Next Article in Journal
Development and Validation of a Reversed-Phase HPLC Method with UV Detection for the Determination of L-Dopa in Vicia faba L. Broad Beans
Next Article in Special Issue
Unravelling the 2e ORR Activity Induced by Distance Effect on Main-Group Metal InN4 Surface Based on First Principles
Previous Article in Journal
Synthesis of New 2,3-Dihydroindole Derivatives and Evaluation of Their Melatonin Receptor Binding Affinity
Previous Article in Special Issue
Tailoring Electrochemical Performance of Co3O4 Electrode Materials by Mn Doping
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimization of Sintering Conditions to Enhance the Dielectric Performance of Gd3+ and Ho3+ Codoped BaTiO3 Ceramics

1
State Key Lab of Inorganic Syntheses and Preparative Chemistry, College of Chemistry, Jilin University, Changchun 130012, China
2
Key Laboratory for Special Functional Materials at Jilin Provincial Universities, Jilin Institute of Chemical Technology, Jilin 132022, China
3
Innovation Center for Chemical Science, College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou 215123, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(21), 7464; https://doi.org/10.3390/molecules27217464
Submission received: 23 September 2022 / Revised: 30 October 2022 / Accepted: 31 October 2022 / Published: 2 November 2022
(This article belongs to the Special Issue Energy-Relevant Advanced Materials)

Abstract

:
BaTiO3 dielectric capacitors, one of the important energy storage devices, play critical roles in storing electricity from renewable energies of water, wind, solar, etc. The synthesis of BaTiO3 ceramics with weak temperature dependence and a high dielectric constant at room temperature (εRT′) is an urgent problem to meet the miniaturization and large capacity of dielectric capacitors. Doping rare earth elements into BaTiO3 can solve this problem, but it is still challenging. In this work, we adopt a synergistic strategy of increasing εRT′ and improving the temperature stability by codoping Gd3+ and Ho3+, respectively, to address this challenge. By carefully adjusting the synthesis conditions in the solid-state reaction, codoping 7% Gd3+ and 7% Ho3+ in BaTiO3 (BGTH7) ceramics were synthesized. The temperature-dependent dielectric constant reveals that the obtained optimal BGTH7 ceramic satisfies the X7U specification and displays a stable ε′ in the temperature range of −55~125 °C. The optimal BGTH7 ceramic after sintering at 1400 °C for 6 h exhibits a high dielectric constant of 5475 and low dielectric loss (tan δ) of 0.0176, hitherto exhibiting the best performance in X7U ceramics. The findings in this work are conducive to the miniaturization and stabilization of dielectric energy storage devices.

1. Introduction

The limited amount of fossil fuels raise much attention about the development of renewable resources [1,2]. The conversion of renewable energy sources into electrical energy and efficient power storage are essential to solve the energy crisis [3,4]. Commonly used electrical energy storage devices are lithium batteries, supercapacitors, and dielectric capacitors. The low release power of lithium batteries and the instability of supercapacitors restrict their application in energy storage [5]. Among the dielectric energy storage materials, lead-free BaTiO3 dielectric capacitors are attracting attention for overcoming the above-mentioned disadvantages, exhibiting ultra-fast charge/discharge capability and high temperature tolerance [6,7].
The capacitance (C) of dielectric capacitors is governed by the dielectric constant (ε′), capacitance plate area (S), and distance between two pole plates (d), as shown in the formulas below:
C = ε S / 4 π k d
k is the electrostatic force constant (k = 8.987551 × 109 N·m2/C2). BaTiO3 dielectric ceramics are a key component of dielectric energy storage capacitors. However, the unstable and low capacity of BaTiO3 dielectric capacitors cannot meet the demand for miniaturization and the high performance of dielectric capacitors [8]. Therefore, obtaining a stable and high dielectric constant near the room temperature (εRT) for BaTiO3 ceramics can meet the requirements of miniaturization and high performance of dielectric capacitors, which are more widely used in miniaturized and integrated circuit systems [9,10].
In previous studies, significant efforts have been made to address the issues mentioned above. Tetragonal BaTiO3 has large spontaneous polarization and high ε′, eventually increasing the energy density [11,12]. Ma et al. prepared BaTiO3-La2O3-SiO2 particles. More La3+ and Si4+ ions incorporated into the BaTiO3 lattice with increasing SiO2 enhanced spontaneous polarization and resulted in increased energy density [13].
During the tetragonal to the cubic transition of BaTiO3 at 120 °C, ε′ increases sharply, forming a dielectric peak [14]. Shifting the dielectric peak temperature (Tm) to room temperature contributes to high εRT. Gd-doped BaTiO3 can significantly reduce the Tm and increase εRT′ [15,16]. Gd-doped 0.7BaTiO3-0.3Sr0.2Bi0.7TiO3 ceramics possess high εRT (εRT ≈ 4000) and Tm near −19 °C, exhibiting excellent energy storage. The difference in the valence of ions at the A-site induces a decreased tolerance factor, resulting in a decrease in Tm [17].
Moreover, the incorporation of Ho in BaTiO3 can broaden the dielectric peak and obtain stable dielectric constant [18]. In recent research reports, Hu [19] successfully synthesized Bi2O3 and Ho2O3 codoped BaTiO3-based ceramics by a solid-phase method, with a stable ε′ between 2400 and 2600 and satisfactory temperature stability of capacitance with an X8R dielectric specification. La-, Pr-, Nd-, Sm-, and Ho- codoped BaTiO3 all had wide dielectric peaks, satisfying the dielectric specifications of X5T, X6T, X7T, and X8S, respectively [20].
Although the dielectric performance was enhanced using the method reported above for BaTiO3, doping Gd or Ho alone is not sufficient to increase εRT and stable ε′ in the temperature range of operation. The dielectric performance of doped ceramics is also dependent on sintering temperature (Ts) and dwell time (td). Therefore, we propose a synergistic strategy codoping 7% Gd3+ and 7% Ho3+ in BaTiO3 (BGTH7), carefully adjusting the synthesis conditions by the solid-phase method to obtain high εRT and stable ε′, achieving stable and high capacities for BaTiO3 dielectric capacitors. The specific roadmap was shown in Figure 1.
In this work, the phase structure of BGTH7 ceramics was studied by X-ray diffraction and Raman spectroscopy. The improved performance with dielectric diffuse phase transition behavior can be obtained from dielectric measurement. The shift of Tm was caused by the charge-coupled substitution of Gd3+ and Ho3+ at both Ba and Ti sites, which was confirmed by X-ray photoelectron spectroscopy (XPS), electron paramagnetic resonance (EPR), and Raman spectrum. The scanning electric microscope (SEM) images show that optimizing Ts can improve the density and average grain size of the ceramic samples. Eventually, high εRT and stable ε′ at work temperature can be obtained, which is promising for miniaturization and high performance of devices.

2. Results and Discussion

2.1. Effect of Sintering Temperature and Dwell Time on Phase Purity and Morphology

Gd3+ and Ho3+ codoped BaTiO3 ceramics were synthesized by a solid-state method to achieve a high εRT and stable ε′ in the operating temperature range. Usually, the sintering temperature (Ts) and dwell time (td) greatly affect the phase composition in solid-phase reactions. Thus, we systematically adjusted Ts and td to optimize the sintering conditions (Figure 2a). Two BGTH7 ceramics after sintering at Ts = 1200 and 1300 °C contained impurity phases Ba12Ho4.67Ti8O35 (PDF # 43−0420) and Ho2O3 (PDF # 43−1018) (marked with * and # in Figure 2b, respectively). Sintering at low temperatures makes it difficult to incorporate Ho3+ into the perovskite structure owing to the large radius of Ho3+ (0.901 Å) in octahedral coordination compared to that of Ti4+ (0.605 Å) [21]. And as shown in Figure S1, ceramics sintered at 1200 and 1300 °C exhibited low density and small grain size. High Ts can change the boundary of the phase diagram in multiple composition compounds but also increase the solution limit of doped ions in perovskite [22]. Thus, at Ts = 1400 °C, BGTH7 ceramics exhibited a pure phase of perovskite.
The phase structure of BGTH7 ceramics sintered at 1400 °C with different td was examined by XRD. BGTH7 ceramics sintered for different td (1, 3, 6, 12 h) are pure phases, as shown in Figure 2c. As td increased, the XRD peaks gradually narrowed, and their intensities were enhanced, demonstrating an improvement in crystallinity. To accurately determine the shift of the peak, the internal standard KCl was added to the BGTH7 ceramic powder. The enlarged peak at approximately 38° slightly shifted toward a higher 2θ angle as td increased, confirming a slight contraction in unit cell volume (V0). Notably, the diffraction peak (200) at approximately 45° for all ceramics did not show any splitting, which may be attributed to the cubic structure, similar to cubic BaTiO3 (PDF # 31−0174). Certainly, a slight distortion in the tetragonal phase structure or broadening of the XRD peaks can also obscure such a splitting, and this obscuring effect is difficult to detect by XRD. Therefore, in addition to XRD data, spectroscopic evidence could be helpful to prove the accurate phase structure of BGTH7 ceramics sintered at 1400 °C.
Raman spectroscopy can effectively distinguish the tetragonal or cubic phase of perovskite. Ideal cubic perovskite with Pm 3 m does not give any Raman signal due to a centrosymmetric structure belonging to the O1h space group [23]. Once the BaTiO3 structure deviates from cubic symmetry, such as tetragonal with P4mm, four degenerate 3F1u (IR) + F2u (inactive) modes for the O1h space group split into eight Raman active 3A1 + B1 + 4E modes [24]. Thus, the Raman signal can be observed.
As shown in Figure 3a, one peak at approximately 310 cm−1, usually from the B1 and E modes, was seen in the Raman spectra of sintered BGTH7 ceramics (td = 1 3, 6, 12 h), indicating that BGTH7 ceramics sintered at 1400 °C had a tetragonal structure [25,26]. Different from the sharp peak for the ceramics that underwent long dwell times, the peak at 310 cm−1 was not obvious for the BGTH7 ceramic sintered at td = 1 h due to incomplete grain growth. With these observations in mind, we refined the XRD data of BGTH7 ceramics sintered at 1400 °C by the tetragonal phase structure. The obtained lattice parameters in Table S1 show that the values of a and c are very close (a = bc, α = β = γ = 90°), i.e., the structure of BGTH7 ceramics sintered at 1400 °C slightly deviates from cubic symmetry. On the other hand, the Raman peak at approximately 520 cm−1 in Figure 3a corresponds to distorted (TiO6)4− octahedra, confirming the presence of a small structure distortion. A similar distortion was also observed in previous reports [27,28]. The tetragonal phase structure of BGTH7 ceramics sintered at 1400 °C can have large spontaneous polarization and a high ε′, which is desirable for BaTiO3 dielectric energy storage capacitors [29,30].
The dwell time greatly affected the morphology of intact disc BGTH7 ceramics sintered at 1400 °C. According to the grain size analysis in Figure S2, the average grain sizes of the BGTH7 ceramics were approximately 0.71, 1.15, 2.11, and 3.17 μm for td = 1, 3, 6, and 12 h, respectively. As illustrated by the SEM images in the inset of Figure 3b–e, the BGTH7 ceramic after dwelling for td = 1 h exhibited tiny grains and pores, showing incomplete grain growth. The grain size was improved by prolonging the dwell time. These gradually increased grains filled the pores, resulting in a denser microstructure. The relative density (ρr) of the BGTH7 ceramics was 73, 79, 86, and 93%, sintered at 1400 °C for td = 1, 3, 6, and 12 h, respectively, in which ρr was estimated using the following formula:
ρ r = ρ a / ρ 0
ρ0 is the theoretical density obtained from Rietveld refinement, and ρa is the actual bulk density that can be measured using the Archimedes method.

2.2. Effect of Dwell Time on the Site Occupation of BGTH7 Ceramics Sintered at 1400 °C

Rare earth ions, such as Gd3+ and Ho3+, have different ion radii than Ba2+ and Ti4+, which often leads to their preferential occupation once they are doped in BaTiO3. Such a preferential occupation has an important influence on the dielectric property. In addition to the ionic radius, some other factors can also change the occupation of rare earth ions in the perovskite, such as the valence state of Ti and oxygen vacancies, which can be determined by XPS, EPR, and photoluminescence analysis.
The valence state of Ti ions was investigated by XPS. Figure 4a and Figure S3 shows the Ti 2p core level spectra of BGTH7 sintered at Ts = 1400 °C. Well spin-orbital splitting peaks Ti2p3/2 (at approximately 457.8 eV) and Ti2p1/2 (at approximately 463.6 eV) were observed, and the splitting energy of 2p3/2 and 2p1/2 was 5.8 eV, confirming the presence of Ti4+ ions [31,32]. Generally, the characteristic peak of Ti3+ 2p3/2 appears at approximately 455 eV [33]. Deconvolution analysis showed that there was the absence of a peak at 455 eV, proving the absence of Ti3+, i.e., only one oxidation state Ti4+ in the BGTH7 ceramics sintered at 1400 °C.
Due to the charge difference between dopant ions Gd3+/Ho3+ and host ions Ba2+ and Ti4+ in BaTiO3, the incorporation of Gd3+/Ho3+ could introduce a positive or negative charge center. These charge mismatches can be compensated by vacancies, described as follows by the Kröger–Vink notation:
Ln 2 O 3 BaTiO 3 2 Ln Ba + V Ba + 3 O O x
Ln 2 O 3 BaTiO 3 2 Ln Ti   + V O + 3 / 2 O 2 ( g )
Here, Ln represents Gd3+ or Ho3+. V Ba   and   V O usually displays EPR signals at g = 1.974 and 1.955 [34,35]. In Figure 4b, only a broad and symmetrical signal with g = 1.990 was observed for BGTH7 ceramics sintered at 1400 °C. This signal is attributed to Gd3+ with an electron configuration of 4f7 (ground state is 8S7/2) [36]. The EPR spectrum of Gd3+ is highly dependent on the symmetry of ion coordination in Figure S4. New weak signals appeared with g = 5.953 and 2.451 at temperatures below 223 K in Figure 4c, which is associated with the change in Gd3+ symmetry from slightly distorted cubic coordination to intermediate coordination because of the increased structure distortion at a low temperature [37,38]. The absence of a signal related to V Ba   or   V O indicates the absence of vacancies. As is well known, performing EPR operations under ultrahigh vacuum and at low temperatures can increase the detection limit. Variable temperature EPR spectroscopy was investigated for the BGTH7 ceramic sintered at 1400 °C for 6 h. As illustrated in Figure 4c, the vacancy signals were not activated when the temperature was changed through Tm (270.36 K). Therefore, the incorporation of Gd3+ and Ho3+ could adopt a charge-coupled substitution on both the Ba and Ti sites:
Ln 2 O 3 BaTiO 3 Ln Ba + Ln Ti + 3 O O x
Ho3+ ions in some perovskites exhibit a strong photoluminescence emission. The Raman spectrum of Ho3+ doped BaTiO3 under 532 nm excitation is often used to probe the site occupations of Ho3+ ions because the concentration of Ho3+ ions on the Ba sites is highly dependent on the transition intensities of 5F4/5S25I8 at 545 nm, 5F55I8 at 653 nm, and 5F4/5S25I7 at 755 nm [21,39]. Figure 4d and Figure S5 display the Raman spectra of BGTH7 ceramics sintered at 1400 °C under 532 nm excitation. Taking the transition intensity at 755 nm as a comparison for the inset in Figure 4d, the concentration of Ho3+ at the Ba site can be estimated [17]. For BGTH7 ceramics with td values of 1, 3, 6, and 12 h, the contents of Ho3+ substituted for Ba2+ were 0.0059, 0.0050, 0.0022, and 0.0018, respectively. Increasing the dwell time at a high temperature (1400 °C) could reduce the incorporation of Ho3+ into the Ba site. The cross-site occupancy between Gd3+ and Ho3+ is described in Equation (5) and occurred in BGTH7 ceramics sintered at 1400 °C. Therefore, for BGTH7 ceramics with td values of 1, 3, 6, and 12 h, the contents of Ho3+ substituted for Ti4+ were 0.0641, 0.065, 0.0678, and 0.0682, respectively. From the ionic radius in Table S2, Gd3+ ions preferentially incorporated into the Ba site, while Ho3+ preferentially incorporated into the Ti site, forming a Gd Ba - Ho Ti complex. On the other hand, a small quantity of Ho3+ and Gd3+ change their occupation on the Ba site and Ti site to Ho Ba - Gd Ti according to photoluminescence analysis. When the sintering conditions were Ts = 1400 °C and td = 6 h, the real molecular formula of the obtained ceramic was (Ba0.93Gd0.0678Ho0.0022)(Ti0.93Ho0.0678Gd0.0022) O3, having more Gd Ba - Ho Ti complexes.

2.3. Dielectric Properties of BGTH7 Ceramics Sintered at 1400 °C

The Electronic Industry Association (EIA) classifies class II ceramic capacitors into two categories: capacitance-stable and high-capacitance ceramic capacitors. The most commonly used capacitance-stable capacitors are R-type capacitors. Over the operating temperatures, the capacitance change of R-type capacitors does not exceed 15%. However, these capacitors have relatively low dielectric constant, which are not suitable for miniaturized applications in electronic products. Another capacitor is the Y-type capacitor. Even though the Y-type capacitor has a high capacitance compared to the R-type capacitor, it is difficult to apply because of its unstable dielectric constant. A U-type capacitor combines the advantages of temperature stability and high capacitance, eventually having a stable and high dielectric constant over the entire operating temperature range [40,41,42,43].
Table S3 lists EIA classification of Class II capacitors. Figure 5a–d shows the temperature dependence of the ε′ and dielectric loss (tan δ) measured at 1 kHz for BGTH7 ceramics sintered at 1400 °C. According to Table S3, BGTH7 ceramics (td = 1, 3, 6, 12 h) satisfy the X8R, X7U, X7U, and X6U specifications, respectively. The dielectric properties for all BGTH7 ceramics are shown in Table S4. In particular, the BGTH7 ceramic sintered for 6 h at 1400 °C exhibits high εRT (εRT > 5000) and ultralow tan δ (tan δ < 0.02), with the location of the dielectric peak near the room temperature. This finding is very important for solving the problem of the location of the dielectric peak for conventional BaTiO3 ceramics being too high (close to 125 °C) to satisfy the requirement of the development and application of small electronics at room temperature.
ε′ and tan δ of BGTH7 ceramics sintered at 1400 °C with different td exhibited the same changes with frequency (Figure 5e,f). The value of ε′ in the high frequency region is lower than the low frequency region. The variation in ε′ with frequency can be interpreted by the Maxwell–Wagner’s two-layer models. The total polarization contributes to a high ε′ value at low frequency, while in the high-frequency region, dipole inversion fails to keep up with the changing electric field, causing ε′ to decrease.
Ceramics sintered at 1200 and 1300 °C exhibited low dielectric constant and high tan δ, as shown in Figures S6 and S7. Dielectric measurements demonstrated that the optimal BGTH7 ceramic sintered at Ts = 1400 °C and td = 6 h possesses high εRT (εRT = 5475) and low tan δ (tan δ = 0.0176) at 1 kHz, as well as enhanced frequency stability from 1 to 107 Hz. In addition, this BGTH7 ceramic satisfied the X7U specification according to Table S4, having stable ε′ in the operating temperature range (−55~125 °C). As shown Table S5 and Figure 5g, the BGTH7 ceramic is superior to the X7U specification reported previously, having potential application in microminiature and temperature-stable BaTiO3 dielectric energy storage capacitors.

2.4. Understanding the Excellent Dielectric Properties of BGTH7 Ceramics Sintered at 1400 °C

2.4.1. High Dielectric Constant at Room Temperature

εRT gradually increases from 1762 to 6580 as the average grain size changes from 0.71 to 3.17 μm (ρr from 73% to 93%, respectively), as shown in Figure S8. The reason for increasing εRT can be understood from the following two points: (1) Air acts as a stress snubber and reduces the entrapment forces between grains in the BaTiO3 ceramic, resulting in the formation of pores. Polarization is a critical factor in improving the dielectric constant. Normally, polarization is weakened by the depolarization field that occurs near the pores, resulting in decreasing dielectric constant [56]. High ρr shields the depolarization field, resulting in a high ε′ of ceramics. Among the four examined BGTH7 ceramics sintered at 1400 °C, the one sintered at td = 12 h, having ρr as high as 93%, gives the highest εRT of 6580. (2) An easier and more regular motion of the domain wall can also increase εRT [57,58]. Generally, large grains are conducive to forming orderly dipole alignment and promoting regular domain wall motion, thus increasing ε′. As the grain sizes increased, the dielectric peak was gradually enhanced, resulting in an increased dielectric constant (εRT) at room temperature. Therefore, a denser structure and large grain size are beneficial to obtaining high εRT in BGTH7 ceramics sintered at 1400 °C.

2.4.2. Shift of the Dielectric Peak

The Goldschmidt tolerance factor (t) can predict the symmetry of ABO3 perovskites that significantly affect their dielectric performance and is defined as the ratios of the constituent ionic radii of A, B, and O:
t = r A + r O   /   2   ( r B + r O )
For pure BaTiO3, rBa2+, rTi4+, and rO2− are 1.61 (in 12 coordinate), 0.605 (in octahedral coordination), and 1.4 Å; thus, t ≈ 1.06. A previous investigation has shown that the dielectric peak temperature (Tm) of BaTiO3 is highly dependent on the tolerance factor (t). Reducing t can shift the Tm to a lower temperature [59]. When some Gd3+ ions are incorporated into Ba sites, the decrease in t shifts Tm to a low temperature. Instead, the substitution of Ho3+ for Ti4+ ions leads to Tm moving toward higher temperatures, similar to the incorporation of Ca2+ [21]. In other words, the charge-coupled substitution of Gd3+ and Ho3+ at both Ba and Ti sites (Equation (5)) has different effects on Tm, which was caused by the different effects of substitution for Gd3+ and Ho3+ on Ti-O octahedral distortion [60]. The photoluminescence analysis in Figure 4d indicated that the Ho3+ content in the Ti site increases. Consequently, Tm first decreased and then increased, as shown in Figure S9.
All studied BGTH7 ceramics sintered at 1400 °C exhibit dielectric diffusion phase transition (DPT) behavior. The broadening dielectric peak was related to the grain size, which is consistent with numerous past reports [61,62]. In addition, Gd3+ and Ho3+ codoped BaTiO3 ceramics can also broaden the dielectric peak, similar to La-, Pr-, Nd-, Sm-, and Ho- codoped BaTiO3 ceramics [20]. Codoping of Ho3+ with Gd3+ has an important effect on the stability of ε′ over a wide temperature range.
The detailed mechanisms for improving dielectric properties is illustrated in Figure 6. Charge-coupled substitution of Gd3+ and Ho3+ at both Ba and Ti sites, decrease of tolerance factor, distortion of octahedral, as well as increased density can be optimized by adjusting sintering temperature, and time. Eventually, the optimal BGTH7 ceramic not only exhibited high εRT but also dielectric DPT, which can ensure stable operation for BaTiO3 dielectric energy storage capacitors.

3. Materials and Methods

BaCO3 (99.5%), TiO2 (99.5%), Ho2O3 (99.95%), and Gd2O3 (99.90%) were used as raw materials, weighed according to 7% Gd3+ and 7% Ho3+ codoping in BaTiO3 (BGTH7), and milled. After drying and calcining at 1100 °C for 5 h for decarburization, the obtained powder was reground and pressed into discs with a diameter of 12 mm and a thickness of 2 mm at 150 MPa using an aqueous solution of PVA (12% by mass) as a binder. The final sintering conditions to form the BGTH7 ceramics were chosen as heating up directly from room temperature to different Ts (1200 °C ≤ Ts ≤ 1400 °C, td = 12 h) and different td (1 h ≤ td ≤ 12 h, Ts = 1400 °C) in the air at a heating rate of 100 °C/h, a cooling rate of −200 °C/h to 700 °C, and then furnace cooling to room temperature.
The crystallographic structures were studied by using X-ray diffraction (XRD: DX-2700, Dandong Haoyuan, Dandong, China) with Cu Kα1 radiation (λ = 1.540562 Å). The microstructure of the sintered samples was observed using an EVOMA 10 scanning electric microscope (SEM: EVO MA10, Zeiss, Oberkochen, Germany) operated at 15 kV. The conducting Au atoms were sputtered on the specimen surface for SEM observations. XRD Rietveld refinements were conducted with the General Structure Analysis System (GSAS) program. The 532 nm and 785 nm lasers were used for excitation to obtain the Raman spectra (RS) of the ceramics and photoluminescence (PL) of Ho3+ using a LabRAM XploRA Raman spectrometer (Horiba Jobin Yvon, Longjumeau, France). X-ray photoelectron spectroscopy (XPS, ESCA-LAB250) measurements were performed to study the valence states of cations. The measurement curves of XPS data were fitted by a mixed Gaussian–Lorentzian function, and Shirley-type background subtraction was used. EPR measurements were carried out with an X-band (≈9.4 GHz) spectrometer (A300, Bruker, Rheinstetten, Germany) at 90–298 K. The gyromagnetic constant (g) was calculated using hv = 0H, where h is the Planck constant (h = 6.626 × 10−34 J·s), v is the microwave frequency, μ0 is the Bohr magnetron (μ0 = 9.262 × 10−24 J/T), and H is the magnetic field strength. The dielectric properties of the ceramic samples were measured on a broadband dielectric spectrometer (Concept 41, Novocontrol Technologies, Montabaur Germany) in a temperature range of 198–473 K and a frequency range of 1 Hz–10 MHz.

4. Conclusions

Converting renewable energy to electricity and efficient electricity storage are keys to addressing the energy crisis. BaTiO3 dielectric capacitors are critical energy storage devices due to their ultrafast charge/discharge ability, exceptional cycle life and high-temperature tolerance. However, the unstable and low dielectric constant (εRT = 1600) near the room temperature of BaTiO3 ceramics cannot meet the needs of miniaturization and the high capacity of dielectric capacitors. In this work, Gd3+ and Ho3+ codoped tetragonal BaTiO3 ceramics were prepared by carefully adjusting the synthesis conditions to address this issue. The optimal ceramic sintered at 1400 °C for 6 h had an ultrahigh room-temperature dielectric constant (εRT = 5475) with low dielectric loss (tan δ = 0.0176). In addition, the BGTH7 ceramic exhibits good X7U dielectric properties in the temperature range of −55~125 °C, which is superior to other X7U BaTiO3-based ceramics reported previously. The mechanisms for improving dielectric properties can be attributed to the tetragonal perovskite structure, large grain size, and the formation of the Gd Ba - Ho Ti complex. The high capacitance of this ceramic helps reduce the capacitor size, improve the efficiency, and enable miniature applications. The BGTH7 ceramic can be prepared as dielectric capacitors with high capacitance, is widely used in pulse power devices, hybrid automotive power supplies, and other fields, and will be beneficial to many portable electronic applications. In addition, the ceramic exhibits wide temperature range stability and is suitable for applications in extreme environments such as polar, high altitude, and underground tunnels.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27217464/s1, Figure S1: SEM images of the surface for BGTH7 ceramics with different Ts (a) Ts = 1200 °C, (b) Ts = 1300 °C, (c) Ts = 1400 °C for 12 h.; Figure S2: The grain size analysis of the BGTH7 ceramics of BGTH7 ceramics sintered for (a) 1 h (b) 3 h and (c) 12 h at 1400 °C; Figure S3: Ti 2p XPS spectra of BGTH7 ceramics sintered for (a) 1 h (b) 3 h (c) 12 h at 1400 °C; Figure S4: EPR spectra of BGTH7 ceramics sintered at different Ts for 12 h measured at room temperature; Figure S5: Raman spectra under 532 nm excitation of ceramics sintered at different Ts for 12 h measured at room temperature; Figure S6: Temperature dependence of the dielectric constant (ε′) and the dielectric loss (tan δ) for ceramics sintered at Ts = 1200 °C, 1300 °C and 1400 °C; Figure S7: Frequency dependence of the dielectric permittivity (ε′) and the dielectric loss (tan δ) for BGTH7 ceramics sintered at Ts = 1200 °C, 1300 °C and 1400 °C; Figure S8: ε’RT and tan δ at room temperature vary with grain size (GS) and relative density (ρr) for BGTH7 ceramics sintered at 1400 °C; Figure S9: Dielectric peak temperature varies with dwell time for BGTH7 ceramics sintered at 1400 °C; Table S1: Rietveld Refined lattice parameters of BGTH7 ceramics sintered at 1400 °C; Table S2: Ionic radii as a function of coordinate number (CN); Table S3: EIA Definition of Class II Capacitors; Table S4: Dielectric properties of all BGTH7 ceramics; Table S5: The dielectric properties of this work are compared with those of previously reported BaTiO3-based dielectric ceramics with X7U specifications.

Author Contributions

Conceptualization, J.B.; methodology, L.L., J.B. and Q.L.; investigation, J.B.; resources, L.L. and Q.L.; writing—original draft preparation, J.B.; writing—review and editing, J.B., L.L., X.L., Q.L. and X.W.; supervision, L.L. and X.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant Nos. 22175070 and 21771075), and the Jilin Provincial Department of Science and Technology (Grant No. 20200801034GH).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The first author expresses their deep gratefulness to Dayong Lu for his supervision of the design of the initial experimental process.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhou, M.; Liang, R.; Zhou, Z.; Dong, X. Novel BaTiO3-based lead-free ceramic capacitors featuring high energy storage density, high power density, and excellent stability. J. Mater. Chem. C 2018, 6, 8528–8537. [Google Scholar] [CrossRef]
  2. Wang, X.; Ren, P.; Wang, J.; Xu, J.; Xi, Y. Multi-phase coexistence and temperature-stable dielectric properties in BaTiO3/ZnO composite ceramics. J. Eur. Ceram. Soc. 2020, 40, 1896–1901. [Google Scholar] [CrossRef]
  3. Tawade, B.; Apata, I.; Pradhan, N.; Karim, A.; Raghavan, D. Recent Advances in the Synthesis of Polymer-Grafted Low-K and High-K Nanoparticles for Dielectric and Electronic Applications. Molecules 2021, 26, 2942. [Google Scholar] [CrossRef] [PubMed]
  4. Boonlakhorn, J.; Manyam, J.; Srepusharawoot, P.; Krongsuk, S.; Thongbai, P. Effects of charge compensation on colossal permittivity and electrical properties of grain boundary of CaCu3Ti4O12 ceramics Substituted by Al3+ and Ta5+/Nb5+. Molecules 2021, 26, 3294. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, C.; Li, F.; Ma, L.P. Advanced materials for energy storage. Adv. Mater. 2010, 22, E28–E62. [Google Scholar] [CrossRef]
  6. Zhou, H.Y.; Zhu, X.N.; Ren, G.R.; Chen, X.M. Enhanced energy storage density and its variation tendency in CaZrxTi1-xO3 ceramics. J. Alloy Compd. 2016, 688, 687–691. [Google Scholar] [CrossRef]
  7. Bokov, A.; Ye, Z.-G. Recent progress in relaxor ferroelectrics with perovskite structure. J. Mater. Sci. 2006, 41, 31–52. [Google Scholar] [CrossRef]
  8. Hong, K.; Lee, T.H.; Suh, J.M.; Yoon, S.-H.; Jang, H.W. Perspectives and challenges in multilayer ceramic capacitors for next generation electronics. J. Mater. Chem. C 2019, 7, 9782–9802. [Google Scholar] [CrossRef]
  9. Jiang, B.; Iocozzia, J.; Zhao, L.; Zhang, H.; Harn, Y.-W.; Chen, Y.; Lin, Z. Barium titanate at the nanoscale: Controlled synthesis and dielectric and ferroelectric properties. Chem. Soc. Rev. 2019, 48, 1194–1228. [Google Scholar] [CrossRef]
  10. Saini, P.; Arora, M.; Gupta, G.; Gupta, B.K.; Singh, V.N.; Choudhary, V. High permittivity polyaniline-barium titanate nanocomposites with excellent electromagnetic interference shielding response. Nanoscale 2013, 5, 4330–4336. [Google Scholar] [CrossRef]
  11. Jia, H.; Chen, J. Tailoring the tetragonal distortion to obtain high Curie temperature and large piezoelectric properties in BiFeO3-PbTiO3-BaTiO3 solid solutions. J. Eur. Ceram. Soc. 2020, 41, 2443–2449. [Google Scholar] [CrossRef]
  12. Xie, Y.; Yin, S.; Hashimoto, T.; Tokano, Y.; Sasaki, A.; Sato, T. Low temperature synthesis of tetragonal BaTiO3 by a novel composite-hydroxide-mediated approach and its dielectric properties. J. Eur. Ceram. Soc. 2010, 30, 699–704. [Google Scholar] [CrossRef]
  13. Ma, R.; Cui, B.; Shangguan, M.; Wang, S.; Wang, Y.; Chang, Z.; Wang, Y. A novel double-coating approach to prepare fine-grained BaTiO3@La2O3@SiO2 dielectric ceramics for energy storage application. J. Alloy. Compd. 2017, 690, 438–445. [Google Scholar] [CrossRef]
  14. Jin, S.H.; Lee, H.W.; Kim, N.W.; Lee, B.-W.; Lee, G.-G.; Hong, Y.-W.; Nam, W.H.; Lim, Y.S. Sonochemically activated solid-state synthesis of BaTiO3 powders. J. Eur. Ceram. Soc. 2021, 41, 4826–4834. [Google Scholar] [CrossRef]
  15. Ben, L.; Sinclair, D.C. Anomalous Curie temperature behavior of A-site Gd-doped BaTiO3 ceramics: The influence of strain. Appl. Phys. Lett. 2011, 98, 092907. [Google Scholar] [CrossRef]
  16. Han, F.; Bai, Y.; Qiao, L.-J.; Guo, D. A systematic modification of the large electrocaloric effect within a broad temperature range in rare-earth doped BaTiO3 ceramics. J. Mater. Chem. C 2016, 4, 1842–1849. [Google Scholar] [CrossRef]
  17. Huang, Q.; Si, F.; Tang, B. The effect of rare-earth oxides on the energy storage performances in BaTiO3 based ceramics. Ceram. Int. 2022, 48, 17359–17368. [Google Scholar] [CrossRef]
  18. Jeon, S.-C.; Kang, S.-J.L. Coherency strain enhanced dielectric-temperature property of rare-earth doped BaTiO3. Appl. Phys. Lett. 2013, 102, 112915. [Google Scholar] [CrossRef] [Green Version]
  19. Hu, W.; Chen, Z.; Lu, Z.; Wang, X.; Fu, X. Effect of Bi2O3 and Ho2O3 co-doping on the dielectric properties and temperature reliability of X8R BaTiO3-based ceramics. Ceram. Int. 2021, 47, 24982–24987. [Google Scholar] [CrossRef]
  20. Lu, D.Y.; Guan, D.X.; Li, H.B. Multiplicity of photoluminescence in Raman spectroscopy and defect chemistry of (Ba1−xRx)(Ti1−xHox)O3 (R = La, Pr, Nd, Sm) dielectric ceramics. Ceram. Int. 2018, 44, 1483–1492. [Google Scholar] [CrossRef]
  21. Lu, D.-Y.; Gao, X.-L.; Wang, S. Abnormal Curie-temperature shift in Ho-doped BaTiO3 ceramics with the self-compensation mode. Results Phys. 2018, 12, 585–591. [Google Scholar] [CrossRef]
  22. Yuan, L.; Hu, W.; Fang, S.; Li, G.; Wang, X.; Wu, X.; Li, L. CdO-CuO-TiO2 ternary dielectric systems: Subsolidus phase diagram and the effects of Cu segregation. J. Eur. Ceram. Soc. 2018, 38, 4978–4985. [Google Scholar] [CrossRef]
  23. Tsuji, K.; Ndayishimiye, A.; Lowum, S.; Floyd, R.; Wang, K.; Wetherington, M.; Maria, J.-P.; Randall, C.A. Single step densification of high permittivity BaTiO3 ceramics at 300 °C. J. Eur. Ceram. Soc. 2019, 40, 1280–1284. [Google Scholar] [CrossRef]
  24. Wang, H.; Li, G.; Li, L. Molten-salt-mediated synthesis and low-temperature electrical conduction of LnCoO3 (Ln=Pr, Nd, Sm, and Gd). J. Alloy Compd. 2014, 612, 227–232. [Google Scholar] [CrossRef]
  25. Wang, X.; Deng, X.; Wen, H.; Li, L. Phase transition and high dielectric constant of bulk dense nanograin barium titanate ceramics. Appl. Phys. Lett. 2006, 89, 162902. [Google Scholar] [CrossRef]
  26. Manika, G.C.; Andrikopoulos, K.S.; Psarras, G.C. On the ferroelectric to paraelectric structural transition of BaTiO3 micro-/nanoparticles and their epoxy nanocomposites. Molecules 2020, 25, 2686. [Google Scholar] [CrossRef] [PubMed]
  27. Naik, R.; Nazarko, J.J.; Flattery, C.S.; Venkateswaran, U.D.; Naik, V.M. Temperature dependence of the Raman spectra of polycrystalline Ba1-xSixTiO3. Phys. Rev. B 2000, 61, 11367–11372. [Google Scholar] [CrossRef]
  28. Lin, M.-F.; Thakur, V.K.; Tan, E.J.; Lee, P.S. Dopant induced hollow BaTiO3 nanostructures for application in high performance capacitors. J. Mater. Chem. 2011, 21, 16500–16504. [Google Scholar] [CrossRef]
  29. Petrovsky, V.; Petrovsky, T.; Kamlapurkar, S.; Dogan, F. Dielectric Constant of Barium Titanate Powders Near Curie Temperature. J. Am. Ceram. Soc. 2008, 91, 3590–3592. [Google Scholar] [CrossRef]
  30. Ma, N.; Zhang, B.P.; Yang, W.G.; Guo, D. Phase structure and nano-domain in high performance of BaTiO3 piezoelectric ceramics. J. Eur. Ceram. Soc. 2012, 32, 1059–1066. [Google Scholar] [CrossRef]
  31. Alkathy, M.S.; Hezam, A.; Manoja, K.; Wang, J.; Cheng, C.; Byrappa, K.; Raju, K.J. Effect of sintering temperature on structural, electrical, and ferroelectric properties of lanthanum and sodium co-substituted barium titanate ceramics. J. Alloy Compd. 2018, 762, 49–61. [Google Scholar] [CrossRef]
  32. Song, Y.; Wang, X.; Zhang, X.; Sui, Y.; Zhang, Y.; Liu, Z.; Lv, Z.; Wang, Y.; Xu, P.; Song, B. The contribution of doped-Al to the colossal permittivity properties of AlxNb0.03Ti0.97-xO2 rutile ceramics. J. Mater. Chem. C 2016, 4, 6798–6805. [Google Scholar] [CrossRef]
  33. Gao, B.; Xu, S.; Xu, Q. CO2 Induced Exposure of the Intrinsic Magnetic Surface of BaTiO3 to Give Room-Temperature Ferromagnetism. Angew. Chem. Int. Ed. Engl. 2022, 61, e202117084. [Google Scholar] [CrossRef]
  34. Lu, D.Y.; Liu, T.T. Dielectric properties and defect chemistry of (Ba1−xLax)(Ti1−xLux)O3 ceramics. J. Alloys Compd. 2017, 698, 967–976. [Google Scholar] [CrossRef]
  35. Kolodiazhnyi, T.; Petric, A. Analysis of point defects in polycrystalline BaTiO3 by electron paramagnetic resonance. J. Phys. Chem. Solids 2003, 64, 953–960. [Google Scholar] [CrossRef]
  36. Munteanu, A.-C.; Badea, M.; Olar, R.; Silvestro, L.; Dulea, C.; Negut, C.-D.; Uivarosi, V. Synthesis and Structural Investigation of New Bio-Relevant Complexes of Lanthanides with 5-Hydroxyflavone: DNA Binding and Protein Interaction Studies. Molecules 2016, 21, 1737. [Google Scholar] [CrossRef] [Green Version]
  37. Singh, V.; Singh, N.; Watanabe, S.; Rao, T.K.G.; Pathak, M.S.; Srivastava, A.K.; Singh, P.K.; Dhoble, S.J. Gadolinium-Activated CaZr4O9 Ultraviolet-B-Emitting Phosphor: A Luminescence and EPR Study. J. Electron. Mater. 2017, 46, 1943–1947. [Google Scholar] [CrossRef]
  38. Aminov, L.K.; Kurkin, I.N.; Malkin, B.Z. Superhyperfine structure in the EPR spectra and optical spectra of impurity f ions in dielectric crystals: A review. Phys. Solid State 2013, 55, 1343–1363. [Google Scholar] [CrossRef]
  39. Lu, D.-Y.; Guan, D.-X. Photoluminescence associated with the site occupations of Ho3+ ions in BaTiO3. Sci. Rep. 2017, 7, 6125. [Google Scholar] [CrossRef]
  40. Pan, M.J.; Randall, C.A. A brief introduction to ceramic capacitors. Electr. Insul. M. 2010, 26, 44. [Google Scholar] [CrossRef]
  41. Tavernor, A.W.; Lia, H.-P.S.; Stevens, R. Production and characterisation of composite relaxor ferroelectric multi-layer structures. J. Eur. Ceram. Soc. 1999, 19, 1859. [Google Scholar] [CrossRef]
  42. Furukawa, O.; Harata, M.; Imai, M.; Yamashita, Y.; Mukaeda, S. Low firing and high dielectric constant X7R ceramic dielectric for multilayer capacitors based on relaxor and barium titanate composite. J. Mater. Sci. 1991, 26, 5838. [Google Scholar] [CrossRef]
  43. Hino, T.; Matsumoto, N.; Nishida, M.; Araki, T. PLD of X7R for thin film capacitors. Appl. Surf. Sci. 2008, 254, 2638. [Google Scholar] [CrossRef]
  44. Lu, D.Y.; Peng, Y.Y.; Yu, X.Y.; Sun, X.Y. Dielectric properties and defect chemistry of La and Tb co-doped BaTiO3 ceramics. J. Alloys Compd. 2016, 681, 128–138. [Google Scholar] [CrossRef]
  45. Zhao, Y.; Wang, Q.; Lv, J.-H.; Zhao, X.; Wang, C.-M. Dielectric properties and electrocaloric effect of yttrium-modified BaTiO3 ceramics. Ceram. Int. 2021, 47, 18610–18618. [Google Scholar] [CrossRef]
  46. Lu, D.Y.; Liang, Y. Valence states and dielectric properties of fine-grained BaTiO3 ceramics co-doped with double valence-variable europium and chromium. Ceram. Int. 2018, 44, 14717–14727. [Google Scholar] [CrossRef]
  47. Liu, M.; Hao, H.; Zhen, Y.; Wang, T.; Zhou, D.; Liu, H.; Cao, M.; Yao, Z. Temperature stability of dielectric properties for xBiAlO3-(1−x)BaTiO3 ceramics. J. Eur. Ceram. Soc. 2015, 35, 2303–2311. [Google Scholar] [CrossRef]
  48. Turcan, I.; Lukacs, V.A.; Curecheriu, L.; Padurariu, L.; Ciomaga, C.E.; Airimioaei, M.; Stoian, G.; Lupu, N.; Mitoseriu, L. Microstructure and dielectric properties of Ag-BaTiO3 composite ceramics. J. Eur. Ceram. Soc. 2018, 38, 5420–5429. [Google Scholar] [CrossRef]
  49. Li, J.; Pu, Y.; Shi, Y.; Shi, R.; Wang, X.; Yang, M.; Wang, W.; Guo, X.; Peng, X. Dielectric, multiferroic and magnetodielectric properties of (1-x)BaTiO3-xSr2CoMoO6 solid solution. Ceram. Int. 2019, 45, 16353–16360. [Google Scholar] [CrossRef]
  50. Haily, E.; Bih, L.; El Bouari, A.; Lahmar, A.; Elmarssi, M.; Manoun, B. Effect of BaO-Bi2O3-P2O5 glass additive on structural, dielectric and energy storage properties of BaTiO3 ceramics. Mater. Chem. Phys. 2020, 241, 123034. [Google Scholar] [CrossRef]
  51. Fan, T.; Ji, C.; Chen, G.; Cai, W.; Gao, R.; Deng, X.; Wang, Z.; Fu, C. Enhanced the dielectric relaxation characteristics of BaTiO3 ceramic doped by BiFeO3 and synthesized by the microwave sintering method. Mater. Chem. Phys. 2020, 250, 123034. [Google Scholar] [CrossRef]
  52. Wan, Y.; Tang, L.; Dang, X.; Ren, P.; Ma, M.; Song, K.; Zhao, G. High temperature dielectrics based on Bi1/2Na1/2TiO3-BaTiO3-Sr0.53Ba0.47Nb2O6 ceramics with high dielectric permittivity and wide operational temperature range. Ceram. Int. 2019, 45, 2596–2601. [Google Scholar] [CrossRef]
  53. Chen, X.; Huang, G.; Ma, D.; Liu, G.; Zhou, H. High thermal stability and low dielectric loss of BaTiO3-Bi(Li1/3Zr2/3)O3 solid solution. Ceram. Int. 2017, 43, 926–929. [Google Scholar] [CrossRef]
  54. Ma, J.P.; Chen, X.M.; Ouyang, W.Q.; Wang, J.; Li, H.; Fang, J.-L. Microstructure, dielectric, and energy storage properties of BaTiO3 ceramics prepared via cold sintering. Ceram. Int. 2018, 44, 4436–4441. [Google Scholar] [CrossRef]
  55. Chen, W.; Hao, H.; Yang, Y.; Chen, C.; Appiah, M.; Yao, Z.; Cao, M.; Yu, Z.; Liu, H. Dielectric properties and impedance analysis of BaTiO3 -based ceramics with core-shell structure. Ceram. Int. 2017, 43, 8449–8458. [Google Scholar] [CrossRef]
  56. Meng, K.; Li, W.; Tang, X.G.; Liu, Q.X.; Jiang, Y.P. A review of a good binary ferroelectric ceramic: BaTiO3–BiFeO3. ACS Appl. Electron. Mater. 2021, 4, 2109–2145. [Google Scholar] [CrossRef]
  57. Ghayour, H.; Abdellahi, M. A brief review of the effect of grain size variation on the electrical properties of BaTiO3-based ceramics. Powder Technol. 2016, 292, 84–93. [Google Scholar] [CrossRef]
  58. Buscaglia, V.; Randall, C.A. Size and scaling effects in barium titanate. An overview. J. Eur. Ceram. Soc. 2020, 40, 3744–3758. [Google Scholar] [CrossRef]
  59. Freeman, C.L.; Dawson, J.A.; Harding, J.H.; Ben, L.B.; Sinclair, D.C. The Influence of A-Site Rare Earth Ion Size in Controlling the Curie Temperature of Ba1−xRExTi1−x/4O3. Adv. Funct. Mater. 2013, 23, 491–495. [Google Scholar] [CrossRef]
  60. Khedhri, M.H.; Abdelmoula, N.; Khemakhem, H.; Douali, R.; Dubois, F. Structural, spectroscopic and dielectric properties of Ca-doped BaTiO3. Appl. Phys. A 2019, 125, 1–13. [Google Scholar] [CrossRef]
  61. Martirenat, H.T.; Burfoot, J.C. Grain-size effects on properties of some ferroelectric ceramics. J. Phys. C Solid State Phys. 1974, 7, 3182–3192. [Google Scholar] [CrossRef] [Green Version]
  62. Zhang, Q.W.; Zhai, J.W.; Shen, B.; Zhang, H.J.; Yao, X. Grain size effects on dielectric properties of barium strontium titanate composite ceramics. Mater. Res. Bull. 2013, 48, 973–977. [Google Scholar] [CrossRef]
Figure 1. Target roadmap for high and stable capacitance achievement.
Figure 1. Target roadmap for high and stable capacitance achievement.
Molecules 27 07464 g001
Figure 2. (a) Schematic illustration of the synthesis of BGTH7 ceramic samples. (b) X-ray diffraction patterns of the BGTH7 ceramics sintered at different sintering temperatures (Ts) for 12 h. (c) X-ray diffraction patterns of BGTH7 ceramics sintered for different dwell times (td) at 1400 °C. The right panel shows the enlarged diffraction peaks of (111) and (200) in the 2θ ranges of 38–39.5° and 44–46°, respectively.
Figure 2. (a) Schematic illustration of the synthesis of BGTH7 ceramic samples. (b) X-ray diffraction patterns of the BGTH7 ceramics sintered at different sintering temperatures (Ts) for 12 h. (c) X-ray diffraction patterns of BGTH7 ceramics sintered for different dwell times (td) at 1400 °C. The right panel shows the enlarged diffraction peaks of (111) and (200) in the 2θ ranges of 38–39.5° and 44–46°, respectively.
Molecules 27 07464 g002
Figure 3. (a) Room-temperature Raman spectra in the low-wavenumber region of 100−1000 cm−1 for BGTH7 powder sintered for different td at Ts = 1400 °C. The results of the XRD refinement and SEM images for BGTH7 ceramics sintered with different td at Ts = 1400 °C. (b) td = 1 h, (c) td = 3 h, (d) td = 6 h, (e) td = 12 h.
Figure 3. (a) Room-temperature Raman spectra in the low-wavenumber region of 100−1000 cm−1 for BGTH7 powder sintered for different td at Ts = 1400 °C. The results of the XRD refinement and SEM images for BGTH7 ceramics sintered with different td at Ts = 1400 °C. (b) td = 1 h, (c) td = 3 h, (d) td = 6 h, (e) td = 12 h.
Molecules 27 07464 g003
Figure 4. (a) Ti 2p XPS spectrum of BGTH7 sintered at Ts = 1400 °C for 6 h. (b) EPR spectra of BGTH7 sintered at Ts = 1400 °C measured at room temperature. (c) Temperature-dependent EPR spectra of BGTH7 sintered at Ts = 1400 °C for 6 h. (d) Raman spectra under 532 nm excitation of BGTH7 ceramics sintered at Ts = 1400 °C for 1, 3, 6 and 12 h.
Figure 4. (a) Ti 2p XPS spectrum of BGTH7 sintered at Ts = 1400 °C for 6 h. (b) EPR spectra of BGTH7 sintered at Ts = 1400 °C measured at room temperature. (c) Temperature-dependent EPR spectra of BGTH7 sintered at Ts = 1400 °C for 6 h. (d) Raman spectra under 532 nm excitation of BGTH7 ceramics sintered at Ts = 1400 °C for 1, 3, 6 and 12 h.
Molecules 27 07464 g004
Figure 5. Temperature dependence of the dielectric constant (ε′) and dielectric loss (tan δ) measured at 1 kHz for BGTH7 ceramics sintered at Ts = 1400 °C for different td. (a) td = 1 h, (b) td = 3 h, (c) td = 6 h, (d) td = 12 h. The frequency dependence of the dielectric constant (ε′) and dielectric loss (tan δ) was measured at room temperature for BGTH7 ceramics sintered at Ts = 1400 °C for different td (e) and (f). (g) The dielectric properties of this work are compared with those of previously reported BaTiO3-based dielectric ceramics with X7U specifications [2,44,45,46,47,48,49,50,51,52,53,54,55].
Figure 5. Temperature dependence of the dielectric constant (ε′) and dielectric loss (tan δ) measured at 1 kHz for BGTH7 ceramics sintered at Ts = 1400 °C for different td. (a) td = 1 h, (b) td = 3 h, (c) td = 6 h, (d) td = 12 h. The frequency dependence of the dielectric constant (ε′) and dielectric loss (tan δ) was measured at room temperature for BGTH7 ceramics sintered at Ts = 1400 °C for different td (e) and (f). (g) The dielectric properties of this work are compared with those of previously reported BaTiO3-based dielectric ceramics with X7U specifications [2,44,45,46,47,48,49,50,51,52,53,54,55].
Molecules 27 07464 g005
Figure 6. The mechanisms for improving dielectric properties.
Figure 6. The mechanisms for improving dielectric properties.
Molecules 27 07464 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bai, J.; Liu, Q.; Li, X.; Wei, X.; Li, L. Optimization of Sintering Conditions to Enhance the Dielectric Performance of Gd3+ and Ho3+ Codoped BaTiO3 Ceramics. Molecules 2022, 27, 7464. https://doi.org/10.3390/molecules27217464

AMA Style

Bai J, Liu Q, Li X, Wei X, Li L. Optimization of Sintering Conditions to Enhance the Dielectric Performance of Gd3+ and Ho3+ Codoped BaTiO3 Ceramics. Molecules. 2022; 27(21):7464. https://doi.org/10.3390/molecules27217464

Chicago/Turabian Style

Bai, Jianghui, Qiaoli Liu, Xia Li, Xin Wei, and Liping Li. 2022. "Optimization of Sintering Conditions to Enhance the Dielectric Performance of Gd3+ and Ho3+ Codoped BaTiO3 Ceramics" Molecules 27, no. 21: 7464. https://doi.org/10.3390/molecules27217464

Article Metrics

Back to TopTop