[go: up one dir, main page]

Next Article in Journal
A Molecular Dynamics Simulation Study of Crystalline and Liquid MgO
Next Article in Special Issue
Tunable Optical Properties and Relaxor Behavior in Ni/Ba Co-Doped NaNbO3 Ceramics: Pathways Toward Multifunctional Applications
Previous Article in Journal
Rheological Behavior of an Algerian Natural Kaolin: Effect of Dispersant
Previous Article in Special Issue
Excellent Energy Storage and Photovoltaic Performances in Bi0.45Na0.45Ba0.1TiO3-Based Lead-Free Ferroelectricity Thin Film
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of (Ba1/3Nb2/3)4+ Substitution on Microstructure, Bonding Properties and Microwave Dielectric Properties of Ce2Zr3(MoO4)9 Ceramics

1
Engineering Technology Research Center of Preparation and Application of Industrial Ceramics of AnHui Province, Chaohu University, Hefei 238000, China
2
School of Environmental and Material Engineering, Yantai University, Yantai 264005, China
3
School of Materials Science and Engineering, University of Jinan, Jinan 250022, China
*
Authors to whom correspondence should be addressed.
Ceramics 2024, 7(3), 1172-1186; https://doi.org/10.3390/ceramics7030077
Submission received: 27 June 2024 / Revised: 5 August 2024 / Accepted: 17 August 2024 / Published: 29 August 2024
(This article belongs to the Special Issue Advances in Electronic Ceramics)
Figure 1
<p>XRD patterns of CZ<sub>1−<span class="html-italic">x</span></sub>N<span class="html-italic"><sub>x</sub></span> ceramics sintered at the optimal temperature with different <span class="html-italic">x</span> values.</p> ">
Figure 2
<p>Rietveld refinement of CZ<sub>1−<span class="html-italic">x</span></sub>N<span class="html-italic"><sub>x</sub></span> ceramics at the optimal sintering temperature with different <span class="html-italic">x</span> values.</p> ">
Figure 3
<p>Lattice parameter changes (<b>a</b>) <span class="html-italic">a</span> and <span class="html-italic">b</span>, (<b>b</b>) <span class="html-italic">c</span>, and (<b>c</b>) <span class="html-italic">V<sub>m</sub></span> of CZ<sub>1−<span class="html-italic">x</span></sub>N<span class="html-italic"><sub>x</sub></span> ceramics as a function of <span class="html-italic">x</span> values.</p> ">
Figure 4
<p>Crystal structure diagram of CZ<sub>1−<span class="html-italic">x</span></sub>N<span class="html-italic"><sub>x</sub></span> ceramics.</p> ">
Figure 5
<p>(<b>a</b>) Diameter shrinkage, (<b>b</b>) apparent density (the relative density at the optimal sintering temperature as a function of <span class="html-italic">x</span> values are shown in the inset) of CZ<sub>1−<span class="html-italic">x</span></sub>N<span class="html-italic"><sub>x</sub></span> ceramics at 600 to 700 °C.</p> ">
Figure 6
<p>Microstructure of CZ<sub>1−<span class="html-italic">x</span></sub>N<span class="html-italic"><sub>x</sub></span> ceramics at optimal sintering temperature with different <span class="html-italic">x</span> values.</p> ">
Figure 7
<p>CZ<sub>1−<span class="html-italic">x</span></sub>N<span class="html-italic"><sub>x</sub></span> ceramics (<b>a</b>) <span class="html-italic">ε<sub>r</sub></span> at a sintering temperature of 600 ℃ to 700 ℃, (<b>b</b>) <span class="html-italic">ε<sub>r</sub></span>, (<b>c</b>) <span class="html-italic">ε<sub>corr.</sub></span>, (<b>d</b>) <span class="html-italic">α<sub>theo.</sub></span>, and (<b>e</b>) <span class="html-italic">f<sub>iave.</sub></span><sub>(Ce-O)</sub> at the optimal sintering temperature as a function of <span class="html-italic">x</span> values.</p> ">
Figure 8
<p>CZ<sub>1−<span class="html-italic">x</span></sub>N<span class="html-italic"><sub>x</sub></span> ceramics (<b>a</b>) <span class="html-italic">Q</span>×<span class="html-italic">f</span> at a sintering temperature of 600 ℃ to 700 ℃, (<b>b</b>) <span class="html-italic">Q</span>×<span class="html-italic">f</span>, and (<b>c</b>) <span class="html-italic">U<sub>ave.</sub></span><sub>(Mo-O)</sub> at the optimal sintering temperature as a function of <span class="html-italic">x</span> values.</p> ">
Figure 9
<p>CZ<sub>1−<span class="html-italic">x</span></sub>N<span class="html-italic"><sub>x</sub></span> ceramics (<b>a</b>) <span class="html-italic">τ<sub>f</sub></span>, (<b>b</b>) <span class="html-italic">α<sub>ave.</sub></span><sub>(Ce-O)</sub>, and (<b>c</b>) <span class="html-italic">E<sub>ave.</sub></span><sub>(Mo-O)</sub> at the optimal sintering temperature as a function of <span class="html-italic">x</span> values.</p> ">
Figure 10
<p>(<b>a</b>) The design model and dimensions, (<b>b</b>) simulated S11 parameters, (<b>c</b>) 3D radiation pattern, (<b>d</b>) E-plane, and (<b>e</b>) H-plane of the antenna.</p> ">
Versions Notes

Abstract

:
In this study, Ce2[Zr1−x(Ba1/3Nb2/3)x]3(MoO4)9 (0.02 ≤ x ≤ 0.1, CZ1−xNx) ceramics were sintered at 600 °C and 700 °C using the traditional solid-state method. An analysis conducted through XRD and Rietveld refinement confirmed that all the CZ1−xNx ceramics displayed a single phase with a trigonal structure (space group R-3c). The observed increases in cell volume with increasing x values indicate the successful substitution of (Ba1/3Nb2/3)4+. The high densification of the synthesized phase was validated by the density and SEM results. Additionally, the P-V-L theory demonstrates a strong correlation between the Ce-O bond and εr, as well as τf, and between the Mo-O bond and Q×f. Notably, the CZ0.98N0.02 ceramics demonstrated superior performance at 675 °C, exhibiting εr = 10.41, Q×f = 53,296 GHz, and τf = −23.45 ppm/°C. Finally, leveraging CZ0.98N0.02 ceramics as substrate materials enabled the design of a patch antenna suitable for the 5G communication band, demonstrating its significant potential in this field.

1. Introduction

The advent of 5G has precipitated the rapid advancement of novel technologies such as autonomous driving, intelligent transportation, and smart manufacturing, marking the onset of an era characterized by the intelligent interconnection of all things. Microwave dielectric ceramics play a pivotal role in the evolution of 5G technology due to their substantial market size. Furthermore, 5G technology necessitates lower signal latency, increased information-carrying capacity, and broader coverage, imposing heightened requirements on related microwave components [1,2,3,4,5]. It is noteworthy that materials with low dielectric constants (εr) offer advantages such as enhanced signal transmission speed, reduced coupling loss between ceramic substrates and metal electrodes, and a high-quality factor (Q×f), rendering them the preferred choice for 5G technology. The nearly zero temperature coefficient of resonant frequency (τf) exhibited by ceramic substrates also constitutes an essential factor in maintaining signal stability within communication technology [6,7,8,9]. Furthermore, the utilization of low-temperature co-fired ceramics (LTCC) technology contributes to enhanced circuit density, facilitating device miniaturization and compliance with current environmental standards. This renders LTCC technology a highly promising multi-layer ceramic packaging solution. In LTCC technology, the firing temperature is typically maintained at a lower range of 800 °C to 900 °C to accommodate the sintering of ceramic materials with Au (melting temperature 1063 °C), Ag (melting temperature 961 °C), and Cu (melting temperature 1084 °C) electrode materials [10]. Notably, Mo-based ceramics have garnered attention from researchers due to their comparatively low sintering temperature and reduced losses [11,12].
In the double-molybdate system, Re2Zr3(MoO4)9 (Re = rare earth elements) has been reported to exhibit superior performance and can be synthesized at temperatures below 900 °C, showing great potential for development in LTCC applications [13,14,15]. Within this system, the Eu2Zr3(MoO4)9 ceramics demonstrate exceptional performance (εr = 10.75, Q×f = 74,900 GHz, τf = −8.88 ppm/°C), and they can be synthesized at 600 °C [15]. It is noteworthy that Tao et al. [12] and Shi et al. [16] have documented the low sintering temperature of Ce2Zr3(MoO4)9 (CZM) ceramics (575 °C and 750 °C respectively), along with their lowest τf value (−1.29 ppm/°C and −7.10 ppm/°C), indicating significant commercial application potential. However, the relatively low Q×f value (19,062 GHz and 24,720 GHz) of the CZM ceramics limits further scalability. Significant progress has been achieved in the doping and modification of CZM ceramics in recent years, representing an exciting development. Zheng et al. conducted studies on the substitution of Sn and Ti at the Zr site, resulting in Q×f value increases to 72,390 GHz and 84,200 GHz, respectively, while maintaining the τf value around −10 ppm/°C [17]. Furthermore, research on complex ion substitution at the Zr sites of CZM ceramics has expanded the range of substitution ions for performance enhancement. For instance, Xu et al. successfully elevated the Q×f value to 100,954 GHz at 750 °C by replacing the Zr site of CZM ceramics with (Zn1/3Sb2/3)4+. Xu et al. optimized the Q×f value to 59,381 GHz at 675 °C by substituting (Zn1/3Nb2/3)4+ for Zr in CZM ceramics [18]. Pan et al. also achieved an increased Q×f value of 82,696 GHz at 775 °C by replacing the Zr site with (Sr1/3Sb2/3)4+ [19]. It is evident that when using the pentavalent ion Nb for complex ion substitution in CZM ceramics, generally, a lower firing temperature is observed. Therefore, further research is warranted on (Ba1/3Nb2/3)4+ as a highly promising substitution ion in CZM ceramics.
This study presents a novel enhancement of the performance of CZM ceramics through the incorporation of (Ba1/3Nb2/3)4+ for the first time and, further, an examination of its impact on the microstructure and sintering characteristics of CZM ceramics. The Phillips-Van Vechten-Levine (P-V-L) valence bond theory is an analytical approach for investigating the correlation between properties and bonding characteristics. Therefore, the structure and properties of Ce2[Zr1−x(Ba1/3Nb2/3)x]3(MoO4)9 (0.02 ≤ x ≤ 0.1, CZ1−xNx) ceramics were elucidated using the P-V-L theory. Additionally, a patch antenna tailored for the 5G communication band was developed based on the CZ0.98N0.02 ceramic substrate.

2. Experimental Section

Following the stoichiometric ratio required for CZ1−xNx ceramics, high-purity CeO2 (>99.9%, Macklin, Shanghai, China), ZrO2 (>99.99%, Macklin, Shanghai, China), BaCO3 (>99.99%, Macklin, Shanghai, China), Nb2O5 (>99.99%, Macklin, Shanghai, China), and MoO3 (>99.5%, Macklin, Shanghai, China) powders were precisely weighed and synthesized via the conventional solid-phase reaction route. Initially, the weighed powders were a ball mill with zirconia balls and ethanol for 15 h of milling. Subsequently, the homogeneous slurry was extracted and subjected to drying at 80 °C in an oven. The resulting dried powders were then sieved through a 200-mesh screen and transferred to an alumina crucible. The precursor was formed by sintering the powder at 600 °C in a muffle furnace for 2 h. This precursor was then blended with 8 wt.% PVA before being pressed into cylindrical green bodies in a cylindrical mold under 200 MPa pressure. Finally, the cylindrical green bodies were placed in an alumina crucible and subjected to sintering at 600 to 700 °C in an air atmosphere for 4 h.
The phase of CZ1−xNx ceramics was determined using an X-ray diffractometer (XRD) model BRUKER D8 ADVANCE in ambient air at room temperature, and the data were refined employing the Rietveld method. The apparent density of CZ1−xNx ceramics was measured by the Archimedes displacement method at a test water temperature of 25.5 °C. The surface morphology of CZ1−xNx ceramics was characterized using a scanning electron microscope (SEM) model JEOL JSM-IT200. The εr and τf values of CZ1−xNx ceramics were measured utilizing a vector network analyzer, model N5234A, from Agilent for TE011 mode. The τf value was determined by measuring the resonant frequency of the sample at 25 °C (f25) and 85 °C (f85) using a temperature chamber, and then calculated using Equation (1). The Q×f value of CZ1−xNx ceramics was obtained through the closed-cavity method in TE01δ mode [9]. Furthermore, the patch antenna of CZ1−xNx ceramics was designed by initially calculating the values for radiation patch length (Lp), width (Wp), and feed position (Xf and Yf) based on εr and the designed center frequency (f0), followed by optimization using ANSYS HFSS software [20,21].
τ f = f 85 f 25 60 f 25
W p = c 2 f 0 ε r + 1 2 0.5
L p = c 2 f 0 ε e 2 Δ L
Δ L = 0.412 h ( ε e + 0.3 ) ( W p h + 0.264 ) ( ε e 0.258 ) ( W p h + 0.8 )
ε e = ε r + 1 2 + ε r 1 2 1 + 12 h W p 0.5
ξ re ( L ) = ε r + 1 2 + ε r 1 2 1 + 12 h L p 0.5
X f = L 2 ξ r e ( L )
Y f = 0
where c represents the speed of light, ΔL denotes the corrected length, εe signifies the effective dielectric constant, and h stands for the thickness of the CZ1−xNx ceramics substrate.

3. Results and Discussion

The XRD patterns of CZ1−xNx ceramics sintered at the optimum temperature are depicted in Figure 1. All diffraction peaks of the CZ1−xNx ceramics exhibit identical positions to those of Pr2Zr3(MoO4)9 (PDF# 51-1851), with no additional diffraction peaks observed [17,18]. This observation suggests that (Ba1/3Nb2/3)4+ was successfully incorporated into the crystal lattice of the CZM ceramics, forming a trigonal structure (space group R-3c) within the CZ1−xNx solid solution. Furthermore, as the value of x increases, all diffraction peaks of CZ1−xNx ceramics display a leftward shift, which is attributable to the larger ion radius of (Ba1/3Nb2/3)4+ (1.41 Å, CN = 6) compared to that of Zr4+ (0.72 Å, CN = 6) [22]. This finding further validates the successful substitution by (Ba1/3Nb2/3)4+ at Zr sites in the CZM ceramics. To investigate the influence of the (Ba1/3Nb2/3)4+ on the crystal structure of the CZM ceramics, the diffraction peaks of the CZ1−xNx ceramics were refined using the Rietveld method, as illustrated in Figure 2 and Table 1. Figure 2 demonstrates a good match between the measured and fitted diffraction peaks. Furthermore, the Rp, Rwp, and χ2 values obtained after the refinement are all within 10, indicating the reliability of the refined structure [23]. The crystal lattice parameters of the CZ1−xNx ceramics are depicted in Figure 3. Owing to the disparity in ionic radii, the collective values of a, b, c, and Vm exhibit an increasing trend with higher substitution levels, corresponding to the shift of the peaks observed in the XRD pattern.
The crystal structure of the CZ1−xNx ceramics is illustrated in Figure 4. The CZ1−xNx ceramics consist of [CeO9] polyhedra, [ZrO6] polyhedra, and [MoO4] tetrahedra interconnected with each other. Ce atoms are coordinated with 9 × O atoms (3 × O(1), 3 × O(2), and 3 × O(6)) to form [CeO9] polyhedra. Both Zr(1) and Zr(2) atoms are coordinated with 6 × O atoms (Zr(1): 6 × O(4), Zr(2): 3 × O(1) and 3 × O(5)) to form [ZrO6] polyhedra. Similarly, both Mo(1) and Mo(2) atoms are coordinated with four O atoms (Mo(1): O(1), O(2), O(3), O(4), Mo(2): 2 × O(5) and 2 × O(6)) to form [MoO4] tetrahedra. In the CZ1−xNx ceramics, the [CeO9] polyhedra and [ZrO6] polyhedra are not directly connected; they are linked through [MoO4] tetrahedra serving as a bridge [18,24].
The performance of ceramics is influenced by both intrinsic and extrinsic factors. Extrinsic factors, such as second phases and density, play a significant role. The XRD results confirmed the purity of the CZ1−xNx ceramics phases, thus necessitating an investigation into the relationship between sintering behavior and the performance of CZ1−xNx ceramics. Figure 5 illustrates the diameter shrinkage behavior and density changes of the CZ1−xNx ceramics with temperature. With the increase in the sintering temperature, the shrinkage rate of the CZ1−xNx ceramics with varying x values exhibited a pattern of initial augmentation followed by diminishment. This phenomenon is attributable to grain growth and pore elimination during sintering. However, excessive temperatures can induce secondary crystallization in the grains, resulting in reduced shrinkage rates. The trends observed in the apparent density align closely with those of the shrinkage rate. Furthermore, at the optimal sintering temperature, the relative density of the CZ1−xNx ceramics consistently exceeded 90%, underscoring their favorable densification.
Figure 6 illustrates the microstructural images of the CZ1−xNx ceramics at the optimal sintering temperature for each dopant content. The presence of atomic arrangement disorder and vacancies at grain boundaries has a detrimental effect on performance. Moreover, the well-developed regular grain morphology indicates favorable grain growth, contributing to performance enhancement [25,26]. Consequently, the CZ0.98N0.02 ceramics exhibit the largest grain size and most regular grain shape, positively impacting performance improvement. It is noteworthy that all the CZ1−xNx ceramics demonstrated high densities without discernible pores, which is consistent with the relative density results.
Figure 7a illustrates the temperature-dependent variation in the εr for the CZ1−xNx ceramics. The consistency between the εr and apparent density suggests that density is the primary factor influencing the dielectric constant of the ceramic. As evidenced by the presence of pores in the ceramic, where air has an εr value of 1, it was observed that the εr value of the CZ1−xNx ceramics significantly exceeded this baseline, indicating that increased porosity leads to a reduction in dielectric constant [19]. Consequently, considering the presence of pores, Equation (9) was employed to correct for the εr (εcorr.) of the CZ1−xNx ceramics, as depicted in Figure 7c, where P (P = 1 − ρrelative) is the porosity. Moreover, the relevant numerical values of the CZM ceramics matrix were derived from the reported findings of Tao et al. and Shi et al. [12,16]. The values of εr and εcorr. increased proportionally with x, a trend explicable by Equation (10) [27]. As per the Clausius–Mossotti equation, it is evident that εr exhibits a positive correlation with the polarizability (α). Additionally, based on the α values reported by Shannon et al. [28], the theoretical polarizability (αtheo.) of the CZ1−xNx ceramics was calculated using Equation (11).
ε c o r r . = ε r 1 + 1.5 P
ε r 1 ε r + 2 = 4 π α 3 V m
αtheo.(CZ1−xNx) = 2α(Ce3+) + 3(1−x)α(Zr4+) + (Ba2+) + 2(Nb5+) + 9α(Mo6+) + 36α(O2) = 2 × 6.15 + 3(1−x) × 3.25 + x × 6.04 + 2x × 3.97 + 9 × 3.28 + 36 × 2.01
Figure 7. CZ1−xNx ceramics (a) εr at a sintering temperature of 600 ℃ to 700 ℃, (b) εr, (c) εcorr., (d) αtheo., and (e) fiave.(Ce-O) at the optimal sintering temperature as a function of x values.
Figure 7. CZ1−xNx ceramics (a) εr at a sintering temperature of 600 ℃ to 700 ℃, (b) εr, (c) εcorr., (d) αtheo., and (e) fiave.(Ce-O) at the optimal sintering temperature as a function of x values.
Ceramics 07 00077 g007
Figure 7d illustrates the continuous increase in the αtheo. values of the CZ1−xNx ceramics as the x value increases, aligning with the observed trend in the εr values. This observation suggests that α plays a significant role as an internal factor contributing to the enhancement of εr values. Furthermore, the P-V-L theory serves as a widely adopted approach for investigating the correlation between material structure and performance [29,30]. Within CZ1−xNx ceramics, an analysis of the interna influence of (Ba1/3Nb2/3)4+ on the CZM ceramics was conducted through calculations of chemical bond variations using the P-V-L theory. Initially, complex chemical formulas were decomposed into binary bond formulas, as depicted in Equation (12).
Ce 2 [ ( Zr 1 x ( Ba 1 / 3 Nb 2 / 3 ) x ] 3 ( MoO 4 ) 9 Ce 2 / 3 O ( 1 ) 3 + Ce 2 / 3 O ( 2 ) 3 + Ce 2 / 3 O ( 6 ) 3 + Zr / ( Ba 1 / 3 Nb 2 / 3 ) ( 1 ) O ( 4 ) 3 + Zr / ( Ba 1 / 3 Nb 2 / 3 ) ( 2 ) O ( 3 ) 3 + Zr / ( Ba 1 / 3 Nb 2 / 3 ) ( 2 ) O ( 5 ) 3 + Mo ( 1 ) 3 / 2 O ( 1 ) 3 + Mo ( 1 ) 3 / 2 O ( 2 ) 3 + Mo ( 1 ) 3 / 2 O ( 3 ) 3 + Mo ( 1 ) 3 / 2 O ( 4 ) 3 + Mo ( 2 ) 3 / 2 O ( 5 ) 3 + Mo ( 2 ) 3 / 2 O ( 6 ) 3
In the P-V-L theory, the correlation between ionicity (fi) and εr was established, and the impact of (Ba1/3Nb2/3)4+ on the εr of the CZM ceramics was determined by evaluating the change in fi for each chemical bond. As depicted in Equation (13), there was a positive correlation between fi and εr. Furthermore, Equations (14)–(17) were utilized to compute the fi of each chemical bond in the CZ1−xNx ceramics [31,32].
ε r = n 0 2 1 1 f i + 1
f i μ = C μ 2 E g μ 2
E g μ 2 = E h μ 2 + C μ 2
E h μ 2 = 39.74 d μ 2.48
E h μ 2 = 39.74 d μ 2.48 C μ = 14.4 b μ exp k s μ r 0 μ Z A μ * n m Z B μ * / r 0 μ
where n0, dμ, and bμ denote the refractive index, bond length, and periodic correction factor, respectively. ( Z A μ )* and ( Z B μ )* represent the effective valence electron numbers of the cation and anion, respectively. Additionally, exp(− k s μ r 0 μ ) represents the Thomas-Fermi factor. The calculated fi values for each chemical bond in the CZ1−xNx ceramics are summarized in Table 2. Notably, the fi value of the Ce-O bond surpasses that of other chemical bonds. Moreover, the average fi value (fiave.(Ce-O)) for this bond was computed, as shown in Figure 7e. It was observed that the εr exhibited a similar upward trend to the fiave.(Ce-O), suggesting that the Ce-O bond exerts a dominant influence on the εr of CZ1−xNx ceramics.
Figure 8a illustrates the variation in the Q×f values for the CZ1−xNx ceramics at different temperatures. The trend in the Q×f values mirrors that of the apparent density, suggesting that density is the primary external factor influencing Q×f values. Lattice energy (U) is commonly linked to the Q×f values of ceramics. The magnitude of the U value reflects the stability of the ceramic, while higher U values indicate greater compound stability, thereby reducing nonharmonic vibration-induced losses and, consequently, enhancing the Q×f value of the ceramic. According to the P-V-L theory, the U values for the chemical bonds in the CZ1−xNx ceramics were computed using Equations (18)–(21) [33].
U c a l = μ U b μ
U b μ = U b c μ + U b i μ
U b c μ = 2100 m P A B 1.64 d 0.75 f c μ
U b i μ = 1270 ( m + n ) P A B P B A d ( 1 0.4 d ) f i μ
where U b c μ and U b i μ represent the covalent and ionic bond energies of the material, respectively. f c μ and f i μ represent the bond covalency and ionicity, respectively. P A B and P B A represent the valence states of the cation and anion, respectively. Table 3 presents the specific U values for each chemical bond. The Mo-O bond exhibits a notably higher U value compared to the other two chemical bonds, indicating its significant influence on the Q×f values of the ceramic. Additionally, Figure 8b,c illustrates the Q×f and average U (Uave.(Mo-O)) of the CZ1−xNx ceramics at optimal temperature. The sample with x = 0.02 demonstrates the highest CZ1−xNx value (53,296 GHz), significantly surpassing that of the CZM ceramics, suggesting that (Ba1/3Nb2/3)4+ serves as an effective substituted complex ion. Furthermore, Uave.(Mo-O) displays a similar trend to the Q×f value, underscoring the major influence of the Mo-O bond on the Q×f in the CZ1−xNx ceramics.
The τf of a ceramic material characterizes its stability in diverse environments, and the closer τf is to 0, the greater the stability of the ceramic. As per Equation (22), an inverse relationship exists between the thermal expansion coefficient (α) and the τf value of the ceramics. The α value of the CZ1−xNx ceramics was determined using Equations (23)–(26) derived from the P-V-L theory [34,35].
τ f = τ ε 2 + α
α = μ F m n μ α m n μ
α mn μ = 3.1685 + 0.8376 γ m n
γ m n = k Z A μ N C A μ U b μ A β m n
β m n = m m + n 2 n
where τε represents the temperature coefficient of the dielectric constant, N C A μ denotes the coordination number of the cation, Z A μ indicates the valence state of the cation, and k and ΔA stand for the Boltzmann constant and the ionic periodicity constant, respectively. The α values of the chemical bonds in the CZ1−xNx ceramics are detailed in Table 4. Among these ceramics, it was observed that the Ce-O bond exhibited the highest α value, thus exerting a significant influence on the τf value. As depicted in Figure 9, the average α value (αave.(Ce-O)) of the Ce-O bond demonstrated a similar trend to the τf. This suggests a strong correlation between the Ce-O bond and the τf value in the CZ1−xNx ceramics.
It is noteworthy that bond energy (E) is frequently employed to characterize the stability of materials. A higher E value indicates a stronger connection between chemical bonds. The E value of the CZ1−xNx ceramics was determined using Equations (27)–(31) [36].
E μ = t c E c μ + t i E i μ
E c μ = r c A + r c B d μ E A A E B B 1 / 2
E i μ = 1389.088 d μ
t i = S A S B 6
t c + t i = 1
Specifically, EA–A and EB–B represent the bond energies of atoms A and B, respectively, while rcA and rcB denote the covalent radii of ions A and B. The calculation results are presented in Figure 9c and Table 5. It is noteworthy that despite not displaying the highest E value, the Ce-O bond was indicated by the α results as linked to the τf value. Moreover, the average E value (Eave.(Mo-O)) for the Ce-O bond aligned with the trend of the τf, suggesting the significant role played by the Ce-O bond in ceramic stability, which is consistent with the α calculation outcomes.
Microstrip antennas, which are widely utilized in wireless communication and radar fields due to their simple design, ease of production, relatively low cost, and very narrow bandwidth, were the focus of this study. A rectangular microstrip antenna with a center frequency of 3.47 GHz was designed using a coaxial feeding method and CZ0.98N0.02 ceramics as the basic materials for the antenna. It is noteworthy that 3.47 GHz falls within the S-band of the microwave radio frequency commonly employed in wireless communication and radar fields. Furthermore, it is situated within the frequency range of the 5G communication band (3.3 GHz~4.2 GHz), specifically the n77 and n78 frequency bands [37,38,39].
Figure 10a illustrates the optimized antenna model and corresponding design dimensions centered at 3.47 GHz, with radiation patches and ground parts designed using silver (Ag). The return loss parameters (S11) based on the designed model are depicted in Figure 10b. It is evident from the figure that the antenna’s center frequency is 3.47 GHz, exhibiting a bandwidth of 75 MHz and an S11 value of −33.15 dB. Notably, when S11 falls below −10 dB, the minimal reflection impact on the transmission system allows for normal antenna operation. Furthermore, Figure 10c–e presents simulations of the antenna’s radiation in both three-dimensional and electromagnetic field views. The distribution of the main lobe and side lobe radiation from the antenna is clearly delineated, with a maximum radiation gain of 4.32 dB observed. Additionally, unidirectional and symmetrical maximum electric field and magnetic field gain directions indicate favorable radiation characteristics for this antenna design. Consequently, CZ0.98N0.02 ceramics exhibit significant potential for application in 5G communication.

4. Conclusions

In this work, CZ1−xNx (0.02 ≤ x ≤ 0.1) ceramics were synthesized via the conventional solid-state method. The XRD and Rietveld refinement analyses confirmed that all the CZ1−xNx ceramics exhibited a trigonal structure (space group R-3c) as a single phase. Furthermore, the increase in the unit cell volume with the increasing x value substantiated the successful substitution of (Ba1/3Nb2/3)4+. The density and SEM results validated the high density of the synthesized ceramics. Notably, the CZ0.98N0.02 ceramics exhibited the most uniform grain size, which enhances performance. Specifically, at 675 °C, the CZ0.98N0.02 ceramics demonstrated superior performance, with εr = 10.41, Q×f = 53,296 GHz, and τf = −23.45 ppm/°C. The P-V-L theory was employed to establish a correlation between structure and performance characteristics. The Ce-O bond predominantly influenced the εr and τf of the CZ1−xNx ceramics. Furthermore, the Mo-O bond was strongly correlated with Q×f. Finally, a patch antenna tailored for 5G communication bands (n77 and n78) was designed using the CZ0.98N0.02 ceramics as the substrate materials. This study represents an initial exploration of the practical applications of CZM ceramics.

Author Contributions

Conceptualization, H.G. and X.X.; methodology, H.G. and X.X.; software, H.G. and X.X.; validation, H.G., X.X., X.L. and X.Z.; formal analysis, H.G., X.X., X.L. and X.Z.; investigation, H.G., X.X., X.L. and X.Z.; resources, M.L., J.D. and H.W.; data curation, H.G.; writing—original draft preparation, H.G.; writing—review and editing, H.G., X.X., M.L., J.D. and H.W.; visualization, H.G. and X.X.; supervision, M.L., J.D. and H.W.; project administration, M.L., J.D. and H.W.; funding acquisition, M.L., J.D. and H.W. All authors have read and agreed to the published version of the manuscript.

Funding

The project was supported by the National Natural Science Foundation of China (52302116) and Graduate Innovation Foundation of Yantai University, GIFYTU. The authors are also thankful to Zeming Qi and Chuansheng Hu in the IR beamline workstation of the National Synchrotron Radiation Laboratory (NSRL) for the IR measurement. Thanks zkec (www.zkec.cc) for XRD and SEM examination.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shafi, M.; Molisch, A.F.; Smith, P.J.; Haustein, T.; Zhu, P.; Silva, P.D.; Tufvesson, F.; Benjebbour, A.; Wunder, G. 5G: A tutorial overview of standards, trials, challenges, deployment, and practice. IEEE J. Sel. Areas Commun. 2017, 35, 1201–1221. [Google Scholar] [CrossRef]
  2. Li, S.; Xu, L.D.; Zhao, S. 5G Internet of Things: A survey. J. Ind. Inf. Integr. 2018, 10, 1–9. [Google Scholar] [CrossRef]
  3. Yao, G.; Yan, J.; Tan, J.; Pei, C.; Liu, P.; Zhang, H.; Wang, D. Structure, chemical bond and microwave dielectric characteristics of novel Li3Mg4NbO8 ceramics. J. Eur. Ceram. Soc. 2021, 41, 6490–6494. [Google Scholar] [CrossRef]
  4. Luo, W.; Yan, S.; Zhou, J. Ceramic-based dielectric metamaterials. Interdiscip. Mater. 2022, 1, 11–27. [Google Scholar] [CrossRef]
  5. Tan, K.; Song, T.; Shen, T.; Yu, H.; Zhang, Y.; Cui, K.; Xu, X.; Li, W.; Wang, H. Research progress of low permittivity microwave dielectric ceramics. Adv. Ceram. 2022, 43, 11–29. [Google Scholar] [CrossRef]
  6. Dai, Y.; Chen, J.; Tang, Y.; Xiang, H.; Li, J.; Fang, L. Relationship between bond characteristics and microwave dielectric properties of REVO4 (RE = Yb, Ho) ceramics. Ceram. Int. 2023, 49, 875–881. [Google Scholar] [CrossRef]
  7. Pei, C.; Hou, C.; Li, Y.; Yao, G.; Ren, Z.; Liu, P.; Zhang, H. A low εr and temperature-stable Li3Mg2SbO6 microwave dielectric ceramics. J. Alloys Compd. 2019, 792, 46–49. [Google Scholar] [CrossRef]
  8. Pei, C.; Tan, J.; Li, Y.; Yao, G.; Jia, Y.; Ren, Z.; Liu, P.; Zhang, H. Effect of Sb-site nonstoichiometry on the structure and microwave dielectric properties of Li3Mg2Sb1−xO6 ceramics. J. Adv. Ceram. 2020, 9, 588–594. [Google Scholar] [CrossRef]
  9. Huang, X.; Guo, H.; Zhu, P.; Liu, L.; Xiao, J.; Tang, D.; Lin, C.; Wu, X.; Zheng, X. Microwave dielectric properties of CaCu3Ti4O12 ceramics: A clue to its intrinsic dielectric response. J. Adv. Dielectr. 2023, 13, 2344001–2344006. [Google Scholar] [CrossRef]
  10. An, Z.; Lv, J.; Wang, X.; Xu, Y.; Zhang, L.; Shi, F.; Guo, H.; Zhou, D.; Liu, B.; Song, K. Effects of LiF additive on crystal structures, lattice vibrational characteristics and dielectric properties of CaWO4 microwave dielectric ceramics for LTCC applications. Ceram. Int. 2022, 48, 29929–29937. [Google Scholar] [CrossRef]
  11. Zhou, D.; Pang, L.; Wang, D.; Guo, H.; Yang, F.; Qi, Z.; Li, C.; Jin, B.; Reaney, I.M. Crystal structure, impedance and broadband dielectric spectra of ordered scheelite-structured Bi(Sc1/3Mo2/3)O4 ceramic. J. Eur. Ceram. Soc. 2018, 38, 1556–1561. [Google Scholar] [CrossRef]
  12. Tao, B.J.; Xing, C.F.; Wang, W.F.; Wu, H.T.; Zhou, Y.Y. A novel Ce2Zr3(MoO4)9 microwave dielectric ceramic with ultra-low firing temperature. Ceram. Int. 2019, 45, 24675–24683. [Google Scholar] [CrossRef]
  13. Liu, W.; Zuo, R. Low temperature fired Ln2Zr3(MoO4)9 (Ln = Sm, Nd) microwave dielectric ceramics. Ceram. Int. 2017, 43, 17229–17232. [Google Scholar] [CrossRef]
  14. Liu, W.; Zuo, R. A novel low-temperature firable La2Zr3(MoO4)9 microwave dielectric ceramic. J. Eur. Ceram. Soc. 2018, 38, 339–342. [Google Scholar] [CrossRef]
  15. Zhang, Y.H.; Sun, J.J.; Dai, N.; Wu, Z.C.; Wu, H.T.; Yang, C.H. Crystal structure, infrared spectra and microwave dielectric properties of novel extra low-temperature fired Eu2Zr3(MoO4)9 ceramics. J. Eur. Ceram. Soc. 2019, 39, 1127–1131. [Google Scholar] [CrossRef]
  16. Shi, L.; Liu, C.; Zhang, H.; Peng, R.; Wang, G.; Shi, X.; Wang, X.; Wang, W. Crystal structure, Raman spectroscopy, metal compatibility and microwave dielectric properties of Ce2Zr3(MoO4)9 ceramics. Mater. Chem. Phys. 2020, 250, 122954. [Google Scholar] [CrossRef]
  17. Zheng, J.; Liu, Y.; Tao, B.; Zhang, Q.; Wu, H.; Zhang, X. Crystal structure and optimised microwave dielectric properties of Ce2(Zr1−xTix)3(MoO4)9 solid solutions. Ceram. Int. 2021, 47, 5624–5630. [Google Scholar] [CrossRef]
  18. Xu, X.; Xi, Z.; Feng, Z.; Zhang, X.; Du, W.; Shan, L.; Du, J.; Wu, H.; Wangsuo, X. Microstructure, bonding characteristics, far-infrared spectra and microwave dielectric properties of co-substituted Ce2[(Zr1−x(Zn1/3Sb2/3)x]3(MoO4)9 ceramics. Ceram. Int. 2024, 50, 24769–24780. [Google Scholar] [CrossRef]
  19. Pan, H.; Yan, S.; Zhang, Y.; Du, J.; Zhang, X.; Gong, P.; Wu, H.; Tian, H.; Wübbenhorst, M. Crystal structure, bond characteristics and microwave dielectric properties of Ce2[Zr1−x(Sr1/3B2/3)x]3(MoO4)9 (B = Ta, Sb) solid solutions. Ceram. Int. 2023, 49, 24038–24046. [Google Scholar] [CrossRef]
  20. Du, K.; Zhou, M.; Li, C.; Yin, C.; Cai, Y.; Cheng, M.; Zhu, W.; Wei, G.; Wang, S.; Lei, W. Ultralow-permittivity and temperature-stable Ba1−xCaxMg2Al6Si9O30 dielectric ceramics for C-Band patch antenna applications. ACS Appl. Mater. Interfaces 2024, 16, 23505–23516. [Google Scholar] [CrossRef]
  21. Siragam, S. Synthesis and fabrication of proto-type microstrip patch antenna by using microwave dielectric ceramic nanocomposite material for C-band application. Optik 2024, 306, 171787. [Google Scholar] [CrossRef]
  22. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Found. Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  23. McCusker, L.; Von Dreele, R.; Cox, D.; Louër, D.; Scardi, P. Rietveld refinement guidelines. J. Appl. Crystallogr. 1999, 32, 36–50. [Google Scholar] [CrossRef]
  24. Yang, H.; Chai, L.; Liang, G.; Xing, M.; Fang, Z.; Zhang, X.; Qin, T.; Li, E. Structure, far-infrared spectroscopy, microwave dielectric properties, and improved low-temperature sintering characteristics of tri-rutile Mg0.5Ti0.5TaO4 ceramics. J. Adv. Ceram. 2023, 12, 296–308. [Google Scholar] [CrossRef]
  25. Bao, J.; Zhang, Y.; Kimura, H.; Wu, H.; Yue, Z. Crystal structure, chemical bond characteristics, infrared reflection spectrum, and microwave dielectric properties of Nd2(Zr1−xTix)3(MoO4)9 ceramics. J. Adv. Ceram. 2023, 12, 82–92. [Google Scholar] [CrossRef]
  26. Wang, X.; Liu, T.; Cao, Z.; Li, Z.; Xu, Y.; Shi, F.; Zhang, L.; Qi, Z.-m. Lattice vibrational characteristics and structure-property relationships of Ca(Mg1/2W1/2)O3 microwave dielectric ceramics with different sintering temperatures. Ceram. Int. 2022, 48, 1415–1422. [Google Scholar] [CrossRef]
  27. Li, J.; Wang, Z.; Guo, Y.; Ran, S. Influences of substituting of (Ni1/3Nb2/3)4+ for Ti4+ on the phase compositions, microstructures, and dielectric properties of Li2Zn[Ti1−x(Ni1/3Nb2/3)x]3O8 (0 ≤ x ≤ 0.3) microwave ceramics. J. Adv. Ceram. 2023, 12, 760–777. [Google Scholar] [CrossRef]
  28. Shannon, R.D. Dielectric polarizabilities of ions in oxides and fluorides. J. Appl. Phys. 1993, 73, 348–366. [Google Scholar] [CrossRef]
  29. Phillips, J.C.; Van Vechten, J.A. Charge redistribution and piezoelectric constants. Phys. Rev. Lett. 1969, 23, 1115–1117. [Google Scholar] [CrossRef]
  30. Levine, B.F. Bond susceptibilities and ionicities in complex crystal structures. J. Chem. Phys. 1973, 59, 1463–1486. [Google Scholar] [CrossRef]
  31. Wu, Z.J.; Meng, Q.B.; Zhang, S.Y. Semiempirical study on the valences of Cu and bond covalency in Y1−xCaxBa2Cu3O6+y. Phys. Rev. B 1998, 58, 958–962. [Google Scholar] [CrossRef]
  32. Stepan, S.B. Dielectric methods of studying the chemical bond and the concept of electronegativity. Russ. Chem. Rev. 1982, 51, 684. [Google Scholar] [CrossRef]
  33. Berkov, D.V. Evaluation of the energy barrier distribution in many-particle systems using the path integral approach. J. Phys. Condens. Matter 1998, 10, L89. [Google Scholar] [CrossRef]
  34. Fang, W.; Chen, J.; Yang, Y.; Ao, L.; Tang, Y.; Li, J.; Fang, L. Anomalous microwave dielectric behaviour induced by the orthorhombic-tetragonal phase transition in CaLaGaO4 ceramics. J. Eur. Ceram. Soc. 2022, 42, 1474–1479. [Google Scholar] [CrossRef]
  35. Sun, Y.; Xiang, H.; Tang, Y.; Li, J.; Fang, L. Constructing the cationic rattling effect to realize the adjustability of the temperature coefficient in Nd2-xSmxO3 microwave dielectric ceramics. J. Eur. Ceram. Soc. 2024, 44, 2859–2865. [Google Scholar] [CrossRef]
  36. Wang, G.; Zhang, D.; Li, J.; Gan, G.; Rao, Y.; Huang, X.; Yang, Y.; Shi, L.; Liao, Y.; Liu, C.; et al. Crystal structure, bond energy, Raman spectra, and microwave dielectric properties of Ti-doped Li3Mg2NbO6 ceramics. J. Am. Ceram. Soc. 2020, 103, 4321–4332. [Google Scholar] [CrossRef]
  37. Li, F.; Li, Y.; Li, S.; Luo, Y.; Lu, Y.; Tang, T.; Liao, Y.; Zhang, J.; Wen, Q. All-ceramic array patch for 5G signal enhancement based on B-site substituted zinc-cobalt molybdate low temperature co-fired ceramics. Chem. Eng. J. 2023, 466, 143325. [Google Scholar] [CrossRef]
  38. Wu, F.F.; Zhou, D.; Du, C.; Xu, D.M.; Li, R.T.; Shi, Z.Q.; Darwish, M.; 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]
  39. Tian, H.; Zhang, X.; Zhang, Z.; Liu, Y.; Wu, H. Low-permittivity LiLn(PO3)4 (Ln = La, Sm, Eu) dielectric ceramics for microwave/millimeter-wave communication. J. Adv. Ceram. 2024, 13, 602–620. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of CZ1−xNx ceramics sintered at the optimal temperature with different x values.
Figure 1. XRD patterns of CZ1−xNx ceramics sintered at the optimal temperature with different x values.
Ceramics 07 00077 g001
Figure 2. Rietveld refinement of CZ1−xNx ceramics at the optimal sintering temperature with different x values.
Figure 2. Rietveld refinement of CZ1−xNx ceramics at the optimal sintering temperature with different x values.
Ceramics 07 00077 g002
Figure 3. Lattice parameter changes (a) a and b, (b) c, and (c) Vm of CZ1−xNx ceramics as a function of x values.
Figure 3. Lattice parameter changes (a) a and b, (b) c, and (c) Vm of CZ1−xNx ceramics as a function of x values.
Ceramics 07 00077 g003
Figure 4. Crystal structure diagram of CZ1−xNx ceramics.
Figure 4. Crystal structure diagram of CZ1−xNx ceramics.
Ceramics 07 00077 g004
Figure 5. (a) Diameter shrinkage, (b) apparent density (the relative density at the optimal sintering temperature as a function of x values are shown in the inset) of CZ1−xNx ceramics at 600 to 700 °C.
Figure 5. (a) Diameter shrinkage, (b) apparent density (the relative density at the optimal sintering temperature as a function of x values are shown in the inset) of CZ1−xNx ceramics at 600 to 700 °C.
Ceramics 07 00077 g005
Figure 6. Microstructure of CZ1−xNx ceramics at optimal sintering temperature with different x values.
Figure 6. Microstructure of CZ1−xNx ceramics at optimal sintering temperature with different x values.
Ceramics 07 00077 g006
Figure 8. CZ1−xNx ceramics (a) Q×f at a sintering temperature of 600 ℃ to 700 ℃, (b) Q×f, and (c) Uave.(Mo-O) at the optimal sintering temperature as a function of x values.
Figure 8. CZ1−xNx ceramics (a) Q×f at a sintering temperature of 600 ℃ to 700 ℃, (b) Q×f, and (c) Uave.(Mo-O) at the optimal sintering temperature as a function of x values.
Ceramics 07 00077 g008
Figure 9. CZ1−xNx ceramics (a) τf, (b) αave.(Ce-O), and (c) Eave.(Mo-O) at the optimal sintering temperature as a function of x values.
Figure 9. CZ1−xNx ceramics (a) τf, (b) αave.(Ce-O), and (c) Eave.(Mo-O) at the optimal sintering temperature as a function of x values.
Ceramics 07 00077 g009
Figure 10. (a) The design model and dimensions, (b) simulated S11 parameters, (c) 3D radiation pattern, (d) E-plane, and (e) H-plane of the antenna.
Figure 10. (a) The design model and dimensions, (b) simulated S11 parameters, (c) 3D radiation pattern, (d) E-plane, and (e) H-plane of the antenna.
Ceramics 07 00077 g010
Table 1. The lattice parameters and reliability factors of CZ1−xNx ceramics sintered at the optimal sintering temperatures.
Table 1. The lattice parameters and reliability factors of CZ1−xNx ceramics sintered at the optimal sintering temperatures.
xLattice ParameterReliability Factors
a = b (Å)c (Å)α = β (°)γ (°)Vm3)Rp (%)Rwp (%)χ2
0.029.840558.9013901204939.566.178.072.52
0.049.835658.8950901204934.085.256.651.70
0.069.833758.8883901204931.715.607.061.85
0.089.840458.9261901204941.584.966.311.49
0.109.844558.9552901204948.084.766.081.49
Table 2. The fi of the CZ1−xNx ceramics sintered at optimal temperature.
Table 2. The fi of the CZ1−xNx ceramics sintered at optimal temperature.
Bond Typex = 0.02x = 0.04x = 0.06x = 0.08x = 0.10
Ce-O(1) a0.83930.84900.84760.85060.8517
Ce-O(1) b0.83930.84900.84760.85060.8518
Ce-O(1) c0.83930.84900.84760.85070.8518
Ce-O(2) a0.83820.85090.84850.84930.8533
Ce-O(2) b0.83820.85090.84850.84930.8533
Ce-O(2) c0.83820.85100.84850.84940.8533
Ce-O(6) a0.86380.85610.85110.85340.8538
Ce-O(6) b0.86380.85610.85110.85340.8538
Ce-O(6) c0.86390.85610.85110.85340.8539
Zr(BaNb)1-O(4) × 60.78650.79130.79520.79230.7913
Zr(BaNb)2-O(3) a0.76660.78170.77810.78210.7835
Zr(BaNb)2-O(3) b0.76660.78170.77810.78210.7836
Zr(BaNb)2-O(3) c0.76660.78170.77820.78200.7836
Zr(BaNb)2-O(5) a0.76430.78760.77620.78030.7836
Zr(BaNb)2-O(5) b0.76440.78770.77620.78030.7836
Zr(BaNb)2-O(5) c0.76440.78770.77630.78030.7837
Mo1-O(1)0.70210.71700.70460.71030.7117
Mo1-O(2)0.70950.71910.70620.71790.7198
Mo1-O(3)0.71610.73700.72890.73120.7390
Mo1-O(4)0.70480.73150.72220.72670.7280
Mo2-O(5) × 20.71040.72510.72680.73070.7317
Mo2-O(6) × 20.62810.70370.70060.70300.7104
ac Three different bonds.
Table 3. The U (kJ/mol) of the CZ1−xNx ceramics sintered at optimal temperature.
Table 3. The U (kJ/mol) of the CZ1−xNx ceramics sintered at optimal temperature.
Bond Typex = 0.02x = 0.04x = 0.06x = 0.08x = 0.10
Ce-O(1) a10951100108010761085
Ce-O(1) b10951100108010751085
Ce-O(1) c10951100108010751085
Ce-O(2) a11031086107410841074
Ce-O(2) b11021085107410841074
Ce-O(2) c11021085107410841074
Ce-O(6) a9061048105510561070
Ce-O(6) b9061048105510561070
Ce-O(6) c9061048105510561070
Zr(BaNb)1-O(4) × 610,16110,65310,14510,47610,702
Zr(BaNb)2-O(3) a37263711367636653696
Zr(BaNb)2-O(3) b37263709367636653696
Zr(BaNb)2-O(3) c37263709367536663695
Zr(BaNb)2-O(5) a37613613370736943695
Zr(BaNb)2-O(5) b37613612370736943695
Zr(BaNb)2-O(5) c37603611370636943694
Mo1-O(1)43,58744,24545,02144,81045,295
Mo1-O(2)42,31443,91944,77343,58544,009
Mo1-O(3)41,12140,77740,86441,25740,646
Mo1-O(4)43,13341,78642,08842,07542,641
Mo2-O(5) × 242,15742,90041,25441,34341,993
Mo2-O(6) × 253,63146,31045,64045,94145,491
ac Three different bonds.
Table 4. The α (10−6/K) of the CZ1−xNx ceramics sintered at optimal temperature.
Table 4. The α (10−6/K) of the CZ1−xNx ceramics sintered at optimal temperature.
Bond Typex = 0.02x = 0.04x = 0.06x = 0.08x = 0.10
Ce-O(1) a10.222710.161910.408710.459210.3461
Ce-O(1) b10.222710.161910.408710.471910.3461
Ce-O(1) c10.222710.161910.408710.471910.3461
Ce-O(2) a10.125610.333710.484610.358610.4846
Ce-O(2) b10.137710.346110.484610.358610.4846
Ce-O(2) c10.137710.346110.484610.358610.4846
Ce-O(6) a13.016310.823310.730410.717310.5356
Ce-O(6) b13.016310.823310.730410.717310.5356
Ce-O(6) c13.016310.823310.730410.717310.5356
Zr(BaNb)1-O(4) × 63.84453.51033.83403.60243.4493
Zr(BaNb)2-O(3) a3.20643.22243.27343.28283.2189
Zr(BaNb)2-O(3) b3.20643.22583.27343.28283.2189
Zr(BaNb)2-O(3) c3.20643.22583.27513.28113.2207
Zr(BaNb)2-O(5) a3.14713.39573.21953.23223.2207
Zr(BaNb)2-O(5) b3.14713.39763.21953.23223.2207
Zr(BaNb)2-O(5) c3.14883.39943.22123.23223.2224
Mo1-O(1)−0.4054−0.4465−0.4934−0.4808−0.5096
Mo1-O(2)−0.3223−0.4263−0.4786−0.4053−0.4319
Mo1-O(3)−0.2397−0.2150−0.2213−0.2494−0.2055
Mo1-O(4)−0.3763−0.2863−0.3070−0.3061−0.3441
Mo2-O(5) × 2−0.3117−0.3612−0.2492−0.2554−0.3005
Mo2-O(6) × 2−0.9229−0.5679−0.5297−0.5470−0.5211
ac Three different bonds.
Table 5. The E (kJ/mol) of the CZ1−xNx ceramics sintered at optimal temperature.
Table 5. The E (kJ/mol) of the CZ1−xNx ceramics sintered at optimal temperature.
Bond Typex = 0.02x = 0.04x = 0.06x = 0.08x = 0.10
Ce-O(1) a406.8505410.6544401.4945399.9166404.4490
Ce-O(1) b406.7539410.5723401.4004399.8233404.3694
Ce-O(1) c406.6895410.5067401.3377399.7611404.3058
Ce-O(2) a410.0645404.6082398.7994403.8610399.6523
Ce-O(2) b410.0155404.5445398.7530403.8133399.6057
Ce-O(2) c409.9500404.4967398.6911403.7499399.5435
Ce-O(6) a326.0845388.2832390.5555391.2096397.8577
Ce-O(6) b326.0742388.2539390.5109391.1798397.8269
Ce-O(6) c326.0224388.1952390.4516391.1053397.7653
Zr(BaNb)1-O(4) × 6459.2762489.8926461.1076481.5940496.2735
Zr(BaNb)2-O(3) a518.9672518.3938512.9597512.3371519.4946
Zr(BaNb)2-O(3) b518.8108518.2381512.9597512.3623519.3393
Zr(BaNb)2-O(3) c518.7847518.2122512.7823512.5390519.3134
Zr(BaNb)2-O(5) a525.3543500.8407518.6251517.7693519.3134
Zr(BaNb)2-O(5) b525.2741500.7680518.5733517.7177519.2358
Zr(BaNb)2-O(5) c525.1139500.6227518.4178517.5632519.1066
Mo1-O(1)592.0846602.8673619.4640614.6348623.9548
Mo1-O(2)567.3968596.3924614.4154590.1899598.0154
Mo1-O(3)544.9440536.5641538.9721545.7791534.0395
Mo1-O(4)583.1916555.2862561.8104561.0780571.3516
Mo2-O(5) × 2564.4201576.4860546.1831547.3699559.0432
Mo2-O(6) × 2820.3593645.2379632.2044637.9052627.9761
ac Three different bonds.
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

Gao, H.; Xu, X.; Liu, X.; Zhang, X.; Li, M.; Du, J.; Wu, H. Effect of (Ba1/3Nb2/3)4+ Substitution on Microstructure, Bonding Properties and Microwave Dielectric Properties of Ce2Zr3(MoO4)9 Ceramics. Ceramics 2024, 7, 1172-1186. https://doi.org/10.3390/ceramics7030077

AMA Style

Gao H, Xu X, Liu X, Zhang X, Li M, Du J, Wu H. Effect of (Ba1/3Nb2/3)4+ Substitution on Microstructure, Bonding Properties and Microwave Dielectric Properties of Ce2Zr3(MoO4)9 Ceramics. Ceramics. 2024; 7(3):1172-1186. https://doi.org/10.3390/ceramics7030077

Chicago/Turabian Style

Gao, Huamin, Xiangyu Xu, Xinwei Liu, Xiaoyu Zhang, Mingling Li, Jialun Du, and Haitao Wu. 2024. "Effect of (Ba1/3Nb2/3)4+ Substitution on Microstructure, Bonding Properties and Microwave Dielectric Properties of Ce2Zr3(MoO4)9 Ceramics" Ceramics 7, no. 3: 1172-1186. https://doi.org/10.3390/ceramics7030077

Article Metrics

Back to TopTop