Next Article in Journal
Improving of Thermoelectric Efficiency of Layered Sodium Cobaltite Through Its Doping by Different Metal Oxides
Next Article in Special Issue
Structural and Dielectric Impedance Studies of Mixed Ionic–Electronic Conduction in SrLaFe1−xMnxTiO6 (x = 0, 0.33, 0.67, and 1.0) Double Perovskites
Previous Article in Journal / Special Issue
Study on the Absorbing Properties of V-Doped MoS2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Ni2+ Doping on the Crystal Structure and Properties of LiAl5O8 Low-Permittivity Microwave Dielectric Ceramics

Naval University of Engineering, Wuhan 430030, China
*
Authors to whom correspondence should be addressed.
Ceramics 2025, 8(3), 85; https://doi.org/10.3390/ceramics8030085
Submission received: 6 May 2025 / Revised: 26 June 2025 / Accepted: 1 July 2025 / Published: 4 July 2025
(This article belongs to the Special Issue Advances in Electronic Ceramics, 2nd Edition)

Abstract

Low-permittivity microwave dielectric ceramics are essential for high-frequency communication and radar systems, as they minimize signal delay and interference, thereby enabling compact and high-performance devices. In this study, LiAl5−xNixO8−0.5x (x = 0.1–0.5) ceramics were synthesized via a solid-state reaction method to investigate the effects of Ni2+ substitution on crystal structure, microstructure, and dielectric properties. X-ray diffraction and Rietveld refinement reveal a phase transition from the P4332 to the Fd 3 m spinel structure at x ≈ 0.3, accompanied by a systematic increase in the lattice parameter (7.909–7.975 Å), attributed to the larger ionic radius of Ni2+ compared to Al3+. SEM analysis confirms dense microstructures with relative densities exceeding 95% and grain size increases from less than 1 μm at x = 0.1 to approximately 2 μm at x = 0.5. Dielectric measurements show a decrease in permittivity (εr) from 8.24 to 7.77 and in quality factor (Q × f) from 34,605 GHz to 20,529 GHz with increasing Ni content, while the temperature coefficient of the resonant frequency (τf) shifts negatively from −44.8 to −69.1 ppm/°C. Impedance spectroscopy indicates increased conduction losses and reduced activation energy with higher Ni2+ concentrations.

1. Introduction

Microwave dielectric ceramics are essential components in modern communication and radar systems, enabling high-frequency signal transmission with minimal loss and interference. With the rapid advancement of wireless communication networks, satellite broadcasting, and next-generation radar technologies, there is an increasing demand for materials with well-tailored dielectric properties [1]. Among these, low-permittivity (εr) ceramics are particularly critical for applications requiring high signal propagation speeds, reduced electromagnetic interference, and compact device design [2]. Such materials are widely employed in high-frequency electronic components, including substrates, resonators, filters, and antennas. Their ability to reduce signal delay and dispersion makes them indispensable in cutting-edge technologies such as 5G technology, satellite communications, and phased-array radar systems. Therefore, the development of new low-permittivity microwave dielectric ceramics is vital to meet the stringent performance requirements of emerging telecommunication and sensing applications [3,4].
LiAl5O8 belongs to the family of lithium aluminates and typically crystallizes in a spinel-related structure. It has attracted considerable attention due to its interesting structural, electrochemical, luminescent, and thermal properties. LiAl5O8 exhibits polymorphism, with two distinct structural forms depending on the synthesis temperature and conditions. At high temperatures (above approximately 1295 ± 5 °C), LiAl5O8 adopts a cubic spinel-like phase (space group: Fd 3 m) with a unit cell parameter of approximately 7.921 Å. Upon cooling, it undergoes a reconstructive first-order transition to a primitive cubic phase (space group: P4332), with a slightly smaller unit cell parameter of about 7.907 Å. The formation of ordered or disordered polymorphs is highly sensitive to processing parameters such as temperature and duration, making LiAl5O8 a structurally tunable material. And it has attracted significant attention in various advanced applications. In lithium-ion and all-solid-state batteries, LiAl5O8 acts as an effective coating material because of its low lithium-ion migration barrier and broad electrochemical stability window. Its favorable ionic conductivity facilitates lithium transport across electrode interfaces, thereby improving cycling stability and overall battery performance [5,6]. Additionally, when processed into thin membranes through techniques such as flame spray pyrolysis and tape casting, LiAl5O8-based composites exhibit promising ionic conductivities, making them suitable candidates for ceramic electrolytes [7]. In nuclear applications, LiAl5O8 forms as a secondary phase within γ-LiAlO2 ceramics used in tritium-producing burnable absorber rods. First-principles studies of its surfaces have revealed important insights into tritiated water formation and tritium release under irradiation [8]. Furthermore, doping LiAl5O8 with transition metal ions such as Fe3+ has led to the development of luminescent nanothermometers operating within the first biological optical window, with temperature-sensitive emission properties ideal for non-contact thermal sensing [9].
Recently, LiAl5O8 ceramics [10] have been reported as a promising microwave dielectric material due to their low εr and relatively good microwave dielectric performance (εr  =  8.43, Q × f  =  49,300 GHz and τf = −38 ppm/°C), although they require a high sintering temperature of 1600 °C. Subsequently, Ao et al. [11] investigated LiGa5O8, in which the Al3+ lattice site was occupied by the isovalent and chemically similar Ga3+ ion, and demonstrated that this substitution also yields excellent microwave dielectric properties. Yang et al. [12] then optimized the microwave dielectric performance of LiGa5O8 by replacing the Ga3+ site with the mixed-valence (Zn0.5Ti0.5)3+ ion. Prior studies have shown that, owing to the similar ionic radii of Ni2+ and Al3+, partial replacement of Al3+ by Ni2+ in both MgAl2O4 and LiGa5O8 matrices significantly enhances their luminescent behavior [13,14]. Moreover, Li et al. [15] explored the substitution of Al3+ with Ni2+ in BaAl2Si2O8 microwave dielectric ceramics as a strategy to improve their microwave dielectric properties. Inspired by these findings, in this work, LiAl5−xNixO8−0.5x (x = 0.1–0.5) microwave dielectric ceramics were synthesized via a solid-state reaction by partially substituting Al3+ with Ni2+ in the LiAl5O8 material.

2. Materials and Methods

LiAl5−xNixO8−0.5x (x = 0.1–0.5) microwave dielectric ceramics were prepared by a solid-state reaction route combined with prolonged mechanical grinding to enhance particle activation before high-temperature sintering. Stoichiometric amounts of Li2CO3, Al2O3, and NiO reagents were mixed with deionized water and milled at 360 rpm for 5 h using zirconia balls of 2 mm, 3 mm, and 5 mm diameter, in a weight ratio of 5:3:1 (250 g:150 g:50 g). The ball-to-powder weight ratio was maintained at 15:1 to ensure effective homogenization. The dried powder was calcined in a furnace of 1100 °C for 3 h. The calcined powder was mixed with 10% PVA and sieved. The granulated powders were pressed into disks and then sintered at 1500 °C for 6 h.
The crystal structures of LiAl5−xNixO8−0.5x (x = 0.1–0.5) ceramics were characterized using a Shimadzu XRD-7000 X-ray diffractometer (XRD, Shimadzu Corporation, Kyoto, Japan) with Cu Kα radiation (λ = 1.5406 Å). The diffraction patterns were collected over a 2θ range of 10–80°. For phase identification, data were acquired at a scan speed of 2°/min. For Rietveld full-pattern fitting, high-resolution data were collected with a step size of 0.2°/min to ensure accurate structural refinement. Lattice parameters were refined using the GSAS software package (1st version) [16]. The bulk densities of the ceramic samples were measured by the Archimedes principle, and the relative densities were calculated based on the theoretical densities obtained from the refined lattice parameters. Impedance spectroscopy of the samples was analyzed in the frequency range of 100 Hz to 10 MHz using an impedance analyzer. The microstructures of the sintered ceramics were observed by scanning electron microscopy (SEM, FEI Sirion 200, Eindhoven, The Netherlands). Dielectric properties, including the εr and the Q × f values, were measured using the Hakki–Coleman method. In this technique, cylindrical dielectric samples were placed between two parallel conducting plates and operated in the TE01δ resonant mode using a network analyzer (Agilent E8362B, Santa Clara, CA, USA) [17]. Moreover, the τf value of the ceramics was calculated using Formula (1):
τ f = [ f ( T 1 ) f ( T 0 ) ] / [ f ( T 0 ) · ( T 1 T 0 ) ] ,
where f(T1) and f(T0) denote the resonant frequencies at T1 (80 °C) and T0 (30 °C), respectively.

3. Discussion

The central graph in Figure 1 presents the XRD pattern of LiAl5−xNixO8−0.5x (x = 0.1–0.5) microwave dielectric ceramics under sintering at 1500 °C. Comparison with standard cards indicates that when 0.1 ≤ x ≤ 0.2, the samples crystallize in the P4332 space group. When 0.3 ≤ x ≤ 0.5, a significant number of diffraction peaks disappear, as clearly seen in Figure 1b,c. The leftmost panel in Figure 1 shows an enlarged view of the diffraction region of 10–30° for x = 0.2 and x = 0.3. Based on the standard cards, a phase transition is observed at x ≈ 0.3, with samples of 0.3 ≤ x ≤ 0.5 exhibiting a spinel structure (Fd 3 m space group). Therefore, in LiAl5O8 materials, Al3+ ions are substituted by Ni2+ ions, and as the substitution content increases, a phase transition in LiAl5O8 materials is induced, occurring at approximately x = 0.3. The rightmost panel of Figure 1 presents the results of the main peaks measured at diffraction angles between 37° and 38° for different substitution contents. The peaks shift to lower angles with increasing Ni2+ content. This is attributed to the fact that the ionic radius of Al3+ is smaller than that of Ni2+. According to Bragg’s equation, 2 d sin θ = n λ , an increase in the lattice constant leads to a larger interplanar spacing d, which results in a lower diffraction angle θ, shifting the peaks to lower angles.
To investigate the effect of Ni2+ substitution for Al3+ ions on the crystal structure of LiAl5O8 materials, Rietveld full-pattern refinement was performed on the X-ray diffraction data. Figure 2 presents the Rietveld full-pattern fitting results for LiAl5−xNixO8−0.5x materials at x = 0.1 and x = 0.3. In the figure, red dots represent the experimental data, the black line corresponds to the calculated diffraction pattern, vertical bars indicate the positions of Bragg peaks, and the purple line shows the difference between the calculated and experimental values. The figure clearly demonstrates that, for samples with different space groups, the fitted values are in excellent agreement with the experimental measurements.
Table 1 summarizes the structural parameters of LiAl5−xNixO8−0.5x (x = 0.1 and x = 0.3) microwave dielectric ceramics, obtained from Rietveld full-pattern refinement. These parameters include atomic positions, site occupancies, and the R factor and χ2 values from the fitting process. The low R factor and χ2 values, together with the visual results in Figure 2, clearly indicate that the Rietveld full-pattern refinement method produces highly reliable results.
In addition, the Rietveld full-pattern refinement method enables the determination of the lattice parameters for the LiAl5−xNixO8−0.5x microwave dielectric ceramics. Given the cubic symmetry of the structure, the unit cell volume was calculated directly from the lattice parameter. Then the theoretical (calculated) densities were obtained from the refined crystal parameter. The actual densities of the ceramic samples, measured using the Archimedes principle, are listed in Table 2. Based on these values, the relative densities were determined as the ratio of the measured density to the calculated density. Since the LiAl5−xNixO8−0.5x ceramics are single-phase, the relative density values are considered reliable. The results presented in Table 2 show that the lattice parameters of the LiAl5−xNixO8−0.5x ceramics increase with the substitution level of Ni2+ ions. Specifically, when x increases from 0.1 to 0.5, the lattice parameter a expands from 7.909 to 7.975 Å. Based on the ionic radii of Al3+ (0.535 Å) and Ni2+ (0.69 Å) reported in reference [18], the substitution of smaller Al3+ ions with larger Ni2+ ions leads to an expansion of the crystal lattice. Furthermore, all samples exhibit relative densities exceeding 95%, indicating that dense ceramic structures were successfully obtained under sintering at 1500 °C.
Figure 3 shows the SEM micrographs and corresponding grain size distributions of the LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics. The micrographs clearly demonstrate that all samples sintered at 1500 °C exhibit minimal porosity, resulting in dense microstructures consistent with the high relative densities reported in Table 2. Additionally, the micrographs indicate that the grain size in the sample with x = 0.1 is less than 1 μm. As the Ni2+ content increases, a clear trend toward larger grain sizes is observed; for instance, at x = 0.2, the grain size reaches approximately 2 μm. According to previous studies [10], LiAl5O8 ceramics sintered at 1600 °C retain a certain degree of porosity, whereas LiAl5−xNixO8 (0.1 ≤ x ≤ 0.5) ceramics sintered at 1500 °C exhibit negligible porosity and achieve a significantly higher densification. Therefore, the partial substitution of Al3+ by Ni2+ in LiAl5O8 not only reduces the required sintering temperature but also promotes densification. Moreover, the average grain size increases progressively with higher Ni2+ content.
Figure 4 presents the microwave dielectric properties of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) ceramics, with the black line indicating the variation in the εr as a function of x. As the Ni2+ substitution level increases from 0.1 to 0.5, the εr decreases gradually from 8.24 to 7.77. According to previous studies [19], εr is typically affected by factors such as porosity, microstructure, polarizability, and the presence of secondary phases. In this work, the XRD data, relative density measurements, and SEM observations confirm that the ceramics prepared via the solid-state reaction are single-phase and highly densified. Therefore, the variation in the εr of LiAl5−xNixO8−0.5x microwave dielectric ceramics is closely related to their microstructural characteristics. Moreover, according to Shannon’s findings [20], the total dielectric polarizability ( α D T ) of a material can be expressed as the sum of the individual ionic polarizabilities (αD). Consequently, the overall dielectric polarizability ( α D T ) for LiAl5−xNixO8−0.5x can be calculated using the following Equation (2):
α D T ( L i A l 5 x Zn x O 8 0.5 x ) = α ( L i + ) + ( 5 x ) α ( A l 3 + ) + x α ( Z n 2 + ) + ( 8 0.5 x ) α ( O 2 )
Subsequently, employing the Clausius–Mossotti equation (Equation (3)), the calculated dielectric constant εc can be calculated by the following expression:
ε c = 1 + 2 b α D T / V m 1 b α D T / V m
where b = 4π/3, and Vm is the molar volume obtained from Rietveld refinement. The ionic polarizabilities of Li+, Al3+, Ni2+, and O2− are 1.20 Å3, 0.78 Å3, 1.23 Å3, and 2.00 Å3, respectively [19]. The data in Table 3 reveal that the calculated values are in good agreement with the experimental measurements and exhibit a similar decreasing trend.
Figure 4 illustrates the variation in the Q × f value of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics as a function of the Ni2+ substitution level, as indicated by the green curve. It is evident that with increasing Ni2+ substitution from x = 0.1 to x = 0.5, the Q × f value significantly decreases from 34,605 GHz to 20,529 GHz. The quality factor serves as an indicator of the dielectric loss in ceramic materials, and impedance spectroscopy is a widely adopted technique for analyzing such losses. Figure 5 shows the complex impedance plot for the LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) ceramic samples. Figure 6 presents the impedance analysis for the sample with x = 0.1, where the data are fitted using an equivalent circuit composed of two parallel R-CPE (resistor–constant phase element) units connected in series. Each R-CPE unit represents a different electrical response: one corresponds to the bulk (grain) contribution and is denoted with the subscript b, while the other corresponds to the grain boundary contribution and is denoted with the subscript gb. Since the circuit elements do not behave as ideal capacitors, a constant phase element (CPE) is used to replace the ideal capacitor in the model. The CPE introduces a parameter, n, which reflects the deviation from ideal capacitive behavior and varies between 0 and 1; n = 1 corresponds to an ideal capacitor, and n = 0 corresponds to a pure resistor. The fitted values for grain resistance (Rb), grain boundary resistance (Rgb), and the CPE exponent (n) are listed in Figure 6 for each temperature. Subsequently, using Equation (4), the Rb is converted into the corresponding bulk conductivity σb.
σ b = 1 R b 4 t π D 2
Here, t and D denote the sample thickness and diameter, respectively. The activation energy Ea is determined by fitting the temperature dependence of σb to the Arrhenius equation, as shown in Figure 7. As x increases from 0.1 to 0.5, activation energy (Ea) decreases from 2.420 eV to 1.134 eV. For these materials, the intrinsic band gap (Eg) and the Ea are related by Eg ≈ 2Ea [21,22,23]. Therefore, impedance spectroscopy confirms that increasing the acceptor doping of Ni2+ ions leads to higher conduction losses, which in turn reduce the Q × f values. Moreover, increasing the Ni2+ doping concentration also leads to a higher concentration of oxygen vacancies, thereby contributing to increasing the dielectric loss.
Figure 4 shows the variation in τf of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics with increasing x, as indicated by the purple line. As the Ni2+ content increases from 0.1 to 0.5, τf clearly shifts in the negative direction from −44.8 ppm/°C to −69.1 ppm/°C. Figure 8 further presents the relationship between τf and the unit cell volume (V) for these ceramics. The trend in τf for Ni2+ substitution is analogous to that for the Zn2+ substitution of a previous study [10]. This similarity arises because τf is inherently linked to the crystal structure, and the structural changes induced by Ni2+ are comparable to those caused by Zn2+. Therefore, the increasingly negative τf with higher Ni2+ content is consistent with the observed experimental data.

4. Conclusions

In this work, a series of Ni-doped LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics were synthesized via a solid-state reaction method. XRD and Rietveld refinement revealed a distinct phase boundary at x ≈ 0.3: compositions with x ≤ 0.2 retain the ordered P4332 structure of LiAl5O8, whereas higher Ni contents stabilize the Fd 3 m spinel phase. The progressive shift of diffraction peaks toward lower angles, along with a linear increase in the lattice parameter from 7.909 Å to 7.975 Å, confirms the successful substitution of larger Ni2+ ions for the Al3+ sublattice. SEM observations and density measurements indicate that all compositions sintered at 1500 °C exhibit high densification (>95% relative density) with minimal porosity. Dielectric measurements show that Ni2+ doping systematically reduces the permittivity (εr decreases from 8.24 to 7.77) and deteriorates the quality factor (Q × f drops from 34,605 GHz to 20,529 GHz) with increasing x. Meanwhile, the τf becomes more negative (−44.8 to −69.1 ppm/°C), indicating increased sensitivity to thermal expansion and polarizability variations in the Ni-substituted lattice. Impedance spectroscopy further reveals that higher Ni content promotes conduction losses and reduces activation energy, thereby accounting for the observed decline in Q × f. Overall, the controlled Ni2+ incorporation into LiAl5O8 provides an effective strategy for tuning crystal structure, microstructure, and dielectric performance, rendering these ceramics promising candidates for next-generation microwave substrates and resonator components.

Author Contributions

Writing—original draft preparation, X.L.; investigation, X.L.; data curation, X.L. and H.T.; methodology, B.T.; writing—review and editing, B.C.; supervision, B.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was sponsored by the talent plan of Naval University of Engineering.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sun, Y.; Xiang, H.C.; 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]
  2. Bao, J.; Zhang, K.H.; Wang, W.; Liu, Z.Y.; Fang, Z.; Wang, X.; Wang, C.H.; Li, Y.C.; He, G.Q.; Zhou, T.; et al. A Series of Ultra-low Permittivity ALaP4O12 (A = Li, Na, K) Metaphosphate Microwave Dielectric Ceramics for Ultra-wideband Dielectric Resonant Antenna Application. ACS Appl. Mater. Interfaces 2024, 16, 58898–58911. [Google Scholar] [CrossRef] [PubMed]
  3. Huang, Y.W.; Yang, X.H.; Zhang, Y.C. Novel single-phase Li2SiO3 microwave dielectric ceramic with low permittivity. J. Eur. Ceram. Soc. 2025, 45, 116940. [Google Scholar] [CrossRef]
  4. Wang, H.Y.; Li, X.Y.; Xu, X.W.; Cheng, J.; Guo, B.; Chen, X.Q.; Wang, H. MgF2-based microwave dielectric ceramics with ultra-low sintering temperature and high thermal expansion coefficient. J. Eur. Ceram. Soc. 2024, 44, 2150–2156. [Google Scholar] [CrossRef]
  5. Mo, S.; Zhang, B.K.; Zhang, K.C.; Li, S.N.; Pan, F. LiAl5O8 as a potential coating material in lithium-ion batteries: A first principles study. Phys. Chem. Chem. Phys. 2019, 21, 13758–13765. [Google Scholar] [CrossRef]
  6. Miyakawa, S.; Matsuda, S.; Tanibata, N.; Takeda, H. Computational studies on defect chemistry and Li-ion conductivity of spinel-type LiAl5O8 as coating material for Li-metal electrode. Sci. Rep. 2022, 12, 16672. [Google Scholar] [CrossRef]
  7. Temeche, E.; Indris, S.; Laine, R. LiAlO2/LiAl5O8 membranes derived from flame-synthesized nanopowders as a potential electrolyte and coating material for all-solid-state batteries. ACS Appl. Mater. Interfaces 2020, 12, 46119–46131. [Google Scholar] [CrossRef]
  8. Roy, A.; Senso, D.; Casella, A.; Devanathan, R. Molecular dynamics simulations of radiation response of LiAlO2 and LiAl5O8. J. Nucl. Mater. 2023, 576, 154280. [Google Scholar] [CrossRef]
  9. Kniec, K.; Tikhomirov, M.; Tikhomirov, M.; Pozniak, B.; Pozniak, K.; Marciniak, L. LiAl5O8: Fe3+ and LiAl5O8: Fe3+, Nd3+ as a new luminescent nanothermometer operating in 1st biological optical window. Nanomaterials 2020, 10, 189. [Google Scholar] [CrossRef]
  10. Lan, X.K.; Li, J.; Li, J.P.; Wang, F.; Lu, W.Z.; Wang, X.C.; Lei, W. Phase evolution and microwave dielectric properties of novel LiAl5-xZnxO8-0.5x-based (0 ≤ x ≤ 0.5) ceramics. J. Am. Ceram. Soc. 2020, 103, 1105–1112. [Google Scholar] [CrossRef]
  11. Ao, L.Y.; Tang, Y.; Li, J.; Fang, W.S.; Duan, L.; Su, C.X.; Sun, Y.H.; Liu, L.J.; Fang, L. Structure characterization and microwave dielectric properties of LiGa5O8 ceramic with low-εr and low loss. J. Eur. Ceram. Soc. 2020, 40, 5498–5503. [Google Scholar] [CrossRef]
  12. Yang, J.; Ao, L.Y.; Luo, X.F.; Chen, C.Q.; Zhang, X.; Fang, L.; Tang, B.; Tang, X.Z. Influence of (Zn0.5Ti0.5)3+ substitution on structural evolution, bond characteristics, and microwave dielectric properties of spinel Li(Zn0.5Ti0.5)xGa5-xO8 solid solutions. Ceram. Int. 2023, 49, 27965–27974. [Google Scholar] [CrossRef]
  13. Donegan, J.F.; Bergin, F.J.; Glynn, T.J.; Imbusch, G.F.; Remeika, J.P. The optical spectroscopy of LiGa5O8: Ni2+. J. Lumin. 1986, 35, 57–63. [Google Scholar] [CrossRef]
  14. Izumi, K.; Miyazaki, S.; Yoshida, S.; Mizokawa, T.; Hanamura, E. Optical properties of 3d transition-metal-doped MgAl2O4 spinels. Phys. Rev. B 2007, 76, 075111. [Google Scholar] [CrossRef]
  15. Li, C.; Ding, S.H.; Zhang, Y.; Zhu, H.; Song, T.X. Effects of Ni2+ substitution on the crystal structure, bond valence, and microwave dielectric properties of BaAl2-2xNi2xSi2O8-x ceramics. J. Eur. Ceram. Soc. 2021, 41, 2610–2616. [Google Scholar] [CrossRef]
  16. Toby, B.H. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 2001, 34, 210–213. [Google Scholar] [CrossRef]
  17. Hakki, B.W.; Coleman, P.A. Dielectric resonator method of measuring inductive capacities in the millimeter range. IEEE T. Microw. Theory 1960, 8, 402–410. [Google Scholar] [CrossRef]
  18. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  19. Zhang, G.; Guo, J.; Yuan, X.F.; Wang, H. Ultra-low temperature sintering and microwave dielectric properties of a novel temperature stable Na2Mo2O7-Na0.5Bi0.5MoO4 ceramic. J. Eur. Ceram. Soc. 2018, 38, 813–816. [Google Scholar] [CrossRef]
  20. Shannon, R.D. Dielectric polarizabilities of ions in oxides and fluorides. J. Appl. Phys. 1993, 73, 348–366. [Google Scholar] [CrossRef]
  21. Li, M.; Pietrowski, M.; Souza, R.; Zhang, H.R.; Reaney, I.M.; Cook, S.N.; Kilner, J.A.; Sinclair, D.C. A family of oxide ion conductors based on the ferroelectric perovskite Na0.5Bi0.5TiO3. Nat. Mater. 2014, 13, 31–35. [Google Scholar] [CrossRef]
  22. Zang, J.; Li, M.; Sinclair, D.C.; Jo, W.; Rodel, J. Impedance spectroscopy of (Bi1/2 Na1/2)TiO3–BaTiO3 ceramics modified with (K0.5Na0.5)NbO3. J. Am. Ceram. Soc. 2014, 97, 1523–1529. [Google Scholar] [CrossRef]
  23. Choi, G.K.; Kim, J.R.; Yoon, S.H.; Hong, K.S. Microwave dielectric properties of scheelite (A = Ca, Sr, Ba) and wolframite (A= Mg, Zn, Mn) AMoO4 compounds. J. Eur. Ceram. Soc. 2007, 27, 3063–3067. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of LiAl5−xNixO8−0.5x microwave dielectric ceramics: (a) x = 0.1; (b) x = 0.2; (c) x = 0.3; (d) x = 0.4; (e) x = 0.5. The insets show enlarged views of selected peaks in panels (b) and (c) for better comparison. The reference diffraction lines correspond to the P4332 phase (PDF#01-087-1278), observed in samples with x = 0.1–0.2, and the Fd 3 m phase (PDF# 01-087-0341), observed in samples with x = 0.3–0.5.
Figure 1. XRD patterns of LiAl5−xNixO8−0.5x microwave dielectric ceramics: (a) x = 0.1; (b) x = 0.2; (c) x = 0.3; (d) x = 0.4; (e) x = 0.5. The insets show enlarged views of selected peaks in panels (b) and (c) for better comparison. The reference diffraction lines correspond to the P4332 phase (PDF#01-087-1278), observed in samples with x = 0.1–0.2, and the Fd 3 m phase (PDF# 01-087-0341), observed in samples with x = 0.3–0.5.
Ceramics 08 00085 g001
Figure 2. Rietveld refinement of LiAl5−xNixO8−0.5x microwave dielectric ceramics: (a) x = 0.1; (b) x = 0.3.
Figure 2. Rietveld refinement of LiAl5−xNixO8−0.5x microwave dielectric ceramics: (a) x = 0.1; (b) x = 0.3.
Ceramics 08 00085 g002
Figure 3. SEM micrographs and corresponding grain size distributions of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics: (a) x = 0.1; (b) x = 0.2; (c) x = 0.3; (d) x = 0.4; (e) x = 0.5.
Figure 3. SEM micrographs and corresponding grain size distributions of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics: (a) x = 0.1; (b) x = 0.2; (c) x = 0.3; (d) x = 0.4; (e) x = 0.5.
Ceramics 08 00085 g003
Figure 4. Variations in εr, Q × f and τf values of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) ceramics.
Figure 4. Variations in εr, Q × f and τf values of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) ceramics.
Ceramics 08 00085 g004
Figure 5. Impedance spectroscopy data of LiAl5−xNixO8−0.5x microwave dielectric ceramics: (A) x = 0.1; (B) x = 0.2; (C) x = 0.4; (D) x = 0.5.
Figure 5. Impedance spectroscopy data of LiAl5−xNixO8−0.5x microwave dielectric ceramics: (A) x = 0.1; (B) x = 0.2; (C) x = 0.4; (D) x = 0.5.
Ceramics 08 00085 g005
Figure 6. Impedance spectroscopy measurement and fitting results for the LiAl5−xNixO8−0.5x (x = 0.1) sample: (A) 575 °C; (B) 600 °C; (C) 675 °C; (D) 700 °C.
Figure 6. Impedance spectroscopy measurement and fitting results for the LiAl5−xNixO8−0.5x (x = 0.1) sample: (A) 575 °C; (B) 600 °C; (C) 675 °C; (D) 700 °C.
Ceramics 08 00085 g006
Figure 7. Activation energy (Ea) results derived from the Arrhenius fitting of the bulk conductivity (σb) for LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics.
Figure 7. Activation energy (Ea) results derived from the Arrhenius fitting of the bulk conductivity (σb) for LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics.
Ceramics 08 00085 g007
Figure 8. The relationship between the τf and the unit cell volume of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics.
Figure 8. The relationship between the τf and the unit cell volume of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics.
Ceramics 08 00085 g008
Table 1. Rietveld-refined structural parameters of LiAl5−xNixO8−0.5x microwave dielectric ceramics. The subscript 1 denotes the sample with x = 0.1, while the subscript 2 denotes the sample with x = 0.3.
Table 1. Rietveld-refined structural parameters of LiAl5−xNixO8−0.5x microwave dielectric ceramics. The subscript 1 denotes the sample with x = 0.1, while the subscript 2 denotes the sample with x = 0.3.
AtomxyzFractionRpRwpχ2
Li110.62500.62500.62501.000.0530.0814.869
Li21−0.87561.12560.12501.00
Al110.00240.00240.00241.00
Al210.62500.62500.62501.00
Al310.37250.12250.12501.00
Ni11−0.0024−0.0024−0.00241.00
Ni210.62500.62500.62501.00
O110.11350.12310.38711.00
O210.40230.40230.40231.00
Li20.50000.50000.50001.000.0480.0714.235
Al120.12500.12500.12500.94
Al220.50000.50000.50000.94
Ni120.12500.12500.12500.06
Ni220.50000.50000.50000.06
O20.25590.25590.25591.00
Table 2. Lattice parameters and relative densities of LiAl5−xNixO8−0.5x microwave dielectric ceramics.
Table 2. Lattice parameters and relative densities of LiAl5−xNixO8−0.5x microwave dielectric ceramics.
xLattice Parameter a (Å)Unit Cell Volume (Å3)Measured Density
(g/cm3)
Calculated Density (g/cm3)Relative Density (%)
0.17.909494.7263.5283.65696.5
0.27.944501.3233.5523.63997.6
0.37.952502.8393.5653.66097.4
0.47.963504.9293.5363.67696.2
0.57.975507.2153.5823.69097.1
Table 3. The calculated dielectric polarizability ( α D T ), measured dielectric constant εr, and calculated dielectric constant εc value of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics.
Table 3. The calculated dielectric polarizability ( α D T ), measured dielectric constant εr, and calculated dielectric constant εc value of LiAl5−xNixO8−0.5x (0.1 ≤ x ≤ 0.5) microwave dielectric ceramics.
x α D T 3)Vm3)εrεc
0.184.18494.7268.248.43
0.283.96501.3238.138.04
0.383.74502.8398.067.94
0.483.52504.9297.927.76
0.583.30507.2157.777.60
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

Lan, X.; Tang, H.; Chen, B.; Tian, B. Effect of Ni2+ Doping on the Crystal Structure and Properties of LiAl5O8 Low-Permittivity Microwave Dielectric Ceramics. Ceramics 2025, 8, 85. https://doi.org/10.3390/ceramics8030085

AMA Style

Lan X, Tang H, Chen B, Tian B. Effect of Ni2+ Doping on the Crystal Structure and Properties of LiAl5O8 Low-Permittivity Microwave Dielectric Ceramics. Ceramics. 2025; 8(3):85. https://doi.org/10.3390/ceramics8030085

Chicago/Turabian Style

Lan, Xuekai, Huatao Tang, Bairui Chen, and Bin Tian. 2025. "Effect of Ni2+ Doping on the Crystal Structure and Properties of LiAl5O8 Low-Permittivity Microwave Dielectric Ceramics" Ceramics 8, no. 3: 85. https://doi.org/10.3390/ceramics8030085

APA Style

Lan, X., Tang, H., Chen, B., & Tian, B. (2025). Effect of Ni2+ Doping on the Crystal Structure and Properties of LiAl5O8 Low-Permittivity Microwave Dielectric Ceramics. Ceramics, 8(3), 85. https://doi.org/10.3390/ceramics8030085

Article Metrics

Back to TopTop