Next Article in Journal
Structural and Dielectric Impedance Studies of Mixed Ionic–Electronic Conduction in SrLaFe1−xMnxTiO6 (x = 0, 0.33, 0.67, and 1.0) Double Perovskites
Previous Article in Journal
Effect of Ni2+ Doping on the Crystal Structure and Properties of LiAl5O8 Low-Permittivity Microwave Dielectric Ceramics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improving of Thermoelectric Efficiency of Layered Sodium Cobaltite Through Its Doping by Different Metal Oxides

by
Natalie S. Krasutskaya
1,
Ekaterina A. Chizhova
1,
Julia A. Zizika
2,
Alexey V. Buka
3,
Hongchao Wang
4 and
Andrei I. Klyndyuk
1,5,*
1
Department of Physical, Colloid and Analytical Chemistry, Organic Substances Technology Faculty, Belarusian State Technological University, Sverdlova Str. 13a, 220006 Minsk, Belarus
2
Structural Analysis Sector, Physical-Technical Institute of the National Academy of Science of Belarus, Academician Kuprevich Str. 10, 220084 Minsk, Belarus
3
Department of Glass, Ceramics and Binding Materials Technology, Chemical Technology and Engineering Faculty, Belarusian State Technological University, Sverdlova Str. 13a, 220006 Minsk, Belarus
4
School of Physics, State Key Laboratory of Crystal Materials, Shandong University, Jinan 250100, China
5
Department of Physical Chemistry and Electrochemistry, Chemical Faculty, Belarusian State University, Nezavisimosti Av. 4, 220030 Minsk, Belarus
*
Author to whom correspondence should be addressed.
Ceramics 2025, 8(3), 86; https://doi.org/10.3390/ceramics8030086
Submission received: 8 May 2025 / Revised: 19 June 2025 / Accepted: 1 July 2025 / Published: 5 July 2025

Abstract

Na0.89Co0.90Me0.10O2 (Me = Cr, Ni, Mo, W, Pb, and Bi) ceramic samples were prepared using a solid-state reaction method, and their crystal structure, microstructure, and electrical, thermal, and thermoelectric properties were investigated. The effect of the nature of the doping metal (Me = Cr, Ni, Mo, W, and Bi) on the structure and properties of layered sodium cobaltite Na0.89CoO2 was analyzed. The largest Seebeck coefficient (616 μV/K at 1073 K) and figure-of-merit (1.74 at 1073 K) values among the samples studied were demonstrated by the Na0.89Co0.9Bi0.1O2 solid solution, which was also characterized by the lowest value of the dimensionless relative self-compatibility factor of about 8% within the 673–873 K temperature range. The obtained results demonstrate that doping of layered sodium cobaltite by transition and heavy metal oxides improves its microstructure and thermoelectric properties, which shows the prospectiveness of the used doping strategy for the development of new thermoelectric oxides with enhanced thermoelectric characteristics. It was also shown that samples with a higher sodium content (Na:Co = 0.89:1) possessed higher chemical and thermal stability than those with a lower sodium content (Na:Co = 0.55:1), which makes them more suitable for practical applications.

Graphical Abstract

1. Introduction

Approximately two-thirds of the energy consumed by factories, transport, and households is dissipated into the environment and, in fact, lost for humanity [1]. This high-potential waste heat can be partially converted into electrical energy using thermoelectrogenerators (TEGs). To produce TEGs, one needs so-called thermoelectric materials, which should possess a complex of unique properties, such as low electrical resistivity (ρ) and thermal conductivity (λ), high values of Seebeck coefficient (S), etc. [2,3,4,5]. Traditional thermoelectrics based on layered chalcogenides of bismuth and lead possess several drawbacks, including the presence of toxic, rare, and expensive components, as well as instability in air at elevated temperatures. These drawbacks are absent for oxide thermoelectrics [6,7,8,9,10]. One of the well-known representatives of them is layered sodium cobaltite NaxCoO2, first described by Jansen and Hoppe [11] and characterized as thermoelectric by Terasaki et al. [12]. Its structure consists of [CoO2] layers (CdI2 structure) with sodium atoms in between [13]. According to Viciu et al. [14], the concentration of oxygen vacancies in the [CoO2] layers is negligible; therefore, the oxidation state of cobalt ions is determined only by the sodium content (xNa). Due to its unique transport properties, NaxCoO2 is considered as potential material for p-branches of TEGs and ceramic (oxide) thermocouples [15,16], cathode material of sodium-ion batteries (SIBs) [17,18,19,20], and its crystal hydrate, NaxCoO2·yH2O, below 4 K indergoes into a superconducting state [21]. The crystal structure, electrotransport, thermal, and functional characteristics of NaxCoO2 strongly depend on the sodium content [22,23,24,25].
The functional characteristics of the ceramic samples of the layered sodium cobaltite are worse than those of monocrystals; however, they can be improved using different approaches: (i) doping in sublattices of cobalt [26,27,28] or/and sodium [29,30,31], (ii) modification by particles of metals [32,33] or semiconductors [34], (iii) using so-called soft synthesis methods [7,35,36], and (iv) using special sintering techniques [37], resulting in the improvement of the ceramic microstructure which enhances its electrotransport and mechanical properties.
Partial substitution of cobalt by copper in NaCo2−xCuxO4 ceramics improves its sinterability, lowers electrical resistivity and enhances Seebeck coefficient, resulting in essential increase of power factor values of these solid solutions comparing to the unsubstituted sodium cobaltite [38,39], which reache maximal value for NaCo1.8Cu0.2O4 phase—3.08 mW/(m·K2) at 1073 K [38]. Substitution of cobalt by nickel deteriorates sinterability of NaCo2−xNixO4 phases, but decreases their ρ and strongly enlarges S, and p value for NaCo1.9Ni0.1O4 at 1073 K reaches 2.36 mW/(m·K2), which is 8 times larger than for the base NaCo2O4 phase [40,41]. When zinc substitutes cobalt in NaCo2−xZnxO4, the electrical resistivity and Seebeck coefficient of ceramics increase, and the power factor reaches 1.7 mW/(m·K2) at 1073 K for NaCo1.9Zn0.1O4 solid solution, which is 4 times larger than for NaCo2O4 cobaltite [42]. Co-substitution of Na by Bi and Co by Mn in Na0.81CoO2 decreases its ρ and increases its S, which results in a strong increase in the p values of ceramics, reaching 0.632 mW/(m·K2) at 781 K for Na0.66Bi0.15Co0.99Mn0.01O2 compound, which is three times larger than that of the parent Na0.81CoO2 phase [43].
The authors of [44] produced a π-shaped thermoelectric device consisting of Na0.50Co0.85Cu0.15O2 (p-leg) cobaltite and Ba0.2Sr0.8PbO3 (n-leg) plumbate, which produced an output power of 12 mW when the temperatures of the hot and cold sides were 1045 K and 539 K, respectively [44]. The power generation properties of this element did not change after high-temperature operation for seven days [44].
Earlier, we estimated the possibility of improving the thermoelectric performance of NaxCoO2 by increasing the sodium content in it [25] and by doping sodium-poor Na0.55CoO2 cobaltite with oxides of different transition and non-transition metals [16,45].
In this work, the effect of the nature of different transition or heavy metals partially substituting cobalt in the sodium-rich Na0.89CoO2 layered cobaltite on its crystal structure, microstructure, thermophysical, electrophysical, and thermoelectric (functional) properties was studied. The chemical and thermal stabilities of the obtained materials were also studied and compared with those of sodium-poor Na0.55Co0.90Me0.10O2 layered cobaltites [16]. It was found that the modification of NaxCoO2 by different metal oxides, along with an increase in sodium content from x = 0.55 to x = 0.89, resulted in an essential enhancement of its thermoelectric performance and chemical and thermal stabilities. This led us to consider the materials obtained in this work as attractive for practical applications in thermoelectric devices for different purposes. To the best of our knowledge, such complex studies have not been performed earlier, and the results of this work bring new knowledge in the field of thermoelectric materials science compared to previously published works [16,25,28,29,36,37,45].

2. Materials and Methods

Ceramic samples of Na0.89Co0.90Me0.10O2 (Me = Cr, Ni, Mo, W, Pb, and Bi) were obtained using a solid-state reaction method according to the following reaction scheme:
1.335 Na2CO3 + 0.3/x MexOy + 0.9 Co3O4 + (0.5325 − 0.15 y/x) O2 =
= 3 Na0.89Co0.90Me0.10O2 + 1.335 CO2
The substitution level of 10 mol. % was chosen as for lower substitution degree it can be hard to detect the effect of substitution on the structure and properties of the parent phase, but for the higher substitution degrees the solubility level can be exceeded [38,40], which results in the formation of not single-phase but composite ceramics.
The dopants were selected, firstly, to provide neutral (Cr3+), acceptor (Ni2+), and donor (Mo6+, W6+, Pb4+, Bi5+) character of substitution without the size factor effect due to the slight difference in the ionic radii values of substituted (Co3+, Co4+ in CoO6 octahedra (C.N. = 6) of NaxCoO2 crystal structure) and substituting ions [46], secondly, to use both transition (Cr,Ni) and heavy metals (W, Pb, Bi), and, thirdly, to demonstrate the effectiveness of some chosen dopants in enhancing the thermoelectric performance of layered sodium cobaltite [16,38,39,40,41,42,43].
The powders of Na2CO3 (99%), Co3O4 (99%), Cr2O3 (99%), NiO (99.99%), MoO3 (99.99%), WO3 (99.99%), PbO (99.99%), and Bi2O3 (99.99%) were taken in the ratio of Na:Co:Me = 1.2:0.9:0.1 (the excess of Na2CO3 in the initial mixture is taken to compensate for the loss of Na2O from the samples during their thermal treatment and to obtain ceramics of the specified composition [25,46]). The powdered samples were milled in a planetary mill PM 100 Retsch (material of beakers and grinding balls–ZrO2) for 90 min at 300 rpm with the addition of C2H5OH (~3–5 wt.%). The resulting slurries were air-dried at a temperature of 323 K and then pressed into tablets with a diameter of 19 mm and a thickness of up to 5 mm. Subsequently, the obtained samples were calcined in air for 12 h at a temperature of 1133 K. The calcination temperature of 1133 K was chosen to reduce the calcination time, as at this temperature, one of the initial reagents, Na2CO3, melts (Tm = 1127 K [47]). Subsequently, after calcination, the blanks were ground in an agate mortar and re-milled in a planetary mill, then pressed into bars measuring 5 × 5 × 30 mm3 and tablets with a diameter of 15 mm and a thickness of 2–3 mm, which were sintered in air for 12 h at a temperature of 1203 K [46]. A sintering temperature of 1203 K was selected based on the results of previous studies [48], as at lower sintering temperatures, samples possessed high porosity values, and at higher sintering temperatures, they partially melted and reacted with the crucible material (alundum).
The real sodium content (xNa) and average oxidation state of cobalt (Co+Z) in the obtained ceramic materials were determined using iodometric [48] and reverse potentiometric titrations [49], as well as spectrophotometrically, according to the method described in [50] (see Supplementary Information).
The phase composition of the samples and the calculation of the crystal lattice parameters of the Na0.89Co0.90Me0.10O2 layered cobaltites were performed by means of X-ray diffraction analysis (XRD) using a Bruker D8 XRD Advance X-ray diffractometer (Bruker Optik GmbH, Ettlingen (Germany)) (CuKα radiation (λ = 1.5406 Å)).
The values of the coherent scattering area for the Na0.89Co0.90Me0.10O2 ceramics were calculated using the Debye–Scherrer equation (DS) (2) [51] and size-strain method (DSS) (3) [52].
D S = 0.9 · λ β · cos Θ ,
d · β · cos Θ = 0 . 9 · λ D S S · d 2 · β · cos Θ + ε 2 2 ,
where d is the interplanar distance (nm), β is the full width of the reflex at its half maxima (rad), Θ is the diffraction angle (°), and ε is the microstrain.
The degree of crystallographic orientation of the grains of the Na0.89Co0.90Me0.10O2 ceramics was evaluated using the Lotgering factor, as shown in Equation (4) [53]:
f = p p 0 1 p 0 ,
where p = ΣI(00l)/ΣI(hkl) (ΣI is the sum of the counts of X-ray diffraction peaks of the synthesized samples); p0 = ΣI0(00l)/ΣI0(hkl) (ΣI0 is the sum of the counts of X-ray diffraction peaks of the reference phase JCPDC #00-030-1182).
The apparent density of the Na0.89Co0.90Me0.10O2 ceramic samples (dEXP) was determined based on their mass and geometric dimensions, and their total porosity (Πt) was calculated using the following formula:
Π t = 1 d E X P d X R D · 100 % ,
where dXRD is the X-ray density of the sample (g/cm3).
The open porosity of the sintered samples (ΠO) was determined by weighing the ceramic samples of Na0.89Co0.90Me0.10O2, which were saturated with ethyl acetate for 60 min under a vacuum of 25–30 Pa, and then weighed in ethyl acetate, according to Equation (6):
Π o = m 1 m m 1 m 2 · 100 % ,
where m is the mass of the dry ceramic sample (g), m1 and m2 are the masses of the sample saturated with liquid ethyl acetate during weighing in air and that introduced into liquid ethyl acetate, respectively (g).
The microstructure and chemical composition of the Na0.89Co0.90Me0.10O2 samples were studied using scanning electron microscopy (SEM) using a Tescan MIRA 3 LMH scanning electron microscope equipped with an AZtecLIVE Advanced energy dispersive microanalysis system with a nitrogen-free Ultim Max 100 standard detector (Oxford Instruments Analytical Ltd., High Wycombe (UK)).
The thermal expansion, electrical resistivity (ρ), and Seebeck coefficient (S) of the sintered ceramics were measured in air within a temperature range of 323–1073 K using the methodology described in [39,54,55]. The values of the average linear thermal expansion coefficient (α, LTEC) were calculated from the linear parts of the Δl/l0 = f(T) plots.
The thermal diffusivity (η) of the Na0.89Co0.90Me0.10O2 layered cobaltites was studied in a helium atmosphere over a temperature range of 323–1073 K using an LFA 457 Micro-Flash device (NETZSCH, Selb (Germany)). The thermal conductivity (λ) of the Na0.89Co0.90Me0.10O2 sintered ceramics was calculated using Equation (7).
λ = η · d E X P · C p ,
where Cp is heat capacity, calculated by the Dulong–Petit law, (J/(g·K)). Thus, according to the data of works [16,56], the heat capacity of layered sodium cobaltite near 300 K becomes slightly dependent on the temperature and reaches a plateau.
The phonon (λph) and electron (λe) parts of the thermal conductivity of the Na0.89Co0.90Me0.10O2 ceramics were roughly approximated by Equations (8) and (9)
λ = λ e + λ p h
λ e = L · T ρ
where L is the Lorentz number (L = 2.45 × 10−8 W∙Ω∙K−2).
The values of the power factor (P) and figure of merit (ZT) of the Na0.89Co0.90Me0.10O2 solid solutions were calculated using the following formula:
P = S 2 ρ
and
Z T = P · T λ
The self-compatibility factor (s) and the dimensionless relative self-compatibility factor (Δs) [57] for the Na0.89Co0.90Me0.10O2 ceramics were calculated using Equations (12) and (13)
s = 1 + Z T 1 S · T
and
Δ s = s max s min s min · 100 % .
The chemical stability of the Na0.89Co0.90Me0.10O2 and Na0.55Co0.90Me0.10O2 [16] samples to the impact of atmospheric water and carbon dioxide was estimated based on the results of their exposure for ten months in wet air at room temperature and a relative moisture of about 75% using optical microscopy (Levenhuk DTX 500 LCD (Levenhuk Inc., Praha (Czech Republic)), XRD (Bruker D8 XRD Advance X-ray diffractometer), infrared absorption spectroscopy (Thermo Nicolet Nexus ESP IR spectrometer (Nicolet Instrument Corporation, Madison (WI, USA)), and gravimetry (RADWAG AS/C/2/N 220 (Radwag Wagi Elektroniczne, Radom (Poland)). According to [16] and the results of XRD and IR-absorption spectroscopy, surface degradation of layered sodium cobaltite is described by Equations (14) and (15)
Na0.89CoO2 + 0.445 CO2 = 0.445 Na2CO3 + 1/3 Co3O4 + 0.117 O2
Na0.89CoO2 + 0.89 H2O = 0.89 NaOH + 1/3 Co3O4 + 1/3 O2
The intensity of chemical degradation of the samples in air was also estimated by comparing their electrotransport properties (electrical resistivity and Seebeck coefficient) before and after their storage in the measuring set within two weeks.
The thermal stability of the materials was estimated as their durability under operational conditions by comparing their electrical resistivity values after multiple thermocycles within the 323–1073 K temperature interval and after annealing in the measuring set within eight hours at 1073 K in air.

3. Results and Discussion

The sodium content in the obtained samples (xNa), determined by iodometry, reverse potentiometry, and spectrophotometry methods, was approximately 0.89 for all the investigated samples (Table 1 and Table S1). The average oxidation state of cobalt in the Na0.89Co0.90Me0.10O2 (Me = Cr, Co, Ni, Mo, W, and Bi) layered cobaltites varied within 2.78–3.23 (Table 1 and Table S1), increasing with the substitution of an acceptor character (Ni instead of Co) and decreasing with the substitution of a donor character (Mo, W, Pb, or Bi instead of Co), which was also observed for the Na0.55Co0.90Me0.10O2 samples in a previous study [16]. Therefore, Co exists in the Na0.89Co0.90Me0.10O2 compounds as a mixture of Co3+ and Co4+, Co2+ and Co3+, and only Co3+ for Me = Cr, Co, and Ni; Me = Mo, W, and Bi; and Me = Pb.
According to the XRD results, all ceramic samples were single-phase within the XRD accuracy. The ceramic samples of Na0.89Co0.90Me0.10O2 (Me = Cr, Ni, Mo, W, Bi) obtained in this study, similar to the base Na0.89CoO2 cobaltite, exhibited a hexagonal structure corresponding to the γ-NaxCoO2 phase structure [11,29,58,59,60] (Figure 1a). Note that samples of sodium-poor Na0.55Co0.90Me0.10O2 cobaltites with the same thermal prehistory [16] contained Co3O4 impurities formed by the partial degradation of these phases in air. Our results indicate that increasing the sodium content in NaxCoO2 enhances its chemical stability in air. This conclusion is confirmed by the results of long-term exposure of the samples to wet air, according to which the sodium-rich Na0.89CoO2 compound degrades according to Equations (14) and (15) two times slower than the sodium-poor Na0.55CoO2 phase. The degradation degree of NaxCo0.90Me0.10O2 cobaltites doped with transition or heavy metal oxides, calculated through their mass change, was 7–27% lower than that of the parent NaxCoO2 phases.
The values of the lattice constants for the Na0.89Co0.90Me0.10O2 solid solutions varied within the ranges of a = 2.822–2.831 Å  and c = 10.92–10.97 Å (Table 1), which are close to the parameters of the Na0.89CoO2 base cobaltite (a = 2.826 Å, c = 10.94 Å) [26,61,62]. As a result, the values of volume of the unit cell for the phases of Na0.89Co0.90Me0.10O2 with partial cobalt substitution by transition or heavy metals changed very little, but their axial ratio significantly decreased only with the substitution of cobalt by nickel in the Na0.89CoO2 structure. Thus, partial substitution of cobalt with other metals in the Na0.89CoO2 phase does not lead to significant changes in the size and shape of the unit cell for the Na0.89Co0.90Me0.10O2 (Me = Cr, Mo, W, and Bi) solid solutions compared to the unsubstituted phase of layered sodium cobaltite.
The apparent density values of the Na0.89Co0.90Me0.10O2 ceramic ranged within 3.18–3.47 g/cm3 and decreased with the partial substitution of Co with Cr, Mo, W, or Pb, while they increased with the substitution of Co with Ni and Bi (Table 2). The values of open porosity of the ceramics varied within 16–22% and were close for samples with different cationic compositions, but values of closed porosity were minimal for the base Na0.89CoO2 phase (13%) and Na0.89Co0.90Ni0.10O2 solid solution (7%) and for other samples varied within 17–24% and were close each other (Table 2). The values of the apparent density of Na-rich samples were 7–8% less than for Na-poor ones [16], which was probably due to the fact that during heat treatment of ceramics, part of Na2O evaporates into the environment, and this process intensifies with the increase in sodium content in the samples. In fact, porosity values of the Na-rich samples were about two times larger (Table 2) than for Na-poor ones [16].
The Lotgering factor values increased from 0.31 for the Na0.89CoO2 base cobaltite (moderate orientation) to 0.69–0.76 for the Na0.89Co0.90Me0.10O2 solid solutions (Me = Ni and Bi) (good orientation) and 0.87–0.92 for the Na0.89Co0.90Me0.10O2 (Me = Cr, Mo, W, and Pb) ceramics of composition (very good orientation) (Table 1). Thus, doping Na0.89CoO2 with various transition or heavy metal oxides increases the degree of crystallographic orientation of the ceramic grains (the degree of its texturing). The most pronounced effect among the samples synthesized in this work was observed for Na0.89Co0.90W0.10O2, with the partial substitution of cobalt by tungsten in the Na0.55CoO2 phase leading to a similar effect [16,48].
The coherent scattering area of the Na0.89Co0.90Me0.10O2 materials, synthesized in this study, calculated using different approaches (Figure 1b), was larger (up to 6–25% and 3360% according to the Debye–Scherrer and the size-strain method, respectively) comparing to the base layered sodium cobaltite phase (Figure 1b, Table 1). In turn, the values of microstrains for the Na0.89Co0.90Me0.10O2 ceramics were slightly (Me = Cr and Pb) or significantly (Me = Ni, Mo, W, and Bi) lower compared to the initial Na0.89CoO2 cobaltite (Table 1). Therefore, the partial substitution of cobalt by transition or heavy metals in Na0.89CoO2 results in the formation of ceramics with larger-grained and less-strained grains.
According to the results of SEM, the grains of the Na0.89CoO2 base unsubstituted cobaltite had a plate-like shape with dimensions (l) of 6–15 μm and a thickness of 2.5–3 μm (with an average size (lav) of about 7 μm and an aspect ratio (AR) of approximately 3.9) (Figure 2b). The microstructure of the Na0.89Co0.90M0.10O2 materials was similar to that of Na0.89CoO2, but it differed in the size and aspect ratio (shape) of the grains (Figure 2a,c–f).
The grain sizes of the Na0.89Co0.90Me0.10O2 (Me = Cr, Ni, W) ceramics varied within 6–23 μm, with a thickness of 6–10 μm. The average dimensions (lav (AR)) were approximately 17 μm (2.8) for Me = Cr, 22 μm (3.3) for Me = Ni, and 14 μm (3.2) for Me = W. Doping of the layered sodium cobaltite ceramics with bismuth or lead oxides resulted in an increase in grain size to 35–60 μm and a decrease in the thickness of the grains by up to 1–3 μm. For Na0.89Co0.90Pb0.10O2 and Na0.89Co0.90Bi0.10O2, the lav (AR) values were around 43 μm (14.3) and 52 μm (15.6), respectively. Consequently, the anisotropy of the grains in the sodium cobaltite ceramic increased with doping by heavy metal oxides (PbO and Bi2O3).
The enhancement of the Lotgering factor and improvement of the anisotropy degree (AR) of the grains of Na0.89Co0.90Me0.10O2 ceramics may be caused by the differences in the sizes of the substituting and substituted ions (size effect) and their charges (charge redistribution). The first results in localized lattice strain promoting preferential alignment of grains along specific crystallographic planes (in our case, (00l)), but the second alters the charge distribution in the structure of Na0.89Co0.90Me0.10O2 phases, enhances interlayer charge screening, reduces electrostatic repulsion of CoO2-layers, and stabilizes (00l)-oriented stacking. The maximal values of f and AR were observed, in the whole, for the samples of layered sodium cobaltite, doped by heavy metal oxides (Pb, W etc.) (Table 1, Figure 2), as in these cases, the maximal difference in the sizes and charges of the substituting and substituted ions occurs.
According to the EDX results, the cationic composition of the Na0.89Co0.90M0.10O2 synthesized compounds was close to the target values, and the distribution of elements within the ceramic was nearly uniform (Figure 3 and Figure S2 (see Supplementary Information)).
The Temperature dependence of the relative elongation of the Na0.89Co0.9Me0.1O2 ceramic samples was practically linear, which proves the absence of structural phase transitions in the complex oxides investigated within all the temperature intervals studied. LTEC values of the Na0.89Co0.9Me0.1O2 solid solutions varied within (1.25–1.68) × 10−5 K−1 and were lower (for Me = Pb, and Bi), slightly (for Me = W) and essentially larger (for Me = Cr, Ni, and Mo) than for unsubstituted sodium cobaltite Na0.89CoO2 (1.34 × 10−5 K−5) (Table 2).
The increase in the LTEC values of the Na0.89Co0.90Me0.10O2 derivatives compared to those of the Na0.89CoO2 base phase is possibly due to the high values of their porosity (Table 2) as well as to the increase in the anharmonicity of the metal–oxygen vibrations in their structure with partial substitution of cobalt by other metals.
As can be seen from the Figure 4a, the Na0.89CoO2 layered sodium cobaltite and its Na0.89Co0.90Me0.10O2 derivatives exhibited metallic-like conductivity character (∂ρ/∂T > 0) (except Na0.89Co0.90W0.10O2 solid solution, which possesses semiconducting conductivity character within all the temperature interval studied), which for the Na0.89Co0.90Bi0.10O2 and Na0.89Co0.90Mo0.10O2 solid solutions near 923 K changed into a semiconducting one (∂ρ/∂T < 0) like the conductivity crossover in layered calcium cobaltite of Ca3Co4O9+δ [63,64]. A similar effect was also observed in several Na0.55Co0.90Me0.10O2 layered sodium cobaltites earlier [16]. The doping of Na0.89CoO2 layered sodium cobaltite by different metal oxides results in increasing its electrical resistivity values, except NiO, as ρ values of Na0.89Co0.90Ni0.10O2 solid solution were essentially less than those for the unsubstituted layered sodium cobaltite (Figure 4a,d). The metallic-like conductivity and low ρ values of the Ni-doped layered cobaltite are in good agreement with the results of [40,41] and are due to the donor nature of the substitution of Co by Ni, resulting in an increase in the concentration of holes, which are the main charge carriers in the layered sodium cobaltites. As can be seen, electrical resistivity values of ceramics increased with the increase in oxidation state of the substituting cobalt metal Me (ρ(Na0.89Co0.90Ni0.10O2) < ρ(Na0.89CoO2) < ρ(Na0.89Co0.90Pb0.10O2)). This can be explained by a decrease in the concentration of the main charge carriers (holes) as the average oxidation state of the cations in the conducting (Co,Me)O2 layers of the crystal structure Na0.89Co0.90Me0.10O2 phases increased (Figure 4a,d, Table 2).
The electrical resistivity of oxide ceramics is affected by their chemical and phase compositions, as well as their microstructural characteristics, such as porosity. In our case, namely larger Πt values explain the facts, that ρ(Na0.89Co0.90Cr0.10O2) > ρ(Na0.89CoO2), ρ(Na0.89Co0.90W0.10O2) > ρ(Na0.89Co0.90Mo0.10O2) (Figure 4a,d, Table 2) though according average oxidation state of cobalt (Table 1) these pairs of Na0.89Co0.90Me0.10O2 ceramic samples should demonstrate close values of electrical resistivity.
The electrical resistivity values of the Na0.89Co0.90Me0.10O2 ceramic samples did not change after several thermocycles, exposure to wet air for two weeks, and annealing in air at 1073 K for eight hours. Note that the ρ values of Na0.55Co0.90Me0.10O2 compounds [16] aged under the same schemes increased by 15–25% depending on their composition. These facts prove the enhancement of the chemical and thermal stability of layered sodium cobaltite and its derivatives with increasing sodium content, and allow us to consider the sodium-rich Na0.89Co0.90Me0.10O2 compounds as more suitable for different practical purposes than sodium-poor ones.
The Seebeck coefficient of the Na0.89Co0.90Me0.10O2 (Me = Cr, Ni, Mo, W, Pb, and Bi) solid solutions was positive throughout the temperature range studied, indicating that the main charge carriers were holes, and these materials were classified as p-type conductors. Notably, at high temperatures, the change in the thermo-EMF for all materials studied was nearly linear. This relationship of the Seebeck coefficient corresponds to Equation (16), which is commonly used to describe the thermoelectric properties of metals and degenerate semiconductors [65]
S = 8 · π 2 · k B 2 3 · e · h · m · T · π 3 · n 2 / 3 ,
where kB is the Boltzmann constant (J/K), m* is the density of the effective mass (kg), h is the Planck’s constant (J∙s), e is the charge of the electron (C), T is the absolute temperature (K), and n is the charge carrier concentration (cm−3).
The partial substitution of cobalt with ions of various transition or heavy metals in the structure of Na0.89CoO2 led to an increase in the Seebeck coefficient, which was more pronounced at high temperatures for tungsten- or bismuth-substituted solid solutions (Figure 4b,e, and Table 2). This is likely due to the increase in configurational entropy provided by the presence of cobalt ions with different charges (Co2+, Co3+, and Co4+) and spin states (high, intermediate, and low spin). The fact that the Seebeek coefficient values of the Na0.89Co0.90W0.10O2 phase are larger than those of the Na0.89Co0.90Mo0.10O2 cobaltite, in which cobalt possesses the same oxidation state and spin state, may be due to its larger porosity. The highest values of S among all investigated samples were demonstrated by the Na0.89Co0.90W0.10O2 and Na0.89Co0.90Bi0.10O2 oxides (519 and 616 µV/K at 1073 K, respectively), which are 1.18 and 1.40 times larger than that of the parent Na0.89CoO2 phase (Figure 4e and Table 2). Therefore, these phases can be considered as prospective materials for the p-legs of ceramic (oxide) thermocouples. It should also be noted that similar results were obtained when studying the Seebeck coefficient of cobalt-substituted derivatives of the Na0.55CoO2 layered cobaltite [16,38]. The Seebeck coefficient of the Na0.89Co0.90Me0.10O2 phases after exposure to wet air for two weeks remained the same within the accuracy of the measurements. In contrast, aged under the same conditions, Na0.55Co0.90Me0.10O2 samples demonstrated an increase in their S value up to 15–25% depending on their composition, which looks like the results of ρ measurements proved their lower chemical stability.
The weighted mobility (μp) and concentration (p) of the main charge carriers («holes») in the samples studied, calculated based on the experimentally determined ρ and S values according to the method described in [66] (see Supplementary Information, Equations (S8) and (S9)), varied with temperature change and essentially and nonmonotonously changed when Na0.89CoO2 was doped with different metal oxides. For the base Na0.89CoO2 cobaltite at 1073 K, the values of μp and p were ~48 cm2/(V∙s) and ~5 × 10−20 cm−3, respectively, which is in good agreement with the literature data [67]. The weighted mobility values of the Na0.89Co0.90Me0.10O2 solid solutions were less than, close to, and larger than those of Na0.89CoO2 for Me = Cr, Mo, Pb, Me = Ni, W, and Me = Bi, respectively. Charge carriers concentration in Na0.89Co0.90Me0.10O2 cobaltites was large, close to, and less, than for Na0.89CoO2 phase for Me = Cr, Me = Mo, Ni, Pb, and Me = W, Bi respectively, that is in good accordance with the results of measurements of Seebeck coefficient of ceramics, for example, with the sharp increasing of S values of the layered sodium cobaltite at partical substitution of cobalt by bismuth or tungsten in it, and with the scharp decreasing of S at partical substitution of cobalt by chromium in Na0.89CoO2 (Figure 4b,e).
The power factor values of Na0.89Co0.90Me0.10O2 sintered ceramics increased with rising temperature and changed non-monotonically with the changing of the nature of metal substituting cobalt in the Na0.89CoO2 structure (Figure 4c,f, and Table 2). The maximum p value was achieved for the Na0.89Co0.90Ni0.10O2 nickel-substituted solid solution (0.910 mW/(m⋅K2) at 1073 K), which is 1.15 times larger than that of the undoped Na0.89CoO2 layered sodium cobaltite, and was determined by both the low values of its electrical resistivity and high values of its Seebeck’s coefficient.
The thermal diffusivity and thermal conductivity values of the Na0.89Co0.10Me0.10O2 ceramics decreased with increasing of temperature (Figure 5a,b) and changed nonmonotonously at the doping of Na0.89CoO2 by various metal oxides, increasing at substitution of Co by Cr, and Ni, and decreasing at substitution of Co by Mo, W, Pb, and Bi (Figure 5d,e). The first is associated with an increase in the grain size of the ceramics, which leads to a decrease in the density of grain boundaries that serve as effective phonon scattering regions. The latter occurs due to the partial replacement of lighter cobalt ions by heavier ions of molybdenum, tungsten, lead, or bismuth, serving as effective scattering centers. The electronic component of the thermal conductivity of the ceramics was relatively small (λe/λ = 0.03–0.13) and increased with increasing temperature, and for the Na0.89Co0.10Me0.10O2 solid solutions, except Na0.89Co0.90Ni0.10O2, it was lower than that of the Na0.89CoO2 base oxide. Thus, the main part of the heat in the Na0.89Co0.10Me0.10O2 phases was carried by lattice vibrations (phonons) (λph/λ ≈ (0.87–0.97)) (Figure 5c,f). As shown in Figure 5, the thermal diffusivity and conductivity values of the doped cobaltites (except Cr- and Ni-doped) are lower than those of the base Na0.89CoO2 phase. It should be noted that for sodium-poor samples (Na0.55Co0.90Me0.10O2), the converse effect was found, namely an increase in η and λ values of doped with transition or heavy metal oxides ceramics compared to the undoped Na0.55CoO2 phase [16,38].
Although the porosity values of the Na0.89Co0.90Me0.10O2 ceramic samples were similar (Table 2), some effects of Πt on the thermal properties of the layered sodium cobaltite derivatives were observed and should be noted. Therefore, the fact that the Na0.89Co0.90Bi0.10O2 sample demonstrates larger η, λ, and λph values than the Na0.89Co0.90Pb0.10O2 solid solution (Figure 5) is probably due to its lower porosity.
The figure of merit values of the investigated materials sharply increased with increasing temperature, and doping of layered sodium cobaltite with various transition and heavy metal oxides had different effects: ZT increased at substitution in Na0.89CoO2 of cobalt by nickel and bismuth and decreased when cobalt was partially substituted by chromium, molybdenum, tungsten, and lead (Figure 6a and Table 2). The highest thermoelectric characteristics were demonstrated by Na0.89Co0.90Ni0.10O2 and Na0.89Co0.90Bi0.10O2 solid solutions, with ZT values reaching 1.65 and 1.74 at 1073 K, which are about 4% and 9% higher, respectively, than that of the Na0.89CoO2 phase (Figure 6c). It should be noted that although for the Na0.89Co0.90W0.10O2, Na0.89Co0.90Mo0.10O2, and Na0.89Co0.90Pb0.10O2 ceramics ZT values were less than that of the Na0.89CoO2 phase, at 1073 K, they were equal to 1.09, 0.97, and 0.84, respectively, which is close to the theoretical criterion (ZT > 1) that defines materials of practical interest for thermoelectric conversion (Figure 6a and Table 2).
The self-compatibility factor values (s) of the ceramics studied increased at temperature increasing and within temperature interval 673–873 K slightly vary (except for Na0.89Co0.90Cr0.10O2 compound) (Figure 6b), and dimensionless relative self-compatibility factor (Δs) of Na0.89Co0.90Me0.10O2 layered cobaltites within 673–873 K varies within 8–29% (Figure 6d), which is, in the whole, less than for the Mg2Si0.6−ySn0.4Sby (25–40%) thermoelectric alloy [68] and states the good self-compatibility of the obtained in this work thermoelectric oxide ceramics. Note, that Δs values for the doped sodium cobaltites, except Na0.89Co0.90Mo0.10O2, were less (8–21%) than for the base layered sodium cobaltite (26%), which shows the effectiveness of doping of Na0.89CoO2 by transition or heavy metal oxides in increasing of its self-compatibility.
Comparing the results obtained in this study with the data published in [16], we can conclude that the modification of layered sodium cobaltite by different metal oxides similarly affects its structure and properties for both sodium-rich (Na0.89Co0.90Me0.10O2) and sodium-poor (Na0.55Co0.90Me0.10O2) samples. However, for the latter, the effect is more pronounced due to the larger ρ and smaller S values for the Na0.55CoO2 phase. In contrast, sodium-rich Na0.89Co0.90Me0.10O2 compounds demonstrate higher chemical and thermal stabilities than sodium-poor Na0.55Co0.90Me0.10O2 phases under both storage and operating conditions. This fact, along with their superior thermoelectric characteristics, makes Na0.89Co0.90Me0.10O2 cobaltites more suitable for practical applications in different thermoelectric devices (ceramic thermocouples, thermoelectric generators, and industrial waste heat recovery systems).
It should also be noted that layered sodium cobaltite and its derivatives demonstrate significantly larger Seebeck coefficients, power factors, and figure-of-merit values than other layered cobaltites (Ca3Co4O9+δ, Bi2Ca2Co1,7Oy, etc.) [6,9,69,70,71]; therefore, they can be considered more promising thermoelectric materials.

4. Conclusions

Combining the results obtained in this study, one can conclude that doping (up to 10 mol.%) of Na0.89CoO2 sodium-rich layered cobaltite by different transition (Cr2O3, NiO) and heavy metal oxides (MoO3, WO3, PbO, Bi2O3) do not change its crystal structure and slightly affects its lattice constants values, but increases coherent scattering area values, grains size as well as grain orientation degree of Na0.89Co0.10Me0.10O2 solid solutions comparing to the parent Na0.89CoO2 phase.
Such doping leads to a decrease in the electrical resistivity of the samples and complex changes in the Seebeck coefficient and thermal properties of the ceramics, which are determined by the nature of the metal substituting cobalt in Na0.89CoO2 and the peculiarities of the microstructure of the ceramics. The highest values of S possessed by the Na0.89Co0.90W0.10O2 and Na0.89Co0.90Bi0.10O2 cobaltites (519 and 616 µV/K, respectively, at 1073 K), which are 18% and 40% higher than that of the parent Na0.89CoO2 phase.
The maximal power factor was observed for the Na0.89Co0.90Ni0.10O2 compound (0.910 mW/(m⋅K2) at 1073 K), which was 15% higher than that of the base Na0.89CoO2 layered sodium cobaltite, which was attributed to both the low values of its electrical resistivity and high values of its Seebeck coefficient. The largest thermoelectric performance was demonstrated by the Na0.89Co0.90Ni0.10O2 and Na0.89Co0.90Bi0.10O2 phases, with ZT values of 1.65 and 1.74 at 1073 K, which were about 4% and 9% higher, respectively, than that of the Na0.89CoO2 phase.
The obtained results demonstrate the effectiveness of the doping strategy of cobalt by heavy metals in layered sodium cobaltite for the enhancement of the thermoelectric performance of cobaltite derivatives. It was also found that the dimensionless relative self-compatibility factor of Na0.89CoO2 layered sodium cobaltite is essentially reduced by doping with transition or heavy metal oxides.
The high sodium content (Na:Co:Me = 0.89:0.90:0.10) in these phases provides enhanced thermal and chemical stability compared to the phases with low sodium content (Na:Co:Me = 0.55:0.90:0.10), which makes the former more promising candidates for thermoelectric devices. Materials possessing high Seebeck coefficient values (Na0.89Co0.90Me0.10O2 (Me = W, Bi)) are suitable for use in thermocouples, but compounds with large values of power factors and figure-of-merit (Na0.89Co0.90Me0.10O2 (Me = Ni, Bi)) can be used as components of thermoelectric generators and industrial waste heat recovery systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ceramics8030086/s1. Figure S1: Absorption spectra and calibration graphs for determining the cobalt content in the form of complexes [Co(SCN)4]2−(a, c) и [CoY] (b, d); Figure S2: Element mapping images of the Na0.89Co0.90M0.10O2 ceramic samples (M = Cr (a), Ni (b), Pb (c), Bi (d)); Table S1: Sodium content (xNa) and average oxidation state of cobalt (Z) in Na0.89Co0.90M0.10O2 layered cobaltite samples determined by various methods.

Author Contributions

Conceptualization, A.I.K.; methodology, A.I.K., E.A.C. and N.S.K.; software, N.S.K.; validation, A.I.K., E.A.C. and H.W.; formal analysis, N.S.K. and E.A.C.; investigation, N.S.K., J.A.Z. and A.V.B.; resources, A.I.K.; data curation, A.I.K.; writing—original draft preparation, N.S.K. and A.I.K.; writing—review and editing, A.I.K. and H.W.; visualization, N.S.K.; supervision, A.I.K.; project administration, A.I.K.; funding acquisition, A.I.K. and N.S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Education of the Republic of Belarus under project number 20111575 and by the Belarusian Republican Foundation for Fundamental Research under project number 20122443.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analyzed during this study are included in this published article.

Acknowledgments

The authors would like to thank Dzmitry S. Kharytonau for fruitful discussions on the research results and Dzmitry Darashchuk for technical support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Klyndyuk, A.I.; Kharytonau, D.S.; Matsukevich, I.V.; Chizhova, E.A.; Lenčéš, Z.; Socha, R.P.; Zimowska, M.; Hanzel, O.; Janek, M. Effect of cationic nonstoichiometry on thermoelectric properties of layered calcium cobaltite obtained by field assisted sintering technology (FAST). Ceram. Int. 2024, 50, 30970–30979. [Google Scholar] [CrossRef]
  2. Polevik, A.O.; Sobolev, A.V.; Glazcova, I.S.; Presniakov, I.A.; Verchenko, V.Y.; Link, J.; Stern, R.; Shevelkov, A.V. Interplay between Fe(II) and Fe(III) and Its Impact on Thermoelectric Properties of Iron-Substituted Colusites Cu26−xFexV2Sn6S32. Compounds 2023, 3, 348–364. [Google Scholar] [CrossRef]
  3. Yu, J.; Chen, K.; Azough, F.; Alvarez-Ruiz, D.T.; Reece, M.I.; Freer, R. Enhancing the thermoelectric performance of calcium cobaltite ceramics by tuning composition and processing. ACS Appl. Mater. Interfaces 2020, 12, 47634–47646. [Google Scholar] [CrossRef] [PubMed]
  4. Zhang, W.; Zhu, K.; Liu, J.; Wang, J.; Yan, K.; Liu, P.; Wang, Y. Influence of the phase transformation in NaxCoO2 ceramics on thermoelectric properties. Ceram. Int. 2018, 44, 17251–17257. [Google Scholar] [CrossRef]
  5. Li, Z.; Xiao, C.; Xie, Y. Layered thermoelectric materials: Structure, bonding, and performance mechanisms. Appl. Phys. Rev. 2020, 9, 011303. [Google Scholar] [CrossRef]
  6. Kuzanyan, A.S.; Margani, N.G.; Zhghamadze, V.V.; Mumladze, G.; Badalyan, G.R. Impact of Sr(BO2)2 Dopant on Power Factor of Bi2Sr2Co1.8Oy Thermoelectric. J. Contemp. Phys. 2021, 56, 146–149. [Google Scholar] [CrossRef]
  7. Oopathumpa, C.; Boonthuma, D.; Smith, S.M. Effect of Poly(Vinyl Alcohol) on Thermoelectric Properties of Sodium Cobalt Oxide. Key Eng. Mater. 2019, 798, 304–309. [Google Scholar] [CrossRef]
  8. Mallick, M.M.; Vitta, V. Enhancing Thermoelectric Figure-of-Merit of Polycrystalline NayCoO2 by a Combination of Non-stoichiometry and Co-substitution. J. Electron. Mater. 2018, 47, 3230–3237. [Google Scholar] [CrossRef]
  9. Yu, J.; Chang, Y.; Jakubczyk, E.; Wang, B.; Azough, F.; Dorey, R.; Freer, R. Modulation of electrical transport in calcium cobaltite ceramics and thick films through microstructure control and doping. J. Eur. Ceram. Soc. 2021, 41, 4859–4869. [Google Scholar] [CrossRef]
  10. Noudem, J.G.; Xing, Y. Overview of Spark Plasma Texturing of Functional Ceramics. Ceramics 2021, 4, 97–107. [Google Scholar] [CrossRef]
  11. Jansen, M.; Hoppe, R. Notiz zur Kenntnis der Oxocobaltate des Natriums. Z. Für Anorg. Und Allg. Chem. 1974, 408, 104–106. [Google Scholar] [CrossRef]
  12. Terasaki, I.; Sasago, Y.; Uchinokura, K. Large thermoelectric power in NaCo2O4 single crystals. Phys. Rev. B 1997, 56, R12685–R12687. [Google Scholar] [CrossRef]
  13. Koumoto, K.; Terasaki, I.; Funahashi, R. Complex Oxide Materials for Potential Thermoelectric Applications. MRS Bull. 2006, 31, 206–210. [Google Scholar] [CrossRef]
  14. Viciu, L.; Huang, Q.; Cava, R.J. Stoichiometric oxygen content in NaxCoO2. Phys. Rev. B 2006, 73, 212107. [Google Scholar] [CrossRef]
  15. Han, Y.; Ruan, Y.; Xue, M.; Shi, M.; Song, Z.; Zhou, Y.; Teng, J. Effect of Annealing Time on the Cyclic Characteristics of Ceramic Oxide Thin Film Thermocouples. Micromachines 2022, 13, 1970. [Google Scholar] [CrossRef]
  16. Krasutskaya, N.S.; Klyndyuk, A.I.; Evseeva, L.E.; Gundilovich, N.N.; Chizova, E.A.; Paspelau, A.V. Enhanced thermoelectric performance of Na0.55CoO2 ceramics doped by transition and heavy metal oxides. Solids 2024, 5, 267–277. [Google Scholar] [CrossRef]
  17. Molenda, J.; Baster, D.; Milewska, A.; Świerczek, K.; Bora, D.K.; Braun, A.; Tobola, J. Electronic origin of difference in discharge curve between LixCoO2 and NaxCoO2 cathodes. Solid State Ion. 2014, 271, 15–27. [Google Scholar] [CrossRef]
  18. Pohle, B.; Gorbunov, M.; Lu, Q.; Bahrami, A.; Nielsch, K.; Mikhailova, D. Structural and electrochemical properties of layered P2-Na0.8Co0.8Ti0.2O2 cathode in sodium-ion batteries. Energies 2022, 15, 3371. [Google Scholar] [CrossRef]
  19. Nguyen, L.M.; Nguyen, V.H.; Nguyen, D.M.N.; Le, M.K.; Tran, V.M.; Le, M.L.P. Evaluating electrochemical properties of layered NaxMn0.5Co0.5O2 obtained at different calcined temperatures. Chemengineering 2023, 7, 33. [Google Scholar] [CrossRef]
  20. Baster, D.; Zając, W.; Kondracki, Ł.; Hartman, F.; Molenda, J. Improvement of electrochemical performance of Na0.7Co1−yMnyO2–cathode material for rechargeable sodium-ion batteries. Solid State Ion. 2016, 288, 213–218. [Google Scholar] [CrossRef]
  21. Banobre-López, M.; Rivadulla, F.; Caudillo, R.; López-Quintela, M.A.; Rivas, J.; Goodenough, J.B. Role of doping and Dimensionality in the Superconductivity of NaxCoO2. Chem. Mater. 2005, 17, 1965–1968. [Google Scholar] [CrossRef]
  22. Akram, R.; Khan, J.; Rafique, S.; Hussain, M.; Maqsood, A.; Naz, A.A. Enhanced thermoelectric properties of single phase Na doped NaxCoO2 thermoelectric material. Mater. Lett. 2021, 300, 130180. [Google Scholar] [CrossRef]
  23. Assadi, M.H.N.; Katayama-Yoshida, H. Restoration of long range order of Na ions in NaxCoO2 at high temperatures by sodium site doping. Comput. Mater. Sci. 2015, 109, 308–311. [Google Scholar] [CrossRef]
  24. Lee, M.; Viciu, L.; Li, L.; Wang, Y.; Foo, M.L.; Watauchi, S.; Pascal, R.A., Jr.; Cava, R.J.; Ong, N.P. Enhancement of thermopower in NaxCoO2 in the large-x regime (x ≥ 0.75). Phys. B 2008, 403, 1564–1568. [Google Scholar] [CrossRef]
  25. Krasutskaya, N.S.; Klyndyuk, A.I.; Evseeva, L.E.; Tanaeva, S.A. Synthesis and properties of NaxCoO2 (x = 0.55, 0.89) oxide thermoelectric. Inorg. Mater. 2016, 52, 393–399. [Google Scholar] [CrossRef]
  26. Plewa, A.; Kulka, A.; Kondracki, Ł.; Lu, L.; Molenda, J. Phase diagram of NaFeyCo1−yO2 and evolution of its physico- and electrochemical properties with changing iron content. J. Power Sources 2019, 419, 42–51. [Google Scholar] [CrossRef]
  27. Devi Meena, J.; Gayathri, N.; Bharathi, A.; Ramachandran, K. Effect of nickel substitution on thermal properties of Na0.9CoO2. Bull. Mater. Sci. 2007, 30, 345–348. [Google Scholar] [CrossRef]
  28. Zhou, R.; Li, J.; Wei, W.; Li, X.; Luo, M. Atomic substituents effect on boosting desalination performances of Zn-doped NaxCoO2. Desalination 2020, 496, 114695. [Google Scholar] [CrossRef]
  29. Ono, Y.; Kato, N.; Miyazaki, Y.; Kajitani, T. Transport Properties of Ca-doped γ-NaxCoO2. J. Ceram. Soc. Jpn. 2004, 5, S626–S628. [Google Scholar] [CrossRef]
  30. Bos, J.W.G.; Hertz, J.T.; Morosan, E.; Cava, R.J. Magnetic and thermoelectric properties of layered LixNayCoO2. J. Chemisrtry State Chem. 2007, 180, 3211–3217. [Google Scholar] [CrossRef]
  31. Yang, X.; Wang, X.; Liu, J.; Hu, Z. Power factor enhancement in NaxCoO2 doped by Bi. J. Alloys Compd. 2014, 582, 59–63. [Google Scholar] [CrossRef]
  32. Seetawan, T.; Amornkitbamburg, V.; Burinprakhon, T.; Maensiri, S.; Kurosaki, K.; Muta, H.; Uno, M.; Yamanaka, S. Thermoelectric power and electrical resistivity of Ag-doped Na1.5Co2O4. J. Alloys Compd. 2006, 407, 314–317. [Google Scholar] [CrossRef]
  33. Hussein, M.; Assadi, N. Hf doping for enhancing the thermoelectric performance in layered Na0.75CoO2. Mater. Today Proc. 2021, 42, 1542–1546. [Google Scholar] [CrossRef]
  34. Nakhowong, R. Effect of reduced graphene oxide on the enhancement of thermoelectric power factor of γ-NaxCo2O4. Mater. Sci. Eng. B 2020, 261, 114679. [Google Scholar] [CrossRef]
  35. Ito, M.; Nagira, T.; Furumoto, D.; Katsuyama, S.; Nagai, H. Synthesis of NaxCoO2 thermoelectric oxides by the polymerized complex method. Scr. Mater. 2003, 48, 403–408. [Google Scholar] [CrossRef]
  36. Suan, M.S.M.; Farhan, M.A.; Lau, K.T.; Munawar, R.F.; Ismail, S. Structural and Electrical Properties of NaxCoO2 Thermoelectric Synthesized via Citric-Nitrate Auto-Combustion Reaction. J. Phys. Conf. Ser. 2018, 1082, 012033. [Google Scholar] [CrossRef]
  37. Itahara, H.; Fujita, K.; Sugiyama, J.; Nakamura, K.; Tani, T. Highly Textured NaxCoO2–δ Ceramics Fabricated by Both Templated Grain Growth and Reactive Templated Grain Growth Methods Using Single-Crystalline Particles as Templates. J. Ceram. Soc. Jpn. 2003, 111, 227–231. [Google Scholar] [CrossRef]
  38. Park, K.; Jang, K.U.; Know, H.-C.; Kim, J.-G.; Cho, W.-S.Y. Influence of partial substitution of Cu for Co on the thermoelectric properties of NaCo2O4. J. Alloys Compd. 2006, 419, 213–219. [Google Scholar] [CrossRef]
  39. Terasaki, I.; Tsukada, I.; Iguchi, Y. Impurity-induced transition and impurity-enhanced thermopower in the thermoelectric oxide NaCo2−xCuxO4. Phys. Rev. B 2002, 65, 195106. [Google Scholar] [CrossRef]
  40. Park, K.; Choi, J.W.; Lee, G.W.; Kim, S.-J.; Lim, Y.-S.; Choi, S.-M.; Seo, W.-S.; Lim, S.M. Thermoelectric Properties of Solution-Combustion-Processed Na(Co1−xNix)2O4. Mater. Lett. 2012, 18, 1061–1065. [Google Scholar] [CrossRef]
  41. Park, K.; Jang, K.U. Improvement in High-Temperature Properties of NaCo2O4 through Partial Substitution of Ni for Co. Mater. Lett. 2006, 60, 1106–1110. [Google Scholar] [CrossRef]
  42. Park, K.; Lee, J.H. Enhanced Thermoelectric Properties of NaCo2O4 by Adding Zn. Mater. Lett. 2008, 62, 2366–2368. [Google Scholar] [CrossRef]
  43. Tian, Z.; Wang, X.; Liu, J.; Lin, Z.; Hu, Y.; Wu, Y.; Han, C.; Hu, Z. Power factor enhancement induced by Bi and Mn co-substitution in NaxCoO2 thermoelectric materials. J. Alloys Compd. 2016, 661, 161–167. [Google Scholar] [CrossRef]
  44. Shin, W.; Murayama, N.; Ikeda, K.; Sago, S.; Terasaki, I. Thermoelectric Device of Na(Co,Cu)2O4 and (Ba,Sr)PbO3. J. Ceram. Soc. Jpn. 2002, 110, 727–730. [Google Scholar] [CrossRef]
  45. Klyndyuk, A.I.; Krasutskaya, N.S.; Chizhova, E.A.; Evseeva, L.E.; Tanaeva, S.A. Synthesis and properties of Na0.55Co0.9M0.1O2 (M =Sc, Ti, Cr-Zn, Mo, W, Pb, Bi) solid solutions. Glass Phys. Chem. 2016, 42, 100–107. [Google Scholar] [CrossRef]
  46. Shannon, D.; Prewitt, C.T. Effective ionic radii in oxides and fluorides, Acta Crystallographica Section B: Structural Science. Cryst. Eng. Mater. 1969, 25, 925–946. [Google Scholar]
  47. Greenstein, G.R. The Merck index: An Encyclopedia of Chemicals, Drugs, and Biologicals. Ref. Rev. 2025, 21, 2746. [Google Scholar]
  48. Klyndyuk, A.I.; Krasutskaya, N.S.; Dyatlova, E.M. Influence of sintering temperature on the properties of NaxCoO2 ceramics. Proc. BSTU 2010, XVIII, 99–102. (In Russian) [Google Scholar]
  49. Pyatnitsky, I.V. Analitical Chemistry of Cobalt; Science: Moscow, Russia, 1965; 292p. (In Russian) [Google Scholar]
  50. Sopicka-Lizer, M.; Kozłowska, K.; Bobrowska-Grzesik, E.; Plewa, J.; Altenburg, H. Assessment of Co(II), Co(III) and Co(IV) content in thermoelectric cobaltites. Arh. Metall. Mater. 2009, 54, 881–888. [Google Scholar]
  51. Azad, S.; Kumar, N.; Chand, S. Structural, morphological, optical and dielectric properties of Ti1−xFexO2 nanoparticles sythesized using sol-gel method. J. Sol-Gel Sci. Technol. 2022, 105, 163–175. [Google Scholar] [CrossRef]
  52. Marwah, T.J. Using the size strain plot method to specity lattice parameters. Haitham J. Pure Appl. Sci. 2023, 36, 123–129. [Google Scholar] [CrossRef]
  53. Lotgering, F.K. Topotactical reactions with ferrimagnetic oxides having hexagonal crystal structures. J. Inorg. Nucl. Chem. 1959, 9, 113–123. [Google Scholar] [CrossRef]
  54. Klyndyuk, A.I.; Chizhova, E.A.; Latypov, R.S.; Shevchenko, S.V.; Kononovich, V.M. Effect of the addition of copper particles on the thermoelectric properties of the Ca3Co4O9+δ ceramics produced by two-step sintering. Inorg. Mater. 2022, 67, 237–244. [Google Scholar] [CrossRef]
  55. Klyndyuk, A.I.; Petrov, G.S.; Poluyan, A.F.; Bashkirov, L.A. Structure and Physicochemical Properties of Y2Ba1−xMxCuO5 (M = Sr, Ca) Solid Solutions. Inorg. Mater. 1999, 35, 512–516. [Google Scholar] [CrossRef]
  56. Zorkovská, A.; Feher, A.; Sébek, J.; Šantavá, E.; Bradaric, I. Non-Fermi-liquid behavior in the layered NaxCoO2. Low Temp. Phys. 2007, 33, 944–947. [Google Scholar] [CrossRef]
  57. Snyder, G.J.; Ursell, T.S. Thermoelectric Efficiency and Compatibility. Phys. Rev. Lett. 2003, 91, 148301. [Google Scholar] [CrossRef]
  58. Delmas, C.; Braconnier, J.-J.; Fouassier, C.; Hagenmuller, P. Electrochemical intertcalation of sodium in NaxCoO2 bronzes. Solid State Ion. 1981, 3, 165–169. [Google Scholar] [CrossRef]
  59. Liu, P.; Chen, G.; Cui, Y.; Zhang, H.; Xiao, F.; Wang, L.; Nakano, H. High temperature electrical conductivity and thermoelectric power of NaxCoO2. Solid State Ion. 2008, 179, 2308–2312. [Google Scholar] [CrossRef]
  60. Matsubara, I.; Zhou, Y.; Takeuchi, T.; Funahashi, R.; Shikano, M.; Murayama, N.; Shin, W.; Izu, N. Thermoelectric properties of spark-plasma-sintered Na1+xCo2O4 ceramics. J. Ceramic. Soc. Jpn. 2003, 111, 238–241. [Google Scholar] [CrossRef]
  61. Li, N.; Jiang, Y.; Li, G.; Wang, C.; Shi, J.; Yu, D. Self-ignition route to Ag-doped Na1.7Co2O4 and its thermoelectric properties. J. Alloys Compd. 2009, 467, 444–449. [Google Scholar] [CrossRef]
  62. Lei, Y.; Li, X.; Liu, L.; Ceder, G. Synthesis and Stochiometry of Different Layered Sodium Cobalt Oxides. Chem. Mater. 2014, 26, 5288–5296. [Google Scholar] [CrossRef]
  63. Klyndyuk, A.I.; Matsukevich, I.V. Synthesis, structure and properties of Ca3Co3.85M0.15O9+δ (M–Ti–Zn, Mo, W, Pb, Bi). Russ. J. Inorg. Chem. 2015, 51, 1025–1031. [Google Scholar] [CrossRef]
  64. Lin, Y.-H.; Lan, J.; Shen, Z.; Liu, Y.; Nan, C.-W.; Li, J.-F. High-temperature electrical transport behaviors in textured Ca3Co4O9-based polycrystalline ceramics. Appl. Phys. Lett. 2009, 94, 072107. [Google Scholar] [CrossRef]
  65. Pan, L.; Mitra, S.; Zhao, L.-D.; Shen, Y.; Wang, Y.; Fesel, C.; Berardan, D. The Role of Ionized Impurity Scattering on the Thermoelectric Performances of Rock Salt AgPbmSnSe2+m. Adv. Funct. Mater. 2016, 26, 5149–5157. [Google Scholar] [CrossRef]
  66. Snyder, G.J.; Snyder, A.H.; Wood, M.; Gurunathan, R.; Snyder, B.H.; Niu, C. Weighted Mobility. Adv. Mater. 2020, 32, 2001537. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, S.; Chen, S.; Liu, F.; Yan, G.; Chen, J.; Li, H.; Wang, J.; Yu, W.; Fu, G. Laser-induced voltage effects in c-axis inclined NaxCoO2 thin films. Appl. Surf. Sci. 2012, 258, 7330–7333. [Google Scholar] [CrossRef]
  68. Liu, W.; Tang, X.; Sharp, J. Low-temperature solid state synthesis and thermoelectric properties of high-performance and low-cost Sb-doped Mg2Si0.6Sn0.4. J. Phys. D Appl. Phys. 2010, 43, 085406. [Google Scholar] [CrossRef]
  69. Fergus, J.W. Oxide material for high temperature thermoelectric energy conversion. J. Eur. Ceram. Soc. 2012, 32, 525–540. [Google Scholar] [CrossRef]
  70. Krόlicka, A.K.; Piersa, M.; Mirowska, A.; Michalska, M. Effect of sol-gel and solid state synthesis techniques on structural, morphological and thermoelectric performancs of Ca3Co4O9. Ceram. Int. 2018, 44, 13736–13743. [Google Scholar] [CrossRef]
  71. Klyndyuk, A.I.; Krasutskaya, N.S.; Khort, A.A. Synthesis and Properties of Ceramics Based on a Layered Bismuth Calcium Cobaltite. Inorg. Mater. 2018, 54, 509–514. [Google Scholar] [CrossRef]
Figure 1. X-ray powder diffractograms (Cu–radiation) (a), size-strain plots (b) and coherent scatting area (c) for Na0.89CoO2 (1, black) layered sodium cobaltite and Na0.89Co0.9Me0.1O2 (Me = Cr (2, red), Ni (3, blue), Mo (4, pink), W (5, green), Pb(6, orange), Bi (7, purple)) solid solutions.
Figure 1. X-ray powder diffractograms (Cu–radiation) (a), size-strain plots (b) and coherent scatting area (c) for Na0.89CoO2 (1, black) layered sodium cobaltite and Na0.89Co0.9Me0.1O2 (Me = Cr (2, red), Ni (3, blue), Mo (4, pink), W (5, green), Pb(6, orange), Bi (7, purple)) solid solutions.
Ceramics 08 00086 g001
Figure 2. Electronic micrographs of ceramic cleavage of Na0.89Co0.9M0.1O2 (M = Cr (a), Co (b), Ni (c), W (d), Pb (e), Bi (f)).
Figure 2. Electronic micrographs of ceramic cleavage of Na0.89Co0.9M0.1O2 (M = Cr (a), Co (b), Ni (c), W (d), Pb (e), Bi (f)).
Ceramics 08 00086 g002
Figure 3. Element mapping images of the Na0.89CoO2 (a) and Na0.89Co0.9W0.1O2 (b) ceramic samples.
Figure 3. Element mapping images of the Na0.89CoO2 (a) and Na0.89Co0.9W0.1O2 (b) ceramic samples.
Ceramics 08 00086 g003
Figure 4. Temperature dependences of electrical resistivity ρ (a), Seebeck’s coefficient S (b) and power factor P (c) of ceramic samples of Na0.89Co0.9Me0.1O2 (Me = Cr (1, red), Co (2, black), Ni (3, blue), Mo (4, pink), W (5, green), Pb (6, orange), Bi (7, purple)). Insets show the electrical resistivity ρ1073 (d), Seebeck’s coefficient S1073 (e), and power factor P1073 (f) values of Na0.89Co0.9Me0.1O2 phases.
Figure 4. Temperature dependences of electrical resistivity ρ (a), Seebeck’s coefficient S (b) and power factor P (c) of ceramic samples of Na0.89Co0.9Me0.1O2 (Me = Cr (1, red), Co (2, black), Ni (3, blue), Mo (4, pink), W (5, green), Pb (6, orange), Bi (7, purple)). Insets show the electrical resistivity ρ1073 (d), Seebeck’s coefficient S1073 (e), and power factor P1073 (f) values of Na0.89Co0.9Me0.1O2 phases.
Ceramics 08 00086 g004
Figure 5. Temperature dependences of thermal diffusivity η (a), total thermal conductivity λ (b), phonon λph and electron λe contributions (c) in thermal conductivity of Na0.89Co0.9Me0.1O2 (Me = Cr (1, red), Co (2, black), Ni (3, blue), Mo (4, pink), W (5, green), Pb (6, orange), Bi (7, purple)) ceramic samples. Insets show the thermal diffusivity η1073 (d), total thermal conductivity λ1073 (e) and phonon thermal conductivity λph,1073 (f) of Na0.89Co0.9M0.1O2 materials.
Figure 5. Temperature dependences of thermal diffusivity η (a), total thermal conductivity λ (b), phonon λph and electron λe contributions (c) in thermal conductivity of Na0.89Co0.9Me0.1O2 (Me = Cr (1, red), Co (2, black), Ni (3, blue), Mo (4, pink), W (5, green), Pb (6, orange), Bi (7, purple)) ceramic samples. Insets show the thermal diffusivity η1073 (d), total thermal conductivity λ1073 (e) and phonon thermal conductivity λph,1073 (f) of Na0.89Co0.9M0.1O2 materials.
Ceramics 08 00086 g005
Figure 6. Temperature dependences of figure-of-merit ZT (a) and self-compability factor s (b) of Na0.89Co0.9Me0.1O2 (Me = Cr (1, red), Co (2, black), Ni (3, blue), Mo (4, pink), W (5, green), Pb (6, orange), Bi (7, purple)) ceramic samples. Insets show the figure-of-merit ZT1073 (c) and dimensionless relative self-compatibility factor Δs673–873K (d) (within 673–873 K temperature range) of Na0.89Co0.90Me0.10O2 ceramics.
Figure 6. Temperature dependences of figure-of-merit ZT (a) and self-compability factor s (b) of Na0.89Co0.9Me0.1O2 (Me = Cr (1, red), Co (2, black), Ni (3, blue), Mo (4, pink), W (5, green), Pb (6, orange), Bi (7, purple)) ceramic samples. Insets show the figure-of-merit ZT1073 (c) and dimensionless relative self-compatibility factor Δs673–873K (d) (within 673–873 K temperature range) of Na0.89Co0.90Me0.10O2 ceramics.
Ceramics 08 00086 g006
Table 1. Values of sodium content (xNa), the average oxidation state of cobalt (Co+Z), lattice constants (a, c, c/a, V), Lotgering factor (f), size of the coherent scattering area obtained using the Debye–Scherrer equation (Ds) and the size-strain method (DSS), microstrain (ε) and X-ray density (dXRD) of Na0.89Co0.9Me0.1O2 (Me = Cr, Co, Ni, Mo, W, Pb, Bi) ceramic samples.
Table 1. Values of sodium content (xNa), the average oxidation state of cobalt (Co+Z), lattice constants (a, c, c/a, V), Lotgering factor (f), size of the coherent scattering area obtained using the Debye–Scherrer equation (Ds) and the size-strain method (DSS), microstrain (ε) and X-ray density (dXRD) of Na0.89Co0.9Me0.1O2 (Me = Cr, Co, Ni, Mo, W, Pb, Bi) ceramic samples.
MexNaCo+Za, Åc, Åc/aV, Å3fDs, nmDSS, nmε × 104dXRD, g/cm3
Cr0.8933.112.82510.933.87075.530.9069623.434.86
Co0.8913.112.82610.943.87275.710.3163454.564.98
Ni0.8893.232.83110.923.85675.750.6970702.614.88
Mo0.8942.782.82210.963.88375.570.8773701.415.06
W0.8892.782.82510.973.88475.800.9267691.505.43
Pb0.8883.012.82510.943.87375.640.9279603.225.55
Bi0.8942.902.82310.933.87275.450.7677722.495.56
Table 2. Values of apparent (dEXP) density, total (Πt), open (Πo) and closed (Πc) porosity, linear thermal expansion coefficient (α), electrical resistivity (ρ1073), Seebeck coefficient (S1073), power factor (P1073) and figure-of-merit (ZT1073) of Na0.89Co0.9Me0.1O2 (Me = Cr, Co, Ni, Mo, W, Pb, Bi) ceramics.
Table 2. Values of apparent (dEXP) density, total (Πt), open (Πo) and closed (Πc) porosity, linear thermal expansion coefficient (α), electrical resistivity (ρ1073), Seebeck coefficient (S1073), power factor (P1073) and figure-of-merit (ZT1073) of Na0.89Co0.9Me0.1O2 (Me = Cr, Co, Ni, Mo, W, Pb, Bi) ceramics.
MdEXP,
g/cm3
Πt, %Πo, %Πc, %105 × α, K−1104 × ρ1073,
Ω·m
S1073, μV/KP1073,
mW/(m·K2)
λ1073,
W/(m·K)
ZT1073
Cr3.183516191.683.281340.0550.6130.10
Co3.383219131.342.434390.7940.5361.59
Ni3.46292271.421.503690.9100.5911.65
Mo3.223619171.475.734080.2910.3230.97
W3.204117241.3910.95190.3200.3161.09
Pb3.344018221.265.583580.2300.2950.84
Bi3.473818201.255.976160.6360.3921.74
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Krasutskaya, N.S.; Chizhova, E.A.; Zizika, J.A.; Buka, A.V.; Wang, H.; Klyndyuk, A.I. Improving of Thermoelectric Efficiency of Layered Sodium Cobaltite Through Its Doping by Different Metal Oxides. Ceramics 2025, 8, 86. https://doi.org/10.3390/ceramics8030086

AMA Style

Krasutskaya NS, Chizhova EA, Zizika JA, Buka AV, Wang H, Klyndyuk AI. Improving of Thermoelectric Efficiency of Layered Sodium Cobaltite Through Its Doping by Different Metal Oxides. Ceramics. 2025; 8(3):86. https://doi.org/10.3390/ceramics8030086

Chicago/Turabian Style

Krasutskaya, Natalie S., Ekaterina A. Chizhova, Julia A. Zizika, Alexey V. Buka, Hongchao Wang, and Andrei I. Klyndyuk. 2025. "Improving of Thermoelectric Efficiency of Layered Sodium Cobaltite Through Its Doping by Different Metal Oxides" Ceramics 8, no. 3: 86. https://doi.org/10.3390/ceramics8030086

APA Style

Krasutskaya, N. S., Chizhova, E. A., Zizika, J. A., Buka, A. V., Wang, H., & Klyndyuk, A. I. (2025). Improving of Thermoelectric Efficiency of Layered Sodium Cobaltite Through Its Doping by Different Metal Oxides. Ceramics, 8(3), 86. https://doi.org/10.3390/ceramics8030086

Article Metrics

Back to TopTop