Yanan Lia,
Ping Wu
*b,
Shuai Ma
*c,
Mingdi Zhang
d and
Yili Peib
aCollege of Science, North China University of Technology, Beijing 100144, China
bSchool of Mathematics and Physics, University of Science and Technology Beijing, Beijing 100083, China. E-mail: pingwu@sas.ustb.edu.cn
cSchool of Mechanical Engineering, Tianjin University of Commerce, Tianjin 300134, China. E-mail: mashal0611@163.com
dRocket Force University of Engineering, Xi'an 710025, China
First published on 20th August 2025
Composite nanomaterial is a scientific solution to modulating the properties of thermoelectric materials. The thermoelectric properties of Ca3Co4O9 ceramic that remain stable in high-temperature air are relatively low, and how to decouple the relationship between heat and electricity is the focus of the research. This study systematically investigates the thermoelectric transport properties of composite Ca3Co4O9@xZnO materials. The samples do not undergo additional chemical reactions and exhibit lamellar microstructures. At temperatures up to 825 K, the thermal conductivity of pure Ca3Co4O9 was determined to be 2.58 W m−1 K−1, whereas that of the Ca3Co4O9@0.7ZnO was significantly reduced to 1.94 W m−1 K−1, a reduction of about 25%. Based on the effective medium theory analysis, adding ZnO in Ca3Co4O9 introduces interfacial thermal resistance and porosity, which is key in reducing thermal conductivity. Adding ZnO promotes the electrical conductivity enhancement of Ca3Co4O9 with a minor reduction in the Seebeck coefficient. Under the coordinated regulation of electrical and thermal properties, the ZT of Ca3Co4O9@0.7ZnO is enhanced by about 75% compared with that of pure Ca3Co4O9.
The core assessment index for the conversion effectiveness of thermoelectric materials is the dimensionless quality factor, defined as ZT = σS2T/(κc + κl), where σ stands for electrical conductivity; S is the Seebeck coefficient; κc denotes the thermal conductivity of carriers; κl stands for thermal conductivity of the lattice. To achieve high thermoelectric conversion efficiency, the following conditions must be met: high σ to guarantee efficient charge transport, high Seebeck coefficient to produce a significant thermoelectric effect, and low (κc + κl) to reduce heat loss. However, the ZT values of polycrystalline Ca3Co4O9 remain low due to the strong interdependence between the abovementioned parameters. Among the four thermoelectric parameters, the κl is the only one that can be independently tuned, improving the thermoelectric properties. The κl can be diminished through enhanced phonon scattering, and the main strategies include the introduction of point defects, dislocations, and interfacial scattering. In addition to introducing atomic size point defects by common ion doping,12–14 interfacial scattering is another important means of modulation. Interfacial scattering mainly affects low-frequency phonons and involves submicron scattering at grain boundaries and the interface between the second phase and the substrate. Heterogeneous interfaces can be increased by grain refinement, preparation of nano-precipitated phases, or direct composite nano-second phases. The additional barrier generated at the interface produces additional scattering of the carriers, suppressing their transport and reducing the mobility while enhancing the energy dependence of the carrier relaxation time and boosting the scattering factor value, increasing the Seebeck coefficient.15 In addition, introducing the nano-second phase forms a potential barrier, scatters low-energy carriers, produces an energy filtering effect, and raises the average carrier energy, further increasing the Seebeck coefficient.16 The composite of nanomaterials as the second phase into Ca3Co4O9 reduces the grain size and generates grain boundaries between the composite and the matrix. The rise in grain boundary density boosts long-wavelength phonon scattering, reducing lattice thermal conductivity. Therefore, the composite nanosecond phase is of significant research significance and value for optimizing the properties of thermoelectric materials.
Previous studies have revealed the diverse effects of SiO2, ZrO2, SiC, and Co3O4 composites with Ca3Co4O9 on their thermoelectric properties.17–20 In the previous work of our research team, by compositing nano second-phase materials MoSi2 and carbon nanotubes with Ca3Co4O9,21 we have adjusted for the electrical conductivity of Ca3Co4O9. Still, we have also significantly reduced its thermal conductivity. Based on an in-depth analysis of the effective medium theory and the series-parallel model, combined with actual measured thermal conductivity data, we find that the composite nanoscale second-phase material introduces additional scattering centers, interfacial thermal resistance, and pore structure, which together effectively reduce the thermal conductivity. Specifically, the σ of the Ca3Co4O9/0.1MoSi2 composite sample is enhanced to 83.16 S cm−1, a 49.2% increase in electrical conductivity compared to that of the pure Ca3Co4O9. Thermal conductivity of Ca3Co4O9/0.4MoSi2 decreased to 1.26 W m−1 K−1 at 1080 K, showing enhanced performance over pure Ca3Co4O9 (1.74 W m−1 K−1), a substantial thermal conductivity reduction was achieved.21 In addition, the ZT of the composite reaches 0.26, which is nearly 44.4% higher compared to the pure sample, and this result fully demonstrates the remarkable effect of the composite nanoscale second-phase material in optimizing the thermoelectric properties of Ca3Co4O9.
The incorporation of composite nanosecond phases leads to the formation of defects within the material. While these defects are beneficial for lowering thermal conductivity, they simultaneously hinder carrier mobility, adversely impacting electrical conductivity. The primary source of this phenomenon is that the nanoscale second phase can hinder the carrier transport path. Therefore, it is imperative to rationally regulate the distribution and morphology of the nanoscale second phase and its interfacial state with the substrate material. Compared with SiO2, ZrO2, SiC, Co3O4 and other oxides mentioned in the introduction, adding ZnO22 with an energy bandgap of 3.37 eV—a critical feature enabling potential electrical conductivity enhancement while suppressing thermal transport. In addition, ZnO is not only inexpensive and easy to prepare but also has excellent conductivity and good thermal stability, which has shown potential for broad application in a variety of fields, such as supercapacitors, solar cells, sensors, and high-temperature thermoelectric materials. Although Luo's group23 has investigated the Seebeck coefficient (∼119.3 μV K−1) and electrical properties of Ca3Co4O9/ZnO heterostructures by using pulsed laser deposition technique, the study on the composites of ZnO and Ca3Co4O9 ceramic materials and their effects on the thermoelectric properties have not been reported.
This research involved the synthesis of Ca3Co4O9@xZnO through the sol–gel method, employing atmospheric pressure sintering. A comprehensive analysis was conducted to explore how incorporating ZnO influences the thermoelectric performance of Ca3Co4O9. The results show that adding ZnO in suitable amounts reduces thermal conductivity and enhances electrical conductivity in Ca3Co4O9 ceramics, optimizing their thermoelectric properties.
![]() | ||
Fig. 1 (a) XRD pattern, (b–f) SEM pattern, and (g–i) CCO@0.7ZnO main element distribution map of CCO@xZnO. |
Sample | Atomic content | Atomic ratios | Densities (g cm−3) | ||||
---|---|---|---|---|---|---|---|
Ca | Co | Zn | O | Ca![]() ![]() |
Zn![]() ![]() |
||
CCO | 22.62 | 30.06 | 0 | 47.32 | 0.75 | 0 | 3.559 |
CCO@0.1ZnO | 22.26 | 28.67 | 0.37 | 48.70 | 0.78 | 0.01 | 3.729 |
CCO@0.3ZnO | 24.33 | 27.32 | 0.45 | 49.90 | 0.82 | 0.02 | 3.787 |
CCO@0.5ZnO | 23.63 | 28.32 | 1.03 | 47.72 | 0.83 | 0.04 | 3.820 |
CCO@0.7ZnO | 24.44 | 28.53 | 1.53 | 45.50 | 0.86 | 0.05 | 3.827 |
To explore the ionic state of the elements in the samples and the specific effect of the composite ZnO on the ionic ratios, the samples were tested in detail using XPS analysis. Fig. 2a illustrates the XPS full-spectrum scan results for each sample of CCO@xZnO, and the locations of the individual characteristic peaks have been clearly labeled. As shown in Fig. 2b, the XPS spectrum shows the characteristic peak of Zn 2p in ZnO (Zn 2p3/2 at 1021.8 eV and Zn 2p1/2 at 1044.9 eV). The peak of Zn 2p is relatively weak, but the intensity of the peak increases slightly with the increase of ZnO content.
Given that calcium and cobalt are the primary metal ion components in the sample, with cobalt being present in relatively high concentrations, high-resolution XPS spectra of the Co 2p region were selected for detailed analysis (Fig. 3a–e) to obtain precise data. The Co 2p spectrum reveals two distinct spin–orbit splitting peaks: for the undoped sample, the Co 2p3/2 and Co 2p1/2 maxima occur at 779.93 eV and 795.33 eV, respectively. For Co3+, the 2p3/2 and 2p1/2 peaks are observed at 779.6 eV and 794.8 eV, respectively, while for Co4+, these peaks shift to 781.4 eV and 796.8 eV. These binding energies fall within the expected ranges for Co3+ and Co4+. Additionally, a satellite peak occurs near 790.5 eV, attributed to Co3+. To assess the influence of composite ZnO on the properties of the sample, a peak fitting analysis was performed on the high-resolution Co 2p spectra (Fig. 3a–e). The Co3+/Co4+ ratio was determined from the area ratios of their peaks in Fig. 3f. As the ZnO content increased, the Co3+/Co4+ ratio remained stable at approximately 0.9, showing minimal variation. Furthermore, EDS data (Table 1) indicate that the Ca/Co atomic ratio remained nearly constant with increasing ZnO. Since previous XPS analysis confirmed the stability of Co content, it is inferred that the Ca content also remains unchanged, suggesting that ZnO does not participate in substitutional reactions within the sample but instead functions as a nanoscale additive at the pores of Ca3Co4O9. This is consistent with XRD analysis results.
![]() | ||
Fig. 3 (a–e) Co ion split peak fitting curves for CCO@xZnO and (f) relative Co3+ and Co4+ contents in the samples. |
In the Ca3Co4O9 system, the thermal conductivity is denoted as κ = κc + κl. According to the Wiedemann–Franz law, κc can be calculated through κc = LT/ρ (L is the Lorentz constant). The calculation shows relatively tiny values of κc (Fig. 4e). Therefore, k is mainly affected by κl (Fig. 4f). According to Matthiessen's rule, κl is primarily governed by point defect scattering, grain boundary scattering, and non-simple harmonic three-phonon Umklapp scattering. Borrowing from the Debye model for analyzing the heat conduction of gas molecules, the phonon contribution to the heat conduction can be expressed as κl = CVvl, where CV, v, and l represent the constant volume heat capacity per unit volume, the phonon propagation velocity, and the mean free-range, respectively. CV is an intrinsic property of the material.
ZnO doping reduces κl through multiple synergistic mechanisms. While increased density with higher ZnO content (Table 1) indicates reduced macro-porosity, which could potentially weaken pore-induced phonon scattering, the dominant effect arises from multi-scale interface engineering and defect-induced scattering. Specifically, nano-ZnO inclusions introduce dense heterogeneous interfaces between CCO and ZnO, with increasing ZnO content enhancing interface density and strengthening phonon barrier effects. Density mismatch between ZnO and the matrix induces localized crystal distortions, and the ionic radius difference between CCO and ZnO introduces stress field scattering, collectively intensifying mass fluctuation scattering. ZnO incorporation also reduces particle size, enhancing point defect scattering, while increased contact interfaces strengthen grain boundary scattering. Additionally, residual pores redistribute into nanoscale interfacial voids that scatter mid-frequency phonons more effectively than larger, isolated pores in undoped samples. These combined effects reduce phonon mean free path and propagation velocity, leading to significant κl suppression. Thus, ZnO doping effectively lowers the thermal conductivity of Ca3Co4O9, consistent with the observed 25% reduction in κ for the CCO@0.7ZnO sample at 825 K.
The out-of-plane thermal conductivity (c-axis direction) was measured using laser flash method. The study revealed that compared to in-plane thermal conductivity, the out-of-plane thermal conductivity is relatively lower,24,25 primarily attributed to enhanced phonon scattering effects at the interface between the layered structure and zinc oxide matrix. The c-axis direction inherently exhibits weak interlayer coupling characteristics, combined with the dispersed distribution of zinc oxide nanoparticles at grain boundaries, which collectively form additional barriers hindering phonon propagation. Consequently, the experimental results obtained in this study are lower than those measured in-plane.
![]() | (1) |
![]() | (2) |
The κ2 represents the thermal conductivity of Ca3Co4O9, while κ1 corresponds to the thermal conductivity of ZnO. V1 refers to the composite material's volume fraction of the ZnO, where ρcomp is the measured composite density, ρCCO = 3.559 g cm−3 and ρtheo,ZnO = 5.610 g cm−3 are theoretical densities, and ZnO is the weight fraction. The results calculated based on the above formulas are as follows: 0, 0.001, 0.006, 0.011, 0.016. RB indicates the interfacial thermal resistance between the ZnO particles and the Ca3Co4O9 matrix, which is about 4.0 × 10−9 m2 K−1 W−1.27 R denotes the diameter of ZnO. It is significant to mention that the particle size of Ca3Co4O9 in this research is about 700 nm, whereas the ZnO particles are much smaller, with a size of about 50 nm. Consequently, within the analytical framework of this study, ZnO particles can be approximated as small spherical entities. The specific values for the relevant parameters are provided in Table 2 for use in further analyses and calculations. Upon accounting for the interfacial thermal resistance, the thermal conductivity of Ca3Co4O9 with ZnO content introduced at ambient temperature can be evaluated—the effective thermal conductivity κcom, calculated as given in Fig. 5b.
Symbol | Meaning | Value |
---|---|---|
κ1 | Thermal conductivity of matrix material | 25 W m−1 K−1 |
κ2 | Thermal conductivity of the additive | 2.2 W m−1 K−1 |
V1 | Volume fraction of the additive phase | 0; 1.7%; 5.4%; 8.9%; 12.5% |
R | Additive phase diameter of the particles | 50 nm |
RB | Interfacial thermal resistance of ZnO | 4.0 × 10−9 m2 K W−1 |
κair | Air thermal conductivity | 0.023 W m−1 K−1 |
When examining ceramic-like substances, it is essential to account for how porosity influences thermal conductivity. Pang et al.26 extended the modeling approach outlined in eqn (2) to understand better how pores contribute to thermal conductivity. The model depicted the composite material as a single-phase solid characterized by effective thermal conductivity, κcom, while considering the pores as a separate phase. Pang et al. employed this method to create a detailed thermal conductivity model for composites, factoring in the influence of voids and interface thermal impedances.
In this model, the researchers calculated the composite material's effective thermal conductivity (κeffect), capturing the combined impact of factors such as porosity, interfacial thermal resistance, and the material's inherent properties on its overall thermal performance. This approach provides a holistic understanding of how these elements influence thermal conductivity. This research result is of great significance for the deep understanding of the heat conduction mechanism of ceramic-like materials, as well as guiding the development and optimization of related materials. The related derivation process and the final obtained expression of κeffect:26
![]() | (3) |
The thermal conductivity of the porous phase is κair, and the porosity ξ represents the proportion of pores in the composite unit cell. This parameter can be calculated by the formula ξ = (ρ0 − ρi)/ρ0, where ρ0 is the theoretical density value of the matrix material Ca3Co4O9, which was set to 4.68 g cm−3, and ρi represents the measured density of the individual measured density of a specific sample. Its values are detailed in Table 1. We analyze how pores and interfacial thermal resistance influence the thermal conductivity of CCO@ZnO, treating interfacial effects as a unified phase according to eqn (2). In contrast, the pores are regarded as another independent phase. In this context, κair is set to be the thermal conductivity of the air in the stomata (0.023 W m−1 K−1). Using these parameters and eqn (3), we calculated the composite's effective thermal conductivity, κeffect, containing stomata and interfacial thermal resistance. Results of the calculations are shown in Fig. 5b for visual analysis.
From the data in Fig. 5b, considering only the interfacial thermal resistance factor, the κcom of CCO@xZnO shows a decreasing trend as the ZnO addition is increased, and this trend coincides with the experimentally measured κ values. This indicates that adding ZnO introduces more interfaces, which serve as barriers to heat transfer, thereby lowering the thermal conductivity. Moreover, it is essential to recognize how stomata influence the κeffect in our analysis. It is found that the value of κeffect also shows a decreasing trend with the increase of ZnO addition, and this decrease is more significant compared to κcom. Especially at high temperatures, the κeffect value is closer to the experimentally measured k value. This phenomenon strongly indicates that pores significantly affect the effective thermal conductivity of the composite material.
To sum up, by incorporating ZnO into Ca3Co4O9, we successfully introduced two mechanisms to reduce thermal conductivity: interfacial thermal resistance and porosity. This result highlights ZnO as a viable approach to reducing thermal conductivity in Ca3Co4O9 while offering a solid theoretical foundation and practical insights for further optimizing the thermal conductivity characteristics in such composites.
We interpreted the experimental results using a small polariton hopping conduction model, under which the electrical conductivity of CCO@xZnO can be expressed as a function of the constants A, which is tied to the scattering mechanism, the carrier concentration n, the carrier power e, jumps distance a, the Boltzmann constant kB depending on the scattering mechanism . This equation emphasizes that the activation energy is an essential factor in determining the electrical conductivity and the carrier concentration. To quantify the activation energy, we further analyzed the plot of ln(σT) versus 1000/T presented in Fig. 6b and derived the activation energies within different temperature intervals (Region A: 850–450 K; Region B: 450–300 K) by linear fitting (the goodness-of-fit R2 was as high as 0.99).
To quantify the activation energy, we further analyzed the plot of ln(σT) versus 1000/T presented in Fig. 6b and derived the activation energies within different temperature intervals (Region A: 850–450 K; Region B: 450–300 K) by linear fitting (the goodness-of-fit R2 was as high as 0.99). Fig. 6c illustrates the fitting results, showing that the activation energy is stable at about 0.3 eV within Region B. In contrast, in Region A, the activation energy fluctuates with increased ZnO doping. The fluctuation of activation energy in the higher temperature region with increasing ZnO content arises from the competitive interplay of three key mechanisms. First, nonlinear interfacial charge transfer: low ZnO content lowers activation energy via sufficient charge transfer; moderate content raises it due to agglomeration-induced hole over-compensation; high content reverses the trend as agglomeration reduces interfacial efficiency. Second, dynamic defect evolution: ZnO-induced oxygen vacancies and fluctuating Co3+/Co4+ valence states (with differing ionization energies) disrupt stability. Third, temperature-dependent stress scattering: uniform stress at low content eases with heat, lowering activation energy; moderate-to-high content causes stress concentration (raising it) and excessive agglomeration triggers “stress shielding” (slightly lowering it). These coupled effects reflect the transition from lattice-dominated to interface-scattering-dominated transport, driving the fluctuation. These coupled effects reflect the transition from intrinsic lattice-dominated to interface-scattering-dominated carrier transport, resulting in the observed activation energy fluctuations.
Furthermore, the size difference between the nano-ZnO (≈50 nm) and Ca3Co4O9 matrix particles (≈700 nm) is pivotal in regulating the electrical transport properties of the composites: nano-ZnO tends to distribute in the pores of the CCO matrix, optimizing electron transport paths and enhancing heterojunction interfacial interactions. Hall test results (Table 3) confirm hole-dominated conduction (positive Hall coefficients) and reveal a continuous increase in carrier concentration with rising ZnO content, which is primarily attributed to charge transfer at the ZnO-CCO heterojunction interface and defect regulation—an effect unaffected by the mobility variation trend. The carrier mobility exhibits a “first increase then decrease” trend, driven by evolving scattering mechanisms: at low ZnO content, uniform dispersion of nanoparticles optimizes the interface structure, reducing intrinsic defect scattering centers in CCO and suppressing random carrier scattering at grain boundaries, thereby increasing mobility; at excessive ZnO content, a sharp rise in grain boundary density, coupled with local agglomeration and potential lattice distortion, intensifies carrier scattering between multiple grain boundaries, leading to decreased mobility. The overall increasing trend in electrical conductivity with ZnO content arises from the synergistic effect of carrier concentration and mobility (σ = enμ), where the dominant positive contribution of continuously increasing carrier concentration outweighs the negative impact of later-decreasing mobility, resulting in enhanced conductivity.
Sample | σ (S cm−1) | ρ (Ω cm−1) | n (1019 cm−3) | μ (cm2 V−1 S−1) | RH (cm−3 C−1) | f |
---|---|---|---|---|---|---|
CCO | 28.186 | 0.035 | 1.962 | 8.979 | 0.319 | 0.975 |
CCO@0.1ZnO | 33.378 | 0.030 | 2.159 | 9.663 | 0.289 | 0.998 |
CCO@0.3ZnO | 35.677 | 0.028 | 2.247 | 9.923 | 0.278 | 0.997 |
CCO@0.5ZnO | 37.021 | 0.027 | 2.887 | 8.031 | 0.217 | 0.998 |
CCO@0.7ZnO | 37.604 | 0.027 | 3.288 | 7.148 | 0.190 | 0.981 |
Moreover, the electrical conductivity measured by the four-probe method primarily reflects the in-plane (ab-plane) transport behavior. The in-plane electrical conductivity (σ∥) exhibits higher values due to the intrinsic layered structure of the Ca3Co4O9 matrix and the strong texture induced by uniaxial pressing along the ab-plane. In this direction, carrier transport benefits from reduced lattice scattering and continuous intra-layer conduction paths. The introduction of ZnO nanoparticles further optimizes in-plane electrical conductivity by enhancing interfacial charge transfer and reducing defect scattering within the ab-plane, contributing to the observed increasing trend with ZnO content.
The S of ZnO-added samples have a slight decrease compared to pure Ca3Co4O9. It can be analyzed from two aspects. The S of the semiconductor material with the composite of two substances can be expressed as:
![]() | (4) |
On the other hand, from the analysis of micro mechanism, as shown in Table 3, when ZnO exists only as nanoscale inclusions, the carrier concentration n increases with the increase of ZnO content, which is attributed to the interfacial physical effects, among which the interfacial charge transfer is the main factor. According to the band bending effect, when ZnO (work function ≈5.2 eV) contacts with Ca3Co4O9 (≈4.8 eV), a Type-II heterojunction is formed, leading to the redistribution of interfacial charges. Electrons flow from Ca3Co4O9 to ZnO, resulting in the increase of hole concentration in the composite. Meanwhile, the change of carrier concentration has a direct impact on the S (consistent with the Pisarenko relation: S ∝ 1/n). Consequently, the S decreases slightly with the increase of ZnO content, which is mainly caused by the increase of carrier concentration due to the additional holes introduced by interfacial defects.
In Fig. 6f, PF (PF = σS2) curves with temperature are demonstrated for a series of samples of CCO@xZnO. The graph shows a gradual increase in PF as the temperature rises. This phenomenon results from the combined effect of electrical conductivity and the Seebeck coefficient, indicating that the addition of ZnO can effectively regulate the power factor of the sample. The PF value of the CCO@0.7ZnO improved to 1.93 μW cm−1 K−2 at 1080 K, up from 1.85 μW cm−1 K−2 for the undoped pure Ca3Co4O9. This finding provides a new perspective for optimizing the thermoelectric characteristics of Ca3Co4O9 through ZnO doping.
This journal is © The Royal Society of Chemistry 2025 |