeck coefficient), σ is the electrical conductivity, and κ is the thermal conductivity. The energy conversion efficiency from transforming the temperature difference into electricity increases as ZT increases. Thus, to realize efficient thermoelectric energy conversion, three physical properties are needed for thermoelectric materials: (1) low κ, which is needed to introduce a large temperature difference into both ends of the material; (2) high σ, which is needed to reduce the internal resistance of the material; and (3) high S, which is needed to obtain a high voltage.

The ZT values of practical thermoelectric materials, such as Bi2Te3 and PbTe, are ~1, which is the lowest value needed for practical applications4,5. Recently, several high-ZT materials (ZT > 1) have been developed sequentially based on heavy metal alloys, including SnSe, PbSe, GeTe, and oxychalcogenides (BiCuSeO)5,6,7,8,9,10,11,12,13,14,15,16,17. Since thermoelectric devices can directly convert a temperature difference into electricity, some automobile companies have developed thermoelectric-assisted hybrid automobiles18,19,20,21 while considering the outside temperature of the exhaust pipe to be ~700 °C. However, these thermoelectric materials are not appealing, particularly when operating at such high temperatures, because decomposition, vaporization, and melting of the constituents can easily occur. Furthermore, the use of these heavy metals should be limited to specific environments, such as space, because they are mostly toxic, low in abundance as natural resources, and not environmentally benign.

To overcome these issues, metal oxides have attracted much attention as thermoelectric power generation materials at high temperatures based on their potential advantages over heavy metal alloys in terms of chemical and thermal robustness22,23,24. From the 1950s to the 1970s, there was a boom in the search for oxide thermoelectric materials. At this time, researchers in the United States used thermoelectric effects to investigate the intrinsic properties of grown oxide single crystals. Since the 1990s, the search for oxide thermoelectric materials, which originated in Japan, has spread all over the world, including in Europe, the United States, Asia, and India. To date, it has been reported that some oxides could exhibit parameters of ZT > 1, exceeding that of PbTe.

History of thermoelectric oxides

There is a long history of thermoelectric oxides. From the 1950s to the 1970s (1st boom), the thermoelectric properties of many simple conducting oxides, including CdO (1949, Hogarth et al.25), NiO (1957, Parravano26), ZnO (1959, Hutson27), In2O3 (1962, Arvin28), SrTiO3 (1964, Frederikse et al.29), rutile-TiO2 (1965, Thurber et al.30), SnO2 (1965, Marley and Dockerty31), Cu2O (1969, Young and Schwartz32), and Fe3O4 (1970, Griffiths et al.33), were studied to obtain the fundamental physical properties of conducting oxides, such as carrier effective mass. After the discovery of cuprous oxide-based high-Tc superconducting oxides in 198634, the thermoelectric properties of superconducting oxides, including La2CuO4 (1987, Cooper et al.35), La-Ba-Cu-O (1987, Chen et al.36), YBa2Cu3O7−δ (1988, Lee et al.37), and Tl-Ca-Ba-Cu-O (1988, Mitra et al.38), were reported sequentially to clarify the superconducting transition (2nd boom).

After the research boom in the field of high-Tc superconductors, two Japanese researchers, Ohtaki and Terasaki, reported that oxides had good thermoelectric performance, including CaMnO3 (1995, Ohtaki et al.39), Al-doped ZnO (1996, Ohtaki et al.40), and NaxCoO2 (1997, Terasaki et al.41). These reports triggered the 3rd boom of thermoelectric oxide research in the 2000s. As a result of energetic exploratory research on good thermoelectric oxides, Ca3Co4O9 (2000, Masset et al.42 and Funahashi et al.43) and electron-doped SrTiO3 (2001, Okuda et al., La-doped44, 2005, Ohta et al. Nb-doped45,46) were discovered. In 2012, Fergus reviewed the thermoelectric properties of promising oxides (mostly bulk ceramics), including Ca3Co4O9, NaxCoO2, SrTiO3, CaMnO3, and ZnO47 (also see the references therein). Recently, several scholars that have reported rather high ZT values for oxide ceramics. Acharya et al.48 reported that SrTi0.85Nb0.15O3 ceramics sintered with graphite flakes exhibit ZT = 1.4 at 1050 K, and Biswas et al.49 reported that Al-doped ZnO ceramics sintered with reduced graphene oxide exhibit ZT = 0.5 at 1100 K. Although these reported ZTs are very attractive as practical thermoelectric materials, there is still no practical application for them, probably due to the low reliability of the ZT values.

To clarify the intrinsic thermoelectric properties of oxides, we focused on high-quality epitaxial films with stepped and terraced surfaces. As a result, we fabricated high-quality epitaxial films of several thermoelectric oxides, including Na3/4CoO250,51, Sr1/3CoO252, Ca1/3CoO253, Ba1/3CoO254, Ca3Co4O953,55, SrTiO3:Nb46, TiO2:Nb56 and SrO(SrTiO3):Nb57. Among these oxides, we found that Ba1/3CoO2 epitaxial films exhibited a ZT value of ~0.55 at 600 °C in air, which is the highest and most reliable value among the reported thermoelectric oxides54. In this context, we reviewed the epitaxial film growth and thermoelectric properties of four representative p-type layered cobalt oxide films based on our efforts: Na3/4CoO250,51, Sr1/3CoO252, Ca1/3CoO253, and Ba1/3CoO254.

Epitaxial film growth of A xCoO2 (A x = Na3/4, Ca1/3, Sr1/3, and Ba1/3)

Na3/4CoO2 50,51

Figure 1 shows the schematic crystal structure of AxCoO2 (Ax = Na3/4, Ca1/3, Sr1/3, and Ba1/3). In the case of AxCoO2, the rigid CoO2 layer and mobile Ax layer are alternately stacked along the c-axis. In 1997, Terasaki et al. discovered that a NaCo2O4 (≡NaxCoO2, x ~ 3/4) single crystal with a two-dimensional layered structure exhibits a very large power factor S2 ∙ σ of 5 mW m−1 K−2 in the in-plane direction at room temperature41. After this discovery, the electronic structure58,59,60, crystal structure61,62, and Na-composition dependence of the thermoelectric properties62 of NaxCoO2 were energetically studied to understand the origin of the unusually large S. In 2001, Fujita and coworkers fabricated NaxCoO2 single crystals and reported that they exhibited ZT values of ~1.2 at 800 K63. This material has attracted much attention because it can be converted into a superconductor (Tc ~ 4.7 K) by introducing H2O molecules into a layer between the two adjacent CoO2 layers64.

Fig. 1: Schematic illustration of the crystal structure of AxCoO2.
figure 1

Left: Ax = Na3/4, Right: Ax = Ca1/3, Sr1/3, and Ba1/3.

To clarify the intrinsic thermoelectric properties of Na3/4CoO2, we fabricated thin epitaxial films of Na3/4CoO2. First, we tried to fabricate Na3/4CoO2 epitaxial films by the conventional pulsed laser deposition (PLD) technique, but we failed. Several reports have been written on the thin film growth of Na3/4CoO2 by PLD. However, the film quality in terms of crystallographic orientation, surface morphology, and lateral grain size is insufficient. In the case of Na3/4CoO2 film growth by PLD, it is very difficult to control the Na concentrations in the films at high temperatures in a vacuum environment due to the high vapor pressure of the Na species.

To overcome this difficulty, we modified the reactive solid-phase epitaxy (R-SPE)50 method that was developed to fabricate single-crystal films of InGaO3(ZnO)m (m = integer). Figure 2 schematically illustrates the R-SPE procedure. Step 1: A highly (111)-oriented CoO epitaxial film was deposited on a (0001)-α-Al2O3 substrate at 700 °C by the PLD technique using a Co3O4 sintered disk as a target. Step 2: The surface of the PLD-deposited CoO film was fully capped by an yttria-stabilized zirconia (YSZ) single-crystalline plate to keep the surface clean. Step 3: NaHCO3 powder was put on the YSZ plate. Step 4: The sandwich specimen was annealed at 700 °C for 1 h in air. Notably, several researchers have used this R-SPE method for fabricating NaxCoO2 epitaxial films65,66,67 since the resultant film quality, especially in surface morphology, is better than that of PLD-grown films68,69.

Fig. 2: Film growth procedure.
figure 2

Schematic illustration of reactive solid-phase epitaxy (R-SPE) for the preparation of NaxCoO2 epitaxial films.

The out-of-plane (Fig. 3a) and in-plane (Fig. 3b) X-ray diffraction patterns of the resultant NaxCoO2 film clearly indicate that epitaxial growth occurred, showing the effectiveness of the R-SPE method. The chemical composition of the obtained film was evaluated to be x = 0.83 by X-ray fluorescence (XRF) measurements. The Na content in the present film was slightly higher than the reported values in the as-grown bulk sample x = 0.770, likely because the amorphous layer at the interface contained some Na ions51. Figure 3c shows an atomic force microscopy (AFM) image of the NaxCoO2 film. A step-like structure composed of several flake-like domains could be observed. The step increment was approximately 3 nm, which was three times longer than the c-axis length of NaxCoO2, suggesting that step bunching occurred during annealing at 700 °C.

Fig. 3: X-ray diffraction (XRD) patterns and topographic atomic force microscopy (AFM) images of NaxCoO2 epitaxial films.
figure 3

a Out-of-plane XRD. b In-plane XRD. c Topographic AFM image. This figure was reproduced with permission50,82.

As described in the next section, the R-SPE-grown Na0.7CoO2 epitaxial film was very useful for fabricating epitaxial films of LiCoO271, Sr0.5CoO252, Ca0.33CoO255,72, and Ca3Co4O955. Furthermore, the Na0.7CoO2 epitaxial film could be converted into a superconducting sodium cobalt oxyhydrate, Na0.3CoO2·1.3H2O (Tc ~ 4 K), by dipping in HNO3 for oxidation treatment followed by dipping in NaCl aqueous solution for hydration treatment73. Furthermore, peeling-off of the Na0.7CoO2 epitaxial film from the α-Al2O3 substrate was possible, and the peeled film could be pasted on the other substrate51.

Ca1/3CoO2

Powder syntheses of CaxCoO2 (x = 0.3, 0.35 and 0.5) were reported in 1996 by Cushing et al.74,75 The scholars used specimens of the sodium cobalt oxide NaxCoO2 (0.6 ≤ x ≤ 1.0) as precursors and performed multivalent ion-exchange reactions.

$${{\rm{Na}}}_{x}{{\rm{CoO}}}_{2}+x/2{\rm{Ca}}{({{\rm{NO}}}_{3})}_{2}\to {{\rm{Ca}}}_{x/2}{{\rm{CoO}}}_{2}+x\,{{\rm{NaNO}}}_{3}$$

Stoichiometric amounts of the AxCoO2 precursors were combined with anhydrous Ca(NO3)2 in evacuated sealed glass tubes and heated at 350 °C for 48 h.

There is an interesting feature in the CaxCoO2 system: there are two superstructures of cation ordering. In 2006, Yang et al. discovered that there are two common well-defined cation ordered states corresponding to the \(2a\,\times \,\sqrt{3}a\) orthorhombic superstructure at approximately x = 1/2 and the \(\sqrt{3}a\,\times \,\sqrt{3}a\) hexagonal superstructure at approximately x = 1/376. In 2006, Sugiura et al. fabricated high-quality Ca0.48CoO2 epitaxial films by ion-exchange reactions. Na3/4CoO2 epitaxial films were heated together with Ca(NO3)2 powder at 300 °C for 0.5 h in air55. In 2008, Huang et al. directly observed two instances of Ca ordering in Ca1/3CoO2 epitaxial films using high-angle annular dark-field (HAADF) scanning transmission electron microscopy (STEM)77,78.

In 2009, Sugiura et al. found that the structural transformation of Ca1/3CoO2 occurred at approximately 300 °C72. The \(\sqrt{3}a\,\times \,\sqrt{3}a\) hexagonal phase transformed into the \(2a\,\times \,\sqrt{3}a\) orthorhombic phase when heated in air. The scholars found that the orthorhombic phase showed insulating electron transport, whereas the hexagonal phase showed metallic transport. Interestingly, the temperature dependence of the thermopower of Ca1/3CoO2 epitaxial films was similar, independent of the crystallographic phases.

Sr1/3CoO2

The most serious drawback of Na3/4CoO2 is its low chemical stability against water. Na3/4CoO2 is easily decomposed into insulating Co(OH)2 under high-humidity conditions (temperature, 80 °C; humidity, ~80%) because Na+ ions can easily dissolve in water. To address this issue, Sugiura and coworkers hypothesized that modification of the chemical composition improves the chemical stability without degrading the thermoelectric performance.

In 2002, Ishikawa et al. reported that the thermoelectric properties of the Sr1/3CoO2 ceramic synthesized by sintering at 400 °C were quite low relative to those of Na3/4CoO2. Although a high-density and single-crystal Sr1/3CoO2 ceramics are preferable for clarifying the intrinsic properties, this process is extremely difficult because the phase transition of Sr1/3CoO2 occurs at a relatively low temperature (~400 °C). To examine the intrinsic thermoelectric properties of Sr1/3CoO2, Sugiura and coworkers fabricated high-quality epitaxial films of Sr1/3CoO2 because epitaxial films generally exhibit intrinsic carrier transport properties, similar to those of bulk single crystals. The Sr1/3CoO2 epitaxial films exhibits better chemical stability than Na3/4CoO2 while retaining good thermoelectric properties.

Ba1/3CoO2

Recently, we found a reliable high-ZT thermoelectric oxide, Ba1/3CoO2. The crystal structure and electrical properties of the Ba1/3CoO2 epitaxial films were maintained to 600 °C. The power factor gradually increased to ~1.2 mW m−1 K−2, and the thermal conductivity gradually decreased to ~1.9 W m−1 K−1 with increasing temperature to 600 °C. Consequently, the ZT reached ~0.55 at 600 °C in air, which was the highest value among oxides and comparable to those of p-type PbTe and p-type SiGe.

Notably, initial investigations on BaxCoO2-based thermoelectric materials have been conducted by Liu et al.79,80 The scholars fabricated BaxCoO2 (x = 0.19, 0.28, 0.30, 0.33) ceramics by using BaxCoO2 powders synthesized through the ion-exchange technique from Na0.7CoO2. The researchers measured the thermoelectric properties to 800 K and found that ZT values ranging from 0.14 to 0.21 were dependent on Ba content. In their research, thermal conductivity did not experience effective suppression from polycrystalline grain boundaries, while electron transport properties deteriorated substantially. Therefore, high-quality epitaxial films are ideal for elucidating the intrinsic properties of BaxCoO2-based thermoelectric materials, which is essential for developing high-performance thermoelectric oxides.

Thermoelectric properties of A xCoO2 (A x = Na3/4, Ca1/3, Sr1/3, and Ba1/3) epitaxial films

Systematic investigations on the thermoelectric properties of AxCoO2 (Ax = Na3/4, Ca1/3, Sr1/3, and Ba1/3) started from our findings that heavy ion substitution at the A-site of AxCoO2 effectively reduces the in-plane thermal conductivity81. By fabricating AxCoO2 (Ax = Li1, Na0.75, Ca0.33, Sr0.33, La0.3) epitaxial films on (0001) α-Al2O3 substrates by conducting R-SPE and an ion exchange process, we clarify the A-site ion mass-dependent thermal conductivity of AxCoO250,51. As shown in Fig. 4, the in-plane thermal conductivity (κ||) obviously decreases with the A-site ion mass due to the mismatch of the impedance between the cation layer and CoO2 layers. This impedance hinders coupling of the vibrational modes, while the cross-plane thermal conductivity (κ) mainly depends on the interfacial scattering.

Fig. 4: Reduced thermal conductivity through heavy ion substitution at the A-site of AxCoO2.
figure 4

As the mass of the A-site ions increases, the in-plane thermal conductivity (κ||) and cross-plane thermal conductivity (κ) show decreasing tendencies. The inset shows a schematic illustration of the phonon propagation of AxCoO2. Heavy ions display a stronger thermal conductivity suppression effect than light ions. This figure was reproduced with permission81.

By conducting heavy ion substitution to reduce thermal conductivity, we further fabricate AxCoO2 (Ax = Na3/4, Ca1/3, Sr1/3, and Ba1/3) epitaxial films on (0001) α-Al2O3 and (111) YSZ substrates and compare their room temperature thermoelectric properties82. Fig. 5 presents a summary of the room-temperature electrical conductivity (σip), thermopower (Sip), power factor (PFip), thermal conductivity (κip) and figure of merit (ZTip) values along the in-plane direction of AxCoO2 epitaxial films. The electrical conductivity and thermopower values of all the films show stable changing patterns, resulting in a consistent power factor (Fig. 5a–c). This consistent value suggests perfect electron–phonon decoupling between the A-site ion layer and CoO2 layer, where ion substitution has almost no effect on the electrical conductivity of the CoO2 layers. Moreover, the thermal conductivity along the layered direction decreases with the atomic mass of Ax, thereby enhancing the ZT value (Fig. 5d, e). The highest ZT value of ~0.11 can be obtained in the Ba1/3CoO2 epitaxial film, reaching a peak value among layered cobalt oxides. In this research, the in-plane thermal conductivity has been deduced based on the experimental results of cross-plane thermal conductivities for differently oriented epitaxial films by varying the substrate orientations. In our latest report, we have directly confirmed the in-plane thermal conductivity through AC calorimetric measurements by using a freestanding Ba1/3CoO2 single-crystalline film, yielding a consistent result83.

Fig. 5: Thermoelectric properties of AxCoO2 epitaxial films at room temperature.
figure 5

a In-plane electrical conductivity (σip), b in-plane thermopower (Sip), c in-plane power factor (PFip), d in-plane and out-of-plane thermal conductivity (κip, κop), and e in-plane figure of merit (ZTip) as a function of the atomic mass Ax in AxCoO2. The κip shows a clear decreasing tendency, whereas the PF shows a slight increasing tendency. These trends result in an increasing tendency of ZT. ZT reached 0.11 when Ax = Ba1/3. This figure was reproduced with permission82.

As an emerging candidate for high-performance oxide-based thermoelectric materials, Ba1/3CoO2 has promising prospects in applications at elevated temperatures. To elucidate the high-temperature thermoelectric performance, we further conducted high-temperature characterizations of AxCoO2 epitaxial films54. First, the thermal stabilities of the Na3/4CoO2, Ca1/3CoO2, Sr1/3CoO2, and Ba1/3CoO2 epitaxial films were tested by annealing at an elevated temperature for 0.5 h in air. Figure 6 shows the room temperature XRD patterns after heat treatment. The 0002 Na3/4CoO2 diffraction peak shrinks above 450 °C, whereas the 111 Co3O4 peak appears due to the evaporation of Na. In contrast, the 0002 Ca1/3CoO2, 0002 Sr1/3CoO2, and 0002 Ba1/3CoO2 peaks appear below 650 °C (Fig. 6a). However, the in-plane XRD patterns (Fig. 6b) demonstrate that a phase transition from hexagonal to orthorhombic occurs in the Ca1/3CoO2 film when the annealing temperature is above 200 °C72. The Sr1/3CoO2 film shows a hexagonal–orthorhombic hybridized phase below 450 °C and a single orthorhombic phase above ~450 °C. Only the Ba1/3CoO2 film can maintain a stable phase composition to 600 °C, which suggests a strong thermal robustness and a high potential for high-temperature Ba1/3CoO2 applications. We have confirmed a similar temperature-dependent behavior from the resistivity variation after heat treatment.

Fig. 6: Changes in the crystal structures of AxCoO2 (Ax = Na3/4, Ca1/3, Sr1/3, and Ba1/3) films after annealing at high temperatures for 0.5 hours in air.
figure 6

a Out-of-plane XRD patterns measured at room temperature after each annealing step. The 0002 Na3/4CoO2 diffraction peak intensity decreases above 450 °C due to decomposition into Co3O4, whereas those for Ca1/3CoO2, Sr1/3CoO2, and Ba1/3CoO2 remain stable to 650 °C. b In-plane XRD patterns measured at room temperature after each annealing step. The 1/3 and 2/3 diffraction peaks of the Ca1/3CoO2 film disappear at approximately 200 °C, and the 1/2 diffraction peak appears above 200 °C. The 1/3 and 2/3 diffraction peaks of the Sr1/3CoO2 film disappear at approximately 450 °C, and the 1/2 diffraction peak appears above 450 °C. The 1/3 and 2/3 diffraction peaks of the Ba1/3CoO2 film are stable to 600 °C. This figure was reproduced with permission54.

Finally, we calculated the temperature-dependent ZT of AxCoO2 epitaxial films. The ZT values increase with temperature for all films (Fig. 7a). Due to the strongest thermal robustness, the Ba1/3CoO2 epitaxial film displays the highest ZT of ~0.55 at 600 °C, which is higher than those of the Ca1/3CoO2 and Sr1/3CoO2 films. This high ZT value of Ba1/3CoO2 is reproducible and reliable. This value is comparable to those of p-type PbTe and p-type SiGe, indicating that Ba1/3CoO2 is a suitable candidate for high-temperature thermoelectric applications (Fig. 7b).

Fig. 7: Temperature dependence of the ZT values of Ba1/3CoO2 epitaxial films along the in-plane direction.
figure 7

a Comparison among the four AxCoO2 (A = Na3/4, Ca1/3, Sr1/3, and Ba1/3) films. In all cases, the ZT values increase as the temperature increases. The Ba1/3CoO2 epitaxial film exhibits the highest ZT among the four AxCoO2 epitaxial films and reaches a ZT value of ~0.55 at 600 °C. b Comparison against commercially available p-type thermoelectric materials. ZT of the Ba1/3CoO2 epitaxial film at 600 °C is comparable to the values of p-type PbTe and p-type SiGe. This figure was reproduced with permission54.

Summary and prospects

We have reviewed the thermoelectric properties of representative layered cobalt oxides: AxCoO2 (A = Na, Ca, Sr, and Ba) and Ca3Co4O9. Although several high-ZT thermoelectric oxides (ZT > 1) have been reported thus far, their reliability is low due to a lack of careful observation of their stabilities at elevated temperatures. We have explained that Ba1/3CoO2 is stable in air even at 600 °C and exhibits a high ZT value of 0.55, which is comparable to p-type PbTe. Bulk crystals (single crystal and sintered) are essential for incorporating Ba1/3CoO2 into thermoelectric conversion elements. To date, we are researching the growth of large single crystals and are proceeding with the production of sintered bodies. Moreover, better thermoelectric performance may be realized by optimizing the compositions and nanostructures of these crystals.