Open access peer-reviewed chapter

Challenges in Improving Performance of Oxide Thermoelectrics Using Defect Engineering

Written By

Jamil Ur Rahman, Gul Rahman and Soonil Lee

Submitted: 21 December 2020 Reviewed: 28 January 2021 Published: 03 March 2021

DOI: 10.5772/intechopen.96278

From the Edited Volume

Thermoelectricity - Recent Advances, New Perspectives and Applications

Edited by Guangzhao Qin

Chapter metrics overview

452 Chapter Downloads

View Full Metrics

Abstract

Oxide thermoelectric materials are considered promising for high-temperature thermoelectric applications in terms of low cost, temperature stability, reversible reaction, and so on. Oxide materials have been intensively studied to suppress the defects and electronic charge carriers for many electronic device applications, but the studies with a high concentration of defects are limited. It desires to improve thermoelectric performance by enhancing its charge transport and lowering its lattice thermal conductivity. For this purpose, here, we modified the stoichiometry of cation and anion vacancies in two different systems to regulate the carrier concentration and explored their thermoelectric properties. Both cation and anion vacancies act as a donor of charge carriers and act as phonon scattering centers, decoupling the electrical conductivity and thermal conductivity.

Keywords

  • thermoelectrics
  • nonstoichiometry
  • defect
  • phonon scattering
  • conductive oxide
  • charge transport

1. Introduction

The demand for alternate energy of fossil fuel becomes a challenging task for the researcher and scientist. A possible strategy for an alternate energy source is thermoelectric (TE) materials, whereby unwanted heat is changed to useful electrical energy with no harmful emissions compared to other traditional power plants [1, 2]. The performance of a TE materials can be evaluated by the figure of merit, zT = S2σT/κ, where σ represents the electrical conductivity, S is the Seebeck coefficient, T is the temperature, and κ is the thermal conductivity [3, 4]. The high zT can be obtained with high electrical conductivity, high Seebeck coefficient, and low thermal conductivity. However, the correlations between them are complex and cannot be treated independently [5, 6]. For instance, the increase in the carrier concentration, rises the electrical conductivity, drop the Seebeck coefficient, and rise the electronic thermal conductivity [7]. Therefore, the optimization of TE performance is a challenging task.

To date, many innovative TE materials like Bi2Te3, PbTe, CoSb3 XNiSn, and SiGe have been commercially applied because of their high performance as compared to other TE materials [5, 8, 9, 10, 11, 12]. However, these materials have restricted applications due to their high price, instability in oxidizing atmospheres, and the most important toxicity [8, 13, 14, 15]. Therefore, many studies on alternative TE materials with low cost, high efficiency, and environmentally friendly characteristics have been explored [16, 17, 18]. In this relay, oxide materials have been considered as the best alternative, such as layered cobalt oxide (NaCo2O4), and strontium titanium oxide (SrTiO3) due to their low price, thermal stability, and eco-friendly compatibility [19, 20]. For TE applications, n-type and p-type materials should be coupled. So it is important to develop high performance both n-type and p-type materials.

Layered structure p-type NaCo2O4 has been explored and is considered as one of the candidate materials for TE applications [16, 21]. This layered structure contains two layers CoO2 and Na. The CoO2 layers play the role of electron source which results in high electrical conductivity, while the layer of Na ions is sandwiched among nearby CoO2 layers which decreases the thermal conductivity along the stacking direction [22, 23]. Various studies propose that such layered structured CoO2 exhibits low thermal conductivity with metallic-like electrical conductivity, which is very attractive for TE applications [16, 24].

Similar to highly studied NaCoO2 TE material, LiNbO2 has a layered configuration in which the NbO6 trigonal-prismatic layers and Li planes are stacked, as shown in the schematic Figure 1(a). This similarity proposes that LiNbO2 could be a new promising TE material. LiNbO2 is a sub-oxide of the main (LiNbO3), and its Li-intercalated structure Li1-xNbO2 was first explored as a promising superconductor [25, 26, 27]. It has also been considered as a potential candidate material for numerous technological applications [28, 29, 30, 31]. The removal of Li atoms provide additional holes to the valence band, made up of Nb dZ2 states, which affects the oxidation of the Nb-atoms and raises the density of states (DOS) at the Fermi level (just like when holes are introduced into NaCoO2). These Li-vacancies increase the carrier concentration, which will raise the electrical conductivity. Moreover, the intrinsic defects in Li1-xNbO2 would act as a scattering center for thermal conductivity. So, defected Li1-xNbO2 is estimated to be extremely favorable for TE applications because of higher electrical and lower thermal conductivities.

Figure 1.

Schematic illustration of (a) LiNbO2 (b) SrTiO3.

Stoichiometric SrTiO3 has a cubic perovskite structure, where oxygen anions form an octahedron with one Ti4+ atom lying at the center as shown in the schematic Figure 1(b). SrTiO3 have been studied widely and is considered as one of the favorite n-type TE materials due to its high absolute Seebeck coefficient [32, 33]. Virgin SrTiO3 is considered to be an insulator and with a bandgap of 3.25 eV [34]. To boost the power factor (PF), only the electrical conductivity needs to be increased through appropriate doping on A-site and/or B-site [19, 35, 36, 37, 38, 39]. The role of the oxygen vacancy, which also acts as electron dopants, is very important for TE materials. Besides, creating oxygen vacancies offer a chance to decrease thermal conductivity through phonon scattering without remarkably affecting electrical conductivity [40, 41, 42]. The cationic nonstoichiometry and controlling oxygen partial pressure in SrTiO3 can produce cation and oxygen vacancies, which play an important role in the TE performance. To understand the defect chemistry and its consequence on the TE performance, pure SrTiO3 should be considered.

Considering the importance of defects in oxide materials as discussed above, it is important to study cation and anion defect engineering in oxide materials. Here in this chapter we have considered LiNbO2 (p-type) and SrTiO3 (n-type) with different vacancy concentrations and explore the experimental observations and correlate them with density functional theory (DFT). We elucidate that the defect engineering which may provide a new track for enhancing the TE performance.

Advertisement

2. Cation defect engineering

2.1 Experimental and computational approaches

2.1.1 Preparation of Li1-xNbO2 compounds

Nonstiochiometric Li1-xNbO2 (x = 0–0.6) compounds were prepared by conventional solid-state reaction using commercially available Li2CO3, NbO, and Nb2O5 with purity level more than 99.99%. First, Li3-yNbO4 was prepared by mixing Li2CO3 and Nb2O5 (y = 0, 0.015, 0.3, 0.6, 0.9, 1.2, and 1.8) at 1173 K in air for 50 hrs. Next, Li1-xNbO2 (x = 0–0.6) were prepared by mixing Li3-yNbO4 (y = 0, 0.015, 0.3, 0.6, 0.9, 1.2, and 1.8) for corresponding x = (0, 0.05, 0.1, 0.2, 0.4 and 0.6), and NbO in a ratio of 1: 2 for 24 hrs in ethanol. The dried slurries were crushed, sieved, and consulidated by spark plasma sintering (SPS) at 1323 K under 50 MPa for 15 min [43].

2.1.2 Theoretical calculations of Li1-xNbO2

The electronic structures of Li1-xNbO2 were explored using DFT with the local density approximations [44]. The lattice constants were calculated and all the atomic sites were relaxed till the forces were met to less than 0.01 eV/Å. For Li-vacancies computations various supercells (1x1x1, 1x1x2, 2x2x1, and 2x2x2) were considered. For the TE calculations, the BolzTraP package based on the Boltzmann transport theory and a constant relaxation time were applied [45, 46]. For band structures calculations QE code and the energy dispersion relation E(n,k) as a function of band index n and wave vector k were used.

2.1.3 Characterization of Li1-xNbO2

X-ray diffraction (XRD) analysis was performed by Rigaku D/MAX-2500/PC with Cu Kα emission. The pictographs of the samples were observed by using the scanning electron microscope (SEM, Verios 460 L, FEI). Rectangular specimens were cut for the measurements of the electrical properties using (ZEM-3, ULVAC-RIKO). Temperature-dependent charge transport properties were measured using HT-Hall, Toyo Corporation, ResiTest 8400. Circular discs were used for thermal diffusivity measurement (DLF-1300, TA instrument). All samples have ≥95% of the theoretical density.

2.2 Results and discussion

Figure 2(a) displays the XRD patterns of all samples. The main peaks of the compounds were indexed according to the LiNbO2 hexagonal structure which can be regarded as alternatively arranged close-packed Li-layers inserted among the two O-Nb-O slabs along the c-axis [29]. In addition to the main peaks, all samples showed small impurity peaks of LiNbO3 and NbO2. LiNbO3 peaks are expected due to moderate PO2 level during the consolidation process, which changes the oxidation state from Nb3+ to Nb5+. The additional NbO2 peaks can be described by the following defect reaction (1).

Figure 2.

(a) X-ray diffraction patterns and (b-g) microstructure for Li1-xNbO2 samples [43].

2NbO+Li3NbO43LiNbO2E1

At lower Li-vacancy concentrations, the Li atoms to combine with Nb3+ ([1/3Nb5++2/3Nb2+]3+) make LiNbO2. But, at high Li-vacancy concentrations, there are insufficient Li-atoms to react with Nb3+ ([1/3Nb5++2/3Nb2+]3+) atoms to form LiNbO2 which turn into the NbO2 phase. Furthermore, at lower oxygen partial pressure the Nb5+ in the Li3NbO4 compound is thermally reduced to Nb4+ which is the consequence of the NbO2 phase. Additionally, the Li-vacancies lead to an increase in the repulsive force among the two adjacent oxygen layers which increases the c-lattice of the unit cell. In the meantime, the a lattice decreases due to the shrinkage of Nb-O bonds in NbO6 octahedra. The experimental lattice constant shown in the inset Figure 2(a) is close to reported work and with our DFT results [47, 48].

Figure 2(b)(f) displays the fractured cross-section of Li1-xNbO2 samples. The dense of all the samples support the high relative densities larger than 95%. All grains are homogeneously distributed, no obvious segregations, and are randomly oriented. Moreover, the Li-vacancy concentrations have no significant effects on the size and shape of the grains.

Figure 3(a) shows the temperature-dependent electrical conductivities of Li1-xNbO2 samples. At first, the electrical conductivities of the samples increased with Li-vacancy concentrations, suggesting that more holes are created, as shown in the defect reaction [2]. The decreasing trends in the electrical conductivity measurements with temperature suggesting metallic behavior. However, for higher Li-vacancies concentrations, i.e., x ≥ 0.4, and at high temperatures the metallic conduction changed to semiconducting-like behavior. For high Li-vacancy concentrations, this transition point is moved to lower temperatures, suggesting that hole creation is inhibited by electrons formed by the replacement of Nb with Li atoms. Above this transition point, the electrical conductivity is governed by electrons which can be described by the defect reactions displayed in Eqs. (3)(5).

Figure 3.

(a) Temperature dependence of electrical conductivity and (b) calculated electronic density of states of Li1-xNbO2 (x = 0, 0.2, 0.4, 0.6, and 0.6) [43].

Li2O+Nb2O32VLi+2h.+2NbNbX+4OOXE2
Li2O+Nb2O3VLi+NbNb..+LiLiX+NbNbX+4OOXE3
Li2O+Nb2O3NbLi..+VLi+e+NbNbX+4OOXE4
Li2O+Nb2O3NbLi..+2e+LiLiX+NbNbX+4OOXE5

To deeply understand the above experimental results DFT calculations were also employed and the electronic structures were calculated. The calculated electronic structure suggests that virgin LiNbO2 has a band-gap of 1.65 eV. It should be noted that DFT-LDA generally miscalculates the band-gap. But, the calculated band-gap in our work is consistent with the reported works [45]. Figure 3(b) summarizes the electronic DOS for various holes concentrations. It suggests that virgin LiNbO2 is a semiconductor and suggesting a metallic behavior for holes incorporated samples. Besides, the DOS at the Fermi energy also increases, suggesting that the electrical conductivity of Li1-xNbO2 should increase with increasing holes concentrations.

Figure 4(a) shows the temperature-dependent Seebeck coefficients (S) for all samples. It can be seen that both the calculated and experimentally observed S are very close. The positive sign indicates that holes are the majority carrier which can be connected to the deviation from stoichiometry in the Li- sublattice [49]. The S values of x ≤ 0.1 samples increase with temperature, showing a degenerately doped semiconductor. Sample with x = 0.2 and at high temperatures the increasing trend in the S values is low suggesting that the hole generation is suppressed by an electron. Additional increase in the Li-vacancy concentrations (x > 0.2) and at high temperatures, the S values starts decreasing, and finally, a changeover is observed. This changeover from p-type to n-type could be due to cation disorder which is very similar to the σ-T as displayed in Figure 3(a). The behavior could be clearly understood by the defect reaction (2)(5).

Figure 4.

Temperature dependence of (a) Seebeck coefficient, (b) carrier concentration, (c) carrier mobility, and (d) normalized mobility as a function of Tr1r=12and12,ofLi1xNbO2 [43].

Figure 4(b)(d) shows the temperature-dependent Hall measurements for all Li1-xNbO2 samples. The obtained by the Hall measurements agreed with the electrical conductivity and Seebeck coefficient data. It can be seen clearly that the carrier concentration increases with the creations of Li-vacancies 2VLi2h.. Samples with lower Li-vacancy concentrations (x ≤ 0.2) were not influenced by temperature, suggesting degenerately doped behavior. However, high Li-vacancy concentrations (x ≥ 0.4), and at high temperatures show a slight rise in the carrier concentration. This tendency is similar to the Seebeck coefficients as shown in Figure 4(a). In the case of a degenerately doped p-type semiconductor, the Fermi level lies below the valence band maxima (VBM) and the carrier concentration is temperature independent of up to the intrinsic-extrinsic transition temperature [50]. Therefore, it is expected that Li1-xNbO2 is a heavily doped semiconductor. Figure 4(c) shows the temperature dependence mobility of nonstoichiometric Li1-xNbO2 compound and the carrier mobility decreases with increasing temperature. Together with the Hall measurements and the defect reactions (2)(5), it is expected that the tendency in electrical conductivity is directed by mobility in the region holes are dominant and by carrier concentration in the region where electrons are dominant (x ≥ 0.4 and at high temperature). All samples show negative temperature-dependence carrier mobility, resulting from the phonon scattering (μ ∝ Tr−1), where r=12,12,and32 signify acoustic, optical, and ionized impurity phonon scattering, respectively [51, 52, 53]. To know the scattering mechanism in Li1-xNbO2, mobilities were re-plotted as a function of Tr−1 and shown in Figure 4(d). The samples x ≤ 0.2 shown a linear correlation with T−1/2, suggesting that mobility is dominantly encountered by optical phonon scattering. However, samples with x ≥ 0.4, as shown in Figure 4(d) inset, displayed a linear relationship with T−3/2, suggesting that the mobility is encountered by acoustic phonon scattering. The transition mechanism is not clear, however, it could be due to the dominant defect change with Li-vacancies, corresponding to Equations (2) and (3), to Nb replacements for Li sites, corresponding to Eqs. (4) and (5).

Figure 5(a) represents the temperature-dependent total thermal conductivities (κtot) of Li1-xNbO2 samples. It can be seen that total thermal conductivity is decreased significantly with increasing Li-vacancies. The Wiedemann-Franz relationship was used to estimate the electronic and lattice thermal conductivity as shown in the inset of Figure 5(a) [54]. It is found that the main influence on the total thermal conductivity originates from lattice vibrations and decreased substantially with Li-vacancies. This suggests that Li-vacancies act as a scattering center for phonons. However, for higher vacancy concentrations (x ≥ 0.2), the thermal conductivity increases to some extent. This increase could be due to a change in the scattering mechanism from optical phonon to acoustic phonon by NbLi contribution as described above. This mechanism is schematically illustrated in Figure 5(b)(d). Therefore, it could be suggested that samples with low Li-vacancies concentrations may have localized double phonon (optical and acoustic phonons), which may play a role as a phonon scattering center, showing a substantial decrease of lattice thermal conductivity. However, with higher Li-vacancies concentrations, the lattice vibration is scattered by a single phonon type of acoustic phonon. Also, the increase in the thermal conductivity could be due to secondary phases [55] as detected in XRD (see Figure 2(a)).

Figure 5.

Temperature-dependent (a) total thermal conductivities for Li1-xNbO2 sample and (b-d) schematic illustration of phonon scattering [43].

The figure-of-merit (zT) for all Li1-xNbO2 samples is presented in Figure 6. A substantial improvement in zT is observed in the whole temperature range and reaches a maximum value of 0.125 at 970 K, which is around ~220% (3-times) higher compared to the virgin LiNbO2 sample. The observed tendency suggests that zT could be higher at higher temperatures. The samples after high-temperature measurement were rechecked by XRD and found that all Li1-xNbO2 samples are highly stable as shown in the inset of Figure 6. This confirms that all samples are stable and can be utilized as a new promising material for high-temperature TE applications.

Figure 6.

Temperature dependence of the dimensionless figure of merit of Li1-xNbO2 samples [43].

Advertisement

3. Anion defect engineering

3.1 Experimental and computational methods

3.1.1 Preparation of SrTiO3-δ

Pristine SrTiO3 samples were synthesized by using conventional solid-state reaction techniques, using TiO2 and SrCO3 with a purity level higher than 99.9%. The stoichiometric powders were ball-milled for 24 hr. Next, the ball-milled powders were calcined at 1373 K for 3 hrs. The powders were then sieved and pressed into thicknesses of 3 mm. The pressed pellets were sintered at 1573 K for 30 hrs under the air atmosphere. To create an Anion defect in SrTiO3 (Oxygen vacancies), the samples were annealed by allowing 1%H2/Ar, 5%H2/Ar, 10%H2/Ar, and 20%H2/Ar gases at 1573 K for 30 hrs, and the samples were designated with the prefix “1HAr, 5HAr, 10HAr, and 20HAr”, respectively.

3.1.2 Theoretical calculations of SrTiO3-δ

To understand the anion defect engineering in SrTiO3, DFT with the local density approximations were applied and correlated with experimental work [44, 56]. To properly pronounce the electronic band structures of pristine and O-vacancies in SrTiO3, we also considered the LDA + U methodology by selecting the effective on-site Coulomb modification (U = 5.0 eV) is applied to d-orbital electrons in Ti-atom in agreement with the reported works [57, 58]. To compute the oxygen vacancy in SrTiO3, a 2 × 2 × 2 supercell was considered and the BoltzTraP program was used for TE properties [45, 46].

3.1.3 Charascterization of SrTiO3-δ

For SrTiO3-δ characterizations, the same techniques were followed as described in section 2.1.3.

3.2 Results and discussion

Figure 7 shows the XRD pattern of the reduced SrTiO3-δ samples. The XRD patterns of all reduced samples exhibit a single-phase perovskite cubic structure. The lattice constant of the reduced samples was refined through Rietveld refinements. The lattice constant increased with decreasingPO2levels. The variations in the lattice constant are similar to the reported values [36, 59]. The variations in the lattice constant indicate that during the annealing process, the oxygen in SrTiO3 sublattice combine with H2 which increases the O-vacancies among the cations, and thus the coulombic repulsion force of cations increases. This leads to a rise in the lattice constant as presented in Table 1.

Figure 7.

Refined XRD patterns of oxygen-deficient SrTiO3-δ. ceramics annealed at 1573 K for 30 hrs in (a) 1%H2/Ar (1HAr), (b) 5%H2/Ar (5HAr), (c) 10%H2/Ar (10HAr), and (d) 20%H2/Ar (20HAr) atmospheres [60].

SampleLattice parameter (Å)Carrier concentration
(×1019 cm−3)
Oxygen vacancies (×1019 cm−3)Mobility
(cm2V−1 s−1)
Estimated PO2
(atm)
[1HAr]3.89910.1060.05306.3162.24 × 10−11
[5HAr]3.89990.1980.0995.3737.78 × 10−15
[10HAr]3.90125.242.624.8852.23 × 10−23
[20HAr]3.904710.6605.333.6943.17 × 10−25

Table 1.

The carrier concentration, the lattice constant, oxygen vacancies, mobility, and estimatedPO2of SrTiO3−δ samples [60].

Figure 8 shows the microstructure of thermally etched SrTiO3-δ samples in high vacuum conditions. It shows that all samples have similar microstructures, well densified, and support the measured densities>95%. There is no substantial variance in the size and shape of the grains.

Figure 8.

SrTiO3-δ. ceramics annealed at 1573 K for 30 hrs in (a) 1%H2/Ar (1HAr), (b) 5%H2/Ar (5HAr), (c) 10%H2/Ar (10HAr), and (d) 20%H2/Ar (20HAr) atmospheres [60].

Figure 9 displays the high-resolution transmission electron microscopy (HRTEM) of oxygen-deficient SrTiO3-δ samples. It is interesting to see that the sample reduced under different PO2 levels shows different reduction levels. In the case of low oxygen-deficient samples, we could not see a clear defective area. However, for the high oxygen-deficient samples, we identify two different kinds of defected areas labeled by A1 and A2 in Figure 9(c,d). The area labeled by A1 denotes the local lattice defect which is due to high oxygen vacancy in the lattice and area A2 shows a different lattice structure from the matrix. It is expected that the area A2 is probably related to the Ruddlesden-Popper phases [SrO·(SrTiO3)n] [61], which could be the reason for the metallic conduction in highly reduced samples.

Figure 9.

High-resolution transmission electron microscopy for 1HAr (a, b) and 20HAr (c, d) SrTiO3-δ [60].

Annealing under a reducing atmosphere at ambient temperature leads to different kinds of defects in materials. Oxygen vacancies could be one of the main defects in the SrTiO3 due to the low formation-energy compared to other atoms in the lattice. Below Eq. (6) shows in the Kröger-Vink notation [62].

OOx12O2+VO+2eE6

where VO denotes the doubly ionized oxygen vacancy and OOx denotes the neutral oxygen that exists on its lattice position. The quasi-free electrons generated by oxygen vacancies was determined using the Hall measurements and then using the free electrons, the oxygen partial pressure PO2 was estimated using Eq.(7) (See Table 1) [63].

PO2=M213K013expΔhRe3kBTne6E7

where ∆hRe represents the enthalpy for reduction, M is the number of oxygen atoms in unit volume (cm−3), kB is the Boltzmann’s constant, K0 is a constant including an entropy term, ne is nominal carrier concentration, and T is the annealing temperature.

The observed carrier concentrations in our samples are in the range of 1018 to 1020/cm3, which is a typical range for Mott transition, i.e., insulator–metal transition. This transition can be calculated from the Mott criterion, ne1/3a0 ~ 0.25, where a0 is the Bohr radius related to the carrier and ne which represents the electronic carrier concentration [64]. In the case of 1HAr and 5HAr samples, the calculated Mott transition is on the boundary of Mott transitions, and 10HAr and 20HAr samples are metallic-like, which supports the temperature-dependency of electrical conductivity.

The temperature-dependent electrical conductivity and Seebeck coefficient of reduced SrTiO3-δ samples are represented in Figure 10(a) and (b). As abovementioned that each O-vacancy produces two electrons through charge neutrality conditions, ne2VO, and henceforth the carrier concentrations would increase, causing an enhance in the electrical conductivity [63]. Moreover, as shown in Figure 10(a) that the electrical conductivities for low reduced samples i.e., 1HAr and 5HAr-annealed are low (0.5 to 2 S/cm, correspondingly) in the entire temperature with semiconducting-like. However, in the case of highly reduced samples (increasing O-vacancies (δ) as shown in Table 1) i.e., 10HAr and 20HAr samples, the electrical conductivities are relatively higher (41 and 63 S/cm, correspondingly). This increase of electrical conductivity evidently suggests electron doping due to oxygen vacancy, which is analogous to donor-doped SrTiO3 [65]. This would fill the n-type conduction band (CB), which evident by the negative sign of the Seebeck coefficient as shown in Figure 10b. Additionally, the increase in n-type charge carrier from 1HAr to 20HAr is consistent with the electrical conductivity and Seebeck coefficient data.

Figure 10.

(a) Temperature-dependent electrical conductivity and (b) Seebeck coefficient of O-deficient SrTiO3-δ, (c) band structure, and (d) total and projected density of states of SrTiO3-δ [60].

DFT calculations were also used to understand the effect of O-vacancies in SrTiO3. The electronic band structures of virgin and O-deficient SrTiO3 samples were calculated using DFT + U, and the results are presented in Figure 10(c). Our DFT + U calculated electronic band-gap results suggest that of virgin SrTiO3 at Г-point is 2.31 eV (1.93 eV), which is close reported work [34, 66]. However, in the case of oxygen-deficient SrTiO3, the electronic band-gap is reduced to 2.18 eV at Г-point. The band structure calculations show a band below the conduction band, and an electron pocket which can be seen at the Г- point. Such an electronic pocket suggests electrons in the conduction band, which is largely formed by the Ti-d electrons. This electron pocket not only decreases the bandgap but also raises the carrier concentrations that can additionally improve the electrical conductivity of SrTiO3-δ. The projected density of states in Figure 10(d) shows that the electron pockets are primarily contributed by the Ti-d and Sr-p electrons nearby the oxygen vacancy. We also found that at a large O-vacancy concentration, an insulator–metal transition occurs. In addition to this, the charge transfer study also suggests that O-vacancies change the conduction mechanism from nondegenerate to degenerate. We also calculated the room temperature DOS effective mass md using nondegenerate and degenerate models Eqs. (8) and (9).

S=kBelnNc/ne+AE8
S=8πkB23eh2π3n23mdTE9

where kB represents the Boltzmann constant, h is the Planck constant, n is the carrier concentration, NcT=22πkBTm/h23/2 is the effective density of states in the conduction band, A is the scattering factor, usually ranges between 0 and 4. By fitting Eq. (8) as shown in Figure 11a with A = 1, 2, and 3, it can be seen that the effective mass for low O-vacancies samples suggests a nondegenerate semiconductor. However, the Pisarenko relation fitting Eq. 9 suggests that high O-vacancies samples have degenerate semiconducting like behavior. Furthermore, the DOS effective mass rises with increasing O-vacancies as shown in Figure 11b. This increase from 0.43 m° to 4.4 m° could be a reason for Hall mobilities reduction as shown in Table 1.

Figure 11.

(a) Seebeck coefficients as a function of carrier concentration: The solid lines represent the estimated Seebeck coefficients using a degenerate semiconducting model and dashed lines for a nondegenerate semiconducting model. (b) DOS effective mass mdresulting from Seebeck coefficient and carrier concentration [60].

Figure 12(a) represents the temperature-dependent thermal conductivity of SrTiO3-δ samples. This decreasing tendency suggests that O-vacancies act as phonon scattering centers despite their electrical conductivity increased as shown in Figure 10(a). Using Wiedemann-Franz law relation with Lorentz number of (2.45 × 10−8 V2K−2), the influence of electronic (κe) and phononic (κL) to total thermal conductivity was calculated and observed that the lattice thermal conductivity is dominant as presented in Figure 12(b). The lattice thermal conductivity was also plotted as a function of T−1 as shown in inset Figure 12(b). The linear connection proposes that the lattice thermal conductivity is affected dominantly by Umklapp scattering [67]. Also, one should note that the similar grain sizes as shown in Figure 8, suggests that the reduction in thermal conductivity is mainly due to oxygen vacancy rather than grain boundary scattering.

Figure 12.

Temperature-dependent (a) total thermal conductivity, (b) lattice thermal conductivity, (c) power factor, (d) DFT calculated powder factor for various carriers concentrations, (e) temperature-dependent ratio of the electrical to thermal properties (σ/κ), and (d) ZT values of various oxygen-deficient SrTiO3 samples [60].

Figure 12(c) illustrates the temperature-dependent PF for various oxygen-deficient SrTiO3 samples. The substantial enhancement in the PF is due to the combination of moderate Seebeck coefficient values and high electrical conductivity. The PF obtained in this work is comparable to the PF obtained in doped-SrTiO3 [65]. The DFT calculated PF for various carrier concentrations (1018–1022/cm3) suggests that the highly reduced SrTiO3 samples would have higher PF as shown in Figure 12(d) [68]. Additionally, one of the most remarkable features of this study is the decoupling among electrical and thermal conductivity of SrTiO3-δ through solely oxygen vacancy as shown in Figure 12(e). This opposite behavior led to a significant increase in the ratio of electrical to thermal conductivity (σ/κ), which is a key motive for TE materials. The complete outcome of oxygen vacancies in the TE materials can be represented by the temperature dependence of the ZT as shown in Figure 12(f). It can be observed that the introduction of oxygen vacancy substantially increases the ZT value of 4.7 × 10−2 for [20HAr] reduced samples at 670 K. Therefore, we suggest that carefully adjusting the oxygen vacancy could be an effective strategy for oxide TE materials.

Table 2 shows that the efforts in the layer-structured cobaltites and SrTiO3 based materials. It shows that significant improvements have been achieved in oxide TE materials, which could be of great interest for power generation applications at high operating temperatures. As compared with other oxides, the materials investigated in this study show relatively low figure of merit (zT), which is because the investigated Li1-xNbO2 and SrTiO3-δ are pure, undoped materials to understand the mechanism for cation and anion defects effects.

MaterialsTypeMaximum zT valueWorking temperature (K)References
Cd-doped Ca3Co4O9p0.351000[69]
NaxCoO2p0.31800[70]
NaxCoO2/Ag/Aup0.4–0.51000[71]
Li1-xNbO2p0.125970This work
Nb-doped SrTiO3n0.165900[72]
La-doped SrTiO3single crystaln0.271073[73]
Pr-doped SrTiO3n0.35773[74]
SrTiO3-δn4.7 × 10−2670This work

Table 2.

The dimensionless figure of merit zT of p-type and n-type oxide thermoelectric materials for comparison.

3.3 Conclusion

In conclusion, this work demonstrated that cation and anion vacancies can successfully control the thermoelectric performance of oxide-based thermoelectric materials. These findings suggest that both cation defect and anion defect can be engineered by reducing atmospheres and the defects in oxide thermoelectric materials simultaneously act as a source of charge carriers and phonon scattering centers. This decoupled behavior between electrical conductivity and thermal conductivity can lead to a substantial increase in the thermoelectric performance of oxide materials. The concept applied in this work is generally important and has the possibility of impacting the thermoelectric performance of oxide thermoelectrics and other functional oxide materials.

Advertisement

Acknowledgments

The experimental work was conducted by using the facilities in the Korea Institute of Ceramic Engineering and Technology (KICET). The authors, Jamil Ur Rahman and Soonil Lee are grateful to all colleagues for help and support in the same research group at KICET.

Advertisement

Conflicts of interest

There are no conflicts of interest to declare.

References

  1. 1. G. Mahan, B. Sales, J. Sharp, Thermoelectric materials: New approaches to an old problem. Physics Today50, 42-47 (1997).
  2. 2. M. S. Dresselhaus et al., New Directions for Low-Dimensional Thermoelectric Materials. Adv. Mater.19, 1043-1053 (2007).
  3. 3. L.-D. Zhao, V. P. Dravid, M. G. Kanatzidis, The panoscopic approach to high performance thermoelectrics. Energy Environ. Sci.7, 251-268 (2014).
  4. 4. T. M. Tritt, M. Subramanian, Thermoelectric materials, phenomena, and applications: a bird's eye view. MRS bull.31, 188-198 (2006).
  5. 5. G. J. Snyder, E. S. Toberer, Complex thermoelectric materials. Nat. Mater.7, 105-114 (2008).
  6. 6. A. Ioffe, Semiconductor Thermoelements and Thermoelectric Cooling (London: Infosearch, 1957). Google Scholar, pp. 36-38.
  7. 7. Z. Dughaish, Lead telluride as a thermoelectric material for thermoelectric power generation. Physica B: Condens. Matter322, 205-223 (2002).
  8. 8. G. Nolas, J. Cohn, G. Slack, S. Schujman, Semiconducting Ge clathrates: Promising candidates for thermoelectric applications. Appl. Phys. Lett.73, 178-180 (1998).
  9. 9. J. M. Song et al., Chemically synthesized Cu2Te incorporated Bi-Sb-Te p-type thermoelectric materials for low temperature energy harvesting. Scr. Mater.165, 78-83 (2019).
  10. 10. J. S. Yoon et al., High thermoelectric performance of melt-spun CuxBi0.5Sb1.5Te3 by synergetic effect of carrier tuning and phonon engineering. Acta Mater.158, 8 (2018).
  11. 11. N. Van Du et al., X-site aliovalent substitution decoupled charge and phonon transports in XYZ half-Heusler thermoelectrics. Acta Mater.166, 650-657 (2019).
  12. 12. N. Van Du et al., Synthesis and thermoelectric properties of Ti-substituted (Hf0. 5Zr0. 5) 1-xTixNiSn0. 998Sb0. 002 Half-Heusler compounds. J. Alloys Compd.773, 1141-1145 (2019).
  13. 13. B. Sales, D. Mandrus, R. K. Williams, Filled skutterudite antimonides: a new class of thermoelectric materials. Science272, 1325-1328 (1996).
  14. 14. H. Ohta, Thermoelectrics based on strontium titanate. Mater. Today10, 44-49 (2007).
  15. 15. S. Lee, R. H. T. Wilke, S. Trolier-McKinstry, S. Zhang, C. A. Randall, SrxBa1-xNb2O6-δ Ferroelectric-thermoelectrics: Crystal anisotropy, conduction mechanism, and power factor. Appl. Phys. Lett.96, 031910 (2010).
  16. 16. I. Terasaki, Y. Sasago, K. Uchinokura, Large thermoelectric power in NaCo2O4 single crystals. Phys. Rev. B56, R12685 (1997).
  17. 17. T. Kawata, Y. Iguchi, T. Itoh, K. Takahata, I. Terasaki, Na-site substitution effects on the thermoelectric properties of NaCo2O4. Phys. Rev. B60, 10584 (1999).
  18. 18. K. Fujita, T. Mochida, K. Nakamura, High-Temperature Thermoelectric Properties of NaxCoO2−δ Single Crystals. Jpn. J. Appl. Phys.,40, 4644-4647 (2001).
  19. 19. S. Ohta, T. Nomura, H. Ohta, K. Koumoto, High-temperature carrier transport and thermoelectric properties of heavily La- or Nb-doped doped SrTiO3 single crystals. J. Appl. Phys.97, 34106-034110 (2005).
  20. 20. Y. Wang, N. S. Rogado, R. J. Cava, N. P. Ong, Spin entropy as the likely source of enhanced thermopower in NaxCo2O4. Nature423, 422-425 (2003).
  21. 21. K. Takada et al., Superconductivity in two-dimensional CoO2 layers. Nature422, 53-55 (2003).
  22. 22. M. Roger et al., Patterning of sodium ions and the control of electrons in sodium cobaltate. Nature445, 631-634 (2007).
  23. 23. H. Nakatsugawa, K. Nagasawa, Evidence for the two-dimensional hybridization in Na0.79CoO2 and Na0.84CoO2. J. Solid State Chem.177, 1137-1145 (2004).
  24. 24. K. Takahata, Y. Iguchi, D. Tanaka, T. Itoh, I. Terasaki, Low thermal conductivity of the layered oxide (N a,Ca)Co2O4: Another example of a phonon glass and an electron crystal. Phys. Rev. B61, 12551 (2000).
  25. 25. E. G. Moshopoulou, a. J. J. C. P. Bordet, Superstructure and superconductivity in Li1−xNbO2 (x≈0.7) single crystals. Phys. Rev. B59, 9590-9599 (1999).
  26. 26. N. Kumada, S. Watauchi, I. Tanaka, N. Kinomura, Superconductivity of hydrogen inserted LiNbO2. Mater. Res. Bull.35, 1743-1746 (2000).
  27. 27. E. R. Ylvisaker, W. E. Pickett, First-principles study of the electronic and vibrational properties ofLiNbO2. Phys. Rev. B74, 075104 (2006).
  28. 28. M. J. Geselbracht, T. J. Richardson, A. M. Stacy, Superconductivity in the layered compound LixNbO2. Nature345, 324-326 (1990).
  29. 29. N. Kumada et al., Topochemical reactions of LixNbO2. J. Solid State Chem.73, 33-39 (1988).
  30. 30. J. D. Greenlee, W. L. Calley, W. Henderson, W. A. Doolittle, Halide based MBE of crystalline metals and oxides. Phys. Status Solidi C9, 155-160 (2012).
  31. 31. J. U. Rahman et al., The Synthesis and Thermoelectric Properties of p-Type Li1−xNbO2-Based Compounds. J. Electron. Mater.46, 1740-1746 (2017).
  32. 32. T. Okuda, K. Nakanishi, S. Miyasaka, Y. Tokura, Large thermoelectric response of metallic perovskites: Sr1-xLaxTiO3 (0~x~0.1). Phys. Rev. B63, (2001).
  33. 33. J. U. Rahman et al., Grain Boundary Interfaces Controlled by Reduced Graphene Oxide in Nonstoichiometric SrTiO3-δ Thermoelectrics. Scientific Reports9, 8624 (2019).
  34. 34. K. Van Benthem, C. Elsässer, R. French, Bulk electronic structure of SrTiO3: Experiment and theory. J. Appl. Phys.90, 6156-6164 (2001).
  35. 35. H. Ohta, K. Sugiura, K. Koumoto, Recent progress in oxide thermoelectric materials: p-type Ca3Co4O9 and n-type SrTiO3. Inorg. Chem.47, 8429-8436 (2008).
  36. 36. H. Muta, K. Kurosaki, S. Yamanaka, Thermoelectric properties of reduced and La-doped single-crystalline SrTiO3. J. Alloys Compd.392, 306-309 (2005).
  37. 37. Y. Cui et al., Thermoelectric Properties of Heavily Doped n-Type SrTiO3 Bulk Materials. J. Electron. Mater.38, 1002-1007 (2009).
  38. 38. S. R. Sarath Kumar, A. I. Abutaha, M. N. Hedhili, H. N. Alshareef, Effect of oxygen vacancy distribution on the thermoelectric properties of La-doped SrTiO3 epitaxial thin films. J. Appl. Phys.112, 114104 (2012).
  39. 39. S. Ohta, H. Ohta, K. Koumoto, Grain size dependence of thermoelectric performance of Nb-doped SrTiO3polycrystals. J. Ceram. Soc. Jpn.114, 102-105 (2006).
  40. 40. Y. Lin et al., Thermoelectric power generation from lanthanum strontium titanium oxide at room temperature through the addition of graphene. ACS Applied Mater. Interfaces7, 15898-15908 (2015).
  41. 41. Y. Wang et al., Interfacial thermal resistance and thermal conductivity in nanograined SrTiO3. Appl. Phys Express3, 031101 (2010).
  42. 42. S. S. Kumar, A. Z. Barasheed, H. N. Alshareef, High temperature thermoelectric properties of strontium titanate thin films with oxygen vacancy and niobium doping. ACS Applied Mater. Interfaces5, 7268-7273 (2013).
  43. 43. J. U. Rahman et al., Localized double phonon scattering and DOS induced thermoelectric enhancement of degenerate nonstoichiometric Li1-xNbO2 compounds. RSC Adv.,7, 53255-53264 (2017).
  44. 44. P. Giannozzi et al., QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter21, 395502 (2009).
  45. 45. G. K. Madsen, D. J. Singh, BoltzTraP. A code for calculating band-structure dependent quantities. Comput. Phys. Commun175, 67-71 (2006).
  46. 46. J. M. Ziman, Principles of the Theory of Solids. (Cambridge University Press, 1979).
  47. 47. H. F. Roth, G. Meyer, Z. Hu, G. Kaindl, Synthesis, structure, and X-ray absorption spectra of LixNbO2 and NaxNbO2 (x ≤ 1). Z Anorg. Allg. Chem.619, 1369-1373 (1993).
  48. 48. A. Miura et al., Octahedral and trigonal-prismatic coordination preferences in Nb-, Mo-, Ta-, and W-based ABX2 layered oxides, oxynitrides, and nitrides. J. Solid State Chem.229, 272-277 (2015).
  49. 49. K. W. Lee, J. Kuneš, R. T. Scalettar, W. E. Pickett, Correlation effects in the triangular lattice single-band systemLixNbO2. Phys. Rev. B76, 144513-144511 (2007).
  50. 50. B. G. Streetman, S. Banerjee, Solid State Electronic Devices. (Prentice Hall New Jersey, 1980).
  51. 51. B.-L. Huang, M. Kaviany, Ab initio and molecular dynamics predictions for electron and phonon transport in bismuth telluride. Phys. Rev. B77, 125209 (2008).
  52. 52. H. Kitagawa, A. Kurata, H. Araki, S. Morito, E. Tanabe, Structure and carrier transport properties of hot-press deformed Bi0. 5Sb1. 5Te3. Phys. Status Solidi (a)207, 401-406 (2010).
  53. 53. T. Zhu et al., Hot deformation induced bulk nanostructuring of unidirectionally grown p-type (Bi, Sb) 2 Te 3 thermoelectric materials. J. Mater. Chem. A1, 11589-11594 (2013).
  54. 54. C. Kittel, Introd. Solid State Phys., (Wiley, 2005).
  55. 55. N. H. a. M. E. Fine, The Lorenz ratio and electron transport properties of NbO. J. Appl. Phys.52, 2876 (1981).
  56. 56. W. Kohn, L. Sham, doi: 10.1103/PhysRev. 140. A1133. Phys. Rev. A140, 113 (1965).
  57. 57. T. Mizokawa, T. Mizokawa and A. Fujimori. Phys. Rev. B51, 12 (1995).
  58. 58. S. Okamoto, A. J. Millis, N. A. Spaldin, Lattice relaxation in oxide heterostructures: LaTiO 3/SrTiO 3 superlattices. Phys. Rev. Lett.97, 056802 (2006).
  59. 59. H. Muta, K. Kurosaki, S. Yamanaka, Thermoelectric properties of rare earth doped SrTiO3. J. Alloys Compd.350, 292-295 (2003).
  60. 60. J. U. Rahman et al., Oxygen vacancy revived phonon-glass electron-crystal in SrTiO3. J. Eur. Ceram. Soc.,39, 8 (2019).
  61. 61. C. Xu et al., Formation mechanism of Ruddlesden-Popper-type antiphase boundaries during the kinetically limited growth of Sr rich SrTiO3 thin films. Scientific reports6, 38296 (2016).
  62. 62. F. Kröger, H. Vink, Relations between the concentrations of imperfections in crystalline solids. Solid State Phys.3, 307-435 (1956).
  63. 63. S. Lee, G. Yang, R. H. T. Wilke, S. Trolier-McKinstry, C. A. Randall, Thermopower in highly reduced n-type ferroelectric and related perovskite oxides and the role of heterogeneous nonstoichiometry. Phys. Rev. B79, 134110-134111 (2009).
  64. 64. P. Cox, Transition metal oxides. (Oxford University Press, 1992), pp. 158.
  65. 65. A. V. Kovalevsky, A. A. Yaremchenko, S. Populoh, A. Weidenkaff, J. R. Frade, Effect of A-Site Cation Deficiency on the Thermoelectric Performance of Donor-Substituted Strontium Titanate. J. Phys. Chem. C118, 4596-4606 (2014).
  66. 66. S. Piskunov, E. Heifets, R. Eglitis, G. Borstel, Bulk properties and electronic structure of SrTiO3, BaTiO3, PbTiO3 perovskites: an ab initio HF/DFT study. Computational Materials Science29, 165-178 (2004).
  67. 67. J. M. Ziman, Electrons and phonons: the theory of transport phenomena in solids. (Oxford university press, 1960).
  68. 68. J. Sun, D. J. Singh, Thermoelectric properties of n-type SrTiO3. APL Materials4, 104803 (2016).
  69. 69. S. Butt et al., Enhancement of thermoelectric performance in Cd-doped Ca3Co4O9 via spin entropy, defect chemistry and phonon scattering. J. Mater. Chem. A2, 19479-19487 (2014).
  70. 70. K. Fujita, T. Mochida, K. Nakamura, High-temperature thermoelectric properties of NaxCoO2-δ single crystals. Jpn. J. Appl. Phys.,40, 4644 (2001).
  71. 71. M. Ito, D. Furumoto, Effects of noble metal addition on microstructure and thermoelectric properties of NaxCo2O4. J. Alloys Compd.450, 494-498 (2008).
  72. 72. N. Wang, L. Han, H. He, Y. Ba, K. Koumoto, Effects of mesoporous silica addition on thermoelectric properties of Nb-doped SrTiO3. J. Alloys Compd.497, 308-311 (2010).
  73. 73. S. Ohta, T. Nomura, H. Ohta, K. Koumoto, High-temperature carrier transport and thermoelectric properties of heavily La-or Nb-doped SrTiO3 single crystals. J. Appl. Phys.97, 034106 (2005).
  74. 74. A. M. Dehkordi, S. Bhattacharya, J. He, H. N. Alshareef, T. M. Tritt, Significant enhancement in thermoelectric properties of polycrystalline Pr-doped SrTiO3−δ ceramics originating from nonuniform distribution of Pr dopants. Appl. Phys. Lett.104, 193902 (2014).

Written By

Jamil Ur Rahman, Gul Rahman and Soonil Lee

Submitted: 21 December 2020 Reviewed: 28 January 2021 Published: 03 March 2021