In order to verify the phase structure, the synthesized ceramics were examined by XRD at room temperature. The measured XRD spectra of SrxLa0.1TiO3/Ti composite ceramics sintered in argon and then annealed at 1350°C are shown in Fig. 1. Rietveld refinement (Fig. 2.) was carried out on all XRD spectra to confirm that the main diffraction peaks can be indexed to Sr0.9La0.1TiO3 cubic perovskite phase (PDF#79–0179, space group \(\text{P}\text{m}\overline{3}\text{m}\)). With the increase of A-site cation vacancy content, the diffraction angle of the XRD diffraction peak does not shift significantly, and the lattice parameters of cubic phase calculated by Rietveld analysis also show unconspicuous changes overall, as shown in Table 1. This reflects the subtle and complex effect of defects formed by charge compensation mechanism on lattice expansion/contraction. Generally, the reduction from Ti4+ ions to Ti3+ ions and oxygen vacancies, can promote the expansion of perovskite lattice, while A-site cation vacancies can lead to lattice contraction. The lattice parameters of samples containing A-site cation vacancies significantly increase compared to pure STO (3.905 Å). Considering that the ionic radius of La (1.36 Å) is smaller than that of Sr (1.44 Å), the lattice expansion is mainly due to the generation of oxygen vacancies. However, for the defect formation process under reductive atmosphere conditions, the A-site cation vacancy promotes the formation of oxygen vacancies to a certain extent, while combining with each other to form Schottky defect pairs, which can be been as self compensation for fluctuations in crystal volume. All XRD spectra show several small additional diffraction peaks, which are at ~ 30°, ~ 42° and ~ 51°, respectively. The detected secondary phase peaks can be marked as \({\text{T}\text{i}}_{1-\text{m}}^{4+}{\text{T}\text{i}}_{\text{m}}^{3+}{\text{O}}_{2-\text{m}/2}\) (Ti3O5 and Ti6O11) phases, which are mainly the oxidation products formed by nano-sized metal titanium after sintering. The presence of secondary phases is further confirmed by the contrast between light and dark ceramic microstructure, as shown in Fig. 3. A reasonable explanation for the evolution of phase composition is that during the Ar atmosphere sintering and Ar + C annealing stages, the titanium dioxide formed by the oxidation of nano-sized Ti powder further reacts with SrxLa0.1TiO3 to form a new perovskite phase, and the remaining titanium dioxide is reduced to the secondary phase \({\text{T}\text{i}}_{1-\text{m}}^{4+}{\text{T}\text{i}}_{\text{m}}^{3+}{\text{O}}_{2-\text{m}/2}\) under the reducing atmosphere. Additionally, it should be noticed that the observed diffraction peak intensity of the secondary phase has not changed significantly with A-site composition, indicating that the content of the secondary phase does not increase significantly.
Table 1
Crystal structural information and measured densities for SrxLa0.1TiO3/Ti composites.
Composition code | Nominal composition | Space group | Cell parameter (Å) | Cell Volume (Å3) | Measured density (g/cm3) |
Pristine STO | SrTiO3 | \(\text{P}\text{m}\overline{3}\text{m}\) | 3.905 | 59.5474 | / |
S90 | Sr0.9La0.1TiO3 | \(\text{P}\text{m}\overline{3}\text{m}\) | 3.90928(5) | 59.7437 | 4.77 |
S875 | Sr0.875La0.1TiO3 | \(\text{P}\text{m}\overline{3}\text{m}\) | 3.90956(9) | 59.7567 | 4.78 |
S85 | Sr0.85La0.1TiO3 | \(\text{P}\text{m}\overline{3}\text{m}\) | 3.90962(4) | 59.7592 | 4.81 |
S825 | Sr0.825La0.1TiO2.975 | \(\text{P}\text{m}\overline{3}\text{m}\) | 3.90924(6) | 59.7419 | 4.71 |
To evaluate the microscopic morphology of the annealed sample, the SEM back-scattered electron (BSE) images of SrxLa0.1TiO3/Ti ceramics are present in Fig. 3. No visible pores were observed in the samples, which revealed that uniform and well dense ceramics were obtained. Two kinds of grain phases can be found in all ceramics, gray grain clusters as intergranular phase and lighter colored grains as matrix phase. In terms of size and shape, these grains have no significant difference between different component microstructures. The lighter colored grains are irregular polygon shape with an average size range of 4.5 ~ 6 µm. The concentration of A-site cationic vacancies has no significant regular effect on the size of the grains. The illustration of histograms in Fig. 3 shows the grain size distribution of four component samples. Compared with S85 and S825 samples, the grain size distribution of S90 and S875 is more concentrated. Furthermore, it is observed that the sizes of different gray grains in the matrix of SrxLa0.1TiO3/Ti show noticeable difference. The apparent density of ceramic samples first increases and then decreases with the increase in content of A-site vacancies, indicating a similar trend in relative density, as shown in Table 1. This is because the large-sized ions with 12 coordination occupy the A-site in the ABO3 perovskite structure, where the generation of cationic vacancies facilitates ion transfer and enhances ion diffusion during the ceramic densification process [34]. However, excessive cation vacancies can also lead to an increase in pores in the microstructure, which results in a decrease in density. The distribution of all elements by EDS mapping for secondary phase of the S875 sample showed an evident aggregation of Ti and O and the absence of Sr, as shown in Fig. 4. To identify light and dark gray grains, element composition analysis measured from the corresponding area is performed and listed in Table 2. Therefore, it can be further inferred that the dark gray phase is a non-stoichiometric titanium oxide phase. The concentration of A-site vacancies in S875 sample normalized to Ti level is 16.2%, which can be inferred from the measured concentration of (Sr + La + Bi). EDS quantitative elemental analysis shows that the (Sr + La + Bi)/Ti ratio is lower than the nominal ratio of S875 sample. It should be attributed to the further reaction between TiO2 generated by nano-sized metal titanium and Sr0.875La0.1TiO3, which further increases the concentration of actual A-site vacancies [28].
Table.2. EDS point analysis in the secondary phase and matrix grains of S875 composite sample.
Composition | Element | Content(atom%) |
Point 1# | Sr | 17.55 |
Ti | 25.55 |
O | 53.04 |
La | 2.48 |
Bi | 1.38 |
Point 2# | Ti | 39.67 |
O | 60.33 |
The degree of reduction of Ti in four components was assessed by X-Ray photoelectron spectroscopy to further reveal their electronic structure characteristics. The high-resolution deconvoluted Ti 2p spectra are presented in Fig. 5. The asymmetric Ti 2p peak indicates the simultaneous presence of Ti4+and Ti3+. The ratios of Ti3+ compared to Ti4+ were estimated by the deconvolution characteristic peaks and the calculated relevant peak areas [35]. The [Ti3+] content for SrxLa0.1TiO3/Ti samples decreases with increasing Sr vacancy content (from 21.75–4.68%), as shown in Table 3. It is widely known that the concentration of [Ti3+] is closely correlated with the carrier concentration in SrTiO3-based ceramics [17]. It suggests that A-site vacancy has an effect on carrier concentration, which prohibits the reduction process from Ti4+ to Ti3+. As a matter of fact, the detected Ti3+ in the sample may come from both the titanium oxide intergranular phase and the matrix phase according to the above experimental results. Meanwhile, the generation of Ti3+ in the matrix phase can be divided into two stages, which can be illustrated by taking the S875 sample as an example. The first stage of reduction is electronic defect charge compensation caused by La3+ doping, which is described as:
$$\text{0.875SrC}{\text{O}}_{3}+0.05\text{L}{\text{a}}_{2}{\text{O}}_{3}+\text{T}\text{i}{\text{O}}_{2}\underrightarrow{\text{c}\text{a}\text{l}\text{c}\text{i}\text{n}\text{a}\text{t}\text{i}\text{o}\text{n}}0.875{\text{S}\text{r}}_{\text{S}\text{r}}^{\text{x}}\text{+0.1}{\text{L}\text{a}}_{\text{S}\text{r}}^{{\bullet }}\text{+}$$
$$\text{0.9}\text{5}{\text{T}\text{i}}_{\text{T}\text{i}}^{\text{x}}+{0.025\text{V}}_{\text{S}\text{r}}^{{\prime }{\prime }}+0.05{\text{T}\text{i}}_{\text{T}\text{i}}^{{\prime }} \left[{\text{T}\text{i}}^{3+}\left(\text{d}\right)\right]\text{+3}{\text{O}}_{\text{O}}^{\text{x}}\text{+}\text{0.0125}{\text{O}}_{2}\text{↑}\text{+}\text{0.875C}{\text{O}}_{2}\uparrow$$
1
Accompany by the release of oxygen, the Ti ion in the A-site nonstoichiometric samples will be reduced to electronic defects \({\text{T}\text{i}}_{\text{T}\text{i}}^{{\prime }}\) that can be recorded as Ti3+ (d), and the strontium deficiency partially offsets the excess charge of La3+. The second stage of reduction depends on the generation and ionization of oxygen vacancy defects by the reduction atmosphere treatment. The electronic defect Ti3+ generated during this process can be recorded as Ti3+ (v). The chemical reaction can be presented as:
$$\text{Ti}{\text{O}}_{2}\underrightarrow{\text{r}\text{e}\text{d}\text{u}\text{c}\text{t}\text{i}\text{o}\text{n}}(1-4{\epsilon }){\text{T}\text{i}}_{\text{T}\text{i}}^{\text{x}}+4{\epsilon }{\text{T}\text{i}}_{\text{T}\text{i}}^{{\prime }} \left[{\text{T}\text{i}}^{3+}\left(\text{v}\right)\right]+(2-2{\epsilon }){\text{O}}_{\text{O}}^{\text{x}}+{2{\epsilon }\text{V}}_{\text{O}}^{{\bullet }{\bullet }}+{\epsilon }{\text{O}}_{2}\uparrow$$
2
$${\text{S}\text{r}}_{0.875}{\text{L}\text{a}}_{0.1}{\text{(Ti}}_{0.95}^{4+}{\text{Ti}}_{0.05}^{3+}{\text{)O}}_{3}\underrightarrow{\text{r}\text{e}\text{d}\text{u}\text{c}\text{t}\text{i}\text{o}\text{n}}0.875{\text{S}\text{r}}_{\text{S}\text{r}}^{\text{x}}\text{+}{0.025\text{V}}_{\text{S}\text{r}}^{{\prime }{\prime }}\text{+}\text{0.1}{\text{L}\text{a}}_{\text{S}\text{r}}^{{\bullet }}\text{+(0.95-4θ)}{\text{T}\text{i}}_{\text{T}\text{i}}^{\text{x}}+$$
$$0.05{\text{T}\text{i}}_{\text{T}\text{i}}^{{\prime }}+4{\theta }{\text{T}\text{i}}_{\text{T}\text{i}}^{{\prime }} \left[{\text{T}\text{i}}^{3+}\left(\text{v}\right)\right]\text{+}{2{\theta }\text{V}}_{\text{O}}^{{\bullet }{\bullet }}\text{+(3-2θ)}{\text{O}}_{\text{O}}^{\text{x}}+{\theta }{\text{O}}_{2}\uparrow$$
3
Where ε and θ are respectively defined as oxygen vacancies concentrations of the matrix phase and the second phase titanium dioxide generated during the reductive process. Formula 2 explains how secondary phases \({\text{T}\text{i}}_{1-\text{m}}^{4+}{\text{T}\text{i}}_{\text{m}}^{3+}{\text{O}}_{2-\text{m}/2}\) are produced. Taking into account the contribution of A-site vacancies to charge neutrality, the charge balance equation can be written as:
$${[\text{T}\text{i}}_{\text{T}\text{i}}^{{\prime }}\left]{+2[\text{V}}_{\text{S}\text{r}}^{{\prime }{\prime }}\right]\approx \left[{\text{L}\text{a}}_{\text{S}\text{r}}^{{\bullet }}\right]+2\left[{\text{V}}_{\text{O}}^{{\bullet }{\bullet }}\right]$$
4
The general consensus is that, [Ti3+] is equivalent to the concentration of n-type charge carriers (n) under reduction conditions. Therefore, the introduction of A-site cation vacancies decreases the carrier concentration of SrxLa0.1TiO3/Ti composite samples, which is consistent with the subsequent Hall test results.
Table.3. XPS results for Ti cations in SrxLa0.1TiO3/Ti composite samples.
Composition | Ti 2P (%) |
Ti4+ | Ti3+ |
S90 | 78.25 | 21.75 |
S875 | 84.26 | 15.74 |
S85 | 88.55 | 11.45 |
S825 | 95.32 | 4.68 |
Various electrical transport characteristics were measured for four SrxLa0.1TiO3/Ti ceramics, as shown in Fig. 6. The electrical conductivity of all samples with different A-site deficient levels are plotted as a function of temperature in the range of 300 K to 1073 K. It can be seen that the electrical conductivity shows a similar tendency that first increases substantially until 523 K ~ 573 K, then gradually decreases with a small slope with the rising temperature. The samples show semiconductor behavior at a relatively low temperature range, which indicates that the raising carrier concentration and mobility increasing the electrical conductivity, and the carriers are mainly scattered by ionized impurities (donor ion). At higher temperature ranges, the lattice vibration and acoustic phonon scattering are dominant, and the carrier mobility decreases continuously with rising temperature, leading to a decrease in conductivity [36]. Therefore, the semiconductor-metal (S-M) transition of conductivity is approximately 550 K. Besides, it is clear that Sr vacancies can improve the electrical conductivity, especially at low temperatures. At 1073 K, the electrical conductivity of S875 composite is still slightly higher than that of S90 composite.
In order to investigate the electrical transport mechanism, the electrical conductivity below the S-M transition temperature is analyzed by the thermally excited small polaron hopping (SPH) conduction mechanism, which is expressed by the following equation [37–38]:
$${\sigma }=\frac{{\sigma }_{0}}{T}\text{e}\text{x}\text{p}\left[-\frac{{E}_{\text{H}\text{o}\text{p}}}{{k}_{\text{B}}T}\right]$$
5
where σ0 is a constant, T is the absolute temperature, EHop is the activation energy of the small polaron transition, and KB is the Boltzmann constant. The relationship between lnσT and KBT is depicted in Fig. 6b, indicating that the carriers of all samples followed the SPH conduction model. With increasing Sr vacancy concentration in SrxLa0.1TiO3/Ti samples, the activation energy (EHop) decreases, providing more opportunity for the transition of small phonon, as displayed in Table 4. Therefore, A-site vacancy for SrxLa0.1TiO3/Ti ceramics is beneficial for the electrical transport performance.
The negative Seebeck coefficients of all samples have been obtained in the whole measuring temperature range, as shown in Fig. 6c. This suggests that SrxLa0.1TiO3/Ti ceramics are n-type semiconductor materials with electrons as dominant charge carriers, which also corresponds to the negative value of the carrier concentration (Table 4). The absolute Seebeck value of all samples increases monotonously with increasing temperature from 373 K to 1073 K. The S85 sample reaches the highest absolute Seebeck coefficient among all samples, which is ~ 208 µV/K at 1073 K. In addition, as Sr vacancy concentration increasing, the absolute value |S| of all samples showed a trend of first increasing and then decreasing in the whole measuring temperature range. It is worth noting that the |S| value in 1073 K has been found to be increased by 6.7% from 195 µV/K for S90 sample to 208 µV/K for S875 sample. Obtaining a high Seebeck coefficient and conductivity simultaneously primarily depends on large carrier mobility, as shown in Table 4.
The temperature dependence of the calculated power factor (Sσ2) is shown in Fig. 6d. For all samples, the power factor (PF) value increases with increasing temperature. The PF value follows a similar trend to Seebeck coefficient, initially increasing with Sr vacancy concentration and then decreasing. It was observed that S875 sample exhibits the highest PF value through the combination of the highest electrical conductivity and Seebeck coefficient. At 873 K, the maximum PF value of ~ 426 µW/mK2 is obtained for S875 sample, which is almost 20% increase compared with S90 sample (~ 356 µW/mK2).
Furthermore, the carrier concentration (n) and mobility (µH) of SrxLa0.1TiO3/Ti composites at room temperature were measured using Hall measurement technology, and it was found that the carrier concentration range of all samples is (1.3 ~ 2.3)×1019 cm3, as shown in Table 4. The Seebeck coefficient expression of degenerate semiconductors can be used to estimate the effective mass (m*) of all samples and is displayed as follows [39].
$$S=\frac{8{{\pi }}^{2}{k}_{\text{B}}^{2}}{3e{ℎ}^{2}}{m}^{\ast }T{\left(\frac{{\pi }}{3n}\right)}^{2/3}$$
6
where e is the electric quantity of the electron and h is the Planck constant. It can be seen from the Pisarenko diagram (Fig. 6f) that the effective mass of electrons at room temperature shows a decreasing trend with increasing Sr vacancy concentration. In order to better illustrate the significant role of electron mobility in improving the electrical transmission performance, the weighted mobility (µw) independent of the carrier concentration is calculated using Eq. 6, which is calculated using the data of electrical conductivity and Seebeck coefficient [40–41].
$${\mu }_{w}=\frac{3{ℎ}^{3}\sigma }{8\pi e{\left(2{m}_{e}{k}_{\text{B}}T\right)}^{3/2}}\left[\frac{\text{e}\text{x}\text{p}\left[\frac{\left|S\right|}{{k}_{\text{B}}/e}-2\right]}{1+\text{e}\text{x}\text{p}\left[-5\left(\frac{\left|S\right|}{{k}_{\text{B}}/e}-1\right)\right]}+\frac{\frac{3}{{\pi }^{2}}\frac{\left|S\right|}{{k}_{\text{B}}/e}}{1+\text{e}\text{x}\text{p}\left[5\left(\frac{\left|S\right|}{{k}_{\text{B}}/e}-1\right)\right]}\right]$$
7
where me is the electronic mass. SrxLa0.1TiO3/Ti composites have exhibited an upward trend of the weighted mobility prior to the S-M transition temperature (373–573 K), attributed to the thermal activation of carriers under the action of high temperature (Fig. 6e). Subsequently, in the temperature region of metallic conductivity behavior, driven by the carrier acoustic phonon scattering, µw decreases in a T3/2 dependence with temperature. At about 480 K, the maximum weighted mobility of S85 sample is ~ 15.2 cm2V− 1S− 1, which is 26% higher than that of S90 sample due to the introduction of A-site deficiency.
Table.4. Activation energy (EHop) of the small polaron transition, room-temperature carrier concentration (n) and carrier mobility (µH) for SrxLa0.1TiO3/Ti composite samples.
Composition | EHop (eV) | n (×1019 cm3) | µH (cm2/V/s) |
S90 | -0.0739 ± 0.0038 | 2.24 | 23.28 |
S875 | -0.0646 ± 0.0033 | 2.15 | 49.15 |
S85 | -0.0614 ± 0.0035 | 1.31 | 109.21 |
S825 | -0.0609 ± 0.0032 | 1.40 | 77.10 |
The temperature dependence of the total thermal conductivity of SrxLa0.1TiO3/Ti composite ceramics are presented in Fig. 7a. The total thermal conductivity of all samples within the measured temperature range drop monotonically as temperature rises, originating from the phonon-phonon Umklapp scattering. It is noteworthy that the total thermal conductivity increases slightly and then decreases with the increase of Sr vacancy in the whole temperature range. The electronic thermal conductivity exhibits a similar variation trend with temperature in accord with the electrical conductivity. It is notable that the proportion of κe/κtot is about 7%, indicating that the lattice thermal conductivity plays a dominant role in the total thermal conductivity. Furthermore, The κL reduced following the similar trend of the κtot with the increasing Sr vacancy content and increasing temperature in the whole temperature range. This phenomenon mainly depends on the additional phonon scattering due to the increase of crystal lattice disorder by A-site deficiency. The effect of A-site vacancy on decreasing lattice thermal conductivity is particularly remarkable at 373 K. The lattice thermal conductivity for S90 sample is 3.56 Wm− 1K− 1, while the lattice thermal conductivity for S825 sample decreases to 3.25 Wm− 1K− 1, decreasing by about 9%. As a consequence, it was discovered that with rising temperature, the dimensionless figure-of-merit ZT value for SrxLa0.1TiO3/Ti composite ceramics increased, as shown in Fig. 7d. The variation trend of ZT value with component is the same as PF value. The S875 sample obtained the highest ZT value, 0.16, which was attained at 1023 K. As can be seen, the thermoelectric properties of SrxLa0.1TiO3/Ti composites can be further enhanced by introducing an appropriate amount of A-site vacancies into the perovskite structure.