Next Article in Journal
Indium-Based Micro-Bump Array Fabrication Technology with Added Pre-Reflow Wet Etching and Annealing
Previous Article in Journal
Statistical Analysis of Uniform Switching Characteristics of Ta2O5-Based Memristors by Embedding In-Situ Grown 2D-MoS2 Buffer Layers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Electrical and Thermal Transport Properties of La-Doped SrTiO3 with Sc2O3 Composite

1
School of Physics and Materials Science, Guangzhou University, Guangzhou 510006, China
2
School of Materials Science and Engineering, Shanghai University, 99 Shangda Road, Shanghai 200444, China
3
Materials Genome Institute, Shanghai University, 99 Shangda Road, Shanghai 200444, China
4
School of Materials Science and Engineering, Guilin University of Electronic Technology, Guilin 541004, China
5
Guangxi Key Laboratory of Information Materials, Guilin University of Electronic Technology, Guilin 541004, China
*
Authors to whom correspondence should be addressed.
Materials 2021, 14(21), 6279; https://doi.org/10.3390/ma14216279
Submission received: 21 August 2021 / Revised: 23 September 2021 / Accepted: 23 September 2021 / Published: 21 October 2021
(This article belongs to the Section Advanced and Functional Ceramics and Glasses)

Abstract

:
Donor-doped strontium titanate (SrTiO3) is one of the most promising n-type oxide thermoelectric materials. Routine doping of La at Sr site can change the charge scattering mechanism, and meanwhile can significantly increase the power factor in the temperature range of 423–773 K. In addition, the introduction of Sc partially substitutes Sr, thus further increasing the electron concentration and optimizing the electrical transport properties. Moreover, the excess Sc in the form of Sc2O3 composite suppresses multifrequency phonon transport, leading to low thermal conductivity of κ = 3.78 W·m−1·K−1 at 773 K for sample Sr0.88La0.06Sc0.06TiO3 with the highest doping content. Thus, the thermoelectric performance of SrTiO3 can be significantly enhanced by synergistic optimization of electrical transport and thermal transport properties via cation doping and composite engineering.

1. Introduction

With the sustainable development of global industrialization, the demand for energy is rapidly growing in recent years, which promotes researchers to explore clean and renewable energy technology. Thermoelectric (TE) materials, enabling the direct interconversion between heat and electricity based on the Seebeck effect and the Peltier effect, would play important role in the energy depletion [1,2,3]. The conversion efficiency of TE materials is essentially determined by the dimensionless figure of merit ZT = σS2T/(κlat + κe). where σ, S, T, κlat, and κe represent the electrical conductivity, Seebeck coefficient, absolute temperature, the lattice and electronic components of thermal conductivity κtot, respectively [4]. Accordingly, high thermoelectric properties require synergistic optimization of electrical and thermal transport properties, and thus lattice softening [5], nanostructure engineering [6,7], band convergence [8,9,10], multiscale phonon scattering including dislocation engineering [11,12], point defect and grain boundary scattering [13,14], have been proposed and developed in these years.
Due to the low-cost, excellent thermal stability, environmental compatibility, and unique oxidation-proof features at high temperatures, transition metal oxides such as NaxCoO2 [15,16], and Ca3Co4O9 [17,18,19] are suitable for p-type thermoelectric candidates. Especially, their significantly high thermal stability allows maintaining large temperature differences (ΔT) in thermoelectric devices, making them possible to achieve high output power [20,21]. However, the poor electrical conductivity restricts the power factor (PF = σS2) and thermoelectric figure of merit ZT. Therefore, many researches have been focusing on the strategies for optimizing the electrical transport properties. For example, a PF as high as 18.92 μW·cm−1 K−2 at 1100 K in NaxCoO2 can be realized via Ag composite with large electron density of ~1021·cm−3 [22].
For n-type oxide thermoelectric materials, strontium titanate (STO) undergoing donor-doping has obtained much attention as a result of their promising thermoelectric properties [23,24]. The band structure calculations reveal that there are heavy and light bands around the Fermi level contributing to the electron transport in SrTiO3, favoring large Seebeck coefficients [25]. In this situation, large power factors of 28–36 μW·cm−1 K−2 at room temperature has been achieved in n-type Sr1-xLaxTiO3 single crystal with relatively high carrier concentrations of (0.2–2) × 1021 cm−3 [26]. However, the thermoelectric performance can be further boosted with reduced thermal conductivities. In context of the lattice thermal conductivity κlat contributing 75–100% of the total thermal conductivity, suppression of phonon transport would enable the optimization of thermoelectric performance in perovskite titanate thermoelectrics (ABO3) [27]. The simple and effective strategy is to introduce point defects by disordering A site to strengthen the phonon scattering. It is reported that doping ions with a smaller ion radius at the A site can reduce thermal conductivity well, while doping ions with a closed radius with Sr can significantly improve electrical transport performance [28,29].
In this work, La doping and Sc2O3 composite have been utilized for the synergistic optimization of electrical and thermal transport properties. Substitution Sr with trivalent La aims to increase the electrical conductivity of SrTiO3, while compositing Sc2O3 is expected to reduce the thermal conductivity. The power factor reaches 9.41 μW·cm−1·K−2 at 517 K. In addition, point defect induced the stress and mass fluctuation favor for the enlargement of expansion coefficients and reduction of lattice thermal conductivity. As a result, the ZT = 0.143 has been obtained for the sample Sr0.88Sc0.06La0.06TiO3 at 773 K.

2. Materials and Methods

2.1. Sample Preparation

Undoped and doped strontium titanate powders were prepared by solid state reaction method, using SrCO3 (99.8%), TiO2 (99.8%), La2O3 (99.9%), and Sc2O3 (99.9%) as raw materials. These powers were weighted according to the stoichiometric ratio Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06), and mixed via ball milling at a speed of 200 r/min for 48 h with stainless steel pots and zirconia balls. The as-obtained mixtures were cold-pressed into tablets with ϕ10 mm × 2 mm, which were then placed in a muffle furnace for annealing at 1573 K for 6 h in air. The as-annealed samples were ground into fine powders by ball milling again with 500 r/min for 12 h. Finally, dense ceramic samples (ϕ10 mm × 2 mm) were prepared by spark plasma sintering (SPS) with graphite dies under 1473 K and 30 MPa for 5 min.

2.2. Phase and Microstructure Characterization

The phase purity of the as-prepared samples was examined by powder X-ray diffraction (PXRD, Rigaku, Japan, Cu Kα radiation, λ = 1.541854 Å, 20° < 2θ < 80°, step width 0.02°) at room temperature. The lattice parameters were calculated using the software of WinCSD (version 4.19, L. Akselrud. Kyiv, Ukraine) [30]. The microstructure and composition were characterized by scanning electron microscope (SEM; ZEISS Gemini 300, Jena, Germany), equipped with energy-dispersive X-ray spectroscopy (EDX), which was performed at the accelerating voltage of 15 kV (Oxford X-MAX, Oxford, UK). The average grain sizes were examined from the observed microstructure by image analysis using the Image-Pro program (Plus 6.0, 2018, Media Cybernetics, MD, USA) [31].

2.3. Thermoelectric Performance Measurements

The Seebeck coefficient and electrical conductivity of the samples were simultaneously measured using a ZEM-3 instrument (ULVAC-RIKO, Kanagawa, Japan) under helium atmosphere from room temperature to 773 K. The room temperature Hall coefficient (RH), Hall carrier concentration (nH), and Hall mobility (μH) were collected with a Hall effect test system (Lake Shore 8400, Westerville, OH, USA) using the four-probe van der Pauw method under a reversible magnetic field of 0.9 T. The thermal expansion coefficients were obtained from 500 K to 800 K by a thermomechanical analyzer (NETZSCH, TMA 402F3, Selb, Germany). The thermal conductivity can be calculated according to the equation κ = Cpλd, where Cp is the specific heat capacity, λ is the thermal diffusivity, and d is the density. A laser flash diffusivity (NETZSCH, LFA467, Selb, Germany) was used to measure λ of a tablet sample with a diameter of 10 mm and a typical thickness of 1–2 mm. Prior to the measurement, the samples were coated with a thin graphite layer to minimize the error of material emissivity. The specific heat capacity (Cp) was determined by the experimental measurement with a thermal analyzer (NETZSCH, STA 449F3, Selb, Germany) using sapphire as reference sample. The density d was measured at room temperature by applying the Archimedes method with ethanol as the immersion liquid.

3. Results and Discussion

The PXRD results of Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) samples are shown in Figure 1a. Almost all diffractions are well consistent with cubic perovskite structure (Figure 1d.) despite the fact that a small amount of impurity phase identified as Sc2O3 and Ti1.87O3 can be tracked. Figure 1b displays the diffractions around 33°, which are basically unchanged for single-doped samples. This can be understood from the low solid solubility of Sc in SrTiO3 due to the large difference in ionic radius [32]. As a matter of fact, Sc is usually treated as dopant for Ti in SrTiO3 to tune the physical properties [33]. High-angle shift is observed for La/Sc co-doped samples, demonstrating that La can successfully substitute Sr since the ionic size of La3+ (1.36 Å, 12-coordination) is slightly smaller than that of Sr2+ (1.44 Å, 12-coordination) [34]. The dependences of lattice parameters on the doping contents verify the conclusion presented in Figure 1c. The lattice parameters are constant with single Sc doping, and get smaller when La substitutes Sr in SrTiO3.
Figure 2a presents the SEM image of the surface for the co-doped sample Sr0.9Sc0.04La0.06TiO3. The element distributions of Sr0.9Sc0.04La0.06TiO3 are basically homogeneous (Figure 2b–f), suggesting La and partial Sc can be dissolved into the matrix. However, a small amount of Sc enrichment area can also be observed, indicating that the low solution limit of Sc, which is well in agreement with the XRD results. Table 1 shows the real compositions of Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) detected by EDS, close to the nominal compositions designed in this work.
The average grain sizes were measured from SEM images by image analysis using the Image-Pro program [31], as can be seen in Table 2. With the increase of doping amount, the average grain sizes are almost the same, ranging from 1.35 μm to 1.87 μm. It is reported that the mean free paths of electron and phonon in SrTiO3 are about 1 nm and 2 nm, respectively [35,36]. Thus, the large grains would not result in the essential difference in electron and phonon transport for samples Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06). The measured densities are very close to the ideal value of single crystal SrTiO3, suggesting dense feature for bulk samples.
Figure 3 shows the temperature-dependence of the electrical transport properties for Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) ceramics. As can be seen in Figure 3a, the pristine SrTiO3 and single Sc-doped samples exhibit metal-like conductive behaviors. Meanwhile, the electrical conductivity of Sr0.96Sc0.04TiO3 increases slightly in comparison with undoped SrTiO3, confirming finite substitution of Sc3+ for Sr2+, which introduces extra electrons and increases the electron concentration (Figure 3b). La-doping enhances the electrical conductivity of SrTiO3 significantly, and the conduction behaviors transform from metal to semiconductor before 468 K. The electrical conductivity increases at low temperatures, while decreases at high temperatures with increasing temperature, which are consistent with the results reported in the literatures [23,37]. However, the values of electrical conductivity are lower than the data reported in the literatures, which is probably ascribed from the Sr vacancy since Sc hardly substitutes Sr.
At room temperature, Sr0.96Sc0.04TiO3 sample has the largest electrical conductivity because of its high mobility (Figure 3b). For La/Sc co-doped samples, the carrier mobilities μH descend, resulting in the low electrical conductivity. At high temperature, the electrical conductivity of all samples monotonically increase with the rise of the doping concentration, which is derived from the donor doping effect. Figure 3c shows the Seebeck coefficients of Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) depending on the temperature. The negative S in all measurements indicates that Sr1-x-yScxLayTiO3 ceramics are n-type semiconductors. The absolute values of Seebeck coefficient for Sr1-x-yScxLayTiO3 increases as the temperature rises, and doping suppresses the Seebeck coefficient since it is inversely proportional to the carrier concentration as follows: S~[π/(3n)]2/3 m × T (m is the electron effective mass) [38]. The power factors PF are calculated from σS2 and presented in Figure 3d. The undoped and single-doped Sc samples have the largest power factor at room temperature (9.8 μW·cm−1·K2 at 320 K). For the co-doped samples, the peak values shift to the high temperature and PF = 9.41 μW·cm−1·K−2 has been achieved for Sr0.88Sc0.06La0.06TiO3.
The thermal expansion coefficients (αV) ranging from 500–800 K for these samples are shown in Table 2. With the increase of the doping amount at the Sr site, this parameter increases monotonically, which leads to the reduction in the lattice thermal conductivity (Figure 4b). The lattice thermal conductivity is inversely proportional to the absolute value of the average volumetric thermal expansion coefficients, expressed by Grüneisen’s law and the Slack phonon model as follows:
α V = β T V γ C
κ l a t = A M ¯ θ d 3 δ γ 2 n 2 / 3 T
where A is a constant, M ¯ is the average atomic mass, n is the number of atoms per unit cell, δ3 is the volume per atom, T is the absolute temperature, γ is the average Grüneisen parameter for the acoustic branches, and θd is the Debye temperature, βT is the isothermal compressibility, C is the heat capacity, and V is the molar volume. From Equations (1) [39] and (2) [40], αV is directly proportional to the Grüneisen parameter γ under certain conditions. On the other hand, γ is usually the parameter that characterizes the strength of anharmonic inversely proportional to κlat.
Lattice thermal conductivity κlat is calculated by subtracting κe from κtot, and the electronic contribution κe is estimated according to the Wiedemann-Franz law (κe = LσT) with the Lorentz umber (L) estimated from the single parabolic band (SPB) model [38]. The contribution of κe to the total thermal conductivity κtot is too low, which can be negligible in this work. Thus, the varieties of κtot and κlat on the temperature are basically identified (Figure 4a,b). The pristine SrTiO3 exhibits a high thermal conductivity at room temperature, reaching 9.38 W·m−1·K−1. The relatively high thermal conductivity restricts figure of merit ZT, and additional scattering mechanism should be introduced to lower the lattice thermal conductivity. Figure 4b displays the lattice thermal conductivity decreases with increasing the temperature which basically conforms to the relationship of T−1, indicating that phonon scattering is dominant at high temperatures.
The lattice thermal conductivity drops sharply with La and Sc doping in SrTiO3. Such a significant reduction mainly ascribes to the point defect scattering due to the mass fluctuation and strain fluctuations (described by disorder scattering factors Γ M and Γ S , respectively) between La/Sc and Sr, and thereby giving rise to the reduction of the lattice thermal conductivity at room temperature. The τ P D 1 (phonon-point-defect scattering) and disorder scattering factors can be obtained through [41]:
τ P D 1 = τ S 1 + τ M 1 = V ω 4 4 π v s 3 ( Γ S + Γ M )
Γ = Γ S + Γ M
Γ S = i = 1 n c i ( M i ¯ M ̿ ) 2 f i 1 f i 2 ϵ i ( r i 1 r i 2 r ¯ i ) 2 i = 1 n c i
Γ M = i = 1 n c i ( M i ¯ M ̿ ) 2 f i 1 f i 2 ( M i 1 M i 2 M ¯ i ) 2 i = 1 n c i
where ω is the phonon frequency, and vs is the sound speed. The disorder Γ is related to both mass fluctuation scattering Γ M and strain field Γ S . ci is the relative degeneracy of the site, fi is the fractional occupation, M i ¯ and r ¯ i are the average mass and radii of element, respectively, and M ¯ is the average mass [42]. The difference in the ionic radius between Sr and rare earth elements leads to high distortion into the lattice and thus reduces the lattice thermal conductivity. In addition, Sc2O3 as a composite phase, plays a role in low-frequency phonon scattering, favoring the reduction of lattice thermal conductivity. The total thermal conductivity of the sample Sr0.88Sc0.06La0.06TiO3 significantly reduced to 6.97 W·m−1·K−1 at room temperature, which is 25.6% lower than the pristine sample.
Figure 5 plots the ZT values as a function of temperature from 323 K to 773 K. The ZT values increase with the rise of the temperature, and the largest ZT = 0.143 at 773 K has been achieved for composition Sr0.88Sc0.06La0.06TiO3. In comparison with the value reported in literature with single La doping at the same temperature (ZT~0.2) [43], the thermoelectric performance is uncompetitive in this work. However, the composite engineering turns out to be an effective route to optimize the electrical and thermal transport properties of thermoelectric materials.

4. Conclusions

The polycrystalline bulk Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) samples with promising thermal stability were synthesized by solid-state reaction method and spark plasma sintering. The introduction of trivalent La increases the electron concentration and improves the electrical conductivity. Sc hardly substitutes Sr as a result of the low solid solution, and it distributes in the sample in the form of Sc2O3, which favors for the reduction of lattice thermal conductivity due to the low-frequency phonon scattering. Therefore, the total thermal conductivity decreases from 9.38 W·m−1·K−1 for pristine SrTiO3 to 6.97 W·m−1·K−1 for Sr0.88Sc0.06La0.06TiO3 at room temperature. The enhanced ZT value of 0.143 was achieved at 773 K in nominal sample Sr0.88Sc0.06La0.06TiO3.

Author Contributions

Data curation, F.Y. and T.W.; formal analysis, J.C. and G.R.; investigation, K.G., F.Y. and H.L.; methodology, J.C. and J.Z. (Jiye Zhang).; project administration, J.L.; supervision, K.G., J.C. and J.Z. (Jingtai Zhao); validation, T.W.; writing—original draft, K.G. and F.Y.; writing—review and editing, K.G., J.C., G.R. and J.Z. (Jingtai Zhao). All authors have read and agreed to the published version of the manuscript.

Funding

We like to thank the financial support from the National Key Research and Development Program of China (2018YFA0702100), National Natural Science Foundation of China (21771123), the Program of Introducing Talents of Discipline to Universities (D16002). G.R. is grateful to the foundation for Guangxi Bagui scholars.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data is available within the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tritt, T.M. Holey and unholey semiconductors. Science 1999, 283, 804–805. [Google Scholar] [CrossRef]
  2. Disalvo, F.J. Thermoelectric cooling and power generation. Science 1999, 285, 703–706. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, X.; Zhao, L.D. Thermoelectric materials: Energy conversion between heat and electricity. J. Mater. 2015, 1, 92–105. [Google Scholar] [CrossRef] [Green Version]
  4. Bourgès, C.; Srinivasan, B.; Fontaine, B.; Sauerschnig, P.; Minard, A.; Halet, J.-F.; Miyazaki, Y.; Berthebaud, D.; Mori, T. Tailoring the thermoelectric and structural properties of Cu–Sn based thiospinel compounds [CuM1+xSn1−xS4 (M = Ti, V, Cr, Co)]. J. Mater. Chem. C 2020, 8, 16368–16383. [Google Scholar] [CrossRef]
  5. Muchtar, A.R.; Srinivasan, B.; Le Tonquesse, S.; Singh, S.; Soelami, N.; Yuliarto, B.; Berthebaud, D.; Mori, T. Physical insights on the lattice softening driven mid-temperature range thermoelectrics of Ti/Zr-inserted SnTe—an outlook beyond the horizons of conventional phonon scattering and excavation of Heikes’ Equation for estimating carrier properties. Adv. Energy Mater. 2021, 11, 2101122. [Google Scholar] [CrossRef]
  6. Li, J.F.; Liu, W.S.; Zhao, L.D.; Zhou, M. High-performance nanostructured thermoelectric materials. NPG Asia Mater. 2010, 2, 152–158. [Google Scholar] [CrossRef]
  7. Nan, Y.C.; Minnich, A.J.; Chen, G.; Ren, Z.F. Enhancement of thermoelectric figure-of-merit by a bulk nanostructuring approach. Adv. Funct. Mater. 2010, 20, 357–376. [Google Scholar]
  8. Pei, Y.; Shi, X.; LaLonde, A.; Wang, H.; Chen, L.; Snyder, G.J. Convergence of electronic bands for high performance bulk thermoelectrics. Nature 2011, 473, 66–69. [Google Scholar] [CrossRef]
  9. Liu, X.; Zhu, T.; Wang, H.; Hu, L.; Xie, H.; Jiang, G.; Snyder, G.J.; Zhao, X. Low electron scattering potentials in high performance Mg2Si0.45Sn0.55 based thermoelectric solid solutions with band convergence. Adv. Energy Mater. 2013, 3, 1238–1244. [Google Scholar] [CrossRef]
  10. Tan, G.; Shi, F.; Hao, S.; Chi, H.; Zhao, L.-D.; Uher, C.; Wolverton, C.; Dravid, V.P.; Kanatzidis, M.G. Codoping in SnTe: Enhancement of thermoelectric performance through synergy of resonance levels and band convergence. J. Am. Chem. Soc. 2015, 137, 5100–5112. [Google Scholar] [CrossRef]
  11. Kim, S.; Lee, K.H.; Mun, H.A.; Kim, H.S.; Hwang, S.W.; Roh, J.W.; Yang, D.J.; Shin, W.H.; Li, X.S.; Lee, Y.H.; et al. Dense dislocation arrays embedded in grain boundaries for high-performance bulk thermoelectrics. Science 2015, 348, 109–114. [Google Scholar] [CrossRef] [Green Version]
  12. Liu, Z.; Meng, X.; Qin, D.; Cui, B.; Wu, H.; Zhang, Y.; Pennycook, S.J.; Cai, W.; Sui, J. New insights into the role of dislocation engineering in N-type filled skutterudite CoSb3. J. Mater. Chem. C 2019, 7, 13622–13631. [Google Scholar] [CrossRef]
  13. Hu, L.; Zhu, T.; Liu, X.; Zhao, X. Point defect engineering of high-performance bismuth-telluride-based thermoelectric materials. Adv. Funct. Mater. 2014, 24, 5211–5218. [Google Scholar] [CrossRef]
  14. Zhao, K.; Zhu, C.; Qiu, P.; Blichfeld, A.B.; Eikeland, E.; Ren, D.; Iversen, B.B.; Xu, F.; Shi, X.; Chen, L. High thermoelectric performance and low thermal conductivity in Cu2−yS1/3Se1/3Te1/3 liquid-like materials with nanoscale mosaic structures. Nano Energy 2017, 42, 43–50. [Google Scholar] [CrossRef] [Green Version]
  15. Lee, M.; Viciu, L.; Li, L.; Wang, Y.; Foo, M.L.; Watauchi, S.; Pascal JR, R.A.; Cava, R.J.; Ong, N.P. Large enhancement of the thermopower in NaxCoO2 at high Na doping. Nat. Mater. 2006, 5, 537–540. [Google Scholar] [CrossRef] [Green Version]
  16. Terasaki, I.; Sasago, Y.; Uchinokura, K. Large Thermoelectric power in NaCo2O4 crystals. Phys. Rev. B 1997, 56, 12685–12687. [Google Scholar] [CrossRef]
  17. Masset, A.C.; Michel, C.; Maignan, A.; Hervieu, M.; Toulemonde, O.; Studer, F.; Raveau, B.; Hejtmanek, J. Misfit-layered cobaltite with an anisotropic giant magnetoresistance: Ca3Co4O9. Phys. Rev. B 2000, 62, 166–175. [Google Scholar] [CrossRef]
  18. Liu, Y.; Lin, Y.; Shi, Z.; Nan, C.-W. Preparation of Ca3Co4O9 and improvement of its thermoelectric properties by spark plasma sintering. J. Am. Ceram. Soc. 2005, 88, 1337–1340. [Google Scholar] [CrossRef]
  19. Ohta, H.; Sugiura, K.; Koumoto, K. Recent progress in oxide thermoelectric materials: P-type Ca3Co4O9 and n-type SrTiO3. Inorg. Chem. 2008, 47, 8429–8436. [Google Scholar] [CrossRef] [PubMed]
  20. Boston, R.; Schmidt, W.L.; Lewin, G.D.; lyasara, A.C.; Lu, Z.; Zhang, H.; Sinclair, D.C.; Reaney, I.M. Protocols for the fabrication, characterization, and optimization of n-type thermoelectric ceramic oxides. Chem. Mater. 2016, 29, 265–280. [Google Scholar] [CrossRef] [Green Version]
  21. Fergus, J.W. Oxide materials for high temperature thermoelectric energy conversion. J. Eur. Ceram. Soc. 2012, 32, 525–540. [Google Scholar] [CrossRef]
  22. Ito, M.; Furumoto, D. Microstructure and thermoelectric properties of NaxCo2O4/Ag composite synthesized by the polymerized complex method. J. Alloys Compd. 2008, 450, 517–520. [Google Scholar] [CrossRef]
  23. Muta, H.; Kurosaki, K.; Yamanaka, S. Thermoelectric properties of rare earth doped SrTiO3. J. Alloys Compd. 2003, 350, 292–295. [Google Scholar] [CrossRef]
  24. Ravichandran, J.; Siemons, W.; Oh, D.-W.; Kardel, J.T.; Chari, A.; Heijmerikx, H.; Scullin, M.L.; Majumdar, A.; Ramesh, R.; Cahill, D.G. High temperature thermoelectric response of double-doped SrTiO3 epitaxial films. Phys. Rev. B 2010, 82, 165126. [Google Scholar] [CrossRef] [Green Version]
  25. Shirai, K.; Yamanaka, K. Mechanism behind the high thermoelectric power factor of SrTiO3 by calculating the transport coefficients. J. Appl. Phys. 2013, 113, 188–576. [Google Scholar] [CrossRef]
  26. Okuda, T.; Nakanishi, K.; Miyasaka, S.; Tokura, Y. Large thermoelectric response of metallic perovskites: Sr1-xLaxTiO3 (0 < x < 0.1). Phys. Rev. B 2001, 63, 113104. [Google Scholar]
  27. Daniels, L.M.; Ling, S.; Savvin, S.N.; Pitcher, M.J.; Dyer, M.S.; Claridge, J.B.; Slater, B.; Corà, F.; Alaria, J.; Rosseinsky, M.J. A and B site doping of a phonon-glass perovskite oxide thermoelectric. J. Mater. Chem. A 2018, 6, 15640–15652. [Google Scholar] [CrossRef] [Green Version]
  28. Liu, J.; Wang, C.L.; Li, Y.; Su, W.B. Influence of rare earth doping on thermoelectric properties of SrTiO3 ceramics. J. Appl. Phys. 2013, 114, 223174. [Google Scholar] [CrossRef]
  29. Daniels, L.M.; Savvin, S.N.; Pitcher, M.J.; Dyer, M.S.; Claridge, J.B.; Ling, S.; Slater, B.; Cora, F.; Alaria, J.; Rosseinsky, M. Phonon-glass electron-crystal behaviour by A site disorder in n-type thermoelectric oxides. Energy Environ. Sci. 2017, 10, 1917–1922. [Google Scholar] [CrossRef] [Green Version]
  30. Akselrud, L.; Grin, Y. WinCSD: Software package for crystallographic calculations (Version 4). J. Appl. Crystallogr. 2014, 47, 803–805. [Google Scholar] [CrossRef]
  31. Hu, T.-Y.; Yao, M.-Y.; Hu, D.-L.; Gu, H.; Wang, Y.-J. Effect of mechanical alloying on sinterability and phase evolution in pressure-less sintered TiB2‒TiC ceramics. J. Mater. 2019, 5, 670–678. [Google Scholar] [CrossRef]
  32. Tkach, A.; Vilarinho, P.M. Scandium doped strontium titanate ceramics: Structure, microstructure, and dielectric properties. Bol. Soc. Esp. Artic. Cerámica Y Vidr. 2008, 47, 238–241. [Google Scholar] [CrossRef]
  33. Li, X.; Zhao, H.; Luo, D.; Huang, K. Electrical conductivity and stability of A-site deficient (La, Sc) co-doped SrTiO3 mixed ionic-electronic conductor. Mater. Lett. 2011, 65, 2624–2627. [Google Scholar] [CrossRef]
  34. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chaleogenides. Acta Crystallogr. A 1976, A32, 751–767. [Google Scholar] [CrossRef]
  35. Collignon, C.; Bourges, P.; Fauqué, B.; Behnia, K. Heavy nondegenerate electrons in doped strontium titanate. Phys. Rev. X 2020, 10, 031025. [Google Scholar] [CrossRef]
  36. Wang, Y.; Fujinami, K.; Zhang, R.; Wan, C.; Wang, N.; Ba, Y.; Koumoto, K. Interfacial thermal resistance and thermal conductivity in nanograined SrTiO3. Appl. Phys. Express 2010, 3, 031101. [Google Scholar] [CrossRef]
  37. Ahmed, A.; Hossain, M.S.A.; Islam, S.M.K.N.; Yun, F.; Yang, G.; Hossain, R.; Khan, A.; Na, J.; Eguchi, M.; Yamauchi, Y.; et al. Significant improvement in electrical conductivity and figure of merit of nanoarchitectured porous SrTiO3 by Ladoping optimization. ACS Appl. Mater. Interfaces 2020, 12, 28057–28064. [Google Scholar] [CrossRef]
  38. Snyder, G.J.; Toberer, E.S. Complex thermoelectric materials. Nat. Mater. 2008, 7, 105–114. [Google Scholar] [CrossRef]
  39. Reid, K.M.; Yu, X.; Leitner, D.M. Change in vibrational entropy with change in protein volume estimated with mode Gruneisen parameters. J. Chem. Phys. 2021, 154, 055102. [Google Scholar] [CrossRef]
  40. Bosoni, E.; Sosso, G.C.; Bernasconi, M. Grüneisen parameters and thermal conductivity in the phase change compound GeTe. J. Comput. Electron. 2017, 16, 997–1002. [Google Scholar] [CrossRef]
  41. Ren, G.-K.; Lan, J.-L.; Ventura, K.J.; Tan, X.; Lin, Y.-H.; Nan, C.-W. Contribution of point defects and nano-grains to thermal transport behaviours of oxide-based thermoelectrics. npj Comput. Mater. 2016, 2016, 16023. [Google Scholar] [CrossRef] [Green Version]
  42. Yang, J.; Meisner, G.P.; Chen, L. Strain field fluctuation effects on lattice thermal conductivity of ZrNiSn-based thermoelectric compounds. Appl. Phys. Lett. 2004, 85, 1140–1142. [Google Scholar] [CrossRef]
  43. Lu, Z.; Zhang, H.; Lei, W.; Sinclair, D.C.; Reaney, I.M. High-figure-of-merit thermoelectric La-doped A-site-deficient SrTiO3 ceramics. Chem. Mater. 2016, 28, 925–935. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) PXRD patterns of Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) ceramics, (b) the enlarged peak in the vicinity of 33°, (c) lattice parameter and volume of Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) ceramics as function of the doping contents, (d) crystal structure diagram of SrTiO3.
Figure 1. (a) PXRD patterns of Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) ceramics, (b) the enlarged peak in the vicinity of 33°, (c) lattice parameter and volume of Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) ceramics as function of the doping contents, (d) crystal structure diagram of SrTiO3.
Materials 14 06279 g001
Figure 2. (a) SEM images of typical sample Sr0.9Sc0.04La0.06TiO3; (bf) the corresponding elemental distribution for Ti, O, Sr, La, and Sc.
Figure 2. (a) SEM images of typical sample Sr0.9Sc0.04La0.06TiO3; (bf) the corresponding elemental distribution for Ti, O, Sr, La, and Sc.
Materials 14 06279 g002
Figure 3. (a) Temperature-dependent electrical conductivity (σ). (b) Hall carrier concentration (nH) and mobility (μH) as a function of the Sc/La contents at room temperature. (c) Temperature-dependent Seebeck coefficient (S), and (d) power factor (PF).
Figure 3. (a) Temperature-dependent electrical conductivity (σ). (b) Hall carrier concentration (nH) and mobility (μH) as a function of the Sc/La contents at room temperature. (c) Temperature-dependent Seebeck coefficient (S), and (d) power factor (PF).
Materials 14 06279 g003
Figure 4. The temperature dependences of (a) the total thermal conductivity (κtot), and (b) the lattice thermal conductivity (κlat). The dotted line indicates that the temperature-dependent lattice thermal conductivity satisfies T−1 relation when phonon scattering dominates.
Figure 4. The temperature dependences of (a) the total thermal conductivity (κtot), and (b) the lattice thermal conductivity (κlat). The dotted line indicates that the temperature-dependent lattice thermal conductivity satisfies T−1 relation when phonon scattering dominates.
Materials 14 06279 g004
Figure 5. Temperature-dependence of the figure-of-merit ZT. for samples Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06).
Figure 5. Temperature-dependence of the figure-of-merit ZT. for samples Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06).
Materials 14 06279 g005
Table 1. The real compositions of Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) detected by EDS.
Table 1. The real compositions of Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06) detected by EDS.
Nominal Comp.Ti (%)O (%)Sr (%)La (%)Sc (%)Real Comp.
SrTiO320.01 ± 0.2760.04 ± 0.1219.95 ± 0.25--SrTi0.997O3
Sr0.96Sc0.04TiO320.03 ± 0.1660.05 ± 0.1319.20 ± 0.15-0.72 ± 0.28Sr0.959Sc0.036Ti1.001O3
Sr0.9Sc0.04La0.06TiO319.88 ± 0.2160.18 ± 0.3117.98 ± 0.191.14 ± 0.090.82 ± 0.32Sr0.897Sc0.040La0.058Ti0.99O3
Sr0.88Sc0.06La0.06TiO320.02 ± 0.1859.92 ± 0.2117.69 ± 0.201.21 ± 0.061.16 ± 0.21Sr0.885Sc0.058La0.061Ti1.002O3
Table 2. Average grain sizes, densities, and thermal expansion coefficients of samples Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06).
Table 2. Average grain sizes, densities, and thermal expansion coefficients of samples Sr1-x-yScxLayTiO3 (x = 0, 0.04, 0.06; y = 0, 0.06).
Sr1-x-yScxLayTiO3Average Grain Size (μm)Real Density (g cm−3)Thermal Expansion Coefficients (10−5 K−1, 500–800 K)
SrTiO31.77 ± 0.075.1051.00 ± 0.01
Sr0.96Sc0.04TiO31.87 ± 0.014.9971.01 ± 0.01
Sr0.9Sc0.04La0.06TiO31.35 ± 0.075.1241.02 ± 0.01
Sr0.88Sc0.06La0.06TiO31.35 ± 0.055.1341.08 ± 0.01
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Guo, K.; Yang, F.; Weng, T.; Chen, J.; Zhang, J.; Luo, J.; Li, H.; Rao, G.; Zhao, J. The Electrical and Thermal Transport Properties of La-Doped SrTiO3 with Sc2O3 Composite. Materials 2021, 14, 6279. https://doi.org/10.3390/ma14216279

AMA Style

Guo K, Yang F, Weng T, Chen J, Zhang J, Luo J, Li H, Rao G, Zhao J. The Electrical and Thermal Transport Properties of La-Doped SrTiO3 with Sc2O3 Composite. Materials. 2021; 14(21):6279. https://doi.org/10.3390/ma14216279

Chicago/Turabian Style

Guo, Kai, Fan Yang, Tianyao Weng, Jianguo Chen, Jiye Zhang, Jun Luo, Han Li, Guanghui Rao, and Jingtai Zhao. 2021. "The Electrical and Thermal Transport Properties of La-Doped SrTiO3 with Sc2O3 Composite" Materials 14, no. 21: 6279. https://doi.org/10.3390/ma14216279

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop