Next Article in Journal
Agent Based Fuzzy T-S Multi-Model System and Its Applications
Next Article in Special Issue
A Novel Piezoelectric Energy Harvester Using the Macro Fiber Composite Cantilever with a Bicylinder in Water
Previous Article in Journal
Microwave-Assisted Synthesis and Antifungal Activity of Some Novel Thioethers Containing 1,2,4-Triazole Moiety
Previous Article in Special Issue
Enhanced Magnetoelectric Effect in Permendur/Pb(Zr0.52Ti0.48)O3 Laminated Magnetostrictive/Piezoelectric Composite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Ta2O5 and Nb2O5 Dopants on the Stable Dielectric Properties of BaTiO3–(Bi0.5Na0.5)TiO3-Based Materials

Department of Materials and Mineral Resources Engineering, National Taipei University of Technology, Taipei 106, Taiwan
*
Author to whom correspondence should be addressed.
Current Address: Department of Materials and Mineral Resources Engineering, National Taipei University of Technology, 1, Sec. 3, Chung-Hsiao E. Rd., Taipei 106, Taiwan
Appl. Sci. 2015, 5(4), 1221-1234; https://doi.org/10.3390/app5041221
Submission received: 4 September 2015 / Revised: 23 October 2015 / Accepted: 2 November 2015 / Published: 16 November 2015
(This article belongs to the Special Issue Ferroelectric Ceramics)

Abstract

:
In this study, BaTiO3–(Bi0.5Na0.5)TiO3 ceramics with various amounts of Ta2O5 dopant were investigated for their ability to enhance high-temperature stability to meet X9R specifications. The results were compared to those for ceramics with the common Nb2O5 additive. The best composition appeared to be 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 with 2 mol% Ta2O5 dopant sintered at 1215 °C, which had a dielectric constant of 1386, a tanδ value of 1.8%, temperature coefficients of capacitance (TCCs) of −1.3% and 1.2%, and electrical resistivities of 2.8 × 1012 and 1.5 × 1010 Ω·cm at 25 °C and 200 °C, respectively. Its microstructure consisted of fine equiaxed grains with a perovskite structure and an average grain size of 0.46 μm and some rod-like grains of second-phase Ba6Ti17O40 with a size of approximately 3.2 μm. The Ta2O5 dopant caused a reduction in the grain size and a slight increase in trapped pores. The temperature dependence of the dielectric constant flattened and the Curie point was dramatically suppressed with the addition of Ta2O5 dopant, leading to smooth dielectric temperature characteristics over a relatively broad temperature range. The X9R formulations and their dielectric properties were highly repeatable in this study.

1. Introduction

Multilayer ceramic capacitors (MLCCs) are crucial electronic components used in virtually every area of electronics, and the number and variety of their applications have steadily grown. During the past decade, MLCC manufacturers have been driven to accelerate progress in miniaturization, enhancement of volumetric efficiency, reinforcement of stability and reliability against environment, and cost reduction, none of which can be achieved without an improvement in materials formulation and processing technology. The compositions of most ceramic capacitors are based on ferroelectric barium titanate (BaTiO3) materials. BaTiO3 can be either chemically or physically modified to exhibit the required temperature-stable dielectric behavior [1].
MLCCs used in automobile electronic parts, such as anti-lock brake systems, engine electronic control units, crank angle sensors, and programmed fuel injections, are exposed to high-temperature working conditions of greater than approximately 130 °C in the engine room of an automobile. MLCCs with X8R specifications (X8R: −55 °C to 150 °C, ∆C/C25 °C ≤ ±15%) were developed to satisfy this requirement. Because of the dramatic variation in permittivity of BaTiO3 ceramics above the Curie temperature (Tc ≈ 125 °C), the X8R formulation makes use of the core–shell structure to retain the temperature-stable characteristic by adding rare earth elements with a smaller ionic radius, such as Tm, Yb, Lu, Sc, Er, and Y, to shift the Tc to higher temperatures [2,3,4]. Nevertheless, for aerospace, oil drilling, and various other applications, MLCCs in electronic devices must withstand even harsher working environments, typified by temperatures higher than 150 °C. Accordingly, Yuan et al. and a few other researchers have recently started to work on the development of dielectric formulations capable of satisfying X9R specifications (−55 °C to 200 °C, ∆C/C25 °C ≤ ±15%) [5,6,7,8,9,10,11]. The utilization of an effective Curie-point shifter, such as (Bi0.5Na0.5)TiO3 or PbTiO3, to increase the Tc point along with the formation of the core–shell structure was found to be an effective method for obtaining high-temperature stability [12,13,14,15,16]. Dielectric formulations, including BaTiO3-based compositions, such as BaTiO3–(Bi0.5Na0.5)TiO3 [5,7,8,14,17,18,19], BaTiO3–Bi(Zn0.5Ti0.5)O3 [16,20], and BaTiO3–Pb(Ti0.55Sn0.45)O3 [10] systems, and (Bi0.5Na0.5)TiO3-based compositions, such as (Bi0.5Na0.5)TiO3–BaTiO3–CaTiO3 [6,9,21] systems, have been developed recently to meet the temperature coefficient of capacitance (TCC) requirements of the X9R specifications. Another approach was a multi-phase composite consisting of a high-Tc material and a low-Tc material. For instance, new X9R dielectrics based on the BaTiO3–LiTaO3 compositions reported by Wang et al. can be sintered in a reducing atmosphere and are compatible with the base metal electrode (BME) process [22].
In the BaTiO3–(Bi0.5Na0.5)TiO3 system, (Bi0.5Na0.5)TiO3 shows strong ferroelectricity at room temperature, undergoes a phase transition with a broad dielectric maximum approximately close to 320 °C [8], and is considered as an ideal material to shift the Curie temperature of BaTiO3 above 150 °C [15]. The distortion and deformation of BaTiO3 structure induced by the substitution of Na+ and Bi3+ into Ba sites by (Bi0.5Na0.5)TiO3 addition improves the stable temperature characteristics of the dielectric properties of BaTiO3. With the addition of Co3O4 and Nb2O5, the stable dielectric properties of BaTiO3–(Bi0.5Na0.5)TiO3 were tailored to meet the X9R specification by adjusting the core–shell structure [14,15]. It was also shown in the literature that the dielectric properties of BaTiO3 or bismuth-based compounds can be significantly enhanced with the addition of Ta2O5, through tailoring the microstructure, improving densification, and shifting Curie temperature [23,24]. In this study, BaTiO3–(Bi0.5Na0.5)TiO3 dielectric ceramics incorporating various amounts of Ta2O5 additive were investigated to enhance their high temperature stability to meet the X9R specifications. The effect of the Ta2O5 additive on the densification, crystalline phase, microstructure development, and dielectric properties of BaTiO3–(Bi0.5Na0.5)TiO3 ceramics compared to those of the common Nb2O5 additive under the same processing conditions were investigated and are discussed below.

2. Experimental Procedure

Dielectric powders of (1 − x)BaTiO3x(Bi0.5Na0.5)TiO3 ceramics with various amounts of Ta2O5 and Nb2O5 were prepared using the solid-state reaction technique. Highly pure (>99.9% purity) BaTiO3 (Ferro, reagent grade), Bi2O3 (SHOWA, reagent grade), Na2CO3 (SHOWA, reagent grade), TiO2 (SHOWA, reagent grade), Ta2O5 (Alfa, reagent grade), and Nb2O5 (Alfa, reagent grade) were used as raw materials. The host material (Bi0.5Na0.5)TiO3 was prepared by mixing appropriate amounts of Bi2O3, Na2CO3, and TiO2 and calcining them at 800 °C for 12 h, and the phase was identified using X-ray diffraction (XRD, Rigaku DMX-2200, Rigaku, Tokyo, Japan). Formulations based on compositions of (1 − x)BaTiO3x(Bi0.5Na0.5)TiO3 (x = 0.05, 0.1, and 0.2) with various amounts of Ta2O5 and Nb2O5 (0, 1.0, 1.5, and 2.0 mol%) were mixed and milled in methyl alcohol solution using polyethylene jars and zirconia balls for 24 h and then dried at 80 °C in an oven overnight. The mixed powders were added to 4 wt% of a 15 wt% Polyvinyl Alcohol (PVA) solution and were pressed into disc-shaped compacts under a uniaxial pressure of 120 MPa. The samples were then heat treated at 550 °C for 4 h to eliminate the PVA, followed by sintering at 1200 °C to 1245 °C. The heating and cooling rates were controlled at 5 °C/min for all samples, and the dwell time was set to 2 h.
The apparent density of the sintered ceramics was measured on five samples for each composition using the Archimedes method with de-ionized water. Phase identification on the polished surfaces of the sintered bulk ceramics was performed using X-ray diffraction (XRD, Siemens D5000, Karlsruhe, Germany). The microstructures of the sintered samples were examined with scanning electron microscopy (SEM, JEOL 6500F, Tokyo, Japan) and energy-dispersive spectroscopy (EDS). The average grain sizes of the sintered samples were determined using the linear intercept, which was calculated as the mean value of the measurement diagonals of about 200 grains. The dielectric properties of the sintered samples were measured as a function of temperature using an automatic measurement system incorporating a Delta Design Environmental Chamber with an LCR meter (HP 4284A, Santa Clara, CA, USA) at a frequency of 1 kHz. Before measurement, electrodes were fabricated on the disc samples using silver paste applied on opposing surfaces and fired at 750 °C. The measurements were performed during heating cycles from −55 °C to 130 °C at a rate of 2 °C/min with an accuracy of 0.1 °C. An alternating voltage of 1 V was applied. The electrical resistivity of the bulk ceramics was determined by a high resistance meter (HP 4339B, Santa Clara, CA, USA) at 25 and 200 °C at a DC voltage of 500 V. The measurements of dielectric properties and electrical resistivity were performed on at least three samples in each case, with an accuracy of ±3%.

3. Results and Discussion

To determine the best composition of the host (1 − x)BaTiO3x(Bi0.5Na0.5)TiO3 ceramic, formulations with an x value of 0.05, 0.1, or 0.2 were prepared and characterized. While the sintering temperature of BaTiO3 is typically above 1300 °C, the addition of (Bi0.5Na0.5)TiO3 significantly reduced the sintering temperature to 1230 °C for an x value of 0.05 and to 1215 °C for x values of 0.1 and 0.2, which was caused by the formation of the liquid phase [14,25]. XRD results indicated the presence of the tetragonal BaTiO3 perovskite phase (tetragonal, P4/mmm; JCPDF 75-0462) with trace amounts of second-phase Ba6Ti17O40 (monoclinic, C2/c; JCPDF 77-1566), which increased with the (Bi0.5Na0.5)TiO3 content. This observation is in accord with those reported in the literature [25,26]. In addition, the peaks corresponding to BaTiO3 in the XRD patterns shifted to lower angles due to the incorporation of the smaller ions Na+ (VI; 1.02 Å) and Bi3+ (VI; 1.03 Å) in the Ba2+ (VI; 1.35 Å) sites. The lattice constants calculated from XRD patterns appeared to be a = 3.991 Å and c = 4.025 Å. The dielectric properties of the (1 − x)BaTiO3–x(Bi0.5Na0.5)TiO3 at the maximum density, measured at a frequency of 1 KHz, are listed in Table 1. The Curie point of the (1 − x)BaTiO3x(Bi0.5Na0.5)TiO3 increased with the (Bi0.5Na0.5)TiO3 content and was 145 °C, 170 °C, and >200 °C for x values of 0.05, 0.1, and 0.2, respectively. This increase occurred because Bi–O bonds are weaker than Ba–O bonds, the substitution of Bi3+ into Ba sites leads to strengthened Ti–O bonds, and more energy is needed to maintain the balance of Ti4+ displacement in the TiO6 octahedrons; thus, Tc increased [14,15,19]. The dielectric constant of (1 − x)BaTiO3x(Bi0.5Na0.5)TiO3 at 25 °C decreased from 2066 to 677 while the tanδ value increased from 2.9% to 3.5% as the x value increased from 0.05 to 0.2. The TCC at the cold end (−55 °C) decreased slightly from −20.8% to −24.6%, whereas the TCC at the hot end (200 °C) significantly increased from 27.0% to 641.0%, which was simply due to the increase in Tc. Because a high dielectric constant, a low tanδ value, and a stable TCC areneeded for X9R materials, the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic was selected as the host material to incorporate the Ta2O5 and Nb2O5 additives in this study.
Figure 1a shows the sintered densities of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with 1.0%, 1.5%, and 2.0% Ta2O5 sintered at different temperatures. In all cases, the sintered density increased the sintering temperature, and the maximum densification was reached at 1215 °C. With further increases in sintering temperature, the sintered density gradually declined because of the trapped porosities associated with the fast grain growth and partly the evaporation of the volatile species such as bismuth and sodium. The weight losses of the Ta2O5-doped 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics during sintering were found to be less than 1.5 wt% through careful monitoring in this study, suggesting that the deviations of the sintered compositions from the nominal composition were only slight. The sintered density of the ceramics decreased slightly with an increase in Ta2O5 content, though Ta has a high molecular weight than the substituted Ti. For instance, the maximum densities of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with 1.0%, 1.5%, and 2.0% Ta2O5 were 5.82, 5.81, and 5.79 g/cm3, respectively. The results imply that the addition of Ta2O5 inhibits the sintering of the host material. As shown in Figure 1b, the sintered densities of 0.9BaTiO3–0.1Bi0.5Na0.5)TiO3 ceramics with 1.0%, 1.5%, and 2.0% Nb2O5 dopant showed a trend similar to those with added Ta2O5. Apparently, both Ta2O5 and Nb2O5 additives slightly obstructed the densification of the 0.9BaTi O3–0.1(Bi0.5Na0.5)TiO3 ceramic.
Table 1. Dielectric properties of (1 − x)BaTiO3x(Bi0.5Na0.5)TiO3 ceramics at their maximum sintered densities.
Table 1. Dielectric properties of (1 − x)BaTiO3x(Bi0.5Na0.5)TiO3 ceramics at their maximum sintered densities.
Starting CompositionSintering Condition (°C/h)Sintered Density (g/cm3)Tc (°C)K (25 °C)tanδ (25 °C)TCCResistivity * (Ω·cm)
−55 °C200 °C25 °C200 °C
0.95BaTiO3–0.05(Bi0.5Na0.5)TiO31230/25.7814520662.9%−20.827.04.4 × 10129.2 × 1010
0.9BaTiO3–0.1(Bi0.5Na0.5)TiO31215/25.7817010443.4%−26.2202.07.2 × 10121.4 × 1011
0.8BaTiO3–0.2(Bi0.5Na0.5)TiO31215/25.82>2006773.5%−24.6641.01.9 × 10135.1 × 1010
* Resistivity of the samples was measured at a DC voltage of 500 V.
Figure 1. Apparent densities of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with various amounts of (a) Ta2O5 and (b) Nb2O5 sintered at different temperatures for 2 h.
Figure 1. Apparent densities of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with various amounts of (a) Ta2O5 and (b) Nb2O5 sintered at different temperatures for 2 h.
Applsci 05 01221 g001
Figure 2a shows the XRD patterns of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with various amounts of Ta2O5 sintered at 1215 °C. Only a major phase of tetragonal perovskite was found in the XRD patterns. Several small peaks at around 30° and 43° were due to the Ba6Ti17O40 phase, which was also observed in the pure 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic. The addition of Ta2O5 in the host material did not trigger any additional second phase, implying that the Ta5+ ions were substituted into the B-sites of the perovskite lattices and the excess Ti4+ ions caused the formation of the Ba6Ti17O40 phase. It was also found that the peaks corresponding to the perovskite phase shifted slightly to lower angles because of the ionic radius of Ta5+ (VI; 0.64 Å), which is slightly larger than that of Ti4+ (VI; 0.605 Å). As revealed in Figure 2b, for the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with various amounts of added Nb2O5 sintered at 1215 °C, the XRD patterns are quite similar to those of the sintered ceramics with Ta2O5. However, another second-phase Ba6Ti14Nb2O39 (orthorhombic phase, Bm21b; JCPDF 38-1431) was observed as the Nb2O5 content increased to 2 mol%, in addition to the Ba6Ti17O40 phase. The calculated lattice constants of the sintered ceramics with 2.0 mol% Ta2O5 and Nb2O5 were a = 3.981 Å and c = 4.068 Å and a = 3.981 Å and c = 4.031 Å, respectively. The larger variation in the lattice constants of the Ta2O5-doped ceramic than of the Nb2O5-doped ceramic is correlated with the greater solubility of the former dopant in the perovskite structure than the latter. The limited solubility of Nb2O5 led to the formation of Ba6Ti14Nb2O39 precipitates.
Figure 3 shows the microstructures of the sintered surfaces and fracture surfaces of the pure 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic and those with added Ta2O5 and Nb2O5 sintered at 1215 °C. All the microstructures of the sintered surfaces were composed of two types of grains: fine equiaxed grains and some enlarge rod-like grains. EDS analysis identified the former as the perovskite phase BaTiO3–(Bi0.5Na0.5)TiO3 and the latter as the Ba6Ti17O40 phase. The microstructures of the fracture surfaces indicated that all fractures were transgranular mode and revealed the presence of very little porosity. Figure 3a,b show the sintered surface and fracture surface of the pure 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic sintered at 1215 °C, which included fine equiaxed grains with an average grain size of 0.57 μm and rod-like grains with a size of approximately 3.2 μm, similar to those reported in the literature [14]. The microstructures of the sintered surfaces and fracture surfaces of the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with 1.0 and 2.0 mol% Ta2O5 sintered at 1215 °C are shown in Figure 3c,d, and in Figure 3e,f, respectively. Compared with the average grain size of the pure 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic, the average grain size decreased slightly with the addition of Ta2O5, while the number of intergranular pores appeared to increase. The average grain size of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 with 1.0% and 2.0% Ta2O5 was determined to be 0.55 and 0.46 μm, respectively. The Ta2O5 dopant caused a reduction in the grain boundary mobility, which led to a decrease in grain size and an increase in trapped pores. The microstructures of the sintered surfaces of the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with 1.0 mol% and 2.0 mol% Nb2O5 sintered at 1215 °C are shown in Figure 3g,h, respectively; the equiaxed grains had an average size of 0.51 and 0.50 μm, respectively. The Nb2O5-doped ceramics appeared to have greater porosity than the Ta2O5-doped ceramics. According to Figure 3i,j, which present the microstructures of the fracture surfaces, many fine second-phase precipitates of approximately 50 nm in size were distributed in the microstructure, which was determined to be Ba6Ti14Nb2O39 based on the XRD results shown in Figure 2b. The fine second-phase precipitates suppressed the grain growth because of the pinning effect, which dragged the grain boundaries during sintering.
Overall, it can be concluded that the solubility of Ta2O5 dopant is greater than that of Nb2O5 dopant in the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic. The addition of Ta2O5 resulted in the lattice expansion of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic and reduced the grain boundary mobility and thus slightly inhibited the densification. The average grain size slightly decreased and the number of trapped pores increased with increase of Ta2O5 content. In addition to the pre-existing Ba6Ti17O40 phase, the escalation in addition of Nb2O5 in 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic triggered the formation of Ba6Ti14Nb2O39 precipitates with increased quantity. The distributed second phase suppressed the sintered density and generated porosity in 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic.
Figure 2. XRD patterns of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with various amounts of (a) Ta2O5 and (b) Nb2O5 sintered at 1215 °C for 2 h.
Figure 2. XRD patterns of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with various amounts of (a) Ta2O5 and (b) Nb2O5 sintered at 1215 °C for 2 h.
Applsci 05 01221 g002aApplsci 05 01221 g002b
Figure 3. SEM micrographs of the sintered surfaces (a), (c), (e). (g), and (i) and the fractured surfaces (b), (d), (f), (h), and (j) of 0.9BaTiO3–0.1 (Bi0.5Na0.5)TiO3 ceramic sintered at 1215 °C for 2 h. (a) and (b) are pure 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic; (c) and (d) correspond to the ceramic with 1.0 mol% Ta2O5; (e) and (f) correspond to the ceramic with 2.0 mol% Ta2O5; (g) and (h) correspond to the ceramic with 1.0 mol% Nb2O5; and (i) and (j) correspond to the ceramic with 2.0 mol% Nb2O5.
Figure 3. SEM micrographs of the sintered surfaces (a), (c), (e). (g), and (i) and the fractured surfaces (b), (d), (f), (h), and (j) of 0.9BaTiO3–0.1 (Bi0.5Na0.5)TiO3 ceramic sintered at 1215 °C for 2 h. (a) and (b) are pure 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic; (c) and (d) correspond to the ceramic with 1.0 mol% Ta2O5; (e) and (f) correspond to the ceramic with 2.0 mol% Ta2O5; (g) and (h) correspond to the ceramic with 1.0 mol% Nb2O5; and (i) and (j) correspond to the ceramic with 2.0 mol% Nb2O5.
Applsci 05 01221 g003aApplsci 05 01221 g003b
Figure 4 and Figure 5 show the TCCs and dielectric losses versus temperature, measured at a frequency of 1 kHz, of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with various amounts of added Ta2O5 and Nb2O5 sintered at 1215 °C, respectively. Some of the dielectric properties, including the Curie point, dielectric constant, and tanδ value at 25 °C and TCCs at both 25 °C and 200 °C, are also listed in Table 2. The temperature dependence of the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics became flattened with an increase in the dopant contents of Ta2O5 and Nb2O5. The TCC curves near the Curie point were dramatically suppressed, and a broad and diffuse phase transition occurred in all samples, leading to smooth dielectric temperature characteristics over a relatively broad temperature range (−55 to 200 °C). The temperature corresponding to the maximum of the diffuse phase transition (Tm) increased with the Ta2O5 content, but there was no visible change with an increase in the Nb2O5 content. For instance, Tm corresponding to the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with 1.0, 1.5, and 2.0 mol% Ta2O5 content appeared to be around 145 °C, 150 °C, and 180 °C, respectively, while those with 1.0, 1.5 and 2.0 mol% Nb2O5 had a similar Tm of approximately 140 °C. It is apparent that the Ta2O5 content can effectively elevate the Tm of the sintered ceramics.
Figure 4. (a) TCCs and (b) dielectric losses versus temperature of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with various amounts of Ta2O5 sintered at 1215 °C (measured frequency = 1 kHz; heating rate = 2 °C/min; applied voltage = 1 V).
Figure 4. (a) TCCs and (b) dielectric losses versus temperature of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with various amounts of Ta2O5 sintered at 1215 °C (measured frequency = 1 kHz; heating rate = 2 °C/min; applied voltage = 1 V).
Applsci 05 01221 g004
Figure 5. (a) Temperature coefficients of capacitance and (b) dielectric losses versus temperature of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with addition of various amounts of Nb2O5 sintered at 1215 °C (measured frequency = 1 kHz; heating rate = 2 °C/min; applied voltage = 1 V).
Figure 5. (a) Temperature coefficients of capacitance and (b) dielectric losses versus temperature of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with addition of various amounts of Nb2O5 sintered at 1215 °C (measured frequency = 1 kHz; heating rate = 2 °C/min; applied voltage = 1 V).
Applsci 05 01221 g005
Table 2. Dielectric properties of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic with various amounts of Ta2O5 and Nb2O5 at their maximum sintered densities.
Table 2. Dielectric properties of 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic with various amounts of Ta2O5 and Nb2O5 at their maximum sintered densities.
AdditivesSintering Condition (°C/ h)Sintered Density (g/cm3)Tm (°C)K (25 °C)tanδ (25 °C)TCCResistivity * (Ω·cm)
−55 °C200 °C25 °C200 °C
1.0 mol% Ta2O51215/25.8214517883.6%−38.82.12.6 × 10121.3 × 1011
1.5 mol% Ta2O51215/25.8115018002.3%−16.1−6.43.5 × 10128.2 × 1010
2.0 mol% Ta2O51215/25.7918013861.8%−1.35.22.8 × 10121.5 × 1010
1.0 mol% Nb2O51215/25.7914014713.9%−37.715.79.0 × 10129.1 × 1010
1.5 mol% Nb2O51215/25.7714017903.1%−18.4−16.56.2 × 10125.5 × 1010
2.0 mol% Nb2O51215/25.7414018651.4%−9.4−19.87.8 × 10125.1 × 1010
* Resistivity of the samples was measured at a DC voltage of 500 V.
The TCCs of the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic with 1.0 mol% Ta2O5 at −55 and 200 °C were −38.8% and 2.1%, respectively, which failed to meet the EIA X9R specifications because of the large variation of capacitance between the hot and cold ends (See Figure 4a and Figure 5a). As the Ta2O5 dopant content in the sintered ceramic increased, the capacitance variation at both −55 °C and 200 °C became flattened. The TCCs of the sintered ceramics became X9R-qualified with 2.0 mol% Ta2O5. The TCCs at −55 °C and 200 °C for the sintered ceramics with 1.5 and 2.0 mol% Ta2O5 were, respectively, −16.1% and −6.4% and −1.3% and 5.2%. The dielectric constant of the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic tended to decrease with the Ta2O5 content and appeared to be 1788, 1800, and 1386 for 1.0, 1.5 and 2.0 mol% Ta2O5, respectively. However, the dielectric constant increased from 1471 to 1790 and 1865 as the Nb2O5 content rose from 1.0 to 1.5 and 2.0 mol%, respectively. Apparently, the dielectric constant increased with the addition of Nb2O5 dopant, even though the differences in their microstructures, such as grain size and porosity, seemed to be trivial.
Capacitors for high-temperature applications must possess a low dielectric loss at both room and high temperatures to prevent overheating and failure. As shown in Figure 4b and Figure 5b, all samples revealed a similar trend in their dielectric-loss curves. The dielectric losses of the sintered 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with Ta2O5 and Nb2O5 continued to drop with the measurement temperature. For instance, the dielectric loss of the sintered ceramic with 2.0 mol% Ta2O5 was 3.3%, 1.8%, and 1.7% at temperatures of −55, 25, and 200 °C, respectively. The dielectric loss did not increase at high temperature, unlike that of most ferroelectric ceramics, where the dielectric loss increases because of the increase in ionic mobility or a ferroelectric-paraelectric transition at high temperatures. The dielectric losses of the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic were found to decrease significantly with the addition of Ta2O5 and Nb2O5. For instance, it decreased from 3.6% to 1.8% and from 3.9% to 1.4% at 25 °C when the Ta2O5 and Nb2O5 dopant content increased from 1.0 to 2.0 mol%, respectively, which is close to the values for common X7R and X8R ceramics with a similar thickness. The electrical resistivities of the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramics with added Ta2O5 and Nb2O5 obtained at 25 °C and 200 °C are also listed in Table 1. All samples appeared to be good insulators, with resistivities ranging from 2.6 × 1012 Ω·cm to 9.0 × 1012 Ω·cm at 25 °C and from 1.5 × 1010 Ω·cm to 1.3 × 1011 Ω·cm at 200 °C. There was a decrease of approximately two orders of magnitude in resistivity as the temperature rose from 25 °C to 200 °C.
In this study, it was found that Ta2O5 dopant was more efficient to increase the temperature stability of dielectric constant of the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic, compared with Nb2O5 dopant. Particularly, the host ceramic with 2 mol% Ta2O5 dopant hadthe best composition in terms of dielectric properties. The X9R formulations and their dielectric properties were highly repeatable in this study, which was not the case with the lead-based X9R ceramics reported in the literature [10].

4. Conclusions

BaTiO3–(Bi0.5Na0.5)TiO3 ceramics with various amounts of Ta2O5 and Nb2O5 dopants were explored in the study to comparetheir high-temperature stability for meeting X9R specifications. Addition of Ta2O5 and Nb2O5 dopants resulted in areduction in grain size of the sintered ceramics. The microstructure of the former contained fine equiaxed perovskite grains and some rod-like grains of second-phase Ba6Ti17O40, while that of the latter also included Ba6Ti14Nb2O39 precipitates. The best dielectric properties of the 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 ceramic with 2 mol% Ta2O5 dopant sintered at 1215 °C: a dielectric constant of 1386, a tanδ of 1.8%, and TCCs of −1.3% and 1.2% and electrical resistivities of 2.8 × 1012 and 1.5 × 1010 Ω·cm, respectively, at 25 and 200 °C. The X9R formulations and their dielectric properties were shown to be highly repeatable in this study.

Author Contributions

Sea-Fue Wang and Yung-Fu Hsu conceived and designed the experiments; Yu-Wen Hung and Yi-Xin Liu performed the experiments and analyzed the data; and Sea-Fue Wang wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, S.F.; Dayton, G.O. Dielectric Properties of Fine-Grain BaTiO3 Based X7R Dielectric. J. Am. Ceram. Soc. 1999, 82, 2677–2682. [Google Scholar] [CrossRef]
  2. Tang, B.; Zhang, S.R.; Yuan, Y.; Zhou, X.; Liang, Y. Influence of CaZrO3 on Dielectric properties and Microstructures of BaTiO3-Based X8R Ceramics. Sci. China Ser. E Technol. Sci. 2008, 51, 1451–1456. [Google Scholar] [CrossRef]
  3. Tang, B.; Zhang, S.; Zhou, X.; Wang, D.; Yuan, Y. Regression Analysis for Complex Doping of X8R Ceramics Based on Uniform Design. J. Electron. Mater. 2007, 36, 1383–1388. [Google Scholar] [CrossRef]
  4. Sato, S.; Fujikawa, Y.; Nagai, A.; Terada, Y.; Nomura, T. Effect of Rare-Earth Doping on the Temperature-Capacitance Characteristics of MLCCs with Ni electrodes. J. Ceram. Soc. Jpn. 2004, 112, S481–S485. [Google Scholar]
  5. Yuan, Y.; Zhou, X.H.; Li, B.; Zhang, S.R. Effects of BiNbO4 and Nb2O5 Additions on the Temperature Stability of Modified BaTiO3. Ceram. Silik. 2010, 54, 258–262. [Google Scholar]
  6. Yuan, Y.; Zhou, X.H.; Zhao, C.J.; Li, B.; Zhang, S.R. High-Temperature Capacitor Based on Ca-Doped Bi0.5Na0.5TiO3–BaTiO3 Ceramics. J. Electron. Mater. 2010, 39, 2471–2475. [Google Scholar] [CrossRef]
  7. Tang, B.; Zhang, S.R.; Zhou, X.H.; Yuan, Y.; Yang, L.B. Preparation and Modification of High Curie Point BaTiO3-Based X9R Ceramics. J. Electroceram. 2010, 5, 93–97. [Google Scholar] [CrossRef]
  8. Li, L.X.; Han, Y.M.; Zhang, P.; Ming, C.; Wei, X. Synthesis and Characterization of BaTiO3-Based X9R Ceramics. J. Mater Sci. 2009, 44, 5563–5568. [Google Scholar] [CrossRef]
  9. Yuan, Y.; Li, E.; Tang, B.; Bo, L.; Zhou, X. Effects of ZnO and CeO2 Additions on the Microstructure and Dielectric properties of Mn-Modified (Bi0.5Na0.5)0.88Ca0.12TiO3 Ceramics. J. Mater. Sci. Mater. Electron. 2012, 23, 309–314. [Google Scholar] [CrossRef]
  10. Gao, S.; Wu, S.; Zhang, Y.; Yang, H.; Wang, X. Study on the Microstructure and Dielectric properties of X9R Ceramics Based on BaTiO3. Mater. Sci. Eng. B 2011, 176, 68–71. [Google Scholar] [CrossRef]
  11. Yuan, Y.; Zhang, S.R.; Zhou, X.H.; Tang, B.; Li, B. High-Temperature Capacitor Materials Based on Modified BaTiO3. J. Electron. Mater. 2009, 38, 706–710. [Google Scholar] [CrossRef]
  12. Li, L.; Guo, D.; Xia, W.; Liao, Q.; Han, Y.; Peng, Y. An Ultra-Broad Working Temperature Dielectric Material of BaTiO3-Based Ceramics with Nd2O3 Addition. Am. Ceram. Soc. 2012, 95, 2107–2109. [Google Scholar] [CrossRef]
  13. Zhang, B.; Zhang, B.; Wang, N.; Zhang, J. Low temperature sintering of lead-free BaTiO3-based X9R ceramics with Bi2O3 dopant and assisted by LiF-CaF2 Flux Agent. Adv. Mater. Res. 2014, 906, 31–36. [Google Scholar] [CrossRef]
  14. Yao, G.; Wang, X.; Wu, Y.; Li, L.T. Nb-Doped 0.9BaTiO3–0.1(Bi0.5Na0.5)TiO3 Ceramics with Stable Dielectric Properties at High Temperature. J. Am. Ceram. Soc. 2012, 95, 614–618. [Google Scholar] [CrossRef]
  15. Sun, Y.; Liu, H.; Hao, H.; Zhang, L.; Zhang, S. The role of Co in the BaTiO3–Na0.5Bi0.5TiO3 based X9R ceramics. Ceram. Int. 2015, 41, 931–939. [Google Scholar] [CrossRef]
  16. Wang, T.; Hao, H.; Liu, M.; Zhou, D.; Yao, Z.; Cao, M.; Liu, H. X9R BaTiO3-Based Dielectric Ceramics with Multilayer Core–Shell Structure Produced by Polymer-Network Gel Coating Method. J. Am. Ceram. Soc. 2015, 98, 690–693. [Google Scholar] [CrossRef]
  17. Xu, Q.; Huang, D.P.; Chen, M.; Chen, W.; Liu, H.X.; Kim, B.H. Effect of Bismuth excess on Ferroelectric and Piezoelectric Properties of a (Na0.5Bi0.5)TiO3–BaTiO3 Composition near the Morphotropic Phase Boundary. J. Alloys Compd. 2009, 471, 310–316. [Google Scholar] [CrossRef]
  18. Takeda, H.; Aoto, W.; Shiosaki, T. BaTiO3–(Bi1/2Na1/2)TiO3 Solid-Solution Semiconducting Ceramics with Tc > 130 °C. Appl. Phys. Lett. 2005, 87, 102–104. [Google Scholar]
  19. Gao, L.; Huang, Y.; Liu, L.; Lin, T.; Liu, C.; Zhou, F.; Wan, X. Crystal Structure and Properties of BaTiO3–(Bi0.5Na0.5)TiO3 Ceramic System. J. Mater. Sci. 2008, 43, 6267–6271. [Google Scholar] [CrossRef]
  20. Raengthon, N.; Brown-Shaklee, H.J.; Brennecka, G.L.; Cann, D.P. Dielectric properties of BaTiO3–Bi(Zn1/2Ti1/2)O3–NaNbO3 solid solutions. J. Mater. Sci. 2013, 48, 2245–2250. [Google Scholar] [CrossRef] [Green Version]
  21. Chu, B.J.; Chen, D.R.; Li, G.R.; Yin, Q.R. Electrical Properties of Na1/2Bi1/2TiO3–BaTiO3 Ceramics. J. Eur. Ceram. Soc. 2002, 22, 2115–2121. [Google Scholar] [CrossRef]
  22. Wang, S.F.; Hsu, Y.F.; Li, J.H.; Lai, Y.C.; Chen, M.H. Dielectric Properties and Microstructures of Non-reducible High-Temperature Stable X9R Ceramics. J. Eur. Ceram. Soc. 2013, 33, 1793–1799. [Google Scholar] [CrossRef]
  23. Hou, J.; Kumar, R.V.; Qub, Y.; Krsmanovic, D. B-site doping effect on electrical properties of Bi4Ti3 − 2xNbxTaxO12 ceramics. Scr. Mater. 2009, 61, 664–667. [Google Scholar] [CrossRef]
  24. Islam1, M.F.; Mahbub, R.; Mousharraf, A. Effect of Sintering Parameters and Ta2O5 Doping on the Microstructure and Dielectric Properties of BaTiO3 Based Ceramics. Key Eng. Mater. 2014, 608, 247–252. [Google Scholar] [CrossRef]
  25. Kaneko, Y.; Azough, F.; Kida, T.; Ito, K.; Shimada, T. (Ba1 + xTiO3)–(Bi0.5Na0.5TiO3) Lead-Free, Positive Temperature Coefficient of Resistivity Ceramics: PTC Behavior and Atomic Level Microstructures. J. Am. Ceram. Soc. 2012, 95, 3928–3934. [Google Scholar] [CrossRef]
  26. Datta, K.; Roleder, K.; Thomas, P.A. Enhanced tetragonality in lead-free piezoelectric (1 − x)BaTiO3xNa1/2Bi1/2TiO3 solid solutions where x = 0.05–0.40. J. Appl. Phys. 2009. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Wang, S.-F.; Hsu, Y.-F.; Hung, Y.-W.; Liu, Y.-X. Effect of Ta2O5 and Nb2O5 Dopants on the Stable Dielectric Properties of BaTiO3–(Bi0.5Na0.5)TiO3-Based Materials. Appl. Sci. 2015, 5, 1221-1234. https://doi.org/10.3390/app5041221

AMA Style

Wang S-F, Hsu Y-F, Hung Y-W, Liu Y-X. Effect of Ta2O5 and Nb2O5 Dopants on the Stable Dielectric Properties of BaTiO3–(Bi0.5Na0.5)TiO3-Based Materials. Applied Sciences. 2015; 5(4):1221-1234. https://doi.org/10.3390/app5041221

Chicago/Turabian Style

Wang, Sea-Fue, Yung-Fu Hsu, Yu-Wen Hung, and Yi-Xin Liu. 2015. "Effect of Ta2O5 and Nb2O5 Dopants on the Stable Dielectric Properties of BaTiO3–(Bi0.5Na0.5)TiO3-Based Materials" Applied Sciences 5, no. 4: 1221-1234. https://doi.org/10.3390/app5041221

Article Metrics

Back to TopTop