Next Article in Journal
Modeling and Analysis of the Drying Process of Lithium-Ion Battery Electrodes Based on Non-Steady-State Drying Kinetics
Next Article in Special Issue
Assessment of Microsilica as a Raw Material for Obtaining Mullite–Silica Refractories
Previous Article in Journal
A New Fracturing Method to Improve Stimulation Effect of Marl Tight Oil Reservoir in Sichuan Basin
Previous Article in Special Issue
Oxide Strontium-Barium Perovskites Ceramics: Examinations of Structural Phase Transitions and Potential Application as Oxygen Carriers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of MgO and Rare-Earth Oxides (Y2O3, Yb2O3, Dy2O3) on the Structural Characteristics and Electrical Properties of BaTiO3

Department of Advanced Materials Engineering, Graduate School, Kyonggi University, Suwon 16227, Republic of Korea
*
Author to whom correspondence should be addressed.
Processes 2023, 11(11), 3235; https://doi.org/10.3390/pr11113235
Submission received: 28 October 2023 / Revised: 11 November 2023 / Accepted: 14 November 2023 / Published: 16 November 2023
(This article belongs to the Special Issue Advances in Ceramic Processing and Application of Ceramic Materials)

Abstract

:
This study investigated the impact of MgO and rare-earth oxides (Y2O3, Yb2O3, and Dy2O3) on the structural characteristics and electrical properties of BaTiO3. Specimens sintered at 1350 °C for durations ranging from 1 to 5 h in air exhibited a single phase of BaTiO3 with a tetragonal structure. This was observed for pure BaTiO3 and specimens co-doped with MgO-Y2O3 and/or MgO-Dy2O3. However, a pseudo-cubic structure of BaTiO3 was detected for specimens doped with MgO or co-doped with MgO-Yb2O3. The unit-cell volume of the sintered specimens was found to be dependent on the type of substitution ion for the A/B site of BaTiO3 (ABO3). The dielectric constant (εr) of the sintered specimens decreased with the substitution of MgO and rare-earth oxides due to a decrease in tetragonality (c/a). The electrical resistivities of the sintered specimens were influenced not only by their microstructural characteristics but also by the secondary phases of the sintered specimens. The BaTiO3 specimens co-doped with MgO-Yb2O3 and/or doped with MgO met the EIA X7R and X8R specifications (−55 to 125~150 °C, ΔC/C = ±15% or less), respectively.

Graphical Abstract

1. Introduction

The ongoing advancement of electronic devices has led to an escalating demand for the miniaturisation and enhancement of electronic components. Multilayer ceramic capacitors (MLCCs), due to their high volumetric efficiency and stability, have become a crucial passive component. The increasing application of MLCCs has necessitated a fundamental investigation into their high dielectric constant (εr) and thermal stability. Specifically, the industry has formulated the X7R and X8R standards (ΔC/C ~±15% and −50~125 °C, 150 °C, respectively), as thermal stability limits the application of MLCCs [1,2,3,4].
To achieve a high dielectric constant, it is necessary to thin the dielectric layer of MLCCs by decreasing the particle size of BaTiO3 (barium titanate, BT), which is one of the primary raw materials due to its high dielectric constant (εr) [5]. Therefore, many studies have been conducted to identify methods that can control the morphology of BT particles, such as the sol-gel method, hydrothermal method, and solvothermal method [6,7,8,9]. For example, the sol-gel method provides a high purity of products and a relatively low synthesis temperature [7,8]. However, the calcination step at 700 °C or over induces coarse and chemically bonded aggregates. Zhou et al. [6] reported preparing BT nanoparticles using a hydrothermal method from two different starting materials. BT nanoparticles prepared from titanium hydroxide and barium hydroxide showed a spherical grain size of 65 nm with a narrow size distribution, while those prepared from barium hydroxide and titanium dioxide showed coarse particles arising from lattice defects, which were compensated by the presence of cation vacancies [6]. A solvothermal method for synthesizing BT nanoparticles in large batches at room temperature was reported by Wei et al. [9]. By adjusting the ratio of the solvent mixture and the reaction time, the particle size of the BT can be modulated within the range of 7 nm to 16 nm, showing a narrow size distribution. Based on these considerations, the starting materials and the solvent are important factors when synthesizing BT nanoparticles.
Furthermore, a narrow nanoparticle size distribution of dopant oxides such as NiO, MgO, and SiO2 is important for the performance of MLCCs. Most common methods for obtaining nanosized doping particles for BT, such as MgO, SiO2 and Al2O3, are modified from the Pechini method, which is based on the abilities of organic compounds. Organic compounds such as acids, carbohydrates, and alcohols are used as chelating agents for the metal ions. Due to the formation of a polymer network with a homogeneous structure and a uniform distribution of ions during polyesterification, a final product with a good dispersion can be obtained. The formation of oxide nanoparticles occurs in the reaction space, which is concentrated in the walls of the cellular structure during swelling of the polymer network in the intermediate form stage and is determined by the processing temperature. In addition, the average particle size of nano oxides can be determined by the wall thickness in the polymer network. Kozerozhets et al. obtained MgO, SiO2, and Al2O3 nanoparticle oxides within an average particle size of 20 nm to 30 nm by modifying the Pechini method [10].
Another way to improve the dielectric constant of MLCCs is to substitute proper ions in the BT lattice. There have been many studies to investigate the effect of various additives on the dielectric properties of BT [11,12,13,14,15,16,17], since the dielectric properties of BT could be changed according to the valence state of the dopant and its substitution site in a BT lattice. Especially, many rare-earth oxides are used as dopants due to their amphoteric characteristics to improve the dielectric properties of BT, and previous studies have reported that the ionic radius of rare-earth oxides is a crucial factor when they are substituted into a BT lattice. Smaller ions occupy the Ti site, and larger ones occupy the Ba site. Intermediate ions such as Dy, Y, and Ho could occupy both the Ba and Ti sites depending on the Ba/Ti ratio of BT and the sintering temperature [13,14,15,16,17]. Rare-earth oxides with an intermediate ionic size are known to act as a donor when incorporated into the Ba site and as an acceptor in the Ti site as follows:
Ba   site:   Re 2 O 3 2 R B a · + 3 O O + V B a
Ti   site:   Re 2 O 3 2 R Ti + 2 O O + V O · ·
Donors and acceptors are useful in capturing oxygen vacancies bound to free electrons, reducing carrier concentrations, and thereby improving the anti-reduction performance of BT ceramics [18].
Thermal stability is another important factor for the applications of MLCCs. However, the nonlinear dielectric behaviour of pure BT with temperature, caused by phase transitions (rhombohedral to orthorhombic at −90 °C, orthorhombic to tetragonal at 0 °C, tetragonal to cubic at 125 °C), results in an unstable temperature dependence of the dielectric constant. Therefore, the composition of BT is controlled by doping elements such as Mg2+, Ca2+, and Zr4+ ions to suppress or shift the Curie temperature [4,5,11,12]. Rare-earth oxides also have an influence on the thermal stability of BT, forming core-shell structures in the BT grains. In this structure, the shell comprises a non-ferroelectric pseudo-cubic structure created by the partial substitution of dopants, while the core contains ferroelectric BT with a tetragonal structure. Consequently, the dielectric characteristics and thermal stability of BT can be regulated by varying the amount and type of rare-earth oxides.
Numerous studies have been conducted on BaTiO3 doped with MgO and rare-earth oxides. Kishi et al. reported the formation mechanism of a core-shell structure in the BaTiO3-MgO-Ho2O3 system. At a low temperature, the shell phase is formed by a reaction between MgO and BT, and then Ho2O3 reacts with the shell at a high temperature. The diffusion of Ho2O3 is hindered by MgO, inhibiting grain growth [19]. Pu et al. explored the effect of a grain growth inhibitor of Dy ions for the formula Ba1-xDy2x/3TiO3 and obtained an increased dielectric constant (εr) for x = 0.75 [1]. Kim et al. investigated the role of MgO and Y2O3 in the BT-MgCO3-Y2O3-BaCO3-SiO2 system. They found that Y2O3 plays a dominant role in the formation of the shell by easily dissolving into BT, while MgO tends to stay at the grain boundary [13]. Additionally, Y2O3 flattened the dielectric constant temperature (the εr-T curve) for BT-Y2O3-MgO without shifting the Curie peak of BT [20]. Wang et al. reported the effect of the sintering atmosphere on the microstructure and dielectric properties of specimens co-doped with Yb/Mg [21].
However, most studies have focused on the effect of MgO and rare-earth oxides on the microstructural characteristics and the temperature stability of the dielectric constant. In light of these considerations, this study investigated the effects of MgO and rare-earth oxides (Y2O3, Yb2O3, and Dy2O3) on the structural characteristics and electrical properties of BT. The aim was to propose relationships between the structural characteristics and electrical properties of BT.

2. Materials and Methods

BaTiO3 (170 nm, Fuji Titanium Ind. Co., Osaka, Japan), MgO (Fine Material, Seongnam, Republic of Korea), Dy2O3 (Solvay, Brussels, Belgium), Y2O3 (Solvay, Brussels, Belgium), and Yb2O3 (Solvay, Brussels, Belgium) were utilised as starting materials. These were mixed in stoichiometric ratios as per Table 1. The mixed powders were milled with yttrium-stabilised zirconia balls for 24 h in ethyl alcohol and then dried. The dried powders were isostatically pressed into pellets under a pressure of 500 kg/cm2 using 15 mm diameter disks. The pressed disks were sintered at 1350 °C for durations ranging from 1 to 5 h in an air atmosphere.
The crystalline phases of the sintered specimens were analysed using X-ray diffraction (XRD, D/max-2500V/PC, Rigaku, Tokyo) over the range of 2θ = 10–80°. Rietveld refinement (RIR) measurements were performed on the XRD data using Fullprof software (version 7.95) to obtain structural characteristics such as the lattice parameters, tetragonality (c/a), and unit-cell volume of the sintered specimens. Silver electrodes were coated on both surfaces of the sintered specimens and fired at 550 °C for 30 min. The microstructures of the sintered specimens were observed using a scanning electron microscope (SEM, Su-70, Hitachi, Japan). Apparent densities and relative densities of the sintered specimens were obtained by the Archimedes method and Rietveld refinement, respectively. The dielectric constants of the specimens were determined from their measured capacitances (pF) using a Capacitance Meter (E4981A, KEYSIGHT, Santa Rosa, CA, USA) at 1 KHz. The resistivities (ρ) of the specimens were calculated using a High Resistance Meter (4339B, KEYSIGHT, Santa Rosa, CA, USA) at 500 V for 60 s. The temperature coefficient of capacitance of the specimens was measured using a temperature coefficient of capacitance (TCC) measurement system (TM-100, NANOIONICS KOREA, Gangneung-si, Gangwon-do, Republic of Korea) from −50 °C to 150 °C at 1 KHz.

3. Results and Discussion

3.1. Physical Properties

According to the report of Gong et al. [22], the relative density is a critical factor for dielectric materials. It could affect not only the formation of a core shell for BT substituted with rare-earth oxides but also the thermal stability of the BT. Therefore, the sintering condition of BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (with Re representing Y, Yb, or Dy) was optimised to achieve a higher relative density above 94% of the theoretical value, as listed in Table 2. This was performed to neglect the effect of pores on the dielectric and electrical properties of the sintered specimens, which will be discussed in Section 3.2.
The XRD patterns of the BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (with Re representing Y, Yb, or Dy) sintered at 1350 °C for 1–5 h in air are presented in Figure 1a. All of the sintered specimens, except for those co-doped with MgO-Yb2O3, exhibited a single phase of BaTiO3 with a perovskite structure. However, the secondary phase of pyrochlore Yb2Ti2O7 (ICCD No. 98-017-3750) was detected for the specimens co-doped with MgO-Yb2O3, while a single phase of BT was obtained for the MgO-doped and MgO-Y2O3/MgO-Dy2O3 co-doped specimens. The formation of the Yb2Ti2O7 secondary phase is due to the smaller solubility of Yb2O3 than that of Y2O3 and Dy2O3. Rare-earth oxides are forced to substitute at the Ba site of BT for the BT-MgO-Re2O3 (with Re representing Y, Yb, or Dy) system due to the substitution of the Mg2+ ion for the Ti4+ ion. However, Yb2O3 acts as an acceptor due to its smaller ionic radius than Y3+ and Dy3+ ions [23]. Thus, extra Yb2O3 may react with TiO2, forming a Yb2Ti2O7 phase.
Figure 1b displays the enlarged XRD patterns of the sintered specimens over the range of 2θ = 42–48°. The crystal structure of BaTiO3 ceramics can be determined by the splitting of (002) and (200) peaks over the range of 2θ = 44–46° [24]. The sintered pure BT and the sintered specimens co-doped with MgO-Y2O3 and MgO-Dy2O3 exhibited a tetragonal structure (ICCD No. 01-081-2202) of BT with peak splitting of the (002) and (200) planes. Rare-earth oxides play a dominant role in the formation of a core-shell structure in the BaTiO3 grain, and they exist at a shell of BaTiO3 grain, as described above [13]. The migration of rare-earth oxides into the core grain of ferroelectric BT ought to be minimised to ensure the proper formation of a core-shell structure in the BT grain [25]. Furthermore, it has been observed that as the ionic radius of the rare-earth oxides increases, a greater quantity of MgO is needed to establish a core-shell structure in the MgO-Re2O3-BaTiO3 system [26]. The sintering condition also significantly influences the development of the core-shell structure [22]. According to the XRD patterns presented in Figure 1b, 1 mol% of MgO was insufficient to create a core-shell structure for specimens sintered with MgO–Y2O3 and/or MgO–Dy2O3. As a consequence, the Y3+ and Dy3+ ions might permeate the core grain, undermining the core-shell structure and revealing a tetragonal phase [25]. In contrast, the specimens sintered with MgO and co-sintered with MgO-Yb2O3 exhibited a pseudo-cubic structure (ICCD No. 01-075-0212), evidenced by the fusion of the (002) and (200) planes in the XRD patterns. The diffusion of the Yb3+ ion (0.87 Å, CN = 6), which has a slightly smaller ionic radius than that of the Y3+ (0.9 Å, CN = 6) and Dy3+ ions (0.912 Å, CN = 6), was effectively inhibited by MgO in the specimens co-sintered with MgO–Yb2O3, resulting in the observed pseudo-cubic structure [27]. For the sintered specimens doped with MgO, MgO is substituted at the Ti site of BT due to the ionic radius of Mg2+ ions (0.72 Å) being similar to that of Ti4+ ions (0.605 Å) (Equation (3)), and the positively charged oxygen vacancy induces the deformation of BT from a tetragonal to a pseudo-cubic structure [12,27,28].
MgO + Ti O 2 Ba Ba + Mg Ti + 2 O O + V O · ·
To obtain the lattice parameters of the sintered specimens, Rietveld refinement was performed on the XRD patterns, as shown in Figure 2. In the XRD data, the dotted line represents the observed intensity, while the solid line signifies the calculated intensity. The line at the base highlights the disparity between the observed and calculated values. Table 3 presents the R-factors and the lattice parameters derived from the Rietveld refinements for the sintered specimens. The goodness of fit (GoF) and the Bragg R-factor, serving as indicators of RIR, fell within the ranges of 1.9–2.6 and 1.79–2.87, respectively. These values suggest that the RIR data are highly reliable.
Figure 3a displays the unit-cell volumes of the BT specimens either doped with MgO or co-doped with MgO-Re2O3 (with Re representing Y, Yb, or Dy) and sintered under an optimal sintering condition. The unit-cell volume of specimens co-doped with MgO-Y2O3 and/or MgO-Dy2O3 may be influenced by the substitution sites of Y3+ and Dy3+ ions within the BT lattice when intermediate rare-earth oxides, such as Y2O3 and Dy2O3, are substituted into BT [16]. As previously mentioned, MgO replaces the Ti site of BT, leading to the Y3+ (1.234 Å, CN = 12) and Dy3+ ions (1.255 Å, CN = 12) primarily occupying the Ba site (1.61 Å, CN = 12) in specimens co-doped with MgO-Y2O3 and/or MgO-Dy2O3 [28]. Consequently, specimens co-doped with MgO-Y2O3 and/or MgO-Dy2O3 exhibited a smaller unit-cell volume compared to specimens solely doped with MgO. Conversely, the unit-cell volume of specimens co-doped with MgO-Yb2O3 was akin to that of the MgO-doped specimens, owing to the Yb2O3 substitution at the Ti site. Notably, prior research indicates that the Yb3+ ion (0.87 Å, CN = 6), due to its compact ionic radius, takes the place of the Ti site regardless of the Ba/Ti ratio [11,28].
Figure 3b displays the lattice parameters and tetragonality of the BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (with Re representing Y, Yb, or Dy). The sintered specimens co-doped with MgO-Y2O3 and MgO-Dy2O3, possessing a tetragonal structure, exhibited reduced tetragonality (c/a) in comparison to that of pure BT. This reduction can be attributed to the substitution of Dy3+ and Y3+ at the Ba site, leading to a marked decrease in the c-axis relative to the a-axis [29,30]. In contrast, the sintered specimens doped with MgO and co-doped with MgO-Yb2O3 demonstrated a tetragonality (c/a = 1) lower than that of pure BT. This observation is linked to the presence of positively charged oxygen vacancies for the specimens doped with MgO and the emergence of a shell phase for the specimens co-doped with MgO-Yb2O3. For those sintered specimens with a pseudo-cubic structure, the a-axis and c-axis lattices within the perovskite framework align, negating any XRD pattern separation between the (002) and (200) planes.

3.2. Dielectric and Electrical Properties

Figure 4 shows the dependence of the dielectric constant (εr) on the tetragonality (c/a) of the BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (with Re representing Y, Yb, or Dy). With the substitution of MgO and rare-earth oxides, the sintered specimens showed a lower dielectric constant (εr) than that of pure BT. The dielectric constant (εr) is affected by the magnitude of spontaneous polarisation and the domain wall density and its mobility [31,32,33,34]. Because spontaneous polarisation is proportional to tetragonality (c/a) [35], the dielectric constant (εr) of the sintered specimens doped with MgO and/or rare-earth oxides decreased with the decrease of tetragonality, which restricts the moving space of the Ti4+ ion in the [TiO6] octahedron. Although the tetragonalities of the sintered specimens co-doped with MgO-Yb2O3 and doped with MgO are the same, the εr of the sintered specimens doped with MgO was slightly larger than that of the sintered specimens co-doped with MgO-Yb2O3 due to the larger grain size at the submicron level, as confirmed in Figure 5.
Figure 5 presents SEM micrographs detailing the surface structures of the BT specimens, either doped with MgO or co-doped with MgO-Re2O3 (with Re representing Y, Yb, or Dy). All sintered specimens exhibited dense microstructures, aligning with relative densities exceeding 94%, as documented in Table 2. The pure BT manifested the most prominent grain size, accompanied by abnormal grain growth. Upon introducing MgO and rare-earth oxides, the grain size of the sintered specimens was observed to be more diminutive compared to that of pure BT. This reduction is attributed to the grain growth inhibition properties of MgO and the rare-earth oxides during the sintering process [11,12,13,17]. Specimens co-doped with MgO-Y2O3 and MgO-Dy2O3 displayed comparable grain sizes. Conversely, those doped with MgO and co-doped with MgO-Yb2O3 demonstrated considerably finer grain sizes than their counterparts co-doped with MgO-Y2O3 and MgO-Dy2O3.
Figure 6 illustrates the electrical resistivity (ρ) of the BT specimens either doped with MgO or co-doped with MgO-Re2O3 (with Re representing Y, Yb, or Dy). The sintered specimens doped with MgO, as well as those co-doped with MgO-Y2O3 and MgO-Dy2O3, exhibited an elevated electrical resistivity value of 850 GΩ (500 V, 60 s) or higher, approximately thrice that of pure BT. This enhancement is attributed to the grain size reduction mediated by MgO, Y2O3, and Dy2O3, all of which act as grain growth inhibitors, as previously discussed [36,37]. In contrast, specimens co-doped with MgO-Yb2O3 displayed electrical resistivity analogous to pure BT despite the grain size reduction. This behaviour can likely be ascribed to the pyrochlore phase of Yb2Ti2O7, a noted high ionic conductor [38]. Prior research has also indicated that the pyrochlore phase expedites the migration of oxygen vacancies, a primary factor in insulation degradation [39].
To assess the temperature stability of the sintered specimens, the TCC was gauged from −50 °C to 150 °C at 1 KHz. Figure 7a depicts the temperature-dependent dielectric constant (εr) for BT specimens either doped with MgO or co-doped with MgO-Re2O3 (with Re representing Y, Yb, or Dy). Specimens co-doped with MgO-Y2O3 and MgO-Dy2O3 exhibited a pronounced dielectric peak near 100 °C, with the Curie temperature (Tc) for each specimen marginally dropping to 99 °C and 103 °C, as seen in Figure 7a. For specimens doped with MgO that display a cubic structure, there was a broadening of the dielectric peak accompanied by a shift in the Curie temperature. The TCCs for the BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (with Re representing Y, Yb, or Dy) are presented in Figure 7b. As previously mentioned, the BT specimens with MgO doping exhibited a flattened TCC curve alongside a reduction in Curie temperature. This observed behaviour met the X8R specification (−55~150 °C, ΔC/C approximately ± 15%) of the EIA classification, stemming from the introduction of oxygen vacancies by the substitution of Mg2+ ions at the Ti site. This trend aligns with previous studies on MgO-doped BT [12]. For enhanced temperature stability in rare-earth oxide-doped BaTiO3, a diminished transition peak is essential. The predominant factor influencing the spread of the Curie temperature appears to be the internal microstructural stress caused by the transition from tetragonal to cubic structures in BT [11,40]. It appears that the phase transition was not effectively realised in specimens co-doped with MgO-Y2O3 and MgO-Dy2O3, failing to meet the X8R specification. However, the transition was prominent in specimens co-doped with MgOYb2O3, which adhered to the X7R specification.

4. Conclusions

The effects of MgO and rare-earth oxides (Y2O3, Yb2O3, and Dy2O3) on the structural and electrical properties of BaTiO3 have been investigated. For specimens sintered at 1350 °C for durations ranging from 1 to 5 h in air, a single phase of BaTiO3 with a tetragonal structure was identified in pure BT and BT specimens co-doped with MgO-Y2O3 and/or MgO-Dy2O3. However, for specimens doped with MgO or MgO-Yb2O3, BaTiO3 with a pseudo-cubic structure was detected. A secondary phase of Yb2Ti2O7 was identified in BT specimens co-doped with MgO and Yb2O3, attributed to the limited solubility of the Yb3+ ions in the BT lattice. These results indicate a smaller unit-cell volume in BT specimens co-doped with MgO-Y2O3 and MgO-Dy2O3 compared to that of BT specimens doped solely with MgO, suggesting that the Y3+ and Dy3+ ions are substituted at the Ba site of BT. With the introduction of MgO and rare-earth oxides, the dielectric constant (εr) of the sintered specimens decreased, which was associated with a reduction in tetragonality (c/a) that limited the mobility of the Ti4+ ion within the TiO6 octahedron. The electrical resistivity of the sintered specimens increased with reductions of the grain size in BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y or Dy). BT specimens co-doped with MgO-Yb2O3 and/or MgO met the EIA X7R (−55 to 125 °C, ΔC/C of ±15% or less) and X8R specifications (−55 to 150 °C, ΔC/C of ±15% or less), respectively.

Author Contributions

E.S.K. and J.H.P. designed the experiments. J.H.P. performed the experiments and analysed the specimens. E.S.K. analysed and explained the dependence of the electrical properties on the microstructure behaviours of the specimens. E.S.K. and J.H.P. discussed the data and prepared the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Technology Innovation (20010938, Development of 630 V high capacity MLCC array module for power train), funded by the Ministry of Trade, Industry, and Energy (MOTIE, Sejong City, Republic of Korea).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pu, Y.; Chen, W.; Chen, S.; Langhammer, H.T. Microstructure and dielectric properties of dysprosium-doped barium titanate ceramics. Ceramica 2005, 51, 214–218. [Google Scholar] [CrossRef]
  2. Hao, H.; Liu, M.; Liu, H.; Zhang, S.; Shu, X.; Wang, T.; Yao, Z.; Cao, M. Design, Fabrication and dielectric properties on core-double shell BaTiO3-based ceramics for MLCC application. RSC Adv. 2015, 5, 8868–8876. [Google Scholar] [CrossRef]
  3. Zhang, Y.; Wang, X.; Kim, J.Y.; Tian, Z.; Fang, J.; Hur, K.H.; Li, L. High performance of BaTiO3-Based BME MLCC Nanopowder Prepared by Aqueous Chemical Coating method. J. Am. Ceram. Soc. 2012, 95, 1628–1633. [Google Scholar] [CrossRef]
  4. Hernandez, V.I.; García-Guti, D.I.; Aguilar-Garib, J.A.; Nava-Quintero, R.J. Characterization of precipitates formed in X7R 0603 BME-MLCC during sintering. Ceram. Int. 2021, 47, 310–319. [Google Scholar] [CrossRef]
  5. MYoon, M.S.; Ur, S.C. Effects of A-site Ca and B-site Zr substitution on dielectric properties and microstructure in tin-doped BaTiO3-CaTiO3 composites. Ceram. Int. 2008, 34, 1941–1948. [Google Scholar]
  6. Zhu, X.; Zhu, J.; Zhou, S.; Liu, Z.; Ming, N.; Hesse, D. BaTiO3 nanocrystals: Hydrothermal synthesis and structural characterization. J. Cryst. Growth 2005, 283, 553–562. [Google Scholar] [CrossRef]
  7. Beck, H.P.; Eiser, W.; Haberkorn, R. Pitfalls in the synthesis of nanoscaled perovskite type compounds. Part I: Influence of different sol-gel preparation methods and characterization of nanoscaled BaTiO3. J. Eur. Ceram. Soc. 2001, 21, 687–693. [Google Scholar] [CrossRef]
  8. Shimooka, H.; Kuwabara, M. Crystallinity and Stoichiometry of Nano-Structured Sol-Gel Derived BaTiO3 Monolithic Gels. J. Am. Ceram. Soc. 1996, 79, 2983–2985. [Google Scholar] [CrossRef]
  9. Wei, X.; Xu, G.; Ren, Z.; Shen, G.; Han, G. Room-Temperature Synthesis of BaTiO3 Nanoparticles in Large Batches. J. Am. Ceram. Soc. 2008, 91, 3774–3780. [Google Scholar] [CrossRef]
  10. Kozerozhets, I.; Semenov, E.; Kozlova, L.; Ioni, Y.; Avdeeva, V.; Ivakin, Y. Mechanism to form nanosized oxides when burning aqueous carbohydrate salt solutions. Mater. Chem. Phys. 2023, 309, 128387. [Google Scholar] [CrossRef]
  11. Song, Y.H.; Han, Y.H. Effects of Rare-Earth Oxides on Temperature Stability of Acceptor-Doped BaTiO3. Jpn. J. Appl. Phys. 2005, 44, 6143. [Google Scholar] [CrossRef]
  12. Jeong, J.; Han, Y.H. Effects of MgO-Doping on Electrical Properties and Microstructure of BaTiO3. Jpn. J. Appl. Phys. 2004, 43, 5373. [Google Scholar] [CrossRef]
  13. Kim, C.H.; Park, K.J.; Yoon, Y.J.; Hong, M.H.; Hong, J.O.; Hur, K.H. Role of Yttrium and magnesium in the formation of core-shell structure of BaTiO3 grains in MLCC. J. Eur. Ceram. Soc. 2008, 28, 1213–1219. [Google Scholar] [CrossRef]
  14. Sakabe, Y.; Hamaji, Y.; Sano, H.; Wada, N. Effects of Rare-Earth Oxides on the Reliability of X7R Dielectrics. Jpn. J. Appl. Phys. 2002, 41, 5668. [Google Scholar] [CrossRef]
  15. Saito, H.; Chazono, H.; Kishi, H.; Yamaoka, N. X7R Multilayer Ceramic Capacitors with Nickel Electrodes. Jpn. J. Appl. Phys. 1991, 30, 2307. [Google Scholar] [CrossRef]
  16. Park, K.J.; Kim, C.H.; Yoon, Y.J.; Song, S.M.; Kim, Y.T.; Hur, K.H. Doping behaviors of dysprosium, yttrium and holmium in BaTiO3 ceramics. J. Eur. Ceram. Soc. 2009, 29, 1735–1741. [Google Scholar] [CrossRef]
  17. Li, Y.X.; Yao, X.; Wang, X.S.; Hao, Y.B. Studies of dielectric properties of rare earth (Dy, Tb, Eu) doped barium titanate sintered in pure nitrogen. Ceram. Int. 2012, 38, S29–S32. [Google Scholar] [CrossRef]
  18. Wang, M.; Xue, K.; Zhang, K.; Li, L. Dielectric properties of BaTiO3-based ceramics are tuned by defect dipoles and oxygen vacancies under a reducing atmosphere. Ceram. Int. 2022, 48, 22212–22220. [Google Scholar] [CrossRef]
  19. Kishi, H.; Okino, Y.; Honda, M.; Iguchi, Y.; Imaeda, M.; Takahashi, Y.; Ohsato, H.; Okuda, T. The effect of MgO and rare-earth oxide on formation behavior of core-shell structure in BaTiO3. Jpn. J. Appl. Phys. 1997, 36, 5954. [Google Scholar] [CrossRef]
  20. Yang, W.C.; Hu, C.T.; Lin, I.N. Effect of Y2O3/MgO Co-doping on the electrical properties of base-metal-electroded BaTiO3 materials. J. Eur. Ceram. Soc. 2004, 24, 1479–1483. [Google Scholar] [CrossRef]
  21. Wang, S.; Zhang, S.; Zhou, X.; Li, B.; Chen, Z. Effect of sintering atmospheres on the microstructure and dielectric properties of Yb/Mg co-doped BaTiO3 ceramics. Mater. Lett. 2005, 59, 2457–2460. [Google Scholar] [CrossRef]
  22. Gong, H.; Wang, X.; Zhang, S.; Yang, X.; Li, L. Influence of sintering temperature on core-shell structure evolution and reliability in Dy modified BaTiO3 dielectric ceramics. Phys. Status Solidi 2014, 211, 1213–1218. [Google Scholar] [CrossRef]
  23. Hahn, D.W.; Han, Y.H. Electrical properties of Yb-Doped BaTiO3. Jpn. J. Appl. Phys. 2009, 48, 111406. [Google Scholar] [CrossRef]
  24. Nfissi, A.; Ababou, Y.; Belhajji, M.; Sayouri, S.; Hajji, L.; Bennani, M.N. Investigation of Ba and Ti site occupation effects on structural, optical and dielectric properties of sol gel processed Y-doped BaTiO3 ceramics. Opt. Mater. 2021, 122, 111708. [Google Scholar] [CrossRef]
  25. Liu, R.; Chen, Z.; Lu, Z.; Wang, X. Effects of sintering temperature and Bi2O3, Y2O3 and MgO co-doping on the dielectric properties of X8R BaTiO3-based ceramics. Ceram. Int. 2022, 48, 2377–2384. [Google Scholar] [CrossRef]
  26. Kishi, H.; Kohzu, N.; Sugino, J.; Ohsato, H.; Iguchi, Y.; Okuda, T. The effect of Rare-earth (La, Sm, Dy, Ho and Er) and Mg on the microstructure in BaTiO3. J. Eur. Ceram. Soc. 1999, 19, 1043–1046. [Google Scholar] [CrossRef]
  27. Lin, Y.T.; Ou, S.F.; Lin, M.H.; Song, Y.R. Effect of MgO addition on the microstructure and dielectric properties of BaTiO3 ceramics. Ceram. Int. 2018, 44, 3531–3535. [Google Scholar]
  28. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976, 32, 751–766. [Google Scholar] [CrossRef]
  29. Badapanda, T.; Senthil, V.; Panigrahi, S.; Anwar, S. Diffuse of phase transition behavior of dysprosium doped barium titanate ceramic. J. Electroceram. 2013, 31, 55–60. [Google Scholar] [CrossRef]
  30. Tihtih, M.; Ibrahim, J.E.F.M.; Basyooni, M.A.; En-nadir, R.; Belaid, W.; Hussainova, I.; Kocserha, I. Development of yttrium-doped BaTiO3 for next-generation multilayer ceramic capacitors. ACS Omega 2023, 8, 8448–8460. [Google Scholar] [CrossRef] [PubMed]
  31. Arlt, G.; Hennings, D.; de With, G. Dielectric properties of fine-grained barium titanate ceramics. J. Appl. Phys. 1985, 58, 1619–1625. [Google Scholar] [CrossRef]
  32. Yoon, S.H.; Kim, S.J.; Kim, D.Y. Influence of excess Ba concentration on the dielectric nonlinearity in Mn and V-doped BaTiO3 multilayer ceramic capacitors. J. Appl. Phys. 2013, 114, 224103. [Google Scholar] [CrossRef]
  33. Yoon, S.H.; Kim, M.Y.; Nam, C.H.; Seo, J.W.; Wi, S.K.; Hur, K.H. Correlation between tetragonality (c/a) and direct current (dc) bias characteristics of BaTiO3-based multilayer ceramic capacitors. Appl. Phys. Lett. 2015, 107, 072906. [Google Scholar] [CrossRef]
  34. Tutuncu, G.; Li, B.; Bowman, K.; Jones, J.L. Domain wall motion and electromechanical strain in lead free piezoelectrics: Insight from the model system (1-x)Ba(Zr0.2Ti0.8)O3-x(Ba0.7Ca0.3)TiO3 using in situ high-energy X-ray diffraction during application of electric fields. J. Appl. Phys. 2014, 115, 144104. [Google Scholar] [CrossRef]
  35. Qi, T.; Grinberg, I.; Rappe, A.M. Correlations between tetragonality, A.M.R.; polarization, and ionic displacement in PbTiO3-derived ferroelectric perovskite solid solutions. Phys. Rev. B 2010, 82, 134113. [Google Scholar] [CrossRef]
  36. Gong, H.; Wang, X.; Zhang, S.; Wen, H.; Li, L. Grain size effect on electrical and reliability characteristics of modified fine-grained BaTiO3 ceramics for MLCCS. J. Eur. Ceram. Soc. 2014, 34, 1733–1739. [Google Scholar] [CrossRef]
  37. Yoon, S.H.; Randall, C.A.; Hur, K.H. Influence of Grain Size on Impedance Spectra and Resistance Degradation Behavior in Acceptor (Mg)-Doped BaTiO3 Ceramics. J. Am. Ceram. Soc. 2009, 92, 2944–2952. [Google Scholar] [CrossRef]
  38. Shlyakhtina, A.; Fedtke, P.; Busch, A.; Kolbanev, I.; Barfels, T.; Wienecke, M.; Sokolov, A.; Ulianov, V.; Trounov, V.; Shcherbakova, L. Effect of Ca-doping on the electrical conductivity of oxide ion conductor Yb2Ti2O7. Solid State Ion. 2008, 179, 1004–1008. [Google Scholar] [CrossRef]
  39. Yoon, S.H.; Park, Y.S.; Hong, J.O.; Sinn, D.S. Effect of pyrochlore (Y2Ti2O7) phase on the resistance degradation on yttrium-doped BaTiO3 ceramic capacitors. J. Mater. Res. 2007, 22, 2539–2543. [Google Scholar] [CrossRef]
  40. Hennings, D.; Schnell, A.; Simon, G. Diffuse Ferroelectric Phase Transitions in Ba(Ti1-yZry)O3 ceramics. J. Am. Ceram. Soc. 1982, 65, 539–544. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of the sintered specimens doped with MgO and co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy) and (b) the enlarged XRD patterns of the sintered specimens in the range of 2θ from 44° to 46°.
Figure 1. (a) XRD patterns of the sintered specimens doped with MgO and co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy) and (b) the enlarged XRD patterns of the sintered specimens in the range of 2θ from 44° to 46°.
Processes 11 03235 g001
Figure 2. Rietveld refinement patterns of the sintered specimens: (a) doped with MgO, (b) co-doped with MgO-Yb2O3, (c) pure BT, (d) co-doped with MgO-Y2O3, (e) co-doped with MgO-Dy2O3.
Figure 2. Rietveld refinement patterns of the sintered specimens: (a) doped with MgO, (b) co-doped with MgO-Yb2O3, (c) pure BT, (d) co-doped with MgO-Y2O3, (e) co-doped with MgO-Dy2O3.
Processes 11 03235 g002
Figure 3. (a) Unit-cell volume and (b) lattice parameters of the BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy).
Figure 3. (a) Unit-cell volume and (b) lattice parameters of the BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy).
Processes 11 03235 g003
Figure 4. Dependence of the dielectric constant (εr) on the tetragonality (c/a) of the BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy).
Figure 4. Dependence of the dielectric constant (εr) on the tetragonality (c/a) of the BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy).
Processes 11 03235 g004
Figure 5. Scanning electron micrographs of the sintered specimens: (a) pure BT, (b) co-doped with MgO-Y2O3, (c) co-doped with MgO-Dy2O3, (d) doped with MgO, and (e) co-doped with MgO-Yb2O3 (scale bar: 10 μm).
Figure 5. Scanning electron micrographs of the sintered specimens: (a) pure BT, (b) co-doped with MgO-Y2O3, (c) co-doped with MgO-Dy2O3, (d) doped with MgO, and (e) co-doped with MgO-Yb2O3 (scale bar: 10 μm).
Processes 11 03235 g005
Figure 6. Electrical resistivities of BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy).
Figure 6. Electrical resistivities of BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy).
Processes 11 03235 g006
Figure 7. (a) Temperature dependence of the dielectric constant (εr) and (b) temperature coefficient of capacitance (TCC) of BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy).
Figure 7. (a) Temperature dependence of the dielectric constant (εr) and (b) temperature coefficient of capacitance (TCC) of BT specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy).
Processes 11 03235 g007
Table 1. Composition of the BaTiO3 specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy) [unit: mol].
Table 1. Composition of the BaTiO3 specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy) [unit: mol].
AdditivesAbbreviationBaTiO3MgOY2O3Yb2O3Dy2O3
Pure BaTiO3BT1000000
MgOMg1001000
MgO-Y2O3Mg/Y1001100
MgO-Yb2O3Mg/Yb1001010
MgO-Dy2O3Mg/Dy1001001
Table 2. Relative densities of the BaTiO3 specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy) sintered at 1350 °C for 1–5 h in air.
Table 2. Relative densities of the BaTiO3 specimens doped with MgO and/or co-doped with MgO-Re2O3 (Re = Y, Yb, or Dy) sintered at 1350 °C for 1–5 h in air.
SpecimenSintering Condition
(°C/h)
Apparent
Density
(g/cm3)
Theoretical
Density
(g/cm3)
Relative Density
(%)
MgO-Dy2O31350 °C/1 h5.81406.020096.58
Pure BT5.80306.001896.42
MgO-Y2O35.78456.017096.14
MgO5.66866.008094.35
MgO-Yb2O31350 °C/5 h5.76836.017095.87
Table 3. Lattice parameters of the sintered specimens and R-factors obtained from the Rietveld refinements.
Table 3. Lattice parameters of the sintered specimens and R-factors obtained from the Rietveld refinements.
SpecimenLattice Parameter (Å)Tetragonality
(c/a)
Unit-Cell Volume
3)
RbraggGoF
a-Axisc-Axis
Pure BT3.99354.03501.0103964.34992.872.6
MgO-Dy2O33.99574.02921.0083864.32942.302.3
MgO-Y2O33.99734.02801.0076864.36252.352.4
MgO4.00974.0097164.46562.212.1
MgO-Yb2O34.01004.0100164.47921.791.9
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

Park, J.H.; Kim, E.S. Effects of MgO and Rare-Earth Oxides (Y2O3, Yb2O3, Dy2O3) on the Structural Characteristics and Electrical Properties of BaTiO3. Processes 2023, 11, 3235. https://doi.org/10.3390/pr11113235

AMA Style

Park JH, Kim ES. Effects of MgO and Rare-Earth Oxides (Y2O3, Yb2O3, Dy2O3) on the Structural Characteristics and Electrical Properties of BaTiO3. Processes. 2023; 11(11):3235. https://doi.org/10.3390/pr11113235

Chicago/Turabian Style

Park, Jae Hoon, and Eung Soo Kim. 2023. "Effects of MgO and Rare-Earth Oxides (Y2O3, Yb2O3, Dy2O3) on the Structural Characteristics and Electrical Properties of BaTiO3" Processes 11, no. 11: 3235. https://doi.org/10.3390/pr11113235

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