Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Interrelated lithium-ion transport, NTCR behavior, and dielectric-optical coupling in multifunctional Li3Mg2NbO6 ceramic

Samia Aydia, Mohamed Mounir Bouzayania, Moufida Boukribab, Saber Nasria, Omar Radhi Alzoubic and Abderrazek Oueslati*a
aLaboratory of Spectroscopic and Optical Characterization of Materials (LaSCOM), Faculty of Sciences, University of Sfax, B. P. 1171, 3000 Sfax, Tunisia. E-mail: oueslatiabderrazek@yahoo.fr
bDepartment of Chemistry, College of Science, Qassim University, Buraydah 51452, Saudi Arabia
cComputing Department, College of Engineering and Computing in Al-Qunfudah, Umm Al-Qura University, Kingdom of Saudi Arabia

Received 6th September 2025 , Accepted 28th October 2025

First published on 17th November 2025


Abstract

The rapid progress of 5G/6G communication and advanced radar systems demands dielectric ceramics with excellent microwave performance and multifunctional characteristics. Lithium-based Li3Mg2NbO6 (LMN), crystallizing in a rock-salt-derived orthorhombic structure, is identified as a promising candidate due to its structural stability and tunable dielectric response. LMN ceramics were synthesized by the solid-state route and comprehensively analyzed using X-ray diffraction, UV-Vis spectroscopy, and impedance spectroscopy. The material exhibits a direct bandgap of 3.78 eV, indicating potential for optoelectronic and photocatalytic applications, while an Urbach energy of 0.92 eV reveals structural disorder affecting charge transport. Impedance results confirm thermally activated hopping conduction with an activation energy of 1.16 eV, governed by non-Debye relaxation associated with defect-mediated mechanisms. The strong correlation between structure, optical properties, and charge dynamics highlights LMN as a versatile, high-performance ceramic for next-generation microwave and UV-responsive electronic devices.


1. Introduction

The rapid development of next-generation wireless technologies, including 5G/6G networks, satellite communications, and radar systems, has intensified the search for advanced dielectric ceramics capable of sustaining high-frequency operation with minimal energy dissipation.1–3 Within this context, lithium-based ceramics have emerged as promising candidates due to their chemical stability, low production cost, and environmental compatibility.4–8 Their multifunctional character combining structural robustness with excellent dielectric and thermal properties has positioned them at the forefront of materials research for microwave and electronic applications.9–13

A particularly important subgroup of these ceramics comprises lithium-containing compounds with rock-salt-derived structures, which exhibit tunable permittivity, low dielectric losses, and remarkable thermal stability.14–18 Beyond their established microwave performance, these materials have demonstrated unique functionalities, such as the Negative Temperature Coefficient of Resistance (NTCR) effect, making them attractive for thermistor applications.19,20 Their structural flexibility has also stimulated interest in energy storage and optoelectronic devices, broadening their technological relevance. Representative compounds including Li2Mg3TiO6, Li2MgTiO4, Li2Mg4TiO7, and Li3Mg2NbO6 have consistently shown excellent dielectric performance, enabling their use in resonators, filters, and antennas essential for modern communication systems.21–27

Among these, Li3Mg2NbO6 (LMN) has attracted particular attention owing to its distinctive orthorhombic structure derived from rock-salt supercells.28 In this lattice, Nb5+ occupies well-defined octahedral sites within a close-packed oxygen framework, while Li+ and Mg2+ ions share multiple octahedral positions in a partially ordered arrangement.33–35 This ordering results in distorted [Li/Mg/NbO6] octahedra, which strongly influence the dielectric, thermal, and functional properties of the material.36 Recent studies, have advanced the understanding of LMN's structural evolution and microwave dielectric response, employing spectroscopic techniques to link cation ordering with dielectric behavior.29–32 In the case of Li3Mg2NbO6 (LMN), the term “8a Wyckoff position” refers to the specific cation site within its ordered double perovskite-type structure. The compound crystallizes in the orthorhombic system with the Fddd (no. 70) space group, typical of rock-salt-derived ordered perovskites. In this crystal structure, Li+ ions primarily occupy the 8a Wyckoff sites, while Mg2+ and Nb5+ ions are distributed over the 4c and 4d positions, respectively. Oxygen atoms are located at the 8d sites, completing the framework of corner-shared [MgO6]/[NbO6] octahedra. These atomic distributions are consistent with our Rietveld refinement results and previously reported literature on similar complex oxides.33–36 To assess Li3Mg2NbO6 as an advanced dielectric ceramic, it is crucial to link its intrinsic properties to performance. Its orthorhombic structure and ordered cation distribution enhance polarizability and reduce dielectric loss, while the wide bandgap and controlled defect states limit electronic conduction pathways. Temperature-dependent electrical behavior further ensures thermal stability. Together, these factors justify the comprehensive study of structural, optical, and electrical properties to optimize LMN for high-performance dielectric applications.

Despite this progress, important aspects of LMN remain underexplored. In particular, its charge transport mechanisms, relaxation dynamics, and optical properties have not been fully clarified, even though these factors are critical for tailoring multifunctional ceramics for energy-efficient microwave and electronic applications. Addressing this knowledge gap, the present study provides a systematic investigation of LMN's crystal structure, optical behavior, and frequency- and temperature-dependent electrical transport properties. By correlating intrinsic structural features with charge carrier dynamics, we deliver insights into how lattice distortions and cation distribution govern relaxation processes. These findings not only deepen the scientific understanding of LMN but also highlight its potential as a multifunctional platform material for next-generation dielectric and electronic devices.

2. Experimental section

2.1. Sample preparation of Li3Mg2NbO6

Li3Mg2NbO6 was synthesized via a conventional solid-state reaction route. High-purity starting powders lithium carbonate (Li2CO3, 99%, Sigma-Aldrich), magnesium oxide (MgO, 99%, Sigma-Aldrich), and niobium oxide (Nb2O5, 99.9%, Sigma-Aldrich) were stoichiometrically weighed and homogenized by dry ball-milling in an agate mortar for 2 hours to ensure uniform mixing. The mixed powder was then calcined in an alumina crucible at 900 °C for 8 hours in a muffle furnace to decompose carbonates and initiate phase formation, with intermittent grinding to promote reactivity.

The calcined powder was re-milled for 1 hour to refine particle size and improve homogeneity, followed by uniaxial pressing into pellets (8 mm diameter, ∼1 mm thickness) at 150 MPa. Finally, the pellets were sintered at 900 °C (1173 K) for 2 hours in air to achieve densification and crystallize the desired Li3Mg2NbO6 phase. Phase purity and structural evolution were confirmed by X-ray diffraction (XRD), as shown in Fig. 1.


image file: d5ra06717k-f1.tif
Fig. 1 Process flow chart for the solid-state synthesis of Li3Mg2NbO6.

2.2. Materials characterisation

X-ray diffraction (XRD) analysis of Li3Mg2NbO6 was performed using a Philips powder diffractometer with a Cu Kα radiation source (λ ≈ 1.54187 Å), operating at 40 kV and 40 mA. Measurements were conducted at room temperature in Bragg–Brentano (θ–2θ) geometry, with a Ni filter to suppress Kβ radiation. The diffraction patterns, collected over a 2θ range of 10°–80° with a step size of 0.02°, were analyzed for phase identification, lattice parameters, crystallite size, and microstrain using peak fitting methods.

The optical measurements were conducted with a Shimadzu UV-3101PC spectrophotometer equipped with a 150 W xenon arc lamp as the excitation source. An excitation wavelength of 275 nm was selected, corresponding to the bandgap absorption edge of Li3Mg2NbO6, to ensure efficient electron–hole pair generation and probe direct band-to-band transitions. Emission spectra were acquired across the 200–800 nm range using a high-sensitivity photomultiplier tube (PMT) detector, with a spectral resolution of 1 nm. To minimize interference from scattered excitation light, a long-pass optical filter was employed.

The electrical properties of Li3Mg2NbO6 were analyzed by impedance spectroscopy using a Solartron 1260 Impedance Analyzer over a frequency range of 10 Hz to 5 MHz with an AC amplitude of 500 mV. Measurements were performed on a sintered pellet with silver electrodes, ensuring good electrical contact and minimizing electrode polarization effects. The temperature was varied from 493 K to 673 K using a controlled heating stage to investigate the temperature dependence of the electrical response.

Combined, these techniques provide a comprehensive understanding of the structural framework, optical characteristics, and electrical properties of Li3Mg2NbO6. This multifaceted analysis enables a deeper evaluation of its fundamental behavior and supports its potential suitability for advanced functional applications in optoelectronic and energy-related technologies.

3. Results and discussion

3.1. X-ray diffraction analysis

The phase purity, crystal structure, and phase composition of Li3Mg2NbO6 were characterized by X-ray diffraction (XRD). Fig. 2 presents the Rietveld-refined XRD pattern of the sintered ceramic, obtained at room temperature (2θ = 20°–80°) using FullProf software.37 The analysis confirms a pure orthorhombic phase (space group Fddd), consistent with the reference JCPDS file (#86-0346),38 with no detectable secondary phases. The refined lattice parameters (Table 1) show excellent agreement with previously reported values,39,40 validating the successful synthesis of phase-pure Li3Mg2NbO6.
image file: d5ra06717k-f2.tif
Fig. 2 Rietveld refinement profile carried out through FullProf software of Li3Mg2NbO6.
Table 1 Refined structural parameters of Li3Mg2NbO6 obtained from Rietveld refinement of Li3Mg2NbO6 at room temperature
Compound Li3Mg2NbO6
Space group Fddd
Crystal system Orthorhombic
[thin space (1/6-em)]
Unit cell dimensions
a (Å) 8.552
b (Å) 5.897
c (Å) 17.721
V (Å) 895.055
Rp (%) 5.2
Rwp (%) 6.8
Rexp (%) 5.8
χ2 2.1
DSC (nm) 38
δ (cm−2) 26 × 103
Atom Li(1) Li(2) Mg Nb O(1) O(2)
Wyckoff site 4c 4c 4b 4a 4c 8d
x 0.250 0.750 0.500 0.000 0.250 0.489
y 0.250 0.250 0.000 0.000 0.038 0.250
z 0.000 0.500 0.000 0.000 0.250 0.756
Occupancy 1.0 1.0 1.0 1.0 1.0 1.0


A detailed summary of the refined structural parameters obtained from the Rietveld refinement is presented in Table 1. The table includes the lattice constants, atomic coordinates (x, y, z), site occupancies, and reliability factors (Rwp, Rp, χ2), as well as the calculated cell volume. These results provide a comprehensive overview of the crystallographic characteristics of LMN and confirm the successful fitting between the experimental and simulated XRD patterns, as shown in Fig. 2.

X-ray diffraction (XRD) patterns were utilized to estimate the average crystallite size (DSC) of the Li3Mg2NbO6 sample. The crystallite size was calculated using the Debye–Scherrer equation, expressed as follows:41

 
image file: d5ra06717k-t1.tif(1)
Here, β represents the full width at half maximum (FWHM) of the diffraction peak intensity, measured in radians. The dimensionless shape factor K is typically taken as 0.94. The wavelength of the Cu Kα radiation, λ, is 1.5406 Å, and θ denotes the Bragg angle corresponding to the diffraction peak.

Dislocation density (δ) is a key structural parameter that provides insight into the defect concentration within the sample. For Li3Mg2NbO6 nanoparticles, the dislocation density was calculated using the relation:42

 
image file: d5ra06717k-t2.tif(2)

The crystallite size and dislocation density values obtained via Scherrer's method, as summarized in Table 1, offer valuable information about the material's microstructure. Physically, these parameters reflect the average size of the crystalline domains and the degree of lattice distortion caused by defects or internal stresses. Notably, smaller crystallite sizes generally correspond to stronger materials due to their ability to hinder dislocation motion.

The density of polycrystalline Li3Mg2NbO6 significantly influences its physical properties. The powder density ρX-ray was determined from XRD data using the formula:43

 
image file: d5ra06717k-t3.tif(3)
where M is the molecular weight, NA is Avogadro's number, V is the unit cell volume, and Z = 4Z = 4 is the number of formula units per unit cell. Using the following relations, the bulk density (ρexp), porosity (P), and specific surface area (S) were calculated according to eqn (4), (5), and (6):44
 
image file: d5ra06717k-t4.tif(4)
where r is the pellet radius, and h and m are the thickness and mass of the pellet, respectively. The capacity (C), representing the relative density of the sample, was calculated as:
 
image file: d5ra06717k-t5.tif(5)

Assuming spherical particles, the specific surface area was estimated using:

 
image file: d5ra06717k-t6.tif(6)

For the annealed sample, the lattice parameters obtained from Rietveld refinement yielded bulk and powder densities of 3.853 g cm−3 and 3.987 g cm−3, respectively. The capacity and specific surface area were determined to be 0.966 and 38.543 m2 g−1, respectively.

In addition to the structural observations, the annealing temperature has a significant impact on the densification behavior of the material. Higher annealing temperatures enhance atomic mobility, facilitating diffusion along grain boundaries and lattice sites, which in turn promotes grain growth and reduces residual microstrain, as reflected by the sharper XRD peaks and decreased FWHM values. These structural improvements are physically associated with the elimination of lattice defects and pore shrinkage, leading to a more compact microstructure and improved densification. Consequently, the observed enhancement in crystallinity and reduction in microstrain indicate that thermal treatment provides the necessary energy to overcome diffusion barriers, thereby enabling the material to achieve higher packing density, which influenced in electrical study.45,46

3.2. Electrical properties

3.2.1. Impedance study. Complex Impedance Spectroscopy (IS) is a versatile technique used to study electrical transport, polarization, and charge relaxation in materials. It analyzes how low-resistance grains and high-resistance grain boundaries influence conductivity and dielectric behavior. By measuring impedance across a wide frequency and temperature range, IS distinguishes contributions from grains, boundaries, and electrodes, revealing charge transport, defect states, and thermally activated processes like ionic diffusion.47–50 This study uses IS to investigate material transport properties, correlate electrical behavior with microstructure, and optimize materials for electronics, energy storage, and sensors.

Fig. 3(a) displays the temperature-dependent Nyquist plot (−Zvs. Z′) for Li3Mg2NbO6 between 493 K and 673 K. Unlike conventional dielectric materials that often show three distinct semicircular arcs corresponding to grain, grain boundary, and electrode interface effects,51 our sample exhibits a single flattened semicircular response. This behavior suggests overlapping contributions from both bulk grain and grain boundary conduction mechanisms, with no detectable electrode polarization effects.


image file: d5ra06717k-f3.tif
Fig. 3 Nyquist plots (a), the frequency dependence of the real part (Z′) (b), imaginary part (Z″) of the complex electrical impedance (c), the dependence resistances of grain (Rg) (d) and grains (Rgb) at different temperatures of Li3Mg2NbO6 (e).

The observed semicircular arc becomes progressively smaller with increasing temperature, as seen by the reduction in both its diameter and real-axis intercept (Fig. 3(a)). This temperature-dependent evolution clearly demonstrates enhanced electrical conduction at elevated temperatures. To quantitatively analyze the overlapping grain and grain boundary contributions embedded within this single arc, we utilize the Maxwell–Wagner equivalent circuit approach,52,53 which enables deconvolution of these competing conduction pathways. The spectra were successfully modeled using a two-cell circuit model, consisting of two parallel combinations (inset Fig. 3(a)).

Fig. 3(b) presents the frequency-dependent real impedance Z′(ω) for the Li3Mg2NbO6 sample over a temperature range of 493–673 K, displaying a sigmoidal pattern. At lower frequencies, Z′(ω) exhibits higher values, which decrease with increasing frequency and temperature, aligning with a negative temperature coefficient of resistance.54 This results in increased AC conductivity at elevated temperatures. At higher frequencies, the impact of grain boundaries on impedance weakens, with Z′(ω) converging across temperatures, possibly due to space charge release and diminished barrier effects, enhancing AC conductivity. The corresponding imaginary impedance (Z″) spectrum in Fig. 3(c) reveals temperature-dependent relaxation peaks that broaden and shift with heating. These peaks occur when the applied field frequency matches the characteristic hopping frequency of localized charge carriers. The observed peak broadening, which exceeds that predicted by ideal Debye relaxation, confirms non-Debye behavior in Li3Mg2NbO6. Furthermore, the progressive peak widening with temperature reflects an increase in relaxation time distribution, demonstrating thermally activated charge carrier dynamics.

The temperature-dependent evolution of grain resistance (Rg) and grain boundary resistance (Rgb), derived from impedance spectroscopy (Fig. 3d and e), elucidates the interplay between microstructure and charge transport. Both Rg and Rgb decrease with rising temperature, albeit through distinct mechanisms. For Rg, the reduction stems from thermally activated charge carriers in grain interiors, where increased phonon-assisted hopping and diminished scattering losses enhance bulk conductivity.55 Concurrently, Rgb declines due to the thermal lowering of potential barriers at grain boundaries, which mitigates space-charge effects and defect-mediated trapping. This dual reduction highlights a transition in conduction dominance: at low temperatures, grain boundaries govern transport through barrier-limited processes, while at elevated temperatures, intrinsic grain mobility prevails. The impedance spectra corroborate this behavior, with relaxation peak shifts reflecting accelerated carrier dynamics under heating. Collectively, these trends demonstrate how thermal energy modifies the material's energy landscape reducing resistive losses in both crystalline grains and disordered boundaries to optimize overall charge transport.56

The temperature-dependent impedance spectra were recorded over the range of 493–673 K, chosen to probe the intrinsic electrical response of Li3Mg2NbO6 under conditions of enhanced ionic mobility while maintaining structural stability. Within this temperature window, the material exhibits thermally activated conduction and polarization processes, as reflected in the evolution of the Nyquist plots (Fig. 3). This range provides a comprehensive understanding of the relaxation mechanisms and allows for consistent comparison with other temperature-dependent measurements, including dielectric and conductivity analyses. The observed trends confirm that the conduction and relaxation phenomena are strongly influenced by temperature, highlighting the role of thermally activated charge carriers in LMN.

3.2.2. Complex conductivity spectroscopy. In disordered solids, electrical conductivity stems primarily from charge carriers hopping between localized sites, driven by thermal activation and the applied electric field, shaping both DC and AC conductivity.57 As depicted in Fig. 4, which illustrates conductivity trends, AC conductivity varies with temperature and frequency. At low frequencies, DC contributions dominate, showing minimal dispersion as ions complete long-range hops. At higher frequencies, AC conductivity prevails due to rapid charge carrier responses to the oscillating field, causing significant dispersion. The Jump Relaxation Model (J. Werking model)58 elucidates this: at low frequencies, ions achieve successful long-range hops; at intermediate frequencies, reversible (unsuccessful) hops introduce dispersion; and at very high frequencies, successful hops regain dominance. This frequency dependence is quantitatively captured by Jonscher's universal power law:59
 
σac(ω, T) = σdc(T) + s (7)
where σdc(T) represents DC conduction from long-range motion, and s describes dispersive AC conductivity. The exponent s (0 ≤ s ≤ 1) reflects charge carrier-lattice coupling strength, with s ≈ 1 indicating weak interactions and s < 1 denoting stronger coupling.60,61

image file: d5ra06717k-f4.tif
Fig. 4 The variation of (σ) versus (ω) at different temperatures of Li3Mg2NbO6.

Arrhenius analysis of the DC conductivity (Fig. 5) yields an activation energy Ea = 1.16 ± 0.03 eV for Li3Mg2NbO6, consistent with efficient Li+ migration in rock-salt structured ceramics.


image file: d5ra06717k-f5.tif
Fig. 5 Evolution of ln(σdcT) as a function of (1000/T).

Investigating the mechanisms of charge transport in Li3Mg2NbO6, materials is essential for understanding their electrical characteristics. The temperature-related changes in the exponent s offer key insights into the primary conduction pathways. Various theoretical frameworks have been developed to account for the behavior of s, often integrating quantum mechanical tunneling with charge carrier hopping. In these frameworks, particles such as electrons, polarons, or ions overcome energy barriers, providing clarity on the fundamental conduction processes.

One such model, overlapping large polaron tunneling (OLPT), characterizes the movement of large polarons, where significant overlap between adjacent potential wells facilitates efficient charge transfer. In contrast, non-overlapping small-polaron tunneling (NSPT) describes charge transport via small polarons hopping between isolated states with limited overlap. Another framework, correlated barrier hopping (CBH), depicts conduction as charge carriers moving between localized sites, influenced by Coulomb interactions and energy barriers. Additionally, pure quantum mechanical tunneling (QMT) explains charge carriers passing through barriers without requiring thermal activation.62–65

Analyzing the temperature dependence of the exponent s enables researchers to determine the most appropriate model for describing charge transport in Li3Mg2NbO6, materials. This study not only deepens our understanding of conduction mechanisms but also supports the optimization of material properties for targeted applications, such as energy storage, solid-state batteries, and advanced electronic devices.

Fig. 6 depicts the temperature-dependent variation of s, showing an initial decrease followed by a slight increase at higher temperatures. This pattern aligns with the Non-Overlapping Small Polaron Tunneling (NSPT) model, which describes charge transport as thermally assisted tunneling of polarons (localized carriers coupled with lattice distortions) between defect centers.


image file: d5ra06717k-f6.tif
Fig. 6 Temperature dependence of the exponent “s” of Li3Mg2NbO6.

The NSPT model predicts the following relation for s(T):

 
image file: d5ra06717k-t7.tif(8)
where WH is the polaron hopping energy and τ0 is a characteristic relaxation time which is in the order of an atom vibrational period τ0 ≈ 10−13 s.63 The AC conductivity can also be described microscopically using the expression:
 
image file: d5ra06717k-t8.tif(9)

Whither

 
image file: d5ra06717k-t9.tif(10)

This formulation ties the frequency dependence of AC conductivity to fundamental polaronic parameters, offering a microscopic understanding of the conduction process.

The variation of AC conductivity as a function of temperature at various frequencies is shown in Fig. 7. These curves confirm the thermally activated nature of the charge transport and support the validity of the NSPT model in describing the behavior of Li3Mg2NbO6 over the studied temperature and frequency ranges.66–69 The fitting parameters are summarized in Table 2.


image file: d5ra06717k-f7.tif
Fig. 7 (a–c) Graphs of ln σac versus 1000/T for Li3Mg2NbO6 at specific frequencies.
Table 2 AC conductivity parameters at various frequencies
  Frequency (MHz) WH (eV) ε N (eV−1 cm−1) α−1)
NSPT model 0.15 0.28 0.46 2.9 × 1017 0.43
1.4 0.27 4.8 × 1018 0.51
  15 0.18   5.1 × 1019 0.56


This model not only explains the frequency and temperature dependence of AC conductivity but also quantitatively links microscopic polaronic parameters such as hopping energy (WH), lattice relaxation time (τ0), and defect spacing to macroscopic electrical behavior. For instance, WH, derived from the temperature-dependent exponent s, reflects the energy required to dissociate a charge carrier from its lattice distortion, while τ0 ∼ 10−13 s corresponds to the atomic vibration timescale governing polaron dynamics. These insights enable predictive material optimization: reducing WH (via oxygen vacancies or cation substitution) could enhance ionic mobility, while tuning defect spacing may minimize hopping losses at intermediate frequencies. In Li3Mg2NbO6, the rock-salt framework facilitates small polaron migration by balancing disorder with lattice flexibility, offering practical guidance for solid-state batteries (e.g., designing low-barrier Li+ pathways) and high-frequency electronics (e.g., doping-engineered conductivity control). By connecting atomic-scale polaronic effects to bulk conductivity, this framework accelerates the development of advanced energy storage and electronic devices.

3.2.3. Dielectric measurements. Dielectric relaxation offers essential insights into charge dynamics, polarization mechanisms, and energy losses in materials exposed to alternating electric fields, aiding applications from energy storage to high-frequency devices. By studying dielectric properties, researchers can explore the connection between charge conduction and energy dissipation, critical for material performance. The complex permittivity, ε*(ω) = ε′(ω) − iε″(ω), underpins this analysis: ε′(ω) reflects the material's capacity to store energy via dipole orientation and ionic/electronic polarization,70 while ε″(ω) indicates energy losses due to resistive effects from charge movement or dipole realignment.71

The dielectric constant (ε′) of the material exhibited significant temperature (493–673 K) and frequency dependence, as shown in Fig. 8(a), with exceptionally high low-frequency values (∼107) comparable to advanced lithium-based dielectrics (ε′ ≈ 104 at 493 K/50 Hz),72 suggesting strong potential for energy storage applications. The characteristic ε′ decrease with rising frequency reflects polarization relaxation dynamics, where dipolar and interfacial mechanisms dominant at low frequencies become unable to follow the rapidly oscillating field,73 while the temperature-driven ε′ enhancement indicates thermally activated polarization processes. Four primary mechanisms govern this behavior: electronic and ionic polarization at high frequencies, and dipolar/interfacial polarization at low frequencies,74 with the latter diminishing as increasing frequency prevents dipole alignment, causing ε′ to stabilize at interfacial polarization-dominated values.75 The material's high ε′ and tunable response, particularly in lithium-containing compounds, demonstrates excellent electric field resistance and promising performance for electronic storage systems,76 with frequency- and temperature-dependent characteristics offering optimization pathways for specific applications ranging from high-capacity energy storage to high-frequency electronics.


image file: d5ra06717k-f8.tif
Fig. 8 (a) Frequency-dependent of real part of dielectric constants (ε′), (b) imaginary part (ε″) and (c) tan(δ) at different temperatures of Li3Mg2NbO6.

As illustrated in Fig. 8(b), ε″ rises with temperature, particularly at lower frequencies, due to space charge polarization, resulting in elevated dielectric losses.77 Consequently, tan δ is anticipated to mirror this pattern, with higher values at low frequencies and higher temperatures, driven by increased energy dissipation from interfacial charge buildup and dipole relaxation. At elevated frequencies, both ε″ and tan δ drop considerably, reflecting lower energy losses as charge carriers fail to keep pace with the fast-changing electric field, aligning with observed reductions in capacitance and resistivity (Fig. 8(c)). Capacitance data in Fig. 9 reinforce this trend, revealing higher capacitance at lower frequencies, with values increasing at higher temperatures due to enhanced ionic movement and weakened bond energies.78 This relationship among dielectric constant, loss tangent, and capacitance highlights Li3Mg2NbO6's potential for high-frequency applications requiring minimal dielectric loss (low tan δ), while its temperature-sensitive properties enable adaptability for devices operating under varying thermal conditions.79,80 The material's ordered rock-salt crystal structure, combined with the ability to adjust its composition through methods like Sn4+ doping, allows for fine-tuning of these dielectric properties, optimizing the balance between high capacitance and low loss for enhanced performance in microwave resonators and thermally stable capacitors.


image file: d5ra06717k-f9.tif
Fig. 9 The variation of capacity (C) as a function of frequency at different temperatures.
3.2.4. Modulus study. To investigate the relaxation mechanisms in Li3Mg2NbO6, the electric modulus formalism was employed, offering a clearer perspective on charge carrier dynamics by reducing the influence of electrode polarization.81 The imaginary component of the electric modulus, M″(ω), displayed distinct frequency-dependent behavior (Fig. 10). At low frequencies, the absence of a peak suggests long-range charge carrier migration, facilitating unrestricted hopping between lattice sites.82 Conversely, a prominent peak emerged at higher frequencies, indicating a shift to localized motion where carriers are confined within potential barriers or defect sites. The asymmetric broadening of these peaks confirms non-Debye relaxation behavior, implying a spectrum of relaxation times rather than a single dominant process.83 As temperature increased, the M″ peaks shifted toward higher frequencies, consistent with thermally assisted hopping that accelerates relaxation and enhances carrier mobility.
image file: d5ra06717k-f10.tif
Fig. 10 Frequency dependence of the imaginary part of electric modulus at different temperatures.

From an Arrhenius analysis of the peak frequencies (Fig. 11), the activation energy (Ea) was calculated as 1.16 eV, closely matching the value obtained from DC conductivity measurements. This agreement suggests that ionic relaxation and charge transport in Li3Mg2NbO6 are governed by the same thermally activated hopping mechanism.84


image file: d5ra06717k-f11.tif
Fig. 11 Dependence of ln(ωp) versus reciprocal temperature for Li3Mg2NbO6.

Further correlation with impedance spectroscopy (Fig. 12) revealed additional details about the relaxation process. At lower temperatures (493 K), the mismatch between Z″ and M″ peak positions indicates contributions from both localized and long-range charge dynamics, likely influenced by grain boundary effects. At elevated temperatures (673 K), the convergence of these peaks points to a transition toward localized hopping as the dominant relaxation mechanism.85


image file: d5ra06717k-f12.tif
Fig. 12 Combined M″ and −Z″ spectroscopic plots at (a) 493 K and (b) 673 K.

These findings demonstrate that dielectric relaxation in Li3Mg2NbO6 is driven by thermally activated charge carriers, evolving from extended-range conduction at low frequencies to confined motion at higher frequencies. The non-Debye characteristics and consistent activation energy highlight the intrinsic connection between relaxation and conduction processes, shaped by the material's rock-salt layered structure and space charge interactions. Such insights are crucial for optimizing Li3Mg2NbO6 in applications like microwave resonators and capacitors, where precise control over dielectric response and loss minimization are key to performance and stability.

3.3. Optical properties

Ultraviolet-visible (UV-Vis) spectroscopy offers a powerful, non-invasive approach to probe the optical characteristics of semiconductors, particularly complex oxides like Li3Mg2NbO6, which contain lithium and exhibit intricate electronic structures. This technique yields critical insights into the material's photophysical properties without compromising sample integrity, making it ideal for comprehensive characterization alongside electrical and structural studies. The optical response of Li3Mg2NbO6 is primarily dictated by its optical energy bandgap (Eg), a pivotal parameter that governs its suitability for applications in optoelectronics and photocatalysis. Room-temperature UV-Vis absorption spectra, collected over the 200–800 nm range (Fig. 13(a)), display four distinct absorption peaks at approximately 475, 500, 550, and 750 nm, each linked to specific electronic transitions within the material's lattice. The peaks at longer wavelengths (550 and 750 nm) are attributed to ligand-to-metal charge-transfer (LMCT) transitions within [NbO6] octahedral clusters, where electrons shift from oxygen 2p orbitals to unoccupied niobium 4d orbitals, highlighting the covalent nature of Nb–O bonds in these polyhedral units. In contrast, the shorter-wavelength peaks at 475 and 500 nm are associated with electronic transitions involving Li+ and Mg2+ cations, likely arising from crystal-field-induced d–d transitions or interionic charge-transfer processes shaped by the unique coordination environments within the layered rock-salt structure of Li3Mg2NbO6.
image file: d5ra06717k-f13.tif
Fig. 13 (a) The UV-visible absorbance spectrum, (b) direct and indirect band gap, (c) ln(α) as a function of photon energy (eV) for Li3Mg2NbO6 sample.

The optical bandgap was determined using the Tauc formalism, expressed as:86

(αhν)n = A(Eg)
where α represents the absorption coefficient, is the photon energy, A is a material-specific constant, and n indicates the transition type (n = 1/2 for direct allowed transitions, n = 2 for indirect allowed transitions). By constructing a Tauc plot and extrapolating the linear region of (αhν)(1/n) versus hν to the energy axis (Fig. 13(b)), a direct bandgap of 3.78 eV was calculated. This wide bandgap classifies Li3Mg2NbO6 as a semiconductor comparable to other lithium-containing oxides, rendering it promising for applications such as UV photodetectors, transparent conductive films, and short-wavelength optoelectronic devices. The bandgap, slightly narrower than some lithium-based analogs, may reflect band-filling effects due to intrinsic defects or subtle variations in cationic ordering within the rock-salt lattice. Importantly, the 3.78 eV bandgap surpasses the 1.23 eV threshold required for photoelectrochemical water splitting, suggesting potential for solar-driven hydrogen production, pending further evaluation of charge carrier dynamics and surface catalytic properties.

To quantify structural disorder and defect states, the Urbach energy (Eu) was analyzed, which characterizes the exponential tail near the absorption edge according to ref. 87:

 
image file: d5ra06717k-t10.tif(11)
where α0 is a constant and Eu represents the Urbach energy in electronvolts. By plotting ln(α) against (Fig. 13(c)) and extracting the inverse slope of the linear segment, an Urbach energy of 0.92 eV was determined, equivalent to approximately 24% of the bandgap. This substantial Eu value indicates a high density of localized subgap states, likely stemming from point defects (e.g., oxygen vacancies), lattice distortions, or compositional inhomogeneities within the Li3Mg2NbO6 structure. These disorder-induced states significantly influence charge carrier trapping, recombination kinetics, and space-charge polarization, thereby affecting both optical and electrical performance.88

The optical properties elucidated through UV-Vis spectroscopy provide a critical link to the electrical behavior observed in impedance and modulus spectroscopy studies. The localized states identified by the Urbach energy contribute to the distributed relaxation times observed in dielectric measurements, facilitating charge hopping and space-charge effects that underpin frequency- and temperature-dependent relaxation phenomena. Additionally, the LMCT transitions within [NbO6] clusters, which dominate the absorption spectra, shape the electronic structure that governs ionic conduction pathways and dielectric losses. This interdependence underscores the complex defect chemistry of Li3Mg2NbO6, where the ordered rock-salt framework, with its alternating Li+, Mg2+, and Nb5+ cations, modulates both optical absorption and charge transport.

The wide bandgap and significant Urbach energy highlight LMN's potential for specialized applications while also revealing challenges posed by its inherent disorder. The material's optical characteristics suggest viability for UV-sensitive devices and photocatalytic systems, but the high density of defect states may limit charge carrier mobility or increase recombination losses, necessitating targeted optimization. Defect engineering, such as doping with transition metals (e.g., Sn4+) or adjusting the Li/Mg ratio, could reduce subgap states or tune the bandgap to enhance optical transparency or catalytic efficiency. The structural similarity of Li3Mg2NbO6 to other high-performance oxides, such as Li3Mg2NbO6, invites comparative studies to uncover universal principles for designing wide-bandgap ceramics. Advanced synthesis techniques, including high-pressure sintering or controlled annealing, could further minimize disorder, improving the material's functional properties. These findings establish LMN as a versatile platform for optoelectronic devices, photocatalysts, and microwave dielectrics, where precise control of optical and electrical properties is paramount.89 Future research should focus on correlating optical parameters with structural modifications, exploring the material's performance under operational conditions, and integrating it into prototype devices to fully realize its technological potential.

In conclusion, UV-Vis spectroscopy of our sample reveals a complex interplay of electronic transitions, a wide 3.78 eV bandgap, and significant structural disorder, providing a comprehensive understanding of its photophysical and electrical properties. These insights, combined with electrical characterizations, offer a robust framework for tailoring the material's performance through defect control and compositional tuning, paving the way for its integration into advanced functional devices.

Taken together, the structural, optical, and electrical analyses reveal that the exceptional dielectric behavior of Li3Mg2NbO6 arises from the synergistic interplay of its intrinsic characteristics. The orthorhombic rock-salt-derived lattice ensures a highly ordered ionic arrangement, minimizing defects and promoting uniform dipolar alignment, which underpins low dielectric loss and stable permittivity. The wide optical bandgap (∼3.78 eV) effectively suppresses electronic polarization at microwave frequencies, while shallow defect states modulate space-charge polarization, influencing the frequency-dependent dielectric response. Electrical analysis further indicates thermally activated relaxation processes governed by ionic mobility and lattice vibrations, directly affecting dielectric stability under operational conditions. Together, these properties elucidate the fundamental mechanisms controlling LMN's low-loss, high-stability dielectric behavior, confirming its strong potential for microwave and energy storage applications.

4. Conclusion

Li3Mg2NbO6 has been established as a multifunctional dielectric ceramic that integrates robust structural, electrical, and optical performance. The rock-salt-derived orthorhombic framework endows the material with high permittivity, low dielectric losses, and a wide direct bandgap of 3.78 eV, signifying its suitability for both microwave and UV optoelectronic applications. Electrical investigations reveal non-Debye relaxation and thermally activated charge transport (Ea ≈ 1.16 eV), highlighting the influence of cation distribution and structural disorder on the conduction process. These insights underscore the potential of defect engineering, compositional substitution, and sintering optimization as effective routes to tailor its functional behavior. The combined dielectric, optical, and electrical characteristics position Li3Mg2NbO6 as a versatile candidate for high-frequency resonators, thermally stable capacitors, and photoelectronic components relevant to emerging 5G/6G and sustainable energy technologies.

This work therefore opens new opportunities for the design of next-generation multifunctional ceramics combining structural stability, dielectric reliability, and optical responsiveness.

Author contributions

Samia Aydi: investigation, data curation, formal analysis, visualization, writing – original draft, writing – review & editing. Mohamed Mounir Bouzayani: investigation, data curation, formal analysis, visualization, writing – original draft, writing – review & editing. Moufida Boukriba: investigation, data curation, formal analysis, visualization, writing – original draft, writing – review & editing. Saber Nasri: supervision, resources, validation, methodology, writing – review & editing. Omar Radhi Alzoubi: investigation, formal analysis, visualization, writing – original draft, writing – review & editing. Abderrazek Oueslati: conceptualization, methodology, formal analysis, validation, writing – review & editing, project administration.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Data availability

The authors confirm that the data supporting the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgements

The authors extend their appreciation to Umm Al-Qura University, Saudi Arabia for funding this research work through grant number: 25UQU4360914GSSR01.

References

  1. Q. W. Liao, L. X. Li, X. Ren, X. X. Yu, D. Guo and M. J. Wang, J. Am. Ceram. Soc., 2012, 95, 3363 CrossRef CAS.
  2. C. S. Chen, C. C. Chou, C. S. Chen and I. N. Lin, J. Eur. Ceram. Soc., 2004, 24, 1795 CrossRef CAS.
  3. D. Zou, Q. L. Zhang, H. Yang and S. C. Li, J. Eur. Ceram. Soc., 2008, 28, 2777 CrossRef.
  4. C. Liu, H. Zhang, G. Wang, T. Zhou, H. Su, L. Jia, L. Jin, J. Li and Y. Liao, Mater. Res. Bull., 2017, 93, 16 CrossRef.
  5. M. Guo, S. P. Gong, G. Dou and D. X. Zhou, J. Alloys Compd., 2011, 509, 5988 CrossRef.
  6. G. G. Yao, P. Liu and H. W. Zhang, J. Am. Ceram. Soc., 2013, 96, 1691 CrossRef.
  7. W. Wersing, Curr. Opin. Solid State Mater. Sci., 1996, 1, 715 CrossRef.
  8. M. T. Sebastian and H. Jantunen, Int. Mater. Rev., 2008, 53, 57 CrossRef.
  9. S. J. Fiedziuszko, I. C. Hunter, T. Itoh, Y. Kobayashi, T. Nishikawa, S. N. Stitzer and K. Wakino, IEEE Trans. Microwave Theory Tech., 2002, 50, 706 CrossRef.
  10. Y. L. Liao, F. Xu, D. N. Zhang, T. C. Zhou, Q. Wang, X. Y. Wang, L. J. Jia, J. Li, H. Su, Z. Y. Zhong and H. W. Zhang, J. Am. Ceram. Soc., 2015, 98, 2556 CrossRef.
  11. H. Ren, L. Hao and H. Peng, et al., J. Eur. Ceram. Soc., 2018, 38, 3498 CrossRef.
  12. X. Lu, Y. Zheng and Z. Dong, et al., Mater. Lett., 2014, 131, 1 CrossRef.
  13. H. Ren, S. Jiang and M. Dang, et al., J. Alloys Compd., 2018, 740, 1188 CrossRef.
  14. L. L. Yuan and J. J. Bian, Ferroelectrics, 2009, 387, 123 CrossRef.
  15. P. Zhang, X. Zhao and Y. Zhao, J. Mater. Sci.: Mater. Electron., 2016, 27, 6395 CrossRef.
  16. Y. Qian, Q. Zhang and X. Tang, et al., J. Mater. Sci.: Mater. Electron., 2021, 32, 22450 CrossRef.
  17. P. Zhang, Y. G. Zhao and L. X. Li, Phys. Chem. Chem. Phys., 2015, 17, 16692 RSC.
  18. V. B. Braginsky, V. S. Ilchenko and K. S. Bagdassarov, Phys. Lett. A, 1987, 120, 300 CrossRef.
  19. V. L. Gurevich and A. K. Tagantsev, Adv. Phys., 1991, 40, 719 CrossRef CAS.
  20. J. D. Breeze, J. M. Perkins, D. W. McComb and N. M. Alford, J. Am. Ceram. Soc., 2009, 92, 671 CrossRef CAS.
  21. J. Guo, D. Zhou, H. Wang and X. Yao, J. Alloys Compd., 2011, 509, 5863 CrossRef CAS.
  22. S. George and M. T. Sebastian, Int. J. Appl. Ceram. Technol., 2011, 8, 1400 CrossRef CAS.
  23. T. Zhang and R. Zuo, Ceram. Int., 2014, 40, 15677 CrossRef CAS.
  24. G. Wang, H. Zhang and C. Liu, et al., J. Electron. Mater., 2018, 47, 4672 CrossRef CAS.
  25. P. Zhang, H. Xie and Y. Zhao, et al., J. Alloys Compd., 2017, 690, 688 CrossRef CAS.
  26. X. Zhou, X. Ning and X. Zhang, et al., J. Mater. Sci.: Mater. Electron., 2020, 31, 17029 CrossRef CAS.
  27. M. T. Sebastian and H. Jantunen, Int. Mater. Rev., 2008, 53, 57 CrossRef CAS.
  28. D. Zhou, D. Guo, W. B. Li, L. X. Pang, X. Yao, D. W. Wang and I. M. Reaney, J. Mater. Chem. C, 2016, 4, 5357 RSC.
  29. D. Zhou, W. B. Li, H. H. Xi, L. X. Pang and G. S. Pang, J. Mater. Chem. C, 2015, 3, 2582 RSC.
  30. R. I. Scott, M. Thomas and C. Hampson, J. Eur. Ceram. Soc., 2003, 23, 2467 CrossRef CAS.
  31. B. Jancar, M. Valant and D. Suvorov, Chem. Mater., 2004, 16, 1075 CrossRef CAS.
  32. G. K. Choi, J. R. Kim, S. H. Yoon and K. S. Hong, J. Eur. Ceram. Soc., 2007, 27, 3063 CrossRef CAS.
  33. D. Zhou, L. X. Pang, J. Guo, Z. M. Qi, T. Shao, X. Yao and C. A. Randall, J. Mater. Chem., 2012, 22, 21412 RSC.
  34. L. Fang, Y. Tang and D. Chu, et al., J. Mater. Sci.: Mater. Electron., 2012, 23, 478 CrossRef CAS.
  35. G. Wang, H. Zhang and X. Huang, et al., Ceram. Int., 2018, 44, 19295 CrossRef CAS.
  36. G. Wang, H. Zhang and C. Liu, et al., Mater. Lett., 2018, 217, 48 CrossRef CAS.
  37. W. Wang, C. Liu and L. Shi, et al., Appl. Phys. A: Mater. Sci. Process., 2019, 125, 1 CrossRef.
  38. G. Wang, D. Zhang and Y. Lai, et al., J. Alloys Compd., 2019, 782, 370 CrossRef CAS.
  39. P. Zhang, L. Liu and Y. Zhao, et al., J. Mater. Sci.: Mater. Electron., 2017, 28, 5802 CrossRef CAS.
  40. P. Zhang, J. Liao and Y. Zhao, et al., J. Mater. Sci.: Mater. Electron., 2017, 28, 686690 Search PubMed.
  41. D. W. Kim, D. K. Kwon, S. H. Yoon and K. S. Hong, J. Am. Ceram. Soc., 2006, 89, 3861 CrossRef CAS.
  42. D. Zhou, H. Wang and X. Yao, Mater. Chem. Phys., 2007, 104, 397 CrossRef CAS.
  43. C. L. Huang and M. H. Weng, Mater. Lett., 2000, 43, 32 CrossRef CAS.
  44. S. H. Yoon, D. W. Kim, S. Y. Cho and K. S. Hong, J. Eur. Ceram. Soc., 2006, 26, 2051 CrossRef CAS.
  45. Y. Wang, R. Zuo, C. Zhang, J. Zhang and T. Zhang, J. Am. Ceram. Soc., 2015, 98, 1 CrossRef CAS.
  46. L. X. Pang, H. Wang, D. Zhou and X. Yao, J. Electroceram., 2009, 23, 13 CrossRef CAS.
  47. E. S. Kim and D. H. Kang, Ceram. Int., 2008, 34, 883 CrossRef CAS.
  48. R. Jindal, M. M. Sinha and H. C. Gupta, Spectrochim. Acta, Part A, 2013, 113, 286 CrossRef CAS PubMed.
  49. A. Watanabe, J. Solid State Chem., 2000, 153, 174 CrossRef CAS.
  50. D. Zhou, L. X. Pang, H. Wang, J. Guo, X. Yao and C. A. Randall, J. Mater. Chem., 2011, 21, 18412 RSC.
  51. D. Zhou, L. X. Pang, J. Guo, Z. M. Qi, T. Shao, Q. P. Wang, H. D. Xie, X. Yao and C. A. Randall, Inorg. Chem., 2014, 53, 1048 CrossRef CAS.
  52. A. K. Bhattacharya, K. K. Mallick and A. Hartridge, Mater. Lett., 1997, 30, 7 CrossRef CAS.
  53. P. Zhang, S. Wu and M. Xiao, J. Alloys Compd., 2018, 766, 498 CrossRef CAS.
  54. A. Kudo, K. Omori and H. A. Kato, J. Am. Chem. Soc., 1999, 121, 11459–11467 CrossRef CAS.
  55. J. A. Baglio and O. J. Sovers, Crystal structures of the rare-earth orthovanadates, J. Solid State Chem., 1971, 3, 458 CrossRef CAS.
  56. S. A. Miller, H. H. Caspers and H. E. Rast, Phys. Rev., 1968, 168, 964 CrossRef CAS.
  57. M. Pollet and S. Maranel, J. Eur. Ceram. Soc., 2003, 23, 1925 CrossRef CAS.
  58. R. F. Moran, A. Fernandes, D. M. Dawson, S. Sneddon, A. S. Gandy, N. Reeves-McLaren, K. R. Whittle and S. E. Ashbrook, J. Phys. Chem. C, 2020, 124, 17073 CrossRef CAS.
  59. N. Xu, J. Zhou and H. Yang, et al., Ceram. Int., 2014, 40, 15191 CrossRef CAS.
  60. J. Zhang, Z. Yue and X. Zhang, et al., Ceram. Int., 2015, 41, S515 CrossRef CAS.
  61. R. Zuo, H. Qi, F. Qin and Q. Dai, J. Eur. Ceram. Soc., 2018, 38, 5442 CrossRef CAS.
  62. S. Zhai, P. Liu and Z. Fu, et al., J. Mater. Sci.: Mater. Electron., 2019, 30, 5404 CrossRef CAS.
  63. W. Xia, L. Li and P. Zhang, et al., Mater. Lett., 2011, 65, 3317 CrossRef CAS.
  64. Z. Fu, X. Chen and Y. Zhang, et al., Ceram. Int., 2022, 48, 36638 CrossRef CAS.
  65. R. D. Shannon, J. Appl. Phys., 1993, 73, 348 CrossRef CAS.
  66. Y. Lai, H. Su and G. Wang, et al., J. Am. Ceram. Soc., 2019, 102, 1893 CrossRef CAS.
  67. Y. Lai, C. Hong, L. Jin, X. Tang, H. Zhang, X. Huang, J. Li and H. Su, Ceram. Int., 2017, 43, 16167 CrossRef CAS.
  68. J. Ma, Z. Fu and P. Liu, et al., Mater. Sci. Eng., B, 2016, 204, 15 CrossRef CAS.
  69. R. Vali, Solid State Commun., 2009, 149, 1637 CrossRef.
  70. F. D. Hardcastle and I. E. Wachs, J. Phys. Chem., 1991, 95, 5031 CrossRef.
  71. Y. Lai, C. Hong, L. Jin, X. Tang, H. Zhang, X. Huang, J. Li and H. Su, Ceram. Int., 2017, 43, 16167 CrossRef.
  72. R. D. Shannon, J. Appl. Phys., 1993, 73, 348 CrossRef.
  73. J. Petzelt and S. Kamba, Mater. Chem. Phys., 2003, 79, 175 CrossRef.
  74. D. Zhou, H. Wang, Q. P. Wang, X. G. Wu, J. Guo, G. Q. Zhang, L. Shui, X. Yao, C. A. Randall, L. X. Pang and H. C. Liu, Mater. Lett., 2010, 3, 253 Search PubMed.
  75. A. J. Bosman and E. E. Havinga, Phys. Rev., 1963, 129, 1593 CrossRef.
  76. L. L. Yuan and J. J. Bian, Ferroelectrics, 2009, 387, 123 CrossRef.
  77. L. X. Pang, H. Wang, D. Zhou and X. Yao, J. Alloys Compd., 2010, 493, 626 CrossRef.
  78. D. Zhou, H. Wang, L.-X. Pang, X. Yao and X.-G. Wu, J. Am. Ceram. Soc., 2008, 91, 4115 CrossRef.
  79. X. M. Chen, Y. Xiao, X. Q. Liu and X. Hu, J. Electroceram., 2003, 10, 111 CrossRef.
  80. H.-C. Huang; Y. Wang and X. Jian, Asia-Pacific Conference on Antennas and Propagation (APCAP), 2018, p. 420 Search PubMed.
  81. C. Pecharroman, M. Ocana, P. Tartaj and C. J. Serna, Mater. Res. Bull., 1994, 29, 417 CrossRef.
  82. E. Kobayashi, L. S. Plashnitsa, T. Doi, S. Okada and J. I. Yamaki, Electrochem. Commun., 2010, 12, 894 CrossRef.
  83. A. Rossbach, F. Tietz and S. Grieshammer, J. Power Sources, 2018, 391, 1 CrossRef.
  84. A. R. Armstrong, M. Holzapfel, P. Novák, C. S. Johnson, S.-H. Kang, M. M. Thackeray and P. G. Bruce, J. Am. Chem. Soc., 2006, 128, 8694 CrossRef PubMed.
  85. M. Sathiya, A. M. Abakumov, D. Foix, G. Rousse, K. Ramesha, M. Saubanère, M. L. Doublet, H. Vezin, C. P. Laisa, A. S. Prakash, D. Gonbeau, G. Van Tendeloo and J.-M. Tarascon, Nat. Mater., 2015, 14, 230 CrossRef CAS; I. P. Kondratyuk, B. A. Maksimov and L. A. Muradyan, Dokl. Akad. Nauk SSSR, 1987, 292, 1376 Search PubMed.
  86. A. B. Bykov, A. P. Chirkin, L. N. Demyanets, S. N. Doronin, E. A. Genkina, A. K. Ivanov-Shits, I. P. Kondratyuk, B. A. Maksimov, O. K. Mel'nikov, L. N. Muradyan, V. I. Simonov and V. A. Timofeeva, Solid State Ionics, 1990, 38, 31 CrossRef CAS.
  87. S. A. Anna, B. Kalska, P. Jönsson, L. Häggström, P. Nordblad, R. Tellgren and J. Thomas, J. Mater. Chem., 2000, 10, 2542 RSC.
  88. H. Karami and F. Taala, J. Power Sources, 2011, 196, 6400 CrossRef CAS.
  89. M. Sato, S. Tajimi, H. Okawa, K. Uematsu and K. Toda, Solid State Ionics, 2002, 152, 247 CrossRef.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.