Next Article in Journal
New Insights into Materials for Pesticide and Other Agricultural Pollutant Remediation
Previous Article in Journal
Hydrogels for Neural Regeneration: Exploring New Horizons
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Crystal Structure, Infrared Reflection Spectrum, and Improved Microwave Dielectric Characteristics of Ba4Sm28/3Ti18O54 Ceramics via One-Step Reaction Sintering

1
School of Electronic and Information Engineering, Hubei University of Science and Technology, Xianning 437100, China
2
Key Laboratory of Photoelectric Sensing and Intelligent Control, Hubei University of Science and Technology, Xianning 437100, China
*
Author to whom correspondence should be addressed.
Materials 2024, 17(14), 3477; https://doi.org/10.3390/ma17143477 (registering DOI)
Submission received: 19 June 2024 / Revised: 4 July 2024 / Accepted: 10 July 2024 / Published: 13 July 2024
(This article belongs to the Special Issue Structures, Properties, and Phase Transition in Dielectric Ceramics)

Abstract

:
High-k Ba4Sm28/3Ti18O54 ceramics with improved microwave dielectric characteristics were successfully fabricated using the one-step reaction sintering (RS) route. The sintering characteristics, microstructure, crystal structure, infrared reflection spectrum, and microwave dielectric characteristics of Ba4Sm28/3Ti18O54 ceramics prepared by the RS route were systematically investigated. Samples prepared by the RS route exhibited single-phase orthorhombic tungsten–bronze structure and dense microstructure at optimum sintering temperature. Compared with the conventional solid-state (CS) process, the Ba4Sm28/3Ti18O54 ceramics fabricated by the RS route presented a smaller temperature coefficient (TCF), a higher quality factor (Q × f), and a higher permittivity (εr). The improved microwave dielectric characteristics were highly dependent on the theoretical permittivity, atomic packing fraction, suppression of Ti3+, and Ti-site bond valence. Excellent combined microwave dielectric characteristics (TCF = −7.9 ppm/°C, Q × f = 9519 GHz, εr = 80.26) were achieved for Ba4Sm28/3Ti18O54 ceramics prepared by RS route sintered at 1400 °C, suggesting the RS route was a straightforward, economical and effective route to prepare high-performance Ba4Sm28/3Ti18O54 ceramics with promising application potential.

1. Introduction

Recently, the widespread application of 5G/6G wireless communication techniques has brought increasing requirements for high-performance microwave components (e.g., antennas, duplexers, filters, dielectric resonators, etc.) [1,2,3,4,5]. Being a critical material for the fabrication microwave components, microwave dielectric ceramics have attracted continuing commercial and scientific attention because of their superior properties [6,7,8,9,10]. Generally, high-performance microwave dielectric ceramics are desired to have a small temperature coefficient (τf or TCF, here TCF is a measure of the “drift” with respect to the temperature of the resonant frequency) to meet thermal stability, a high quality factor (the quality factor (Q), which is a function of resonant frequency (f), is sometimes expressed as Q × f) to enable good frequency selectivity for microwave components, and a high permittivity (εr) in order to realize miniaturization [11,12,13,14,15]. However, achieving the three dielectric characteristics mentioned above (high Q × f, small TCF, and high εr) simultaneously in high-k microwave dielectric ceramics for 5G applications is a significant challenge.
The typical high-k microwave dielectric ceramics include BiVO4-based, Bi2O3-Li2O-Ta2O5, lead-based perovskite, SrTiO3-based, LiO2-Nb2O5-TiO2, Na2O-M2O3-TiO2, CaO-LiO2-M2O3-TiO2, (Ca1−xM2x/3)TiO3, and Ba6−3xM8+2xTi18O54 systems (M = Pr, Sm, La, Nd) [16,17,18,19,20,21,22,23,24,25]. As a member of the high-k microwave ceramic ceramics, tungsten–bronze Ba6−3xM8+2xTi18O54-based (M = Nd, Pr, Sm) ceramics have been receiving continued attention by the researchers due to the high-k (εr > 80) and high quality factor (Q × f > 5000 GHz) [26]. Specifically, the Ba4Sm28/3Ti18O54 (BST, x = 2/3) ceramics with outstanding microwave dielectric characteristics (TCF~−10 ppm/°C, Q × f~9000 GHz, and εr~80) have been regarded as a suitable material in 5G applications [27]. Ohsato et al. [28] reported that the phase constitution of BST ceramics is refined as an orthorhombic tungsten–bronze phase with Pbnm (No. 62, superlattice) by Rietveld analysis using high-resolution XRD data. Wu et al. [29] studied the Raman spectroscopy of BST ceramics and suggested that the Raman modes of the Pbnm were 5B2g + 5B3g + 7Ag + 7Bg. The combining of different R3+ (R = Nd, Pr, La) cations is an efficient approach to modify the dielectric performance in BST ceramics, small TCF, high εr and low tanδ are obtained through the variation of Sm/Nd, Sm/Pr and Sm/La ratios [30,31,32]. Santha et al. [33] investigated the effect of BaCu(B2O5) on the the dielectric characteristics and sintering temperature (S.T.) of BST ceramics, and the S.T. of BST ceramics could be lowered to 950 °C with 10 wt% BaCu(B2O5). Different processes have been developed to produce BST ceramics, such as the sol-gel route, classical solid-state (CS) method, and spark plasma sintering (SPS) process [33,34,35]. These approaches are relatively complicated, and the repetitive synthesis process results in fluctuations in the microwave dielectric characteristics of BST ceramics.
Recently, the one-step reaction sintering (RS) route without calcination and secondary ball-milling has been attracting considerable interest because of its high efficiency and process simplification. Du et al. [36] have successfully prepared AlON transparent ceramics with a transmittance of 73.2% at 1600 nm using one-step RS route and tape-casting technology. Liang et al. [37] reported the piezoelectric properties of Bi4Ti3O12 ceramics could be significantly improved using the RS route and Ta/Mn co-doping. In addition, numerous microwave dielectric ceramics have been successfully fabricated using the RS route, such as BaTi5O11, Ba2Ti9O20, Li2MgTi3O8, 0.5Ca(Mg1/3Nb2/3)O3-0.5Ca0.6La0.267TiO3, Li3Mg2NbO6, Ca1.15M0.85Al0.85Ti0.15O4 (M = Y, La, Nd), CaTiO3-LnAlO3 (Ln = La, Nd), Mg4Nb2O9, MgAl2O4-CoAl2O4, AZrNb2O8(A = Zn, Mg, Co), and Ln2Zr3(MoO4)9 (Ln = La, Ce, Nd) [38,39,40,41,42,43,44,45,46,47,48]. Although there have been many studies on tungsten–bronze-structure BST microwave dielectric ceramics, the preparation of BST ceramics by the one-step RS route has not been reported until now.
One problem with using TiO2 as a raw material is the reduction of Ti4+ to Ti3+ in high sintering temperature. This may result in the presence of oxygen vacancies ( V O ), which are regarded as detrimental to microwave dielectric loss (tan δ = 1/Q). During the high-temperature sintering process in air, quasi-free electrons and oxygen vacancies ( V O ) are simultaneously generated in the crystals owing to insufficient oxygen partial pressure [49,50]. F-type color centers can be formed by oxygen vacancies ( V O ) in combination with weakly bound electrons, which have three forms: F2+ ( V O , bare oxygen vacancies), F+ ( V O   e , oxygen vacancies with one trapped electron), and F ( V O   2 e , oxygen vacancies with two trapped electrons) [51,52,53]. The recent efforts to restrain the reduction of Ti4+ and oxygen vacancies ( V O ) have been a significant area of research. Doping is an effective method of reducing dielectric loss, thereby enabling the material to meet the requirements of commercial applications. The dielectric loss of TiO2-based microwave dielectric ceramics (TiO2, BaTi4O9, Na0.5Sm0.5TiO3, Ba2Ti9O20, 0.73ZrTi2O6-0.27MgNb2O6, Ba6−3xSm8+2xTi18O54 and Ba6−3xNd8+2xTi18O54) will be significantly reduced by the introduction of a range of divalent or trivalent acceptor cations with ionic radii between 0.5 Å and 0.95 Å, such as Cr3+, Al3+, Ga3+, Zn2+, Mg2+, Mn2+, Cu2+ and Co2+ [21,49,50,53,54,55,56,57,58]. It is worth noting whether the dielectric loss of BST ceramics can be improved by RS route.
In this study, the BST ceramics with improved microwave dielectric characteristics prepared by the one-step RS route were first introduced. The effects of various sintering routes (CS and RS routes) on the crystal structure, microstructure, valence states of Ti ions, and microwave dielectric characteristics of the BST ceramics were studied. Based on the Rietveld analysis, the theoretical permittivity (εthe), atomic packing fraction (P.F.%), and Ti-site bond valence (VTi) were analyzed to study the structural properties of BST ceramics. Furthermore, infrared reflection spectrum was employed to evaluate intrinsic dielectric characteristics of BST ceramics prepared by RS route.

2. Materials and Methods

Ba4Sm28/3Ti18O54 ceramics were prepared via one-step RS and CS processes employing high-purity Sm2O3 (99.99%, Aladdin, Shanghai, China), TiO2 (99.84%, Zhongxing, Xiantao, China), and BaCO3 (99.8%, Yuanda, Mianyang, China) powders. In particular, Sm2O3 powder was calcined at 950 °C/4h for moisture removal. The oxide and carbonate powders were then weighted to achieve the desired stoichiometric ratio of Ba4Sm28/3Ti18O54 (BST). For the RS process, the blended oxide and carbonate powders were milled in de-ionized water with ZrO2 balls (diameter 2~8 mm) for 2 h. After oven drying at 120 °C/12 h, the drying oxide and carbonate powders were combined with the binder (10 wt% PVA). The ground and sieved powders were molded into cylindrical pellets (Φ8 × 5 mm, and Φ8 × 2 mm). For the CS process, the detailed experimental procedure was described in earlier work [5]. All these pellets prepared by RS and CS route were subsequently sintered in an atmospheric environment at 1300 °C/4 h to 1450 °C/4 h. Figure 1 presents the detailed procedure for the preparation of BST ceramics by the RS route due to the absence of calcination and secondary ball-milling, the RS route is a simple and efficient preparation method compared to the CS route.
The phase structure of sintered BST ceramics prepared by RS and CS routes was detected via aX-ray diffractometer (XRD; BRUKER, Billerica, MA, USA, D8 advance). A Scanning Electron Microscope (SEM; ZEISS, Oberkochen, Germany, GEMINI300) was utilized to evaluate the natural sintered surface of the BST samples. An infrared reflection (IR) spectrum (50~4000 cm−1) was collected in atmospheric (500~5000 cm−1) and vacuum (40~700 cm−1) environments using an FTIR spectrometer (BRUKER, USA, IFS 66v). The titanium (Ti) valence state in BST ceramics was analyzed by an X-ray Photoelectron Spectrometer (XPS; KRATOS, Kyoto, Japan, Axis Ultra DLD). Microwave dielectric characteristics (TCF, εr, and Q × f) of BST ceramics were evaluated at microwave frequency (5–6 GHz) via the Hakki–Coleman technique [59] by a network analyzer (KEYSIGHT, Santa Rosa, CA, USA, E5071C).

3. Results and Discussion

The XRD patterns of BST ceramics prepared using one-step RS and CS routes are displayed in Figure 2. As illustrated in Figure 2, all diffraction peaks of BST ceramics prepared by RS and CS routes corresponded to the tungsten–bronze phase (JCPDS No. 89-4356, Ba3.99Sm9.34Ti18O54) with orthorhombic structure (Pbnm, No. 62) [28], and no diffraction peaks of impurity phases were detected over the entire sintering temperature range. Structural diagrams of tungsten–bronze-structure BST ceramics with VESTA3 projected along [ 001 ] and [ 100 ] are presented in Figure 3 [60]. As shown in Figure 3a, the crystal structure of the tungsten–bronze-type BST ceramics consists of a three-dimensional framework of corner-sharing TiO6 octahedron linked at the corners to form three types of cavities: diamond cavities (A1), large pentagonal cavities (A2), and small triangular cavities (C). Sm is expected to be located in A1 cavities and larger Ba in the A2 cavities. Small triangular cavities are empty [28].
To further analyze the influence of various sintering routes and sintering temperatures on the lattice parameters and crystal structure of BST ceramics produced by the CS and RS routes, the Rietveld approach was applied using the GSAS application [61,62]. The fitted (black solid line) and measured (orange dot) XRD profiles of BST samples fabricated using the CS and RS methods are represented in Figure 4, and the fitted XRD were in reasonable accordance with the measurements. Table 1 presents the calculation of structure parameters (a, b, c), cell volume (V), theoretical density (ρth), and reliability factors (Rp, χ2, Rwp) of all BST ceramic samples. As illustrated in Table 1 and Figure 4, small reliability factors (Rp, Rwp) of less than 10% were observed, indicating that the results of the refinement were acceptable. It should be noted that the cell volume of BST ceramics produced using the CS route (V = 2079.12 Å3, a = 12.1596 Å, b = 22.312 Å, c = 7.6633 Å, Rp = 6.00%, Rwp = 8.01%, χ2 = 1.81) is larger than that of the RS route (V = 2073.65 Å3, a = 12.1476 Å, b = 22.2878 Å, c = 7.659 Å, Rp = 6.03%, Rwp = 7.62%, χ2 = 1.45) at optimal sintering temperature (1400 °C), which is a crucial factor in determining the microwave dielectric characteristics of BST ceramics prepared by different sintering routes.
The as-fired SEM photographs of BST ceramics sintered at 1300 to 1450 °C produced by the RS and CS route are presented in Figure 5. As presented in Figure 5, it is clear that all BST samples prepared by the RS and CS routes have rod-shaped grains of tungsten–bronze structural ceramics [25]. The number of rod-shaped grains increases when it is sintered at an increasing temperature from 1300 to 1450 °C. For the CS route, all ceramic samples sintered at different temperatures present a dense microstructure with well-defined grain boundaries, a gradually increasing grain size with increasing temperature is clearly observed, while for the RS route, the samples exhibit a porous and fine-grained microstructure when they are sintered at 1300 °C. For the samples sintered at 1400 °C and 1450 °C, the grain size increases with increasing sintering temperature, while the number of pores decreases significantly, and a relatively dense microstructure with well-defined grain boundaries is observed. Note that the grain size of the BST samples fabricated by the RS route is smaller than that of the CS route at the same temperature, which can be attributed to the absence of a calcination stage in the RS route.
Figure 6a presents the relative densities of BST ceramics using RS and CS routes sintered at 1300 to 1450 °C. The relative densities (ρre) of BST samples can be evaluated by Equation (1) [47]:
ρ r e = ρ b u ρ t h × 100 %
where ρre and ρbu are theoretical density (Table 1) and bulk density of BST ceramics, respectively. As a result of the increase in sintering temperature from 1300 to 1450 °C, the relative density (ρre) of the BST samples prepared by CS route exhibits a tendency to increase slowly and then decrease, a relative density of 96.88% was achieved for BST ceramic samples calcined at 1150 °C/4 h and sintered at 1400 °C/4 h. For the RS route, a greater increase in the ρre of BST ceramics was observed for the 1300 °C and 1400 °C temperature ranges, i.e., from a relative density (ρre) of 76.96% to 95.57%, indicating that rapid densification should take place at a temperature of over 1300 °C. This result is in accordance with that of SEM (Figure 5a). The improvement in relative density with increased temperature may be associated with porosity reduction.
The permittivity and Q × f values of BST samples sintered at 1300 to 1450 °C using RS and CS routes are presented in Figure 6b and Figure 6c, respectively. The trends in permittivity and Q × f values of all BST ceramics produced by the CS and RS routes varying with sintering temperature is similar to the relative density (ρre). The increased permittivity (εr) is associated with a pore (εr = 1) reduction and an increase in relative density, and the improved Q × f values are related to the grain growth and reduction in grain boundaries. The TCF values of BST ceramics fabricated by RS and CS routes are shown in Figure 6d. This study shows that the sintering temperature is not directly related to TCF as a result of no other additional additives and the single-phase BST ceramic samples. Near-zero TCF values (around −10 ppm/°C) were obtained for all ceramic samples prepared by RS and CS routes.
Interestingly, the BST ceramics produced by the RS route have improved microwave dielectric characteristics than those produced using the CS route at the optimum sintering temperature (1400 °C). The microwave dielectric characteristics of BST ceramics prepared by RS and CS routes sintered at 1400 °C are listed in Table 2. Compared to the CS route (εr = 79.38, Q × f = 8230 GHz, TCF = −12.1 ppm/°C), the BST ceramics fabricated by the RS route present a higher quality factor (Q × f = 9519 GHz), a lower dielectric loss (tanδ = 6.18 × 10−4), a smaller temperature coefficient (TCF = −7.9 ppm/°C), and a higher permittivity (εr = 80.26). The RS route is suggested to improve the microwave dielectric characteristics of BST ceramics. To investigate the effect of various preparation processes on the microwave dielectric characteristics of BST ceramics, atomic packing fraction, theoretical permittivity, bond valence, and vacancy defects were analyzed. The results of theoretical permittivity (εthe), atomic packing fraction (P.F.%), and Ti-site bond valence (VTi) of BST ceramics produced by the CS and RS routes at optimal sintering temperature (1400 °C) are listed in Table 2.
On the basis of previous literature [63,64], the theoretical permittivity (εthe) can be evaluated using the Clausius–Mossotti (CM) formula (Equation (2)):
ε t h e = 3 V m + 8 π α t h e 3 V m 4 π α t h e  
where Vm = V/Z denotes molar volume, and αthe represents theoretical dielectric polarizability. The αthe of Ba4Sm28/3Ti18O54 ceramics can be calculated using Equation (3) [63]:
α the ( Ba 4 Sm 28 / 3 Ti 18 O 54 ) = 4 α the Ba 2 + + 28 3 α the Sm 3 + + 18 α the Ti 4 + + 54 α the O 2
where α(Ba2+) = 6.4 Å3, α(O2−) = 2.01 Å3, α(Ti4+) = 2.93 Å3, and α(Sm3+) = 4.74 Å3. According to the CM formula, the theoretical permittivity (εthe) is inversely proportional to the molar volume (Vm) when the theoretical dielectric polarizability of Ba4Sm28/3Ti18O54 ceramics is constant. As listed in Table 1, due to cell volume (V = 2073.65 Å3) of BST ceramics fabricated by the RS route is smaller than that fabricated by the CS route (V = 2079.12 Å3), the theoretical permittivity calculated from CM formula of RS route (εthe = 43.27) is higher than the CS route (εthe = 41.65), and the tendency of measured permittivity (εr) is similar with the theoretical permittivity (εthe). It is worth noting that the εr is much higher than the εthe, which is attributed to the low symmetry structure of the BST ceramics. The same large deviations are found in other Ti-based microwave dielectric ceramics [30,65,66,67].
A higher Q × f value (9519 GHz) for the RS route prepared ceramic sample was achieved compared to a Q × f value (8230 GHz) for the CS route is shown in Table 2. Generally speaking, the total tanδ (1/Q) of functional ceramics is classified into the extrinsic and intrinsic loss. The extrinsic loss is primarily associated with pores, dopants, impure phases, lattice defects, grain boundaries, vacancy defects, etc., while intrinsic loss is primarily attributed to lattice vibration and structural characteristics. In this work, the effects of pores and impure phases on tanδ could be ignored as the BST ceramics produced by the CS and RS routes at optimum sintering temperature (1400 °C) with high densities (ρre > 95%) and single-phase structure.
As reported by Kim et al. [68], dielectric loss (tanδ) is highly associated with structural characteristics of microwave ceramics. The structural characteristics of microwave ceramics could be assessed by the packing fraction, as this parameter is related to the cell volume (V) and effective ion size. On the basis of the results of Rietveld analysis, the packing fraction (P.F.%) of the BST ceramics prepared by the RS and CS routes is evaluated by means of Equation (4) [68]:
Packing   fraction ( % ) = volume   of   packed   ions volume   of   unit   cell × Z = 4 3 π 4 r B a 3 + 28 3 r S m 3 + 18 r T i 3 + 54 r O 3 V × 2
where V refers to cell volume, r S m , r B a , r T i , and r O are the effective ionic sizes of Sm (1.079 Å), Ba (1.52 Å), Ti (0.605 Å), and O (1.4 Å) [69], respectively. The packing fraction (P.F.%) of the BST ceramics is illustrated in Table 2. As presented in Table 2, a lower Q × f value (8230 GHz) corresponds to a smaller P.F.% (71.69%), and a higher Q × f value (9519 GHz) correlates with a higher P.F.% (71.88%).
In addition, the oxygen vacancy defect in Ti-based microwave dielectric ceramics is widely acknowledged and is believed to be balanced through the generation of Ti3+ ions [49]. The primary reason for the higher dielectric losses is thought to be the Ti3+ ions in Ti-based ceramics. The mechanism could be represented by the following defect equation (Equations (5) and (6)) [50]:
O O × 1 2 O 2 + V O + 2 e  
T i 4 + + e T i 3 +  
It is worth noting that the BST ceramics produced by the RS route at 1400 °C have a light yellow color, while those prepared by the CS route have a dark yellow color, as illustrated in Figure 6c. To explain this difference, valence states of Ti ions in BST ceramics prepared by the RS and CS routes were investigated by XPS analysis. In general, the two peaks of Ti 2p correspond to binding energies (B.E.) of approximately 458 (2p3/2) and 464 (2p1/2) eV [70]. Figure 7 presents the XPS spectroscopy of Ti 2p for BST samples prepared by the RS and CS routes at 1400 °C. The 2p3/2 and 2p1/2 region of Ti 2p in BST ceramics was fitted well to the Gaussian/Lorentzian sub-peak by means of the XPS-PEAK41 software, and the results are given in Figure 7 and Table 3. Table 3 presents that the relative peak area ratio of Ti3+ for the CS route (27.63%) is higher than the RS route (11.87%). This result suggests that the RS route is effective in suppressing the generation of trivalent titanium ions (Ti3+), which improves the quality factor (tanδ) of the BST ceramic. The combination of packing fraction (P.F.%) and XPS analysis explains the higher quality factor (lower dielectric loss) of the BST ceramics prepared by the RS route.
Table 2 shows that a smaller TCF value of −7.9 ppm/°C (1400 °C) for the RS route prepared ceramic sample is obtained compared to the TCF value of −12.1 ppm/°C (1400 °C) for the CS route. It was found that the TCF values of tungsten–bronze-structure Ba6−3xR8+2xTi18O54 (R = Eu, Sm, Nd, Pr, and La) were generally influenced by additives, composition, and structural characteristics [71]. As all ceramic samples have no other extra additives and the unique composition Ba4Sm28/3Ti18O54, the key factor affecting their TCF may be the structural characteristics. Bond valence has been generally accepted as a useful tool for associating structural characteristics with TCF value. The bond valence (Sjk) of the BST ceramics prepared by the RS and CS routes is calculated by means of Equations (7) and (8) [68]:
S j k = k s j k  
s j k = exp R j k d j k 0.37  
where djk represents the bond length, and Rjk denotes the bond valence parameter. The TCF values and the calculated Ti-site bond valence (VTi) of BST ceramics are presented in Table 2. Based on bond valence theory [47], the higher bond valence would lead to a higher restoring force of oxygen octahedrons, ultimately decreasing |TCF|. As shown in Table 2, the VTi of 4.136 v.u. for the RS route is higher than the 4.109 v.u. for the CS route, and the |TCF| of 7.9 ppm/°C for the RS route is smaller than the 12.1 ppm/°C for the CS route. As VTi increases, the |TCF| value decreases.
Microwave dielectric characteristics of high-k BST ceramics prepared by different routes are presented in Table 4 for comparison. Xu et al. [34] produced BST ceramics of Q × f = 11,300 GHz, and εr = 80.8 by a sol-gel technique sintered at 1360 °C/3h. Guo et al. [35] prepared BST ceramics of TCF = −17.2 ppm/°C, Q × f = 10,099 GHz, and εr = 81.2 via the SPS route sintered at 1200 °C/5min. Chen et al. [27] and Santha et al. [33] reported the microwave dielectric characteristics of BST ceramics fabricated by the CS route of TCF = −10.6 ppm/°C, Q × f = 9240 GHz, εr = 81 and TCF = −12 ppm/°C, Q × f = 10,025 GHz, εr = 76 were obtained, respectively. In the present work, excellent microwave dielectric characteristics (TCF = −7.9 ppm/°C, Q × f = 9519 GHz, εr = 80.26) were achieved for Ba4Sm28/3Ti18O54 ceramics sintered at 1400 °C/4 h produced using the one-step RS route. Compared to other routes, the RS route without the calcination and secondary ball-milling processes is a straightforward, economical and effective method for preparing high-performance BST ceramics.
Far-infrared (FIR) reflection spectrum is commonly applied to evaluate intrinsic dielectric characteristics of electronic ceramics by means of the classic oscillator theory (Equation (9)) [72]:
ε ( ω ) = ε + j = 1 n ε j = ω p j 2 ω o j 2 ω 2 j γ j ω
R ( ω ) = 1 ε ( ω ) 1 + ε ( ω ) 2
where γj, ωpj, ε, ωoj, and n are damping factor of the j-th mode, plasma frequency, optical frequency permittivity caused by electronic polarization, eigen frequency and sequence of phonon modes, individually. Equation (10) [72] illustrates the interrelation of complex permittivity (ε*(ω)) and reflectivity (R(ω)).
The room-temperature FIR spectrum of BST ceramics (50~1000 cm−1) sintered at 1400 °C using the RS route is illustrated in Figure 8a. The calculated FIR spectrum (red line) is consistent with the measured spectrum (black circle). Based on Kramers–Kronig (K-K) theory, the calculated FIR spectrum of BST ceramics prepared by the RS route is transformed into ε ( ω ) and ε ( ω ) , and this is presented in Table 5 and Figure 8b. Measured ε ( ω ) (black triangle) and ε ( ω ) (blue triangle) are also shown Figure 8b. Compared to the theoretical permittivity value (43.27) calculated by the CM equation, the permittivity value (73.91) derived from the FIR spectrum is closer to the measured one (80.23), demonstrating that the dominant dielectric contribution of BST ceramics is phonon absorption. As can be seen in Table 5, the FIR spectrum of BST ceramics was fitted by the 24 individual modes. In particular, the eight infrared active modes (j = 1–8) below 220 cm−1 give the majority (84.07%) dielectric contribution to the permittivity. The ε value derived from Equation (10) is 4.69, only 6.3% of the total extrapolated permittivity (73.91), suggesting that ionic polarization plays an important role in the polarization that contributed to the εr of BST ceramics.

4. Conclusions

For the first time, high-permittivity Ba4Sm28/3Ti18O54 ceramics with improved microwave dielectric characteristics have been successfully fabricated by the one-step RS route. The effects of various sintering routes (CS and RS routes) on the crystal structure, microstructure, valence states of Ti ions, and dielectric characteristics of BST ceramics have been investigated. Single-phase orthorhombic tungsten–bronze-structure and dense-microstructure BST ceramics were achieved using RS and CS routes at the optimum sintering temperature (1400 °C). The FIR spectrum analysis presents that the modes below 220 cm−1 give the majority (84.07%) dielectric contribution to the permittivity of BST ceramics prepared by the RS route. Compared to other routes, the RS route without the calcination and secondary ball-milling processes is a straightforward, economical and effective method for preparing high-performance BST ceramics.

Author Contributions

Conceptualization, G.W.; Investigation, Z.L., H.Z. and H.W.; Methodology, Z.L. and G.X.; Supervision, G.W.; Validation, Z.L.; Writing—original draft, G.W.; Writing—review and editing, G.X. All authors have read and agreed to the published version of the manuscript.

Funding

The present study was supported by Foundation of Hubei Provincial Department of Education (D20212803), Hubei Province Natural Science Foundation of China (2022CFB369, 2021CFB418), Ph.D. foundation (Grant No. BK202109), and Innovation and Entrepreneurship Training Programme for University Students (S202110927051).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Reaney, I.M.; Iddles, D. Microwave dielectric ceramics for resonators and filters in mobile phone networks. J. Am. Ceram. Soc. 2006, 89, 2063–2072. [Google Scholar] [CrossRef]
  2. Shehbaz, M.; Du, C.; Zhou, D.; Xia, S.; Xu, Z. Recent progress in dielectric resonator antenna: Materials, designs, fabrications, and their performance. Appl. Phys. Rev. 2023, 10, 021303. [Google Scholar] [CrossRef]
  3. Yang, H.; Zhang, S.; Yang, H.; Wen, Q.; Yang, Q.; Gui, L.; Zhao, Q.; Li, E. The latest process and challenges of microwave dielectric ceramics based on pseudo phase diagrams. J. Adv. Ceram. 2021, 10, 885–932. [Google Scholar] [CrossRef]
  4. Luo, W.; Yan, S.; Zhou, J. Ceramic-based dielectric metamaterials. Interdiscip. Mater. 2022, 1, 11–27. [Google Scholar] [CrossRef]
  5. Wang, G.; Fu, Q.; Guo, P.; Hu, M.; Wang, H.; Yu, S.; Zheng, Z.; Luo, W. Crystal structure, spectra analysis and dielectric characteristics of Ba4M28/3Ti18O54 (M = La, Pr, Nd, and Sm) microwave ceramics. Ceram. Int. 2021, 47, 1750–1757. [Google Scholar] [CrossRef]
  6. Guo, Y.; Wang, Z.; Li, J. Sintering Behavior, Microstructure and Microwave Dielectric Properties of Li2TiO3-Based Solid Solution Ceramics with Lithium Fluoride Addition for Low-Temperature Co-Fired Ceramic Applications. Coatings 2023, 13, 1732. [Google Scholar] [CrossRef]
  7. Liu, K.; Shi, L.; Wang, X.; Liu, C.; Li, J.; Liao, Y.; Jin, L.; Zhang, D.; Zhang, H. Li+ enrichment to improve the microwave dielectric properties of Li2ZnTi3O8 ceramics and the relationship between structure and properties. J. Eur. Ceram. Soc. 2023, 43, 1483–1491. [Google Scholar] [CrossRef]
  8. Zhang, Q.; Su, H.; Peng, R.; Huang, F.; Wu, X.; Tang, X. Effect of phase, chemical bond and vibration characteristics on the microwave dielectric properties of temperature-stable Zn1−x(Li0.5Bi0.5)xMoxW1−xO4 ceramics. J. Eur. Ceram. Soc. 2022, 42, 2813–2819. [Google Scholar] [CrossRef]
  9. Wu, F.; Zhou, D.; Du, C.; Xu, D.M.; Li, R.T.; Shi, Z.Q.; Darwish, M.A.; Zhou, T.; Jantunen, H. Design and Fabrication of a Satellite Communication Dielectric Resonator Antenna with Novel Low Loss and Temperature-Stabilized (Sm1−xCax)(Nb1−xMox)O4 (x = 0.15–0.7) Microwave Ceramics. Chem. Mater. 2023, 35, 104–115. [Google Scholar] [CrossRef]
  10. Zhang, X.; Fang, Z.; Yang, H.; Zhao, P.; Zhang, X.; Li, Y.; Xiong, Z.; Yang, H.; Zhang, S.; Tang, B. Lattice evolution, ordering transformation and microwave dielectric properties of rock-salt Li3+xMg2−2xNb1−xTi2xO6 solid-solution system: A newly developed pseudo ternary phase diagram. Acta Mater. 2021, 206, 116636. [Google Scholar] [CrossRef]
  11. Wang, G.; Fu, Q.; Zha, L.; Hu, M.; Huang, J.; Zheng, Z.; Luo, W. Microwave dielectric characteristics of tungsten bronze-type Ba4Nd28/3Ti18−yGa4y/3O54 ceramics with temperature stable and ultra-low loss. J. Eur. Ceram. Soc. 2022, 42, 154–161. [Google Scholar] [CrossRef]
  12. Yao, G.; Zhao, J.; Lu, Y.; Liu, H.; Pei, C.; Ding, Q.; Chen, M.; Zhang, Y.; Li, D.; Wang, F. Microwave dielectric properties of Li3TiO3F oxyfluorides ceramics. Crystals 2023, 13, 897. [Google Scholar] [CrossRef]
  13. Liu, X.; Wang, Z.; She, X.; Jia, Q.; Li, J. Improved microstructure and high quality factor of Li2Ti0.9(Zn1/3Ta2/3)0.1O3 microwave ceramics with LiF additive for LTCC applications. J. Eur. Ceram. Soc. 2023, 43, 1469–1476. [Google Scholar] [CrossRef]
  14. Wu, F.F.; Zhou, D.; Du, C.; Sun, S.K.; Pang, L.X.; Jin, B.B.; Qi, Z.M.; Varghese, J.; Li, Q.; Zhang, X.Q. Temperature stable Sm(Nb1−xVx)O4 (0.0≤ x ≤ 0.9) microwave dielectric ceramics with ultra-low dielectric loss for dielectric resonator antenna applications. J. Mater. Chem. C 2021, 9, 9962–9971. [Google Scholar] [CrossRef]
  15. Wu, F.F.; Zhou, D.; Du, C.; Jin, B.B.; Li, C.; Qi, Z.M.; Sun, S.; Zhou, T.; Li, Q.; Zhang, X.Q. Design of a Sub-6 GHz Dielectric Resonator Antenna with Novel Temperature-Stabilized (Sm1−xBix)NbO4 (x = 0–0.15) Microwave Dielectric Ceramics. ACS Appl. Mater. Interfaces 2022, 14, 7030–7038. [Google Scholar] [CrossRef]
  16. Zhou, D.; Pang, L.X.; Wang, D.W.; Reaney, I.M. BiVO4 based high k microwave dielectric materials: A review. J. Mater. Chem. C 2018, 6, 9290–9313. [Google Scholar] [CrossRef]
  17. Zhou, D.; Pang, L.X.; Wang, D.W.; Li, C.; Jin, B.B.; Reaney, I.M. High permittivity and low loss microwave dielectrics suitable for 5G resonators and low temperature co-fired ceramic architecture. J. Mater. Chem. C 2017, 5, 10094–10098. [Google Scholar] [CrossRef]
  18. Hu, X.; Chen, X.; Wu, S. Preparation, properties and characterization of CaTiO3-modified Pb (Fe1/2Nb1/2)O3 dielectrics. J. Eur. Ceram. Soc. 2003, 23, 1919–1924. [Google Scholar] [CrossRef]
  19. Luo, T.; Shan, X.; Zhao, J.; Feng, H.; Zhang, Q.; Yu, H.; Liu, J. Improvement of quality factor of SrTiO3 dielectric ceramics with high dielectric constant using Sm2O3. J. Am. Ceram. Soc. 2019, 102, 3849–3853. [Google Scholar] [CrossRef]
  20. Ma, F.S.; Zeng, Q.; Feng, Y.Z.; Yao, C.F.; Guo, C.C. Crystal structure and microwave dielectric properties of LiNb0.6Ti0.5O3 ceramics with Zn and Nb co-doped. Mater. Sci. Eng. B 2022, 285, 115919. [Google Scholar] [CrossRef]
  21. Fang, Z.X.; Tang, B.; Li, E.; Zhang, S.R. High-Q microwave dielectric properties in the Na0.5Sm0.5TiO3+Cr2O3 ceramics by one synthetic process. J. Alloys Compd. 2017, 705, 456–461. [Google Scholar] [CrossRef]
  22. Yu, T.; Yang, Q.; Feng, H.; Yu, H. Phase composition and microwave dielectric properties of Ca0.128Ba0.032Sm0.46Li0.3TiO3 ceramics with alumina addition. J. Eur. Ceram. Soc. 2022, 42, 1480–1485. [Google Scholar] [CrossRef]
  23. Xiong, Z.; Zhang, X.; Tang, B.; Yang, C.; Fang, Z.; Zhang, S. Characterization of structure and properties in CaO-Nd2O3-TiO2 microwave dielectric ceramic modified by Al2O3. Mater. Charact. 2021, 176, 111108. [Google Scholar] [CrossRef]
  24. Ma, Z.; Guo, W.; Yue, Z. Microwave dielectric properties of Al-doped Ba4(Sm, Nd)9.33Ti18O54 ceramics added with TiO2 and sintered in oxygen. Ceram. Int. 2022, 48, 12906–12913. [Google Scholar] [CrossRef]
  25. Li, L.; Wang, X.; Luo, W.; Wang, S.; Yang, T.; Zhou, J. Internal-strain-controlled tungsten bronze structural ceramics for 5G millimeter-wave metamaterials. J. Mater. Chem. C 2021, 9, 14359–14370. [Google Scholar] [CrossRef]
  26. Ohsato, H. Science of tungstenbronze-type like Ba6−3xR8+2xTi18O54 (R = rare earth) microwave dielectric solid solutions. J. Eur. Ceram. Soc. 2001, 21, 2703–2711. [Google Scholar] [CrossRef]
  27. Chen, X.M.; Li, Y. A-and B Site Cosubstituted Ba6−3xSm8+2xTi18O54 Microwave Dielectric Ceramics. J. Am. Ceram. Soc. 2002, 85, 579–584. [Google Scholar] [CrossRef]
  28. Okudera, H.; Nakamura, H.; Toraya, H.; Ohsato, H. Tungsten Bronze-Type Solid Solutions Ba6−3xSm8+2xTi18O54 (x = 0.3, 0.5, 0.67, 0.71) with Superstructure. J. Solid State Chem. 1999, 142, 336–343. [Google Scholar] [CrossRef]
  29. Wu, S.; Li, Y.; Chen, X. Raman spectra of Ba6−3xSm8+2xTi18O54 solid solution. J. Phys. Chem. Solids 2003, 64, 2365–2368. [Google Scholar] [CrossRef]
  30. Xiong, Z.; Tang, B.; Yang, C.; Zhang, S. Correlation between structures and microwave dielectric properties of Ba3.75Nd9.5−xSmxTi17.5(Cr1/2Nb1/2)0.5O54 ceramics. J. Alloys Compd. 2018, 740, 492–499. [Google Scholar] [CrossRef]
  31. Chen, X.; Qin, N.; Li, Y. Microstructures and Microwave Dielectric Characteristics of Ba6−3x(Sm1−yLay)8+2xTi18O54 Solid Solutions (x = 2/3 and 0.75). J. Electroceram. 2002, 9, 31–35. [Google Scholar] [CrossRef]
  32. Wang, G.; Fu, Q.; Shi, H.; Tian, F.; Wang, M.; Yan, L.; Zheng, Z.; Luo, W. Suppression of oxygen vacancies generation in Ba6−3xSm8+2xTi18O54 (x = 2/3) microwave dielectric ceramics through Pr substitution. Ceram. Int. 2019, 45, 22148–22155. [Google Scholar] [CrossRef]
  33. Santha, N.; Sebastian, M. Low temperature sintering and microwave dielectric properties of Ba4Sm9.33Ti18O54 ceramics. Mater. Res. Bull. 2008, 43, 2278–2284. [Google Scholar] [CrossRef]
  34. Xu, Y.; Huang, G.; He, Y. Sol-gel preparation of Ba6−3xSm8+2xTi18O54 microwave dielectric ceramics. Ceram. Int. 2005, 31, 21–25. [Google Scholar] [CrossRef]
  35. Guo, Y.; Kakimoto, K.I.; Ohsato, H. Microwave dielectric properties of Ba6−3xSm8+2xTi18O54 (x = 2/3) ceramics produced by spark plasma sintering. Jpn. J. Appl. Phys. 2003, 42, 7410. [Google Scholar] [CrossRef]
  36. Du, W.; Lu, K.; He, B.; Zhou, X.; Huang, X.; Qi, J.; Lu, T. Direct tape casting of Al2O3/AlN slurry for AlON transparent ceramic wafers via one-step reaction sintering. J. Eur. Ceram. Soc. 2023, 43, 3538–3543. [Google Scholar] [CrossRef]
  37. Liang, Z.; Huang, R.; Xie, T.; Zhang, Y.; Lin, H.T.; Dai, Y. Improvement of piezoelectric properties of Bi4Ti3O12 ceramics by Mn/Ta co-doping and direct reaction sintering. Ceram. Int. 2023, 49, 20920–20928. [Google Scholar] [CrossRef]
  38. Zhou, H.; Wang, H.; Chen, Y.; Li, K.; Yao, X. Microwave Dielectric Properties of BaTi5O11 Ceramics Prepared by Reaction-Sintering Process with the Addition of CuO. J. Am. Ceram. Soc. 2008, 91, 3444–3447. [Google Scholar] [CrossRef]
  39. Luo, T.; Yang, Q.; Yu, H.; Liu, J. Formation mechanism and microstructure evolution of Ba2Ti9O20 ceramics by reaction sintering method. J. Am. Ceram. Soc. 2020, 103, 1079–1087. [Google Scholar] [CrossRef]
  40. Yao, G.; Liu, P.; Zhang, H. Microwave dielectric properties of Li2MgTi3O8 ceramics produced by reaction-sintering method. J. Mater. Sci. Mater. Electron. 2013, 24, 1128–1131. [Google Scholar] [CrossRef]
  41. Li, J.; Xu, P.; Qiu, T.; Yao, L. Sintering characteristics and microwave dielectric properties of 0.5Ca0.6La0.267TiO3-0.5Ca(Mg1/3Nb2/3)O3 ceramics prepared by reaction-sintering process. J. Rare Earths 2018, 36, 404–408. [Google Scholar] [CrossRef]
  42. Wang, G.; Zhang, D.; Gan, G.; Yang, Y.; Rao, Y.; Xu, F.; Huang, X.; Liao, Y.; Li, J.; Liu, C. Synthesis, crystal structure and low loss of Li3Mg2NbO6 ceramics by reaction sintering process. Ceram. Int. 2019, 45, 19766–19770. [Google Scholar] [CrossRef]
  43. Qu, X.; He, S.; Li, Q.; Deng, S.; Xiao, Y.; Liu, K.; Chen, X.; Liang, J.; Zhou, H. Microwave dielectric properties of Ca1.15RE0.85Al0.85Ti0.15O4 (RE = Nd, La, Y) ceramics prepared by the reaction sintering method. Ceram. Int. 2023, 49, 716–721. [Google Scholar] [CrossRef]
  44. Zhou, S.; Wu, Q.; Xu, H.; Luan, X.; Hu, S.; Zhou, X.; He, S.; Wang, X.; Zhang, H.; Chen, X. Synthesis and characterization of reaction sintered CaTiO3-LnAlO3 (Ln = La, Nd) ceramics. Ceram. Int. 2021, 47, 32433–32437. [Google Scholar] [CrossRef]
  45. Yang, T.; Han, Z.; Liu, P.; Guo, B. Microwave dielectric properties of Mg4Nb2O9 ceramics with excess Mg(OH)2 produced by a reaction-sintering process. Ceram. Int. 2015, 41, S572–S575. [Google Scholar] [CrossRef]
  46. Tsai, W.C.; Liou, Y.H.; Liou, Y.C. Microwave dielectric properties of MgAl2O4-CoAl2O4 spinel compounds prepared by reaction-sintering process. Mater. Sci. Eng. B 2012, 177, 1133–1137. [Google Scholar] [CrossRef]
  47. Zhao, Y.; Zhang, P. Preparation for ultra-low loss dielectric ceramics of ZnZrNb2O8 by reaction-sintering process. J. Alloys Compd. 2016, 672, 630–635. [Google Scholar] [CrossRef]
  48. Xing, C.; Zhang, Y.; Tao, B.; Wu, H.; Zhou, Y. Crystal structure, infrared spectra and microwave dielectric properties of low-firing La2Zr3(MoO4)9 ceramics prepared by reaction-sintering process. Ceram. Int. 2019, 45, 22376–22382. [Google Scholar] [CrossRef]
  49. Templeton, A.; Wang, X.; Penn, S.J.; Webb, S.J.; Cohen, L.F.; Alford, N.M. Microwave dielectric loss of titanium oxide. J. Am. Ceram. Soc. 2000, 83, 95–100. [Google Scholar] [CrossRef]
  50. Pullar, R.C.; Penn, S.J.; Wang, X.; Reaney, I.M.; Alford, N.M. Dielectric loss caused by oxygen vacancies in titania ceramics. J. Eur. Ceram. Soc. 2009, 29, 419–424. [Google Scholar] [CrossRef]
  51. Kotomin, E.A.; Eglitis, R.; Popov, A.I. Charge distribution and optical properties of and F centres in crystals. J. Phys. Condens. Matter 1997, 9, L315. [Google Scholar] [CrossRef]
  52. Popov, A.; Kotomin, E.; Maier, J. Basic properties of the F-type centers in halides, oxides and perovskites. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2010, 268, 3084–3089. [Google Scholar] [CrossRef]
  53. He, L.; Yu, H.; Zeng, M.; Luo, T.; Liu, J. 0.73ZrTi2O6-0.27MgNb2O6 microwave dielectric ceramics modified by Al2O3 addition. J. Am. Ceram. Soc. 2018, 101, 5110–5119. [Google Scholar] [CrossRef]
  54. Negas, T. BaTi4O9/Ba2Ti9O20-Based Ceramics Resurrected for Modern Microwave Applications. Am. Ceram. Soc. Bull. 1993, 72, 80–89. [Google Scholar]
  55. Nomura, S.; Tomaya, K.; Kaneta, K. Effect of Mn doping on the dielectric properties of Ba2Ti9O20 ceramics at microwave frequency. Jpn. J. Appl. Phys. 1983, 22, 1125. [Google Scholar] [CrossRef]
  56. Yao, X.; Lin, H.; Chen, W.; Luo, L. Anti-reduction of Ti4+ in Ba4.2Sm9.2Ti18O54 ceramics by doping with MgO, Al2O3 and MnO2. Ceram. Int. 2012, 38, 3011–3016. [Google Scholar] [CrossRef]
  57. Huang, B.; Yan, Z.; Lu, X.; Wang, L.; Fu, Z.; Zhang, Q. “Dark hole” cure in Ba4.2Nd9.2Ti18O54 microwave dielectric ceramics. Ceram. Int. 2016, 42, 10758–10763. [Google Scholar] [CrossRef]
  58. Guo, W.; Zhang, J.; Luo, Y.; Yue, Z.; Li, L. Microwave dielectric properties and thermally stimulated depolarization of Al-doped Ba4(Sm, Nd)9.33Ti18O54 ceramics. J. Am. Ceram. Soc. 2019, 102, 5494–5502. [Google Scholar] [CrossRef]
  59. Hakki, B.; Coleman, P.D. A dielectric resonator method of measuring inductive capacities in the millimeter range. IRE Trans. Microw. Theory Tech. 1960, 8, 402–410. [Google Scholar] [CrossRef]
  60. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  61. Rietveld, H.M. A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr. 1969, 2, 65–71. [Google Scholar] [CrossRef]
  62. Toby, B.H. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef]
  63. Shannon, R.D. Dielectric polarizabilities of ions in oxides and fluorides. J. Appl. Phys. 1993, 73, 348–366. [Google Scholar] [CrossRef]
  64. Guo, H.H.; Zhou, D.; Liu, W.F.; Pang, L.X.; Wang, D.W.; Su, J.Z.; Qi, Z.M. Microwave dielectric properties of temperature-stable zircon-type (Bi, Ce)VO4 solid solution ceramics. J. Am. Ceram. Soc. 2020, 103, 423–431. [Google Scholar] [CrossRef]
  65. Fu, Z.; Liu, P.; Ma, J.; Zhao, X.; Zhang, H. Novel series of ultra-low loss microwave dielectric ceramics: Li2Mg3BO6 (B = Ti, Sn, Zr). J. Eur. Ceram. Soc. 2016, 36, 625–629. [Google Scholar] [CrossRef]
  66. Song, X.Q.; Yin, C.Z.; Zou, Z.Y.; Yang, J.Q.; Zeng, F.F.; Wu, J.M.; Shi, Y.S.; Lu, W.Z.; Lei, W. Structural evolution and microwave dielectric properties of CaTiO3-La(Mg2/3Nb1/3)O3 ceramics. J. Am. Ceram. Soc. 2022, 105, 7415–7425. [Google Scholar] [CrossRef]
  67. Tang, Y.; Shen, S.; Li, J.; Zhao, X.; Xiang, H.; Su, H.; Zhou, D.; Fang, L. Characterization of structure and chemical bond in high-Q microwave dielectric ceramics LiM2GaTi2O8 (M = Mg, Zn). J. Eur. Ceram. Soc. 2022, 42, 4573–4579. [Google Scholar] [CrossRef]
  68. Kim, E.S.; Chun, B.S.; Freer, R.; Cernik, R.J. Effects of packing fraction and bond valence on microwave dielectric properties of A2+B6+O4 (A2+: Ca, Pb, Ba; B6+: Mo, W) ceramics. J. Eur. Ceram. Soc. 2010, 30, 1731–1736. [Google Scholar] [CrossRef]
  69. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. Sect. A Cryst. Phys. Diffr. Theor. Gen. Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  70. Xiong, Z.; Yang, C.; Tang, B.; Fang, Z.; Chen, H.; Zhang, S. Structure-property relationships of perovskite-structured Ca0.61Nd0.26Ti1−x(Cr0.5Nb0.5)xO3 ceramics. Ceram. Int. 2018, 44, 7384–7392. [Google Scholar] [CrossRef]
  71. Ubic, R.; Reaney, I.; Lee, W. Microwave dielectric solid–solution phase in system BaO-Ln2O3-TiO2 (Ln = lanthanide cation). Int. Mater. Rev. 1998, 43, 205–219. [Google Scholar] [CrossRef]
  72. Pang, L.X.; Zhou, D. Modification of NdNbO4 microwave dielectric ceramic by Bi substitutions. J. Am. Ceram. Soc. 2019, 102, 2278–2282. [Google Scholar] [CrossRef]
Figure 1. One-step reaction sintering process for BST ceramics.
Figure 1. One-step reaction sintering process for BST ceramics.
Materials 17 03477 g001
Figure 2. XRD patterns of BST ceramics: (a) RS route sintered at 1300–1450 °C; (b) CS route sintered at 1300–1450 °C.
Figure 2. XRD patterns of BST ceramics: (a) RS route sintered at 1300–1450 °C; (b) CS route sintered at 1300–1450 °C.
Materials 17 03477 g002
Figure 3. Structural diagrams of BST ceramics projected along (a) [ 001 ]; (b) [ 100 ].
Figure 3. Structural diagrams of BST ceramics projected along (a) [ 001 ]; (b) [ 100 ].
Materials 17 03477 g003
Figure 4. The refinement plots of BST ceramics: (a) RS route sintered at 1300–1450 °C; (b) CS route sintered at 1300–1450 °C.
Figure 4. The refinement plots of BST ceramics: (a) RS route sintered at 1300–1450 °C; (b) CS route sintered at 1300–1450 °C.
Materials 17 03477 g004
Figure 5. SEM images of BST ceramics: (a) RS route sintered at 1300–1450 °C; (b) CS route sintered at 1300–1450 °C.
Figure 5. SEM images of BST ceramics: (a) RS route sintered at 1300–1450 °C; (b) CS route sintered at 1300–1450 °C.
Materials 17 03477 g005
Figure 6. (a) Relative densities of BST ceramics; (b) εr of BST ceramics; (c) Q × f of BST ceramics (Inserted images show BST ceramics prepared by RS and CS routes sintered at 1400 °C); (d) TCF of BST ceramics using RS and CS routes.
Figure 6. (a) Relative densities of BST ceramics; (b) εr of BST ceramics; (c) Q × f of BST ceramics (Inserted images show BST ceramics prepared by RS and CS routes sintered at 1400 °C); (d) TCF of BST ceramics using RS and CS routes.
Materials 17 03477 g006
Figure 7. XPS spectra of BST ceramics: (a) Ti 2p for RS route sintered at 1400 °C (b) Ti 2p for CS route sintered at 1400 °C.
Figure 7. XPS spectra of BST ceramics: (a) Ti 2p for RS route sintered at 1400 °C (b) Ti 2p for CS route sintered at 1400 °C.
Materials 17 03477 g007
Figure 8. (a) Calculated and measured infrared reflection spectrum. (b) Fitted complex dielectric response of BST ceramic using RS route.
Figure 8. (a) Calculated and measured infrared reflection spectrum. (b) Fitted complex dielectric response of BST ceramic using RS route.
Materials 17 03477 g008
Table 1. The refinement parameters of Ba4Sm28/3Ti18O54 ceramics for RS and CS process.
Table 1. The refinement parameters of Ba4Sm28/3Ti18O54 ceramics for RS and CS process.
MethodRSCS
S.T.1300 °C1350 °C1400 °C1450 °C1300 °C1350 °C1400 °C1450 °C
a (Å)12.152412.54812.147612.151312.15612.154812.159612.1601
b (Å)22.29822.300122.287822.294622.30722.303722.31222.3134
c (Å)7.66367.66377.6597.66147.66127.66087.66337.6637
V3)2076.9832077.2792073.652075.5672077.4562076.8412079.122079.458
α = β = γ9090909090909090
Rp (%)5.866.266.036.375.795.946.006.14
Rwp (%)7.588.047.628.237.447.698.018.37
χ2 (%)1.381.61.451.711.551.591.812.04
ρth (g/cm3)5.8835.8825.8935.8865.885.8845.8755.876
Table 2. The measured microwave dielectric characteristics (εr, Q × f, tanδ and TCF), and calculated theoretical permittivity (εthe), packing fraction (P.F.) and Ti-site bond valence of BST ceramics sintered at 1400 °C by RS and CS routes.
Table 2. The measured microwave dielectric characteristics (εr, Q × f, tanδ and TCF), and calculated theoretical permittivity (εthe), packing fraction (P.F.) and Ti-site bond valence of BST ceramics sintered at 1400 °C by RS and CS routes.
MethodMeasuredCalculated
εrQ × f (GHz)f
(GHz)
tanδTCF
(ppm/°C)
εtheP.F. (%)VTi
(v.u.)
RS (1400 °C)80.26 ± 0.59519 ± 1005.8826.18 × 10−4 ± 0.07 × 10−4−7.9 ± 143.2771.884.136
CS (1400 °C)79.38 ± 0.58230 ± 1006.2117.55 × 10−4 ± 0.09 × 10−4−12.1 ± 141.6571.694.109
Table 3. Binding energies and relative peak area ratio of Ti 2p spectra of BST ceramics using RS and CS routes.
Table 3. Binding energies and relative peak area ratio of Ti 2p spectra of BST ceramics using RS and CS routes.
MethodBinding Energy (eV)Binding Energy (eV)Relative Peak Area Ratio (%)
Ti2p1/2Ti2p3/2Ti3+/(Ti3+ + Ti4+)
IVIIIIVIII
RS (1400 °C)464463.5458.3457.811.87%
CS (1400 °C)464463.5458.3457.827.63%
Table 4. Microwave dielectric characteristics of BST ceramics prepared by different method.
Table 4. Microwave dielectric characteristics of BST ceramics prepared by different method.
NoMethodS.T.C.T.εrQ × fTCF Ref.
(°C)(°C) (GHz)(ppm/°C)
1Sol–gel 1360 °C for 3 h 1000 °C for 3 h80.810,099 [34]
2Spark plasma sintering (SPS) 1200 °C for 5 min1100 °C for 2 h81.210,099−17.2[35]
3Conventional solid-state (CS) 1360 °C for 3 h1200 °C for 3 h819240−10.6[27]
4Conventional solid-state (CS)1350 °C for 4 h1175 °C for 4 h7610,025−12[33]
5Conventional solid-state (CS)1400 °C for 4 h1150 °C for 3 h79.38 ± 0.58230 ± 100−12.1 ± 1This work
6Reaction sintering (RS)1400 °C for 4 hNo calcining80.26 ± 0.59519 ± 100−7.9 ± 1This work
Table 5. Phonon parameters obtained from the fitting of the infrared reflection spectrum of BST.
Table 5. Phonon parameters obtained from the fitting of the infrared reflection spectrum of BST.
Modeωojωpjγjεj
159.372150.3912.986.42
277.686215.0515.4597.66
394.164265.4512.7377.95
4110.92170.511.8412.36
5119.79143.969.80711.44
6136.27549.3820.39216.3
7180.78608.4447.30311.3
8220.41651.1840.4248.73
9254397.2633.8832.45
10276.38393.129.5632.02
11321.14233.5525.5390.529
12342.62193.125.3980.318
13364.24135.2924.3270.138
14389.39158.9122.8990.167
15415.22210.9244.1790.258
16455.19129.3431.5930.0807
17473.33121.3544.580.0657
18533.19212.2139.590.158
19559.52378.8241.7250.458
20584.43267.5742.1290.21
21617.5623843.3170.149
22732.0695.57642.1650.017
23776.8682.73365.1880.0113
24829.14158.2150.8950.0364
ε = 4.69, ε0 = 69.22.
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

Li, Z.; Zhou, H.; Xiong, G.; Wang, H.; Wang, G. Crystal Structure, Infrared Reflection Spectrum, and Improved Microwave Dielectric Characteristics of Ba4Sm28/3Ti18O54 Ceramics via One-Step Reaction Sintering. Materials 2024, 17, 3477. https://doi.org/10.3390/ma17143477

AMA Style

Li Z, Zhou H, Xiong G, Wang H, Wang G. Crystal Structure, Infrared Reflection Spectrum, and Improved Microwave Dielectric Characteristics of Ba4Sm28/3Ti18O54 Ceramics via One-Step Reaction Sintering. Materials. 2024; 17(14):3477. https://doi.org/10.3390/ma17143477

Chicago/Turabian Style

Li, Zeping, Huajian Zhou, Gang Xiong, Huifeng Wang, and Geng Wang. 2024. "Crystal Structure, Infrared Reflection Spectrum, and Improved Microwave Dielectric Characteristics of Ba4Sm28/3Ti18O54 Ceramics via One-Step Reaction Sintering" Materials 17, no. 14: 3477. https://doi.org/10.3390/ma17143477

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