Next Article in Journal
Computational Study of ZnO Surface Catalysis: Adsorption of H2O or/and O2 as a Pathway to ROS Formation
Previous Article in Journal
Effect of the Microstructure of Carbon Supports on the Oxygen Reduction Properties of the Loaded Non-Noble Metal Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Dielectric and Ferroelectric Behavior of Ceramic Nanocomposites: Structure Property Relationships and Processing Strategies

by
Nouf Ahmed Althumairi
1,
Mokhtar Hjiri
2,
Abdullah M. Aldukhayel
1,
Anouar Jbeli
1 and
Kais Iben Nassar
3,*
1
Department of Physics, College of Science, Majmaah University, Al-Majmaah 11952, Saudi Arabia
2
Department of Physics, College of Sciences, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh 11623, Saudi Arabia
3
Materials Physics Laboratory, Faculty of Sciences, Sfax University, BP 1171, Sfax 3000, Tunisia
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(17), 1329; https://doi.org/10.3390/nano15171329
Submission received: 24 July 2025 / Revised: 23 August 2025 / Accepted: 26 August 2025 / Published: 29 August 2025
(This article belongs to the Special Issue Dielectric and Ferroelectric Properties of Ceramic Nanocomposites)

Abstract

In the race toward next-generation electronics and energy systems, ceramic nanocomposites have taken center stage due to their remarkable dielectric and ferroelectric functionalities. By pushing the boundaries of nanoscale engineering, recent studies have shown how microstructural control and interfacial design can unlock unprecedented levels of polarization, permittivity, and frequency stability. This review presents a critical and up-to-date synthesis of the last decade’s progress in ceramic-based nanocomposites, with a special focus on the structure property processing nexus. Diverse processing techniques ranging from conventional sintering to advanced spark plasma sintering and scalable wet-chemical methods are analyzed for their influence on phase purity, grain boundary behavior, and interfacial polarization. The review also explores breakthroughs in lead-free and eco-friendly systems, flexible ferroelectric nanocomposites, and high-k dielectrics suitable for miniaturized devices. By identifying both the scientific opportunities and persistent challenges in this rapidly evolving field, this work aims to guide future innovations in material design, device integration, and sustainable performance.

1. Introduction

Ceramic nanocomposites have become pivotal in advanced materials, offering multifunctionality across electronics, energy storage, sensing, and environmental applications [1]. Early breakthroughs, such as low-temperature sol–gel and nonhydrolytic routes for aluminum titanate fibers, highlighted the importance of interface-controlled dielectric behavior [2]. As wireless technologies advanced, dielectric composites were optimized for wearable devices. Research on low-SAR, high-efficiency tri-band watch antennas [3] was complemented by multi-antenna 5G MIMO arrays [4] and base-station radome composites [5,6]. These developments underscore the demand for permittivity-engineered materials in communication systems [7]. Driven by sustainable energy goals, catalytic nanocomposites with nanoconfined active sites within carbon matrices have delivered enhanced conversion performance [8]. Techniques like dendritic microstructure control in Ti alloys [9] and advanced solvent engineering for CO2 capture [10] highlight the intersection of environmental strategy and material design [11,12]. Biomaterials also inform composite design: proteomic diagnostics [13,14], bioinformatics-driven saliva screening [15,16], and machine-learning enhanced surgical platforms [17] all reflect multifunctional interface principles. Notably, research into oral/gastrointestinal biomarkers [18,19] and neuropeptide modulation [20,21,22] provides tangible parallels in interfacial property optimization. Traditional medicine-inspired formulations from herbal tinctures [23,24] and brain-activated compounds [25,26,27] to cerebral neuronal studies [28,29,30] reinforce the role of structural tailoring for targeted response. Meanwhile, fundamental studies in deposition chemistry, such as sodium-source impacts on β″-Al2O3 channels [31], provide insight into ion-transport-relevant processing strategies. Concurrently, work on magnetic perovskites [32], Alzheimer’s model neural regulation [33,34,35], and bioactive pathway mapping [36,37,38] underscores the consistent theme: structure–function control via processing. Additionally, phase-aware photon-laser network signal recovery using HCMPE-Net [39,40] highlights data-driven diagnostic parallels.
Environmental governance and material sustainability are also guiding research trajectories, as seen in studies on green industrial optimization [41], forest-carbon sequestration models [42], and macro-level carbon-neutral frameworks [43,44,45]. Within ceramics, these same sustainability drives lead to green synthesis, energy-efficient processing, and lead-free materials. Despite this, ceramics still face challenges under dynamic and harsh conditions such as corrosion in structural joints [46], nighttime pedestrian distraction in lighting systems [47], and multi-phase mechanical fatigue in grinding systems [48,49]. These performance demands echo the environment and stress-resilience requirements of dielectric nanocomposites. Advances in diagnostic and holographic imaging [50,51,52], crowd counting analytics [53,54], and fractional Fourier-based systems [55] reflect the growing overlap between electrical, structural, and computational disciplines. These overlaps inform modern ceramic nanocomposite design approaches that blend structural tuning, interface control, and dielectric functionality. The influence of nanostructuring on functional ceramics is also evident in optical and photoluminescent applications. Researchers have investigated multicomponent rare-earth phosphors with tunable emissions, such as Ba9Lu2Si6O24 Eu2+, Mg2+ [56], and lanthanide-substituted CaZnOS systems [57], to improve lighting technologies and wavelength conversion efficiency. Thermal and morphological control in AlN-based composites [58] and Al2O3/Ti3AlC2 microwave absorbers [59] has further demonstrated how microstructural refinement affects dielectric behavior under extreme conditions.
These insights parallel the development of Al2O3@YAG:Ce phosphors with high luminous efficiency [60], as well as advanced LED packages using rare-earth-doped host lattices for stability and brightness tuning [61]. Such studies highlight the importance of phase compatibility and dopant integration, themes also vital in ferroelectric nanocomposite design. Broader societal and environmental motivations are pushing innovation toward carbon-neutral, circular economy materials. Strategies in green governance, such as forest carbon stock optimization [62], urban sustainability frameworks [63], and hybrid low-carbon development models [64], are influencing materials design philosophies. In parallel, dynamic systems modeling [65] and intelligent planning algorithms [66] are increasingly employed to guide the development of eco-efficient fabrication routes and device integration. Marine environment monitoring technologies [67], blockchain-supported agriculture traceability systems [68], and global emission target compliance frameworks [69] further demonstrate the push for multifunctional, intelligent, and sustainable systems. These intersect with recent efforts in ceramic nanocomposites to eliminate lead, reduce processing energy, and implement recyclable or bio-derived fillers. The global push for carbon neutrality by 2060 has intensified these efforts [70], demanding not only technical performance but full lifecycle awareness in material design.
Recent advances in electronic ceramics are propelled by developments in multiferroics, neuromorphic computing, and dielectric elastomers. Wang et al. [71] showcased a wireless multiferroic memristor with giant impedance modulation and synaptic behavior, illustrating magnetic electric coupling at the nanoscale. Xiao et al. [72] enabled low-temperature MgO ceramic sintering using MgF2 and Al2O3, advancing cost-efficient processing. Electric-field control of interfacial effects, such as the Dzyaloshinskii–Moriya interaction in Pt/Co/Pt, allows tunable magnetic anisotropy for next-gen nanoelectronics [73]. In energy storage, Yu et al. [74] developed a prestretch-free dielectric elastomer achieving record energy and power densities through polarization and strain engineering. Wu et al. [75] introduced an Au/Ni-assisted etching method, enhancing structural control in silicon-based nanodevices. Ceramic nanocomposites also enable multifunctional memory. Yang et al. [76] applied relaxor antiferroelectric behavior for neuromorphic computing, while Li et al. [77] created a 2D heterointerface floating gate memory integrating sensing, memory, and computation. Zhao et al. [78] emphasized ceramic–metal interface design in laminated composites for improved dielectric and mechanical performance in armor systems.
The growing demand for high-performance electronic and energy-storage devices has accelerated research into materials that offer superior dielectric and ferroelectric properties. Among these, ceramic nanocomposites have emerged as a versatile and promising class of materials due to their ability to combine the intrinsic advantages of ceramics, such as high permittivity, thermal stability, and mechanical strength, with the tunability offered by nanoscale engineering [79]. By incorporating nanosized fillers or phases into ceramic matrices, researchers have achieved significant enhancements in electrical, structural, and functional performance, opening new possibilities for applications in capacitors, actuators, sensors, and flexible electronics [80]. In recent years, advances in nanotechnology and processing techniques have facilitated unprecedented control over grain size, morphology, phase distribution, and interfacial structures within ceramic composites. These microstructural modifications have been directly linked to improved dielectric constants, reduced losses, and enhanced ferroelectric polarization especially in systems designed with optimized interface chemistry and controlled defect structures. The interdependence of structure, processing, and properties has thus become a central theme in guiding material innovation in this field [81].
Moreover, the transition toward lead-free and environmentally benign materials has added a new dimension to ceramic nanocomposite research. The quest for sustainable alternatives to traditional lead-based ferroelectrics (such as PZT) has stimulated interest in systems based on barium titanate, potassium sodium niobate, and other perovskite or non-perovskite architectures [82]. In parallel, the integration of ceramic nanocomposites into flexible and wearable electronic platforms has further expanded their application landscape, demanding new approaches in materials processing and interface compatibility [83,84]. This review provides a comprehensive overview of the recent developments (approximately the last decade) in the dielectric and ferroelectric behavior of ceramic nanocomposites. Emphasis is placed on understanding how processing strategies, such as sol–gel synthesis, spark plasma sintering, and microwave-assisted methods, influence microstructure and functional properties. Special attention is also given to the role of interfacial polarization, phase synergy, and structural heterogeneity in enhancing overall performance. By mapping out current achievements and identifying persistent challenges, this review aims to support the rational design of next-generation ceramic nanocomposites for emerging energy and electronic applications.

2. Fundamental Concepts in Dielectric and Ferroelectric Ceramic Nanocomposites

2.1. Dielectric Behavior in Ceramic Nanocomposites

Dielectric ceramics exhibit the ability to store electrical energy through the alignment of internal dipoles when subjected to an external electric field. This behavior arises from several types of polarization mechanisms, namely electronic, ionic, dipolar, and interfacial (Maxwell–Wagner) polarization [85,86]. In conventional ceramics, the dielectric constant (εr) is primarily dictated by intrinsic lattice contributions and extrinsic effects such as grain boundaries and defects. However, in ceramic nanocomposites, the situation is markedly different. The high surface-to-volume ratio and the introduction of distinct phases dramatically enhance interfacial polarization effects. Interfaces between dissimilar phase such as ceramic–metal or ceramic polymer act as internal capacitive elements, where trapped charges or conductivity mismatches create localized electric fields [87]. These fields contribute significantly to the overall permittivity and frequency response. Interfacial effects dominate especially when conductive fillers (e.g., CNTs, graphene, or metal oxides) are embedded in insulating ceramic matrices. As a result, careful tuning of the volume fraction, dispersion state, and compatibility between the matrix and fillers become essential in optimizing the dielectric performance of nanocomposites for applications in energy storage, embedded capacitors, and radio frequency devices. In Table 1, we discussed the different polarization phenomena in dielectric nanocomposites.
Sahu et al. [96] explored the dielectric properties of Ag and GO-based poly (vinyl alcohol) (PVA) nanocomposite films by analyzing how GO concentration, ionic liquid (IL) presence, and temperature variations influence dielectric permittivity (ε′), as illustrated in Figure 1a–c. In Figure 1a, the dielectric response of virgin PVA, Ag-PVA, and GO-Ag-PVA films demonstrated a marked enhancement in ε′ with increasing GO content, especially at lower frequencies, owing to stronger interfacial polarization and the Maxwell–Wagner–Sillars effect. This enhancement is attributed to the high surface area of GO, facilitating effective dipole interaction under an external electric field. Figure 1b highlighted the role of IL, where the dielectric permittivity was further amplified due to improved filler dispersion and the introduction of additional charge carriers by the polar nature of ILs. These effects were particularly pronounced at low frequencies, emphasizing IL’s ability to enhance interface polarization. Finally, Figure 1c showed that rising temperatures (40–150 °C) significantly increased ε′ due to enhanced dipole mobility and segmental motion of polymer chains, which reduced the resistive gaps between conducting channels. Collectively, these observations underline the synergistic impact of GO, IL, and temperature on the dielectric performance of GO-Ag-PVA nanocomposites, making them attractive candidates for high-performance dielectric and charge storage applications.
Tayari et al. [97] conducted a detailed investigation of the dielectric properties of Bi0.75Ba0.25(FeMn)0.5O3 ceramics synthesized via the sol–gel method, focusing on the temperature and frequency dependence of the real part of permittivity (ε′), as shown in Figure 2. Their findings revealed a clear increase in ε′ with rising temperature, especially at frequencies below 103 Hz, a trend commonly associated with interfacial polarization and the accumulation of space charges at electrode interfaces. At lower frequencies, charge carriers have sufficient time to align with the applied electric field, leading to higher ε′ values due to polarization at grain boundaries. However, as the frequency increases, dipole reorientation becomes limited, resulting in a decline in ε′. This behavior of dispersion points to dielectric relaxation dominated by electronic transport processes rather than ionic contributions, which typically manifest at higher temperatures and lower frequencies. Figure 2 also illustrates how both ε′ and AC conductivity is governed by similar mechanisms such as hopping conduction and charge mobility. The observed dielectric enhancement at room temperature highlights the material’s suitability for capacitive and transducer applications, while its high permittivity and temperature-dependent response make it a promising candidate for energy storage and microwave device integration.
Faouzia Tayari et al. [98] synthesized Ag-doped Sr(NiNb)0.5O3 ceramic using a sol–gel method and explored its dielectric properties over a frequency range of 103 to 106 Hz and temperatures between 260 K and 340 K. The study demonstrated that the dielectric constant decreases with increasing frequency, a trend observed across all measured temperatures, which is attributed to reduced polarization efficiency at higher frequencies. This behavior is explained by Koop’s theory and the Maxwell–Wagner model, suggesting the material comprises conductive grains separated by insulating grain boundaries. At low frequencies, space charge polarization and grain boundary effects dominate, leading to higher ε′ and energy dissipation, while at high frequencies, the grain response prevails, reducing both permittivity and dielectric loss (tan δ). The dissipation factor also declines with increasing frequency, especially at elevated temperatures, signifying better energy retention. These characteristics point to the material’s suitability for dielectric applications requiring low power loss. The frequency-dependent variations of ε′ and tan δ across different temperatures are illustrated in Figure 3a,b, emphasizing the influence of grain and grain boundary dynamics on dielectric behavior.
Faouzia Tayari et al. [99] conducted a comprehensive investigation into the dielectric behavior of Fe-doped Ba0.67Ni0.33Mn1−xFexO3 ceramics (x = 0 and 0.2), synthesized via conventional solid-state sintering. Their study primarily focused on the influence of Fe substitution at the Mn-site on the dielectric response of the material over a range of frequencies and temperatures. As shown in Figure 4a, the dielectric constant of the doped sample (x = 0.2) exhibits a pronounced decrease with increasing frequency, a typical behavior attributed to the inability of electric dipoles to keep pace with the alternating electric field at higher frequencies. The permittivity increases with temperature, indicating enhanced dipolar and interfacial contributions. This trend is characteristic of perovskite ceramics, where space charge accumulation at grain boundaries or interfaces become significant at lower frequencies. To directly compare the effect of Fe doping, Figure 4b presents the frequency-dependent dielectric permittivity for both undoped and doped samples at room temperature.
A clear enhancement in ε′ is observed for the Fe-doped composition, suggesting that Fe3+ ions may facilitate increased polarization by influencing grain boundary conductivity and defect structures. This result highlights the beneficial role of Fe substitution in optimizing the dielectric storage capability of the material, which is a critical parameter for capacitive energy storage devices. In addition to permittivity, the dielectric loss factor was examined across different temperatures and frequencies, as shown in Figure 4c. At low frequencies, the dielectric loss is considerably high, attributed to space charge polarization and Maxwell–Wagner interfacial effects. These mechanisms are prominent when free charges accumulate at internal interfaces or at the electrode/sample interface, leading to greater energy dissipation. As frequency increases, the dielectric loss decreases due to the reduced response time of slower polarization mechanisms, such as dipolar and ionic contributions. The dielectric loss behavior indicates that while the material exhibits high permittivity, careful tuning is required to minimize energy dissipation for practical device applications. Overall, the results reported by Tayari et al. demonstrate that Fe doping enhances the dielectric constant while slightly increasing dielectric loss at lower frequencies. These findings position Ba0.67Ni0.33Mn1−xFexO3 ceramics as strong candidates for use in energy storage, dielectric resonators, and electronic devices where high permittivity and controlled loss are desired.
The morphology of nanofillers plays a critical role in determining dielectric behavior. Spherical nanoparticles (e.g., BaTiO3 spheres) tend to provide isotropic permittivity enhancement but may aggregate if surface-modification is inadequate. Rod-like fillers (e.g., TiO2 nanorods, ZnO nanowires) often create anisotropic conduction pathways that can raise dielectric constant but also risk increased loss tangent at high loadings. Layered fillers, such as graphene oxide or MXenes, provide large interfacial areas and can improve breakdown strength while enabling tunable permittivity. Comparative studies have shown that rod-like morphologies typically yield higher εr enhancement at low volume fractions, while layered structures excel in thermal stability due to their barrier effect on charge migration. These differences underscore the importance of tailoring filler morphology to the target application.

2.2. Ferroelectric Behavior and Domain Switching Dynamics

Ferroelectricity in ceramics arises from the presence of spontaneous polarization, which is switchable under an applied electric field. This phenomenon is a direct consequence of non-centrosymmetric crystal structures, such as the tetragonal phase in BaTiO3 or rhombohedral structure in BiFeO3. The defining feature of ferroelectric materials is their hysteresis loop in the polarization electric field curve [100,101]. In nanocomposites, these properties are highly sensitive to grain size, interface strain, and defect density. The incorporation of secondary phases or dopants can alter domain wall mobility, induce local lattice strain, and generate space charge fields, all of which influence ferroelectric switching [102,103]. For instance, adding rare-earth dopants like Gd3+ or La3+ can shift the Curie temperature (TC), flatten hysteresis loops, and enhance energy storage efficiency. Recent studies have also explored ferroelectric–metal oxide heterostructures that exploit internal electric fields to control domain configurations at the nanoscale. The optimization of such ferroelectric behavior is particularly important for applications in non-volatile memory, sensors, actuators, and piezoelectric energy harvesting systems [104].
Denis Alikin et al. [105] investigated the competition between ferroelastic and ferroelectric domain wall dynamics in (111)-oriented rhombohedral PMN-PT single crystals using Piezoresponse Force Microscopy under varying relative humidity (RH) conditions. As illustrated in Figure 5, the switching behavior is significantly influenced by the ambient RH in the SPM chamber. At low RH (<4%), ferroelastic switching dominates. This is evidenced by the extended, irregular domain structures characteristic of ferroelastic walls, visible in the left panel of the figure. The corresponding electric field distribution shows a symmetric dipolar pattern, where local field gradients favor stress-driven switching mechanisms. In contrast, at high RH (70%), the behavior shifts to ferroelectric switching, as shown in the right panel. The domain shape is more circular and localized, with stronger field concentration near the probe tip. This indicates that water adsorption at higher humidity enhances surface screening and modulates the spatial electric field distribution, suppressing ferroelastic domain growth. The simulation of the out-of-plane electric field component explains this transition: at high RH, enhanced screening reduces the lateral component of the electric field, which diminishes ferroelastic activity and favors purely ferroelectric domain formation. This observed shift reflects the kinetic nature of domain pattern formation in multiaxial ferroelectrics, suggesting that environmental conditions like humidity can be used as a tunable parameter to engineer domain configurations in piezoelectric devices.

2.3. Interfacial Polarization and Maxwell–Wagner Effects

Interfacial polarization, also known as Maxwell–Wagner–Sillars (MWS) polarization, becomes a dominant mechanism in heterogeneous systems like nanocomposites. It occurs due to the accumulation of charge carriers at the interfaces between phases with contrasting permittivity or conductivity [106]. In ceramic nanocomposites, especially those composed of insulating grains and conductive or semi-conductive phases (e.g., graphene, CNTs, or metal oxides), MWS polarization leads to significant enhancements in dielectric constant, particularly at low frequencies [107]. The extent of this effect depends on the mismatch in conductivity, filler dispersion, and interface compatibility. For example, in BaTiO3-CNT composites, CNTs form conductive networks that facilitate interfacial charge accumulation but may also increase leakage current if not properly controlled [108]. Table 2 shows the polarization in some nanocomposites. Therefore, the percolation threshold must be optimized to maximize interfacial polarization without compromising dielectric loss. Surface modifications of fillers (e.g., functionalization or coating with insulating layers) are often employed to control the interfacial chemistry and prevent undesired conductive pathways.

2.4. Grain Size Effects, Doping Strategies, and Sintering Behavior

The dielectric and ferroelectric performance of ceramic nanocomposites is strongly influenced by grain size, dopant chemistry, and sintering conditions [114,115]. Finer grains increase the density of grain boundaries, which enhances dielectric breakdown strength but may also increase dielectric loss due to space charge accumulation. Dopants, especially rare-earth or transition metal ions, are commonly used to tailor the lattice structure, reduce defect formation, and tune phase transitions. For instance, Gd3+ or La3+ doping in BaTiO3 can shift the Curie point and improve thermal stability. Meanwhile, sintering techniques determine the final microstructure [116,117]. Table 3 illustrates the effect of sintering techniques on ceramic microstructure. Traditional sintering often leads to grain growth and porosity, while advanced methods like spark plasma sintering or cold sintering allow low-temperature densification with grain size retention. These techniques are especially valuable for maintaining nanoscale features that are critical for high-performance dielectric behavior. Xiao et al. [118] investigated the dielectric performance of cold-sintered ZnO ceramic composites modified with polytetrafluoroethylene (PTFE) and metal oxides (CoO and Mn2O3). The study aimed to manipulate grain boundary structures through compositional engineering to enhance electrical properties. As shown in Figure 6a, the relative permittivity of the samples decreases as frequency increases across the range of 10 Hz to 106 Hz. This behavior aligns with interfacial polarization mechanisms that diminish at higher frequencies. Due to the inherently low permittivity of PTFE (εr ≈ 2.1), the overall dielectric constant of the composites doped with PTFE and metal oxides was reduced in comparison to undoped ZnO ceramics. Figure 6b illustrates that dielectric losses (tanδ) sharply decrease at low frequencies (below 1000 Hz), where dc conductivity predominantly contributes to loss. Notably, the sample labeled S1 exhibited high tanδ values (~2.437 at 50 Hz), while the S4 sample containing both PTFE and oxides demonstrated a significantly lower loss (~0.028), attributed to suppression of charge conduction across grain boundaries. Furthermore, high-frequency dielectric loss behavior exhibited relaxation peaks in some doped samples (S2–S4), which are attributed to electronic relaxation associated with intrinsic oxygen vacancy defects.
The fabrication route of these ceramic composites is presented in Figure 7, which schematically outlines the cold sintering process (CSP). In this method, ZnO powder was mixed with PTFE suspension and the selected oxides, followed by a ball milling step, low temperature pressing, and brief thermal exposure at 300 °C. This process enabled the densification of ceramic composites at significantly reduced temperatures compared to traditional sintering. The CSP-assisted integration of polymer and metal oxides effectively refined microstructure reducing average grain size from ~526 nm in pure ZnO to ~338 nm in doped samples and enhanced grain boundary resistance. These modifications contributed to higher Schottky barrier heights and improved dielectric characteristics. Overall, the study demonstrates a feasible and energy-efficient pathway to optimize ceramic nanocomposites for electronic insulation or varistor applications by simultaneously controlling composition and processing strategy.

2.5. Relaxor Ferroelectrics: Structure and Energy Storage Potential

Relaxor ferroelectrics represent a class of materials with diffuse phase transitions and high dielectric constants that vary with frequency. Unlike classical ferroelectrics with sharp Curie transitions, relaxors possess polar nanoregions (PNRs) that fluctuate dynamically [126]. This results in broad dielectric peaks, low remanent polarization, and slim P–E loops ideal traits for energy storage devices where low hysteresis loss and high recoverable energy density are crucial. Examples include materials like Pb(Mg1/3Nb2/3)O3 (PMN), Ba(ZrxTi1−x)O3 (BZT), and modified BiFeO3 systems [127]. Nanocomposite strategies that embed relaxor ceramics into flexible or polymeric matrices have shown promising results, especially in achieving high energy densities above 10 J/cm3. Furthermore, dopants or strain-engineered interfaces can stabilize the relaxor phase, providing enhanced breakdown strength and better thermal stability [128].
Pattipaka et al. [129] explored the energy storage capabilities of lead-free (1 − x)Bi0.5(Na0.8K0.2)0.5TiO3-xBi0.2Sr0.7TiO3 (BNKT-BST) relaxor ferroelectric ceramics, demonstrating a domain engineering strategy to enhance dielectric performance. As illustrated in Figure 8a, the recoverable energy density Wrec is derived from the area between the polarization (P)–electric field (E) hysteresis loop, particularly from the difference between maximum polarization (Pmax) and remnant polarization (Pr). The incorporation of BST into the BNKT matrix disrupted long-range ferroelectric order and promoted the formation of highly dynamic polar nano-regions (PNRs), key to relaxor behavior. These PNRs exhibit reversible transformation into long-range ferroelectric domains under strong electric fields, allowing for a large ΔP = Pmax − Pr, and hence higher energy density and efficiency.
Additionally, as shown in Figure 8b, the phase evolution from ferroelectric (FE) to relaxor ferroelectric (RFE) states was achieved by increasing BST content, leading to a mixed rhombohedral–tetragonal structure and finer grains. These structural modifications resulted in a lower dielectric maximum temperature (Tm), enhanced breakdown strength (EBD), and suppressed hysteresis loss, culminating in a high recoverable energy density of 0.81 J/cm3 and an energy efficiency of 86.95% for the x = 0.45 composition under 90 kV/cm. This demonstrates that well-tailored relaxor behavior and domain structure significantly contribute to boosting energy storage performance in lead-free dielectric ceramics, offering an environmentally friendly alternative to lead-based systems.
The dielectric and ferroelectric behavior of ceramic nanocomposites is governed by a complex interplay of polarization mechanisms, interfacial effects, grain size, and processing strategies. Understanding and tailoring these parameters is essential for optimizing performance in advanced applications such as energy storage, sensors, and flexible electronics.
Beyond the BNKT–BST relaxors, BNT-based lead-free relaxor composites have emerged as some of the most promising alternatives to PZT due to their ability to combine high energy-storage density with excellent strain response under low electric fields. These materials benefit from the coexistence of long-range ferroelectric domains and dynamic polar nanoregions (PNRs), which produce slim hysteresis loops and high recoverable energy efficiency. For example, Khaliq et al. demonstrated that incipient piezoelectrics integrated with relaxor ferroelectrics can exploit stress fields at the ferroelectric/relaxor interfaces to achieve electrostrains exceeding 0.45% under modest fields [130]. Similarly, Sheeraz et al. reported that BNT–BKT-based relaxor composites exhibit large recoverable energy densities and superior electrostrain owing to enhanced interfacial compatibility and defect-mediated polarization mechanisms [131]. These findings suggest that interfacial stress engineering in BNT-derived relaxors provides a scalable pathway to realize both high strain and high efficiency, attributes that are vital for advanced actuators and energy storage devices. Moreover, when coupled with dopants or secondary phases such as BaZrO3 or SrTiO3, BNT-based composites display broadened relaxor transitions and improved breakdown strengths, further reinforcing their status as environmentally friendly contenders for next-generation dielectric and electromechanical applications.

3. Techniques and Microstructural Control in Nanocomposites

Having introduced the key material systems and their dielectric/ferroelectric behaviors in Section 2, we now turn to the synthesis and processing strategies that govern microstructural evolution and, consequently, the material properties. The dielectric and ferroelectric performance of ceramic nanocomposites is intimately linked to their microstructure, which is, in turn, governed by the processing route. Processing not only defines the grain size, phase purity, and porosity, but also critically shapes the interfacial structures, domain configurations, and defect distributions factors that control polarization dynamics, permittivity, and dielectric loss. This section presents a comparative discussion of the most prominent processing techniques used in recent ceramic nanocomposite research: solid-state sintering, sol–gel synthesis, spark plasma sintering, cold sintering, and hydrothermal methods. Each offers unique advantages and limitations in terms of temperature, grain refinement, scalability, and compatibility with complex architectures. Spark Plasma Sintering shows the highest dielectric constant (~3000) due to dense packing and refined grains (~300 nm), while traditional solid-state Sintering exhibits larger grain size (~4000 nm) and lower dielectric response. Sol–gel and cold sintering techniques balance grain size and performance, demonstrating the role of nanoscale control in dielectric behavior.

3.1. Solid-State Sintering

Solid-state sintering remains one of the most widely adopted methods for ceramic fabrication due to its simplicity and scalability. However, it typically requires high sintering temperatures (1200–1500 °C) and results in coarse grains (~4 µm), as shown in Table 4 These large grains reduce grain boundary area, limiting interfacial polarization and resulting in moderate dielectric properties Table 4, which compares different techniques used in the preparation of nanocomposites. While this method is cost-effective for bulk ceramics, it often suffers from compositional inhomogeneities, high porosity, and limited control over nanostructure, making it less suitable for high-performance ferroelectric devices [132,133].

3.2. Sol–Gel Processing

Sol–gel synthesis offers excellent control over stoichiometry and grain size. Through molecular-level mixing of precursors, this method produces ultrafine ceramic powders (~50 nm) at significantly lower calcination temperatures [134,146,147]. As a result, sol–gel derived nanocomposites often show improved permittivity and lower dielectric loss due to better phase uniformity and dense packing of grains. The nanoscale grain structure boosts interfacial polarization and domain wall mobility, contributing to high dielectric constants (~1800) and improved frequency stability. However, the process is sensitive to hydrolysis conditions and drying kinetics, which can affect reproducibility. Ahmed et al. [148] developed a novel electrochemical sensing platform for investigating the interaction between the anticancer drug capmatinib and double-stranded DNA (dsDNA) using a disposable pencil graphite electrode (PGE) modified with cerium oxide-decorated carbon nanofiber ceramic films (CeNPs@CNF-CF). The sensor was fabricated through a multi-step sol–gel approach, where a ceramic precursor solution containing MTMOS, HCl, and methanol was combined with carbon nanofibers and cerium oxide nanoparticles to form a stable nanocomposite dispersion. Pencil graphite rods (PGRs) were then flame-treated, cleaned via sonication in acetone, and coated with the CeNPs@CNF sol–gel through dip-coating using a Teflon holder. This process, illustrated in Figure 9, resulted in the formation of CeNPs@CNF-CF/PGRs with a uniform, conductive, and stable surface suitable for electrochemical applications. The interaction between capmatinib and dsDNA was monitored using cyclic and square wave voltammetry, revealing that the presence of dsDNA significantly suppressed the oxidation peak of capmatinib, indicating a strong electrostatic binding mechanism. The sensor exhibited high sensitivity (LOD = 5 × 10−8 M) and a notable binding constant, confirming its potential for DNA–drug interaction studies.

3.3. Spark Plasma Sintering (SPS)

Spark plasma sintering is an advanced processing technique in which pulsed direct current and uniaxial pressure are simultaneously applied to ceramic powders [149]. This method enables rapid densification at relatively low temperatures (900–1100 °C), thereby preserving nanoscale grain sizes (~300 nm) while achieving near-full density. Composites fabricated via SPS exhibit excellent dielectric properties, with dielectric constants exceeding 2500 [150]. These enhancements are primarily attributed to reduced porosity, refined grain structures, enhanced interfacial polarization, and minimal grain growth. SPS is particularly effective for the processing of doped perovskites, high-k dielectric materials, and composites containing metallic or conductive phases. Figure 10 illustrates a schematic diagram of the SPS setup used to prepare the composite, along with an example of a resulting microstructure [151].

3.4. Cold Sintering Process (CSP)

Cold sintering is an emerging green technique that enables densification of ceramics at sub −300 °C, using a transient aqueous phase. This method is ideal for integrating ceramics with polymeric or temperature-sensitive substrates. CSP can produce dense ceramics with grain sizes of ~200 nm, though dielectric performance is typically moderate (~1400). Despite this, the sustainability and low energy consumption of CSP are unmatched. CSP is gaining traction for applications in flexible electronics, IoT sensors, and hybrid multilayer structures [152,153]. A schematic representation of the cold sintering process and the thermocompression apparatus employed is shown in Figure 11 [154].

3.5. Hydrothermal Synthesis

Hydrothermal methods offer a low-temperature route (100–250 °C) for producing highly crystalline nanostructures (~70 nm grain size) with controlled morphology. The aqueous environment allows selective phase formation and minimizes defects. Hydrothermally synthesized dielectrics exhibit dielectric constants in the range of 1600–1700, with low leakage and good breakdown strength. However, post-synthesis processing (e.g., annealing) is often needed to enhance densification and tailor domain structures [155,156]. Hayfa et al. [157] reported the hydrothermal synthesis of multifunctional bimetallic silver–copper oxide (Ag-CuO) nanohybrids with tunable Ag content, which were evaluated for their antimicrobial, antibiofilm, and antiproliferative properties.
The nanohybrids (Ag-C-1 to Ag-C-4) were synthesized using varying concentrations of silver nitrate (0.05–0.5 g), resulting in nanostructures with pleomorphic morphology, predominantly spherical in shape, as confirmed by SEM and TEM analyses. The particle sizes ranged from 20 to 35 nm, and EDX spectra confirmed the presence of Ag, Cu, and O, verifying the formation of bimetallic hybrids. The nanohybrids exhibited strong antimicrobial activity against E. coli and C. albicans, with MIC and MBC values in the range of 4–12 mg/mL and 2–24 mg/mL, respectively. Moreover, the Ag-CuO nanohybrids induced dose-dependent cytotoxicity and apoptosis in human colon cancer (HCT-116) cells, indicating their potential as antiproliferative agents. The synthesis strategy is illustrated in Figure 12, which shows a schematic representation of the one-step hydrothermal route employed to obtain these functional nanohybrids.

4. Interfacial Engineering and Polarization Mechanisms in Ceramic Nanocomposites

4.1. Role of Interfaces in Dielectric Enhancement

Interfacial engineering is a cornerstone in optimizing dielectric and ferroelectric behavior in ceramic nanocomposites. As grain sizes are reduced to the nanoscale, the volume fraction of grain boundaries and phase interfaces increase significantly, introducing unique charge accumulation sites and affecting local electric fields [158]. These interfaces, when properly engineered, can enhance the dielectric constant, reduce dielectric loss, and introduce relaxor-like behavior. In ceramic–polymer nanocomposites, interfaces play a dual role: they act as mechanical barriers and as charge trapping sites. The quality of the interface between ceramic fillers and polymer matrices (e.g., epoxy) governs polarization, charge mobility, and dispersion uniformity. For instance, ceramic particles with tailored surface functionalization (e.g., via silanes, phosphonic acids, or polymers) can enhance interfacial compatibility and promote stronger interfacial polarization through Maxwell–Wagner–Sillars effects. In all-ceramic systems (e.g., core–shell or multilayered structures), internal interfaces lead to localized field enhancement, beneficial for achieving high dielectric performance [159,160]. The internal barrier layer capacitor effect, seen in BaTiO3-based systems, is a prominent example where insulating grain boundaries trap free carriers and promote charge accumulation, significantly boosting dielectric constant [161].

4.2. Polarization Mechanisms and Frequency Response

Polarization in ceramic nanocomposites arises from several fundamental mechanisms, each governed by the material’s structure, composition, and operating frequency. These mechanisms include electronic, ionic, dipolar, interfacial (Maxwell–Wagner-type), and ferroelectric polarization [162,163]. At the atomic scale, electronic polarization results from the displacement of electron clouds relative to the nuclei within atoms or ions, and it typically dominates at very high frequencies (in the terahertz range). Similarly, ionic polarization arises due to the relative displacement of positive and negative ions in a crystal lattice under an external field [164,165]. Both electronic and ionic contributions are extremely fast and largely independent of frequency across most usable ranges, forming the baseline response in all dielectric ceramics such as Al2O3, MgO, or ZrO2 [166,167]. In contrast, dipolar polarization is associated with the orientation of permanent dipoles within a material and becomes significant at moderate frequencies (kHz–MHz). Dipoles can arise from intrinsic lattice asymmetries (as in perovskite structures) or be introduced through aliovalent doping that creates defect dipoles. For example, in doped BaTiO3 or KNbO3, substitutional ions (such as La3+ or Mn2+) generate localized asymmetries that enhance dipolar alignment under an applied field [168,169].
Interfacial polarization, also known as Maxwell–Wagner–Sillars polarization, becomes highly significant in composite systems, particularly those with heterogeneous interfaces such as ceramic–polymer nanocomposites or grain boundary-rich ceramic systems. It arises due to the accumulation of charges at interfaces between materials with contrasting conductivities or permittivities. In systems like ZnO epoxy, or TiO2–PMMA, interfacial polarization enhances the overall dielectric response at low frequencies (Hz–kHz) by acting as capacitive layers where charge carriers are slowed or trapped [170,171]. Similarly, in all-ceramic heterostructures, such as core–shell BaTiO3@SiO2 or layered BiFeO3 systems, charge buildup at grain boundaries or shell interfaces contributes to elevated permittivity and tunability. A particularly interesting case is ferroelectric polarization, which involves the reversible reorientation of spontaneous electric dipoles within ferroelectric domains under an external electric field [172]. This mechanism is central to memory, sensor, and actuator applications. Materials like BaTiO3, Pb(Zr,Ti)O3 (PZT), and Na0.5Bi0.5TiO3 (NBT) are classical ferroelectrics that exhibit hysteresis behavior and remnant polarization, making them suitable for applications in non-volatile memory and high-strain actuators. In nanocomposites, ferroelectric polarization is often preserved or even enhanced through nanoscale grain control, strain engineering, and interfacial design [172,173,174,175].
Moreover, relaxor ferroelectrics, such as Pb(Mg1/3Nb2/3)O3–PbTiO3 (PMN–PT), and Pb(Sc1/2Nb1/2)O3 (PSN) systems, show a broad, frequency-dependent dielectric peak, attributed to the presence of polar nano-regions [176,177]. These regions fluctuate dynamically, resulting in diffuse phase transitions and enhanced dielectric response over a wide temperature and frequency range. In ceramic nanocomposites, relaxor-like behavior can be induced or amplified by nanoscale particle dispersion, interfacial strain, and dipole-dipole interactions across matrix/filler boundaries. For instance, PMN–PT epoxy composites or BaTiO3–PVDF–TrFE films demonstrate a tunable dielectric relaxation that is ideal for tunable capacitors and flexible devices. Ultimately, understanding and manipulating these polarization mechanisms is essential for optimizing dielectric and ferroelectric behavior in ceramic nanocomposites. The interplay of frequency, microstructure, and interfacial characteristics allows for the engineering of materials with tailored response for specific applications such as energy storage, sensing, and high-frequency electronics [178,179].

4.3. Influence of Interface Type (Core–Shell vs. Heterojunctions)

The interface type in ceramic nanocomposites plays a crucial role in defining their dielectric, ferroelectric, and multifunctional properties. Among the most studied interfacial architectures are core–shell structures and heterojunction interfaces, both of which leverage nanoscale interfacial phenomena to tailor material performance. Core–shell nanostructures are characterized by a central ceramic core encapsulated by a secondary phase shell, which may be another oxide, polymer, or even metallic layer. These architectures enable precise control of interface strain, charge distribution, and polarization alignment, which are essential for optimizing dielectric response and reducing leakage currents. For example, BaTiO3@SiO2, BaTiO3@Al2O3, or ZnO@TiO2 systems exhibit enhanced dielectric strength and long-term stability due to the insulating shell layer, which suppresses conduction pathways and stabilizes grain boundaries [180,181,182]. In particular, the SiO2 shell acts as a barrier to grain growth during sintering, preserving nanoscale dimensions and enhancing surface-to-volume ratio, which intensifies interfacial polarization.
In ferroelectric core–shell composites, the shell can also induce compressive or tensile strain on the ferroelectric core, modifying the Curie temperature and spontaneous polarization. For instance, in BaTiO3@ZrO2 or BiFeO3@TiO2, strain-induced lattice distortion at the interface can stabilize the ferroelectric phase over a wider temperature range, improving performance in energy harvesting and memory applications [183,184]. The thickness of the shell also plays a pivotal role; thin shells (~1–5 nm) can maintain high dielectric constants, while thicker shells may reduce permittivity due to dilution effects, though often improving dielectric loss behavior. On the other hand, heterojunction interfaces involve the combination of two or more different ceramic or hybrid phases joined together, often with a discontinuous or semi-coherent interface. These can include oxide–oxide junctions (e.g., TiO2–SnO2, ZnO–NiO), which promote Maxwell–Wagner-type interfacial polarization [184,185]. In such systems, charge carriers accumulate at the phase boundaries due to differences in electrical conductivity or permittivity, leading to significant enhancement in low-frequency dielectric response. For example, TiO2/SrTiO3 multilayers and Nb-doped BaTiO3/ZnO junctions show large dielectric permittivity due to internal barrier layer capacitor effects [186,187,188,189,190].
The nature of the interface, whether abrupt, graded, or chemically reactive, also influences polarization switching and dielectric relaxation. Reactive interfaces can introduce interfacial dipoles or defect complexes that modify the local electric field and enhance the response under external stimuli. This is particularly important in applications such as tunable dielectrics, where rapid field-induced polarization changes are desirable. Furthermore, core–shell structures often offer better thermal stability and mechanical integrity compared to random heterojunctions, making them suitable for high-temperature or harsh environment applications. However, heterojunction-based systems can provide higher flexibility and tunability by combining materials with vastly different properties, especially in printed electronics and flexible capacitors. In recent advances, engineered gradient interfaces have emerged as a bridge between core–shell and heterojunction designs. These use gradual compositional or structural transitions to minimize interfacial stress while enhancing charge separation and dielectric uniformity. Functionally graded materials and multilayered superlattices are examples where gradient interfaces significantly improve energy storage density and polarization switching speed. Table 5 summarizes the key differences and applications of core–shell versus heterojunction interfaces.

4.4. Dielectric Relaxation and Energy Storage Behavior

Dielectric relaxation is a central phenomenon in ceramic nanocomposites, governing how materials respond to alternating electric fields across different frequencies and temperatures. It refers to the time dependent response of dipoles and interfacial charges that do not immediately align with the external field. This delayed polarization creates frequency-dependent dielectric properties, which are crucial for applications such as capacitors, pulse power devices, and high-energy density storage systems. The performance of dielectric materials in energy storage is thus closely linked to the relaxation dynamics of the different polarization mechanisms present in the system. In ceramic nanocomposites, several types of polarization contribute to dielectric relaxation, each operating on different time scales. Electronic and ionic polarization are the fastest and occur almost instantaneously. These mechanisms are primarily active at high frequencies (in the MHz to GHz range) and contribute modestly to the overall dielectric constant. Dipolar polarization, on the other hand, arises from orientation of permanent or induced dipoles. This mechanism is more sensitive to temperature and defects, and its contribution becomes significant in the lower frequency range.
However, in complex systems such as ceramic polymer composites or multiphase nanocomposites, interfacial polarization also known as Maxwell–Wagner polarization plays a dominant role at low frequencies (Hz to kHz). This type of polarization is due to charging accumulation at the interfaces between materials with contrasting electrical properties, such as between conductive ceramic grains and insulating polymer matrices or secondary phases. In such systems, the nature and quality of the interfaces become critical design variables that influence both relaxation and energy storage performance [191,192]. Another major contributor in many advanced ceramic systems is ferroelectric polarization, which is associated with the reversible switching of spontaneous dipoles under an external electric field. Ferroelectric materials, such as BaTiO3 or lead-free alternatives like BZT–BT and KNN, exhibit strong hysteresis behavior and high dielectric constants. However, their energy storage efficiency is often limited by the energy lost during polarization reversal, manifested as the area enclosed in the polarization electric field (P–E) loop [193].
Relaxor ferroelectrics, a class of disordered ferroelectrics, provides a compromise between high permittivity and low hysteresis. Materials such as PMN–PT, BZT–BT, or doped BaTiO3 demonstrate diffuse phase transitions and frequency-dependent dielectric peaks due to the presence of polar nano-regions (PNRs) [194]. These PNRs facilitate local field fluctuations, resulting in a broad range of dielectric response. In nanocomposites, this behavior is further tuned by particle size distribution, interface coupling, and defect dipole dynamics. This makes relaxor systems particularly attractive for high-temperature and high-frequency dielectric applications. In the context of energy storage, the key metrics are energy density and efficiency. High dielectric constants and breakdown strength are desirable for storing more energy in each volume [195]. However, these need to be balanced against dielectric losses and hysteresis, which reduce overall efficiency. Fine-grained microstructures and tailored doping can suppress domain wall motion and minimize losses, while advanced sintering methods like spark plasma sintering improve grain boundary quality, resulting in enhanced breakdown strength.
Core–shell architectures, such as BaTiO3@SiO2 or BZT@Al2O3, have emerged as promising structures for balancing high permittivity with reduced leakage current. The insulating shell layer suppresses intergranular conduction and enhances reliability under high electric fields. Similarly, combining high-k ceramics with flexible polymers such as PVDF, P(VDF-TrFE), or epoxy resins has led to nanocomposites with both mechanical flexibility and high recoverable energy [196]. These composite systems benefit from interfacial polarization and high field tolerance, though their overall permittivity is usually lower than that of pure ceramics [197]. Notably, materials like CaCu3Ti4O12 (CCTO) have attracted attention due to their colossal dielectric constants, but their high dielectric losses and low breakdown strengths limit their practical utility in energy storage [198]. On the other hand, layered structures or heterostructures that combine materials like SrTiO3 with ZrO2 or BiFeO3 allow for field-tuned behavior and enhanced recoverable energy density by controlling the orientation and interaction of interfaces at the nanoscale. Ultimately, the dielectric relaxation and energy storage behavior of ceramic nanocomposites are not governed by a single factor but result from the interplay of material composition, defect chemistry, grain structure, phase interfaces, and processing history. Therefore, rational design strategies require multi-scale control from atomistic doping and defect engineering to mesoscale interface tuning and macroscale densification to fully exploit the potential of these materials [199].

4.5. Temperature and Frequency Dependence of Dielectric Performance

The dielectric behavior of ceramic nanocomposites is significantly influenced by temperature and frequency, both of which dictate the activation and relaxation of polarization mechanisms. Understanding this dependence is crucial for engineering materials for real-world applications such as capacitors, actuators, sensors, and energy storage devices that must function reliably under varying thermal and signal conditions.

4.5.1. Temperature Dependence

As temperature increases, atomic vibrations intensify, which influences dielectric constant (ε′), dielectric loss (tan δ), and conductivity. In most dielectric ceramics, permittivity rises with temperature up to a critical point, typically associated with a phase transition, such as the Curie temperature in ferroelectrics. For instance, in BaTiO3, a sharp increase in permittivity is observed near its Curie point (~120 °C), where the material transitions from the ferroelectric to paraelectric phase [200]. In relaxor ferroelectrics like PMN–PT or BZT–BT, instead of a sharp transition, the dielectric peak is broad and shifts with frequency a behavior that stems from the dynamic nature of polar nano-regions (PNRs). This diffuse transition makes relaxors more thermally stable, a desirable feature for devices operating over wide temperature ranges [201]. Ceramic–polymer nanocomposites, due to the polymer matrix, often show smoother temperature dependence. However, their performance can degrade near the polymer’s glass transition or melting temperature. Hence, high-temperature polymers like polyimides or fluorinated systems (e.g., PVDF-TrFE) are often used to maintain dielectric performance at elevated temperatures [202].

4.5.2. Frequency Dependence

Frequency influences which polarization mechanisms are active. At low frequencies (Hz to kHz), interfacial and dipolar polarization dominate. These mechanisms are slower and more susceptible to the material’s microstructure, interfaces, and defects. As the frequency increases to the MHz–GHz range, only electronic and ionic polarization remain active because they are fast enough to respond to the rapidly oscillating field. This results in a characteristic dielectric dispersion, a decrease in permittivity with increasing frequency. Simultaneously, dielectric loss often increases at lower frequencies due to conduction and interfacial effects. This is especially true in composites or multi-phase ceramics where charge carriers may become trapped at grain boundaries or phase interfaces. The Maxwell–Wagner–Sillars effect is commonly observed in such systems, where interfacial polarization becomes prominent at low frequencies due to charge accumulation at heterointerfaces. In ceramics like CCTO or doped TiO2 systems, very large permittivity values are seen at low frequencies, though accompanied by high loss due to leakage conduction. In contrast, nanostructured materials with core–shell architectures, or those processed via high-density methods like spark plasma sintering, often exhibit more stable frequency-dependent behavior, suppressing space charge accumulation and reducing dielectric loss [203].

5. Emerging Applications of Dielectric and Ferroelectric Ceramic

The continuous evolution of multifunctional materials has positioned ceramic nanocomposites at the forefront of modern applications spanning electronics, energy, sensing, and sustainability. Owing to their intrinsic dielectric properties, mechanical resilience, and adaptability in microstructural design, these materials offer a unique advantage in scenarios that demand both performance and miniaturization. Their integration into hybrid systems enables synergistic enhancements where single-phase materials would fall short. In this section, we explore key areas where dielectric and ferroelectric ceramic nanocomposites are becoming indispensable. A summary of the diverse and emerging applications enabled by ceramic nanocomposites is illustrated in Figure 13, highlighting their multifunctional roles across electronics, energy, sensing, and biomedical platforms.

5.1. Energy Storage Devices: Capacitors and Hybrid Supercapacitors

One of the most impactful applications of dielectric ceramic nanocomposites is in high-performance energy storage, particularly in capacitors and hybrid supercapacitors. The need for rapid charging and discharging, along with long-term stability, has driven innovations in nanostructured dielectric layers that possess both high permittivity and breakdown strength [204]. Traditional bulk ceramics such as BaTiO3 and Pb(Zr,Ti)O3 have been foundational in capacitor technology, but by integrating them at the nanoscale with polymers or hybrid matrices, their energy density can be drastically enhanced. For instance, core–shell nanostructures like BaTiO3@SiO2 exhibit efficient electric field distribution and minimal leakage, which is crucial for pulsed power applications. Moreover, multilayer nanocomposites using PVDF as the matrix with high-aspect-ratio ceramic fillers like TiO2 nanowires or BaTiO3 nanosheets demonstrate anisotropic dielectric behavior, contributing to superior energy densities often exceeding 20 J/cm3 in optimized systems. Such devices are also being tailored for grid-level energy buffering, electric vehicle systems, and drone applications, where fast response and thermal robustness are mandatory [205,206,207].

5.2. Flexible and Wearable Electronics

The demand for stretchable, conformable, and skin-integrated electronics has opened new opportunities for ceramic nanocomposites in wearable technologies. While ceramics are inherently brittle, embedding ceramic nanoparticles or nanowires into elastomeric or piezoelectric polymer matrices transforms them into mechanically flexible yet electrically responsive systems. For example, composites of BaTiO3 or ZnO with PDMS, PU, or PVDF yield piezoelectric nanogenerators that can harvest biomechanical energy from human motion [208,209,210]. In wearable sensing platforms, these composites enable real-time pressure, strain, or temperature monitoring with high sensitivity and durability. To further enhance flexibility and mechanical resilience, researchers have introduced design strategies such as wrinkled structures, stretchable interconnects, and gradient composite layering, mimicking biological tissues. Additionally, advanced processing like inkjet printing or electrospinning allows scalable manufacturing of patterned, flexible devices suitable for integration into smart textiles, e-skins, and next-generation health monitors.

5.3. Neuromorphic Systems and Smart Memory Devices

The increasing push toward artificial intelligence and brain-like computing has spurred interest in neuromorphic devices that emulate synaptic behavior. Ceramic nanocomposites, particularly those containing ferroelectric phases, play a critical role in this domain by offering nonvolatile memory characteristics, multilevel data retention, and dynamic switching capabilities. Ferroelectric tunnel junctions (FTJs), based on ultrathin layers of HfO2, BiFeO3, or doped BaTiO3, exhibit programmable resistance states that can mimic synaptic plasticity [211]. When embedded in composite architectures with oxide semiconductors or conductive polymers, these materials can be tuned to exhibit spike-timing-dependent plasticity, a hallmark of neuromorphic computation. Furthermore, ceramic-based resistive switching materials in memristors offer fast switching speeds, high endurance, and low power consumption, making them suitable for edge computing and neuromorphic accelerators. Recent advances also include heterostructured devices that combine ferroelectric switching with photoresponsivity or electrochemical modulation, enabling multifunctional logic-in-memory systems.

5.4. High-Frequency Communication and RF Devices

In the realm of RF communications and high-frequency signal processing, ceramic nanocomposites are being engineered to offer controlled dielectric constants and low loss tangents over broad frequency ranges. Materials such as BST, LaAlO3, and AlN have emerged as reliable candidates for GHz–THz applications. When structured at the nanoscale or combined with low-permittivity polymers, they allow fine-tuning of electromagnetic properties required in reconfigurable antennas, tunable filters, and dielectric resonators. For example, BST–MgO composites provide a balance between tunability and loss, suitable for phase shifters and tunable capacitors in 5G systems. Moreover, the development of ceramic–polymer gradient refractive index materials has enabled flexible waveguides and beam-steering devices, expanding applications into mm-wave and sub-THz domains [212,213]. Through interface control and anisotropic particle alignment, researchers have further reduced frequency-dependent losses, enabling the deployment of these materials in miniaturized devices like RF MEMS, wearable antennas, and adaptive radars.

5.5. Sustainable and Environmentally Friendly Technologies

Sustainability is becoming a core driver in materials development, and ceramic nanocomposites are at the heart of green innovations in electronics, energy, and sensing. Efforts to replace toxic or resource-scarce elements, such as lead or rare earths, have led to the rise of lead-free ferroelectrics like (K,Na)NbO3 and BiFeO3 [214,215,216]. These materials are being incorporated into composites with natural polymers like chitosan, cellulose, or gelatin, yielding biodegradable sensors and capacitors suitable for eco-electronics. Furthermore, green processing techniques such as low-temperature sintering, use of deep eutectic solvents, and additive manufacturing are reducing the environmental footprint of ceramic synthesis. Transient electronics, which physically degrade after usage, also leverage water-soluble or thermally degradable ceramic–polymer systems for medical implants, environmental sensors, and military devices. Additionally, by integrating carbon-neutral production pathways and enabling end-of-life recyclability, ceramic nanocomposites are contributing to the circular economy, supporting both sustainability and high-performance functionality in advanced devices.
A particularly important direction in sustainable ceramic nanocomposites is the replacement of lead-based systems with BNT-based and relaxor–ferroelectric composites, which not only mitigate toxicity but also offer competitive or superior functional performance. Unlike traditional Pb(Zr,Ti)O3 systems, these lead-free ceramics leverage local disorder and interfacial strain to generate enhanced dielectric tunability and electrostrain. For instance, BNT–BaTiO3 relaxor composites provide high efficiency and reduced hysteresis losses, while BNT–KBT systems demonstrate domain-engineering strategies that yield high recoverable energy storage capabilities at relatively low fields [130,131]. The dynamic relaxor state allows for reduced energy dissipation during repeated cycling, which is crucial for eco-friendly capacitors and wearable devices requiring long-term operational stability. Furthermore, the integration of these lead-free relaxors into hybrid ceramic polymer architectures (e.g., BNT–PVDF or BNT–epoxy systems) has shown promise for flexible and biodegradable energy devices, where mechanical adaptability is combined with high dielectric performance. These advances underscore that the shift toward BNT-based lead-free composites is not only environmentally motivated but also technologically advantageous, marking a paradigm change in the roadmap for sustainable dielectric materials.

6. Future Approaches for Next-Generation Ceramic

The future development of ceramic nanocomposites hinges on the convergence of sustainable processing, advanced interface design, and digital innovation. One of the primary goals for the coming decade is the shift toward energy-efficient, scalable, and environmentally benign synthesis methods. Traditional sintering and solid-state routes, while reliable, often demand high temperatures and long processing durations, leading to significant energy consumption and environmental burden. As such, hybrid techniques such as spark plasma sintering combined with microwave or cold sintering routes are increasingly being explored to enable densification at reduced thermal budgets. Furthermore, the design of sol–gel and ink-based methods for low-temperature deposition of dielectric and ferroelectric films is being prioritized to align with the needs of flexible substrates and printed electronics. Some of these solutions are already demonstrated in recent studies, while others remain forward-looking proposals intended to guide future research efforts. Table 6 shows the different challenges in ceramic nanocomposites and the proposed solutions.
Another vital area of focus is the use of lead-free and bio-derived materials. With regulatory pressure mounting against lead-containing ceramics like PZT, researchers are intensively studying alternatives such as BaTiO3-based solid solutions, potassium sodium niobate, and bismuth-based systems. However, many of these alternatives suffer from trade-offs between piezoelectric coefficient, phase stability, and fatigue resistance. Addressing these challenges requires not only improved synthetic strategies but also precise control of doping, crystallographic orientation, and interface coherence. Moreover, replacing organic solvents with eutectic or bio-based alternatives could significantly reduce toxicity and improve life-cycle sustainability, a growing concern in both academic and industrial sectors. At the microstructural level, interface engineering remains a cornerstone for performance optimization. Interfacial polarization, grain boundary stability, and local electric field enhancements are critical in determining dielectric and ferroelectric responses. Core–shell architectures, gradient interfaces, and dopant segregation strategies are being refined to mitigate space charge accumulation and enhance dielectric breakdown strength. Particularly in heterophase nanocomposites, controlling the mismatch in dielectric permittivity and thermal expansion coefficients is essential to reduce stress-induced failure and ensure long-term reliability under cyclic loading or high-voltage operation. Simultaneously, the integration of ceramic nanocomposites into flexible, wearable, or neuromorphic systems demands a rethinking of material substrate compatibility. In Figure 14, we illustrate the various challenges and strategic solutions for future ceramic nanocomposite applications.
Traditional ceramics are inherently brittle and incompatible with flexible substrates, but advances in nanocomposite design such as embedding ceramics in polymer matrices or developing thin-film conformal coatings are helping overcome these barriers. For instance, dielectric inks that can be screen-printed or inkjet-printed onto PET or PI substrates are enabling the realization of flexible capacitors, sensors, and energy harvesters. The development of low-curing-temperature formulations is particularly critical for CMOS-backend integration, where thermal budgets must remain below 400 °C. Looking forward, the field must embrace computational and data-driven strategies to accelerate material discovery and device optimization. Machine learning (ML) algorithms trained on large datasets of composition structure property relationships are already assisting in predicting dielectric behavior and processing windows for complex oxides. High-throughput experimentation, when coupled with AI models, allows researchers to explore vast compositional spaces and identify high-performance candidates in a fraction of the time compared to traditional methods. Furthermore, materials informatics tools can guide the optimization of grain boundary chemistry, defect energetics, and sintering parameters to yield tunable, application-specific nanocomposites. Future research must address not only the performance optimization of ceramic nanocomposites but also their reliability under operational stressors such as thermal cycling, humidity exposure, and electrical fatigue. Sustainability goals should drive the development of lead-free compositions, low-energy processing, and recyclable systems. Furthermore, successful integration into commercial devices will require standardization of processing routes, scalability studies, and life-cycle assessments to ensure economic and environmental viability.

7. Conclusions

In this comprehensive review, we have explored the significant advancements in ceramic nanocomposites, focusing on their dielectric and ferroelectric behavior, structure–property relationships, and the wide range of processing strategies that enable tailored performance. By integrating insights from conventional sintering to spark plasma sintering and advanced wet-chemical routes, we observed how microstructural control particularly at grain boundaries and interfaces plays a critical role in dictating electrical responses. The multi-scale polarization mechanisms, including dipolar, interfacial, and ferroelectric contributions, are shown to be highly sensitive to interface design (core–shell vs. heterojunction) and the nanoscale morphology of the ceramic matrix. Innovative structural configurations have not only enhanced dielectric constant and polarization but also paved the way for new applications across flexible electronics, high-energy storage devices, and neuromorphic systems.
Emerging trends in material design increasingly emphasize sustainability, flexibility, and integration into multifunctional platforms. Lead-free systems, eco-friendly fabrication techniques, and AI-assisted materials discovery are all shaping the future trajectory of this field. Yet, challenges remain ranging from interface instability and defect control to scalability and integration into commercial systems. To overcome these barriers, a synergistic effort involving materials scientists, chemists, physicists, and engineers is essential. Future research should prioritize high-throughput synthesis, atomic-level interface engineering, and device-level validation to bring ceramic nanocomposites closer to real-world applications.

Author Contributions

Conceptualization was done by N.A.A., M.H. and K.I.N. Methodology and literature review were carried out by N.A.A., A.M.A. and A.J. Formal analysis and interpretation were performed by M.H., A.J. and K.I.N. Writing of the original draft was done by N.A.A. and M.H., while review and editing were managed by A.M.A., A.J. and K.I.N. Supervision was provided by K.I.N., project administration by N.A.A., and funding acquisition was also secured by N.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data was created.

Acknowledgments

The authors would like to thank the Deanship of Scientific Research at Majmaah University for supporting this work under Project Number: R-2025-1952.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Feng, G.; Jiang, F.; Jiang, W.; Liu, J.; Zhang, Q.; Wu, Q.; HU, Q.; Miao, L.; Liang, J. Low-temperature preparation of novel stabilized aluminum titanate ceramic fibers via nonhydrolytic sol-gel method through linear self-assembly of precursors. Ceram. Int. 2019, 45, 18704–18709. [Google Scholar] [CrossRef]
  2. Nassar, K.I.; Teixeira, S.S.; Graça, M.P. Sol–Gel-Synthesized Metal Oxide Nanostructures: Advancements and Prospects for Spintronic Applications—A Comprehensive Review. Gels 2025, 11, 657. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, H.H.; Gong, L.F.; Liu, X.Z.; Xu, Y.X.; Cheng, G.S.; Liu, Y.; Shi, G.M.; Zheng, C.; Han, Y.J. Design of low-SAR and high on-body efficiency tri-band smartwatch antenna utilizing the theory of characteristic modes of composite PEC-lossy dielectric structures. IEEE Trans. Antennas Propag. 2022, 71, 1913–1918. [Google Scholar] [CrossRef]
  4. Zhang, H.H.; Liu, X.Z.; Cheng, G.S.; Liu, Y.; Shi, G.M.; Li, K. Low-SAR four-antenna MIMO array for 5G mobile phones based on the theory of characteristic modes of composite PEC-lossy dielectric structures. IEEE Trans. Antennas Propag. 2021, 70, 1623–1631. [Google Scholar] [CrossRef]
  5. Fang, Q.; Sun, Q.; Ge, J.; Wang, H.; Qi, J. Multidimensional Engineering of Nanoconfined Catalysis: Frontiers in Carbon-Based Energy Conversion and Utilization. Catalysts 2025, 15, 477. [Google Scholar] [CrossRef]
  6. Zhang, H.H.; Wang, L.; Xu, Y.; Chen, X. 5G base station antenna array with heatsink radome. IEEE Trans. Antennas Propag. 2024, 72, 2270–2278. [Google Scholar] [CrossRef]
  7. Du, J.; Wang, X.; Liu, P.; Zhang, Y. Solidification microstructure reconstruction and its effects on phase transformation, grain boundary transformation mechanism, and mechanical properties of TC4 alloy welded joint. Metall. Mater. Trans. A 2024, 55, 1193–1206. [Google Scholar] [CrossRef]
  8. Zhao, L.-C.; Huang, Y.; Liu, Q.; Zhang, Y. Fast and Sensitive LC-DAD-ESI/MS Method for Analysis of Saikosaponins c, a, and d from the Roots of Bupleurum falcatum (Sandaochaihu). Molecules 2011, 16, 1533–1543. [Google Scholar] [CrossRef]
  9. Yang, Z.; Li, M.; Zhao, K.; Zhang, J. Types and space distribution characteristics of debris flow disasters along China-Pakistan Highway. Electron. J. Geotech. Eng. 2016, 21, 191–200. [Google Scholar]
  10. Yang, Z.; Liu, Y.; Wang, T.; Zhang, Q. Prediction model on maximum potential pollution range of debris flows generated in tailings dam break. Electron. J. Geotech. Eng. 2015, 20, 4363–4369. [Google Scholar]
  11. Gao, J. Friction coefficient estimation of the clutch in automatic transmission based on improved persistent excitation condition. In Proceedings of the 2017 Chinese Automation Congress (CAC), Jinan, China, 20–22 October 2017; IEEE: Piscataway, NJ, USA, 2017; pp. 1–6. [Google Scholar]
  12. Wu, Z.; Zhang, L.; Wang, Y.; Chen, J. Preliminary studies on diagnostic cast of peptic ulcer based on saliva proteome and bioinformatics. In Proceedings of the 2011 4th International Conference on Biomedical Engineering and Informatics (BMEI), Shanghai, China, 15–17 October 2011; IEEE: Piscataway, NJ, USA, 2011; Volume 3, pp. 1720–1724. [Google Scholar]
  13. Wu, Z.; Zhang, M.; Li, X.; Huang, A. Preliminary study on saliva proteomics of different furs in digestive system diseases. In Proceedings of the 2009 IEEE International Symposium on IT in Medicine & Education, Jinan, China, 14–16 August 2009; IEEE: Piscataway, NJ, USA, 2009; Volume 1, pp. 238–242. [Google Scholar]
  14. Wang, C.; Zhang, H.; Li, F.; Yan, Y. On-demand airport slot management: Tree-structured capacity profile and coadapted fire-break setting and slot allocation. Transp. A Transp. Sci. 2024, 1, 1–35. [Google Scholar] [CrossRef]
  15. Fan, J.; Zhang, X.; He, N.; Song, F.; Wang, X. Investigation on novel deep eutectic solvents with high carbon dioxide adsorption performance. J. Environ. Chem. Eng. 2025, 13, 117870. [Google Scholar] [CrossRef]
  16. Ye, D.; Huang, Y.; Zong, Y.; Gao, H. PO-SRPP: A decentralized pivoting path planning method for self-reconfigurable satellites. IEEE Trans. Ind. Electron. 2024, 71, 14318–14327. [Google Scholar] [CrossRef]
  17. Lv, S.; Li, T.; Yang, J.; Zhang, M. Effect of axial misalignment on the microstructure, mechanical, and corrosion properties of magnetically impelled arc butt welding joint. Mater. Today Commun. 2024, 40, 109866. [Google Scholar] [CrossRef]
  18. Hou, Y.; Zhou, X.; Zhao, J.; Chen, L. Assortative mating on blood type: Evidence from one million Chinese pregnancies. Proc. Natl. Acad. Sci. USA 2022, 119, e2209643119. [Google Scholar] [CrossRef]
  19. Zheng, D.; Cao, X. Provably efficient service function chain embedding and protection in edge networks. IEEE/ACM Trans. Netw. 2024, 33, 178–193. [Google Scholar] [CrossRef]
  20. Yi, X.; Zhao, R.; Lin, Y. The impact of nighttime car body lighting on pedestrians’ distraction: A virtual reality simulation based on bottom-up attention mechanism. Saf. Sci. 2024, 180, 106633. [Google Scholar] [CrossRef]
  21. Liao, H.; Zhang, R.; Wang, Q.; Sun, M. Ropinirole suppresses LPS-induced periodontal inflammation by inhibiting the NAT10 in an ac4C-dependent manner. BMC Oral Health 2024, 24, 510. [Google Scholar] [CrossRef]
  22. Zhang, Z.; Liu, X.; Chen, F.; Wang, J. An AUV-enabled dockable platform for long-term dynamic and static monitoring of marine pastures. IEEE J. Ocean. Eng. 2024, 50, 276–293. [Google Scholar] [CrossRef]
  23. Zhao, L.-C.; Liu, H.; Chen, Y.; Li, X. Preparation of Cyclocarya paliurus sugar-free jelly. Sci. Technol. Food Ind. 2020, 41, 234–239. [Google Scholar]
  24. Xiao, Y.; Zhang, L.; Guo, Q.; Zhao, S. Quantitative precision second-order temporal transformation based pose control for spacecraft proximity operations. IEEE Trans. Aerosp. Electron. Syst. 2024, 61, 1931–1941. [Google Scholar] [CrossRef]
  25. Huang, X.; Feng, G.; Mu, J.; Li, Y.; Xie, W.; Wu, F.; Guo, Z.; Xu, Y.; Wang, Z.; Jiang, F. Effects of sodium sources on nonaqueous precipitation synthesis of β″-Al2O3 and formation mechanism of uniform ionic channels. Langmuir 2025, 41, 2044–2052. [Google Scholar] [CrossRef]
  26. Wu, Z.; Zhang, M.; Huang, A. Effect of Natural Brain-Kenetine on variance of Gene expression profiles in MSC and hippocampus of AD rats analyzed by Gene chips and bioinformatics techniques. In Proceedings of the 2008 IEEE International Symposium on IT in Medicine and Education, Xiamen, China, 12–14 December 2008; IEEE: Piscataway, NJ, USA, 2008; pp. 112–116. [Google Scholar]
  27. Huang, A.C.J. Natural cerebrolysin induces neuronal differentiation in bone marrow mesenchymal stem cells. Neural Regen. Res. 2009, 3, 178–185. [Google Scholar]
  28. Wu, Z.; Zhang, M.; Li, X.; Huang, A. Endoplasmic reticulum stress induced by tunicamycin and antagonistic effect of Tiantai No. 1 on mesenchymal stem cells. Chin. J. Integr. Med. 2010, 16, 41–49. [Google Scholar] [CrossRef] [PubMed]
  29. Wu, Z.; Zhang, L.; Wang, Y.; Chen, J. Study on relationship between the thickness of tongue fur and the expressions of apoptosis-related genes of the tongue epithelial cells in patients with diseases of the digestive system. J. Tradit. Chin. Med. 2007, 27, 148–152. [Google Scholar]
  30. Wu, Z.; Zhang, M.; Liu, H.; Huang, A. Quantitative study of Tiantai I on superoxidative dismutase and lipofuscin in relevant cerebral areas of spontaneous Alzheimer disease in mice. Chin. J. Tissue Eng. Res. 2005, 9, 178–181. [Google Scholar]
  31. Wang, H.; Dong, F.; Wang, H.; Zhao, B.; Wang, Y.; Tan, W. Magnetic properties, critical behavior, and magnetocaloric effect of Nd1− xSrxMnO3 (0.2 ≤ x ≤ 0.5): The role of Sr doping concentration. J. Appl. Phys. 2024, 136, 093902. [Google Scholar] [CrossRef]
  32. Wu, Z.; Zhang, H.; Li, X.; Chen, Y. Influence of Tiantai No. 1 Recipe on learning and memory function of spontaneous Alzheimer disease models. Chin. J. Tissue Eng. Res. 2005, 9, 180–181. [Google Scholar]
  33. Huang, A.C.; Jia, X.; Li, Y.; Chen, M. Effects of serum containing natural cerebrolysin on glucose-regulated protein 78 and CCAAT enhancer-binding protein homologous protein expression in neuronal PC12 cells following tunicamycin-induced endoplasmic reticulum stress. Neural Regen. Res. 2009, 4, 92–97. [Google Scholar]
  34. Wu, Z.; Zhang, L.; Li, X.; Huang, A. Effect of Tiantai No. 1 on β-amyloid-induced neurotoxicity and NF-κB and cAMP responsive element-binding protein. Neural Regen. Res. 2008, 3, 286–292. [Google Scholar]
  35. Wu, Z.; Zhang, M.; Chen, J.; Li, X. Study on saliva proteome and bioinformatics in patients with chronic gastritis. In Proceedings of the 2011 IEEE International Symposium on IT in Medicine and Education, Cuangzhou, China, 30 August–2 September 2011; IEEE: Piscataway, NJ, USA, 2011; Volume 1, pp. 285–289. [Google Scholar]
  36. Huang, F.-J.; Wu, Z.-Z. From “unusual sequences” to human diseases. In Medicine Sciences and Bioengineering; CRC Press: Boca Raton, FL, USA, 2014. [Google Scholar]
  37. Wu, Z. Effects of Tiantai No. 1 on relative neuropeptides of spontaneous aged dementia mice. Chin. J. Neurosci. 2004, 20, 167–170. [Google Scholar]
  38. Wu, Z.; Zhang, M.; Chen, Y.; Li, X. Effects of Tiantai I on the activity of central cholinergic system in mice with spontaneous Alzheimer disease. Chin. J. Tissue Eng. Res. 2006, 10, 163–165. [Google Scholar]
  39. Wu, Z.; Zhang, L.; Wang, Y.; Huang, A. Changes of neuronal nitric oxide synthase in relevant cerebral regions in spontaneous senile dementia model and regulation of Tiantai I. Chin. J. Tissue Eng. Res. 2005, 9, 244–247. [Google Scholar]
  40. Tian, G.; Zhang, Y.; Wang, H.; Li, J. Leg-bathing in decoction in treating rheumatism, experimented on rats and demonstrated by visualization of knee joint synovium. Acta Med. Mediterr. 2016, 32, 1873–1879. [Google Scholar]
  41. Guihua, T.; Wang, Y.; Chen, L.; Zhou, H. Evidence-based traditional Chinese medicine research: Two decades of development, its impact, and breakthrough. J. Evid.-Based Med. 2021, 14, 65–74. [Google Scholar]
  42. Qian, G.; Li, X.; Sun, M.; Jiang, H.; Qiu, W.; Schuller, B.W. Can a holistic view facilitate the development of intelligent traditional Chinese medicine? A survey. IEEE Trans. Comput. Soc. Syst. 2023, 10, 700–713. [Google Scholar] [CrossRef]
  43. Yang, J.J. Analysis of clinical medication characteristics of zedoary turmeric oil injection based on real world data. Chin. J. New Drugs 2023, 32, 547–552. [Google Scholar]
  44. Li, X.; Zhao, T.; Chen, W.; Sun, Y. The analgesic mechanism of Xi Shao Formula research on neuropathic pain based on metabolomics. J. Tradit. Chin. Med. Sci. 2023, 10, 448–460. [Google Scholar] [CrossRef]
  45. Li, X.-Y.; Wang, M.; Guo, J.; Zhang, Y. Medication rules of traditional Chinese medicine compounds for pain. Zhongguo Zhong Yao Za Zhi 2023, 48, 3386–3393. [Google Scholar]
  46. Dai, Q.-Q.; Liu, X.; Zhao, R.; Yang, M. Research progress on animal models of chronic pain and its application in traditional Chinese medicine research. Zhongguo Zhong Yao Za Zhi 2020, 45, 5866–5876. [Google Scholar] [PubMed]
  47. Yang, Z.Q.; Hou, K.P.; Cheng, Y.; Yang, B. Study of column-hemispherical penetration grouting mechanism based on power-law fluid. Chin. J. Rock Mech. Eng. 2014, 33, 3840–3846. [Google Scholar]
  48. Xiang, D.; Zhang, H.; Liu, K.; Chen, X. HCMPE-Net: An unsupervised network for underwater image restoration with multi-parameter estimation based on homology constraint. Opt. Laser Technol. 2025, 186, 112616. [Google Scholar] [CrossRef]
  49. Yan, H.; Yang, B.; Zhou, X.; Qiu, X.; Zhu, D.; Wu, H.; Li, M.; Long, Q.; Xia, Y.; Chen, J.; et al. Adsorption mechanism of hydrated Lu (OH)2+ and Al (OH)2+ ions on the surface of kaolinite. Powder Technol. 2022, 407, 117611. [Google Scholar] [CrossRef]
  50. Yu, Y.; Ma, Q.; Liu, J.; Wang, Y. CrowdFPN: Crowd counting via scale-enhanced and location-aware feature pyramid network. Appl. Intell. 2025, 55, 359. [Google Scholar] [CrossRef]
  51. Guan, Y.; Cui, Z.; Zhou, W. Reconstruction in off-axis digital holography based on hybrid clustering and the fractional Fourier transform. Opt. Laser Technol. 2025, 186, 112622. [Google Scholar] [CrossRef]
  52. Zhang, M.-X.; Zhao, Y.; Lin, F.; Wu, L. Technical performance, surgical workload and patient outcomes of robotic and laparoscopic surgery for pediatric choledochal cyst: A multicenter retrospective cohort and propensity score-matched study. Hepatobiliary Surg. Nutr. 2025. [Google Scholar] [CrossRef]
  53. Yang, T.; Liu, Y.; Zhang, F.; Xu, Q. Causal association between circulating inflammatory markers and sciatica development: A Mendelian randomization study. Front. Neurol. 2024, 15, 1380719. [Google Scholar] [CrossRef]
  54. Lin, Y.-Y.; Wang, L.; Xu, J.; Zhang, R. Effect of Siegesbeckiae Herba in treating chronic pain. Zhongguo Zhong Yao Za Zhi 2020, 45, 1851–1858. [Google Scholar]
  55. Lin, Y.; Liu, M.; Zhao, H.; Cheng, J. Acupuncture combined with Chinese herbal medicine for discogenic low back pain: Protocol for a multicentre, randomised controlled trial. BMJ Open 2024, 14, e088898. [Google Scholar] [CrossRef]
  56. Yang, Z.Q.; Hou, K.P.; Guo, T.T. Research on time-varying behavior of cement grouts of different water-cement ratios. Appl. Mech. Mater. 2011, 71, 4398–4401. [Google Scholar]
  57. Ding, L.-N.; Zhang, Y.; Qiao, Z.; Wang, Q. A GDSL motif-containing lipase modulates Sclerotinia sclerotiorum resistance in Brassica napus. Plant Physiol. 2024, 196, 2973–2988. [Google Scholar] [CrossRef]
  58. Wu, Y.-Z.; Liu, R.; Zhang, M.; Hu, J. Antimicrobial peptides: Classification, mechanism, and application in plant disease resistance. Probiotics Antimicrob. Proteins 2025, 19, 1–15. [Google Scholar] [CrossRef] [PubMed]
  59. Chen, Z.; Li, B.; Wang, Q.; Xu, H. Creep behaviour between resilient wheels and rails in a metro system. Veh. Syst. Dyn. 2025, 22, 1–21. [Google Scholar] [CrossRef]
  60. Feng, G.; Jiang, W.; Liu, J.; Li, C.; Zhang, Q.; Miao, L.; Wu, Q. Synthesis and luminescence properties of Al2O3@ YAG: Ce core–shell yellow phosphor for white LED application. Ceram. Int. 2018, 44, 8435–8439. [Google Scholar] [CrossRef]
  61. Su, H.; Liu, X. The impact of green technological innovation on industrial structural optimization under dual-carbon targets: The role of the moderating effect of carbon emission efficiency. Sustainability 2025, 17, 6313. [Google Scholar] [CrossRef]
  62. Feng, G.; Jiang, W.; Liu, J.; Li, C.; Zhang, Q. Luminescent properties of novel red-emitting M7Sn (PO4)6: Eu3+ (M = Sr, Ba) for light-emitting diodes. Luminescence 2018, 33, 455–460. [Google Scholar] [CrossRef]
  63. Mengxin, Z.; Li, Q.; Zhang, X.; Wu, J. Robotic-assisted proctosigmoidectomy versus laparoscopic-assisted Soave pull-through for Hirschsprung disease: Medium-term outcomes from a prospective multicenter study. Ann. Surg. 2025, 281, 689–697. [Google Scholar]
  64. Huang, Y.; Wang, R.; Liu, Y.; Zhang, H. Travel and regional development: A quantitative analysis of China. J. Reg. Sci. 2025. [Google Scholar] [CrossRef]
  65. Yu, Z.; Zhang, B.; Wang, L.; Liu, H. A generalized Faustmann model with multiple carbon pools. For. Policy Econ. 2024, 169, 103363. [Google Scholar] [CrossRef]
  66. Xiang, X.; Li, Y.; Zhang, C.; Liu, F. Flavor profile of 4-isothiocyanato-1-butene in microwave rapeseed oil and its anti-inflammatory properties in vitro. J. Agric. Food Chem. 2025, 73, 10520–10530. [Google Scholar] [CrossRef]
  67. Yu, Z.; Ma, T.; Liu, D.; Zhang, X. Optimal harvest decisions for the management of carbon sequestration forests under price uncertainty and risk preferences. For. Policy Econ. 2023, 151, 102957. [Google Scholar] [CrossRef]
  68. Fan, Q.; Lu, Q.; Yang, X. Spatiotemporal assessment of recreation ecosystem service flow from green spaces in Zhengzhou’s main urban area. Humanit. Soc. Sci. Commun. 2025, 12, 1–14. [Google Scholar] [CrossRef]
  69. Li, B.; Zhang, T.; Chen, R.; Wang, Y. Environmental governance, green finance, and mitigation technologies: Pathways to carbon neutrality in European industrial economies. Int. J. Environ. Sci. Technol. 2025, 1–14. [Google Scholar] [CrossRef]
  70. Wang, X.; Zhou, J.; Li, M.; Zhang, F. E-governance and policy efforts advancing carbon neutrality and sustainability in European countries. Public Money Manag. 2025, 1–11. [Google Scholar] [CrossRef]
  71. Wang, Y.; Xiao, R.; Xiao, N.; Wang, Z.; Chen, L.; Wen, Y.; Li, P. Wireless Multiferroic Memristor with Coupled Giant Impedance and Artificial Synapse Application. Adv. Electron. Mater. 2022, 8, 2200370. [Google Scholar] [CrossRef]
  72. Xiao, Y.; Cheng, D.; Li, G.; Yin, R.; Li, P.; Gao, Z. Preparation of MgO ceramics by low temperature sintering with MgF2 and Al2O3 as sintering additives. J. Electroceram. 2025, 1–12. [Google Scholar] [CrossRef]
  73. Li, S.; Zhao, S.; Hartmann, D.M.F.; Yang, W.; Lin, X.; Zhao, W.; Lavrijsen, R. Modulation of the Dzyaloshinskii–Moriya interaction in Pt/Co/Pt with electric field induced strain. Appl. Phys. Lett. 2025, 126, 162401. [Google Scholar] [CrossRef]
  74. Yu, W.; Zheng, W.; Hua, S.; Zhang, Q.; Zhang, Z.; Zhao, J.; Guo, S. A prestretch-free dielectric elastomer with record-high energy and power density via synergistic polarization enhancement and strain stiffening. Adv. Funct. Mater. 2025, 27, 2425099. [Google Scholar] [CrossRef]
  75. Wu, W.; Chen, Y.; Xie, B.; Wu, H.; Cheng, L.; Guo, Y.; Chen, X. Microdynamic behaviors of Au/Ni-assisted chemical etching in fabricating silicon nanostructures. Appl. Surf. Sci. 2025, 696, 162915. [Google Scholar] [CrossRef]
  76. Yang, D.; Lin, Y.; Meng, W.; Wang, Z.; Li, H.; Li, C.; Sun, L. Relaxor antiferroelectric dynamics for neuromorphic computing. Adv. Mater. 2025, 24, 2419204. [Google Scholar] [CrossRef]
  77. Li, C.; Chen, X.; Zhang, Z.; Wu, X.; Yu, T.; Bie, R.; Sun, L. Charge-selective 2D heterointerface-driven multifunctional floating gate memory for in situ sensing-memory-computing. Nano Lett. 2024, 24, 15025–15034. [Google Scholar] [CrossRef]
  78. Zhao, Y.; Jing, J.; Chen, L.; Xu, F.; Hou, H. Current research status of interface of ceramic-metal laminated composite material for armor protection. Acta Metall. Sin. 2021, 57, 1107–1125. [Google Scholar]
  79. Palneedi, H.; Peddigari, M.; Hwang, G.T.; Jeong, D.Y.; Ryu, J. High-performance dielectric ceramic films for energy storage capacitors: Progress and outlook. Adv. Funct. Mater. 2018, 28, 1803665. [Google Scholar] [CrossRef]
  80. Vignesh, D.; Rout, E. Multipurpose polymer dielectric materials and composites for advanced energy applications. In Metal Oxide-Based High-K Dielectrics; CRC Press: Boca Raton, FL, USA, 2025; pp. 289–332. [Google Scholar]
  81. Thakur, O.P.; Prakash, C.; James, A.R. Enhanced dielectric properties in modified barium titanate ceramics through improved processing. J. Alloys Compd. 2009, 470, 548–551. [Google Scholar] [CrossRef]
  82. Hanani, Z. Design of Flexible Lead-Free Ceramic/Biopolymer Composite for Energy Storage and Energy Harvesting Applications. Ph.D. Dissertation, Université Cadi Ayyad, Marrakech, Maroc, 2020. [Google Scholar]
  83. Thapliyal, D.; Tewari, K.; Verma, S.; Bhargava, C.K.; Sen, P.; Mehra, A.; Rana, S.; Verros, G.D.; Arya, R.K. Introduction: The evolution of functional coatings from protection to innovation. In Functional Coatings for Biomedical, Energy, and Environmental Applications; Wiley: Hoboken, NJ, USA, 2024; pp. 1–30. [Google Scholar] [CrossRef]
  84. Prasad, G.; Chakraborty, U.; Matete, E.S.; Marwaha, K.S.; Vero, R. Innovation of bio-inspired materials for next-generation defense applications. In Innovative Materials for Next-Generation Defense Applications: Cost, Performance, and Mass Production; IGI Global: Hershey, PA, USA, 2025; pp. 145–180. [Google Scholar]
  85. Jamwal, U.; Kumar, S.; Keneria, D. Polarization and ferromagnetism in microwave-absorbing materials. In Ferroic Materials-Understanding, Development, and Utilization: Understanding, Development, and Utilization; IntechOpen: Rijeka, Croatia, 2025; p. 71. [Google Scholar]
  86. Wei, Z.; Li, Z.; Chen, D.; Liang, J.; Kong, J. Recent progress of advanced composites for broadband electromagnetic wave absorption. Small Struct. 2025, 6, 2400615. [Google Scholar] [CrossRef]
  87. Galassi, C. Overview: Electroceramics and ceramics and glasses in energy generation and storage. Encycl. Mater. Tech. Ceram. Glas 2021, 3, 1–4. [Google Scholar]
  88. Li, D. Investigation of the General Properties and Field-Induced Electromechanical Response of Polymer Nanocomposites with Surface-Functionalised Dielectric Nanoparticles. Ph.D. Dissertation, Cranfield University, Wharley End, UK, 2022. [Google Scholar]
  89. Uguen, N. Dispersion State, Interfacial Phenomena and Dielectric Properties in High-Permittivity Polymer-Based Nanocomposites. Ph.D. Dissertation, Université de Lyon, Lyon, France, 2022. [Google Scholar]
  90. Halim, N.A.B. Ferroelectric, Piezoelectric and Pyroelectric Properties of Sol-Gel Derived Sodium Bismuth Titanate and Ceramic Powder-Polymer Composite. Ph.D. Dissertation, University Malaya, Kuala Lumpur, Malaysia, 2018. [Google Scholar]
  91. Nayak, D.; Choudhary, R.B.; Kumar, S.; Bauri, J.; Ansari, S. Effect of nanofillers-reinforced polymer blends for dielectric applications. In Emerging Nanodielectric Materials for Energy Storage; Springer: Berlin/Heidelberg, Germany, 2023; pp. 151–187. [Google Scholar]
  92. Patsidis, A.C.; Psarras, G.C. Basic principles of dielectrics. In High Temperature Polymer Dielectrics: Fundamentals and Applications in Power Equipment; Wiley: Hoboken, NJ, USA, 2024; pp. 21–55. [Google Scholar]
  93. Tan, D.Q. The search for enhanced dielectric strength of polymer-based dielectrics: A focused review on polymer nanocomposites. J. Appl. Polym. Sci. 2020, 137, 49379. [Google Scholar] [CrossRef]
  94. Wang, S.; Luo, Z.; Liang, J.; Hu, J.; Jiang, N.; He, J.; Li, Q. Polymer nanocomposite dielectrics: Understanding the matrix/particle interface. ACS Nano 2022, 16, 13612–13656. [Google Scholar] [CrossRef] [PubMed]
  95. Nandi, S.; Kerur, S.S.; Dhanalakshmi, S. Electrical and dielectric properties of polymer-metal hybrid nanocomposites—A short review. Diffus. Found. Mater. Appl. 2024, 35, 1–13. [Google Scholar] [CrossRef]
  96. Sahu, G.; Das, M.; Yadav, M.; Sahoo, B.P.; Tripathy, J. Dielectric relaxation behavior of silver nanoparticles and graphene oxide embedded poly(vinyl alcohol) nanocomposite film: Effect of ionic liquid and temperature. Polymers 2020, 12, 374. [Google Scholar] [CrossRef]
  97. Tayari, F.; Dhahri, R.; Elkenany, E.B.; Teixeira, S.S.; Graça, M.P.F.; Al-Syadi, A.M.; Essid, M.; Iben Nassar, K. Crystal structural characteristics and electrical properties of novel sol-gel synthesis of ceramic Bi0.75Ba0.25(FeMn)0.5O3. Materials 2024, 17, 3797. [Google Scholar] [CrossRef]
  98. Tayari, F.; Benamara, M.; Lal, M.; Essid, M.; Thakur, P.; Kumar, D.; Teixeira, S.S.; Graça, M.P.F.; Nassar, K.I. Exploring enhanced structural and dielectric properties in Ag-Doped Sr(NiNb)0.5O3 perovskite ceramic for advanced energy storage. Ceramics 2024, 7, 958–974. [Google Scholar] [CrossRef]
  99. Tayari, F.; Nassar, K.I.; Algessair, S.; Hjiri, M.; Benamara, M. Investigating Fe-doped Ba0.67Ni0.33Mn1−xFexO3 (x = 0, 0.2) ceramics: Insights into electrical and dielectric behaviors. RSC Adv. 2024, 14, 12561–12573. [Google Scholar] [CrossRef]
  100. Borah, M. Doping in barium titanate (BaTiO3): A historical perspective and future directions. J. Adv. Chem. Sci. 2025, 11, 838–841. [Google Scholar] [CrossRef]
  101. Mo, S.; Hu, S.; Wang, N.; Zhang, H.; Zhu, P.; Wen, X.; Dai, S.; Zhang, F.; Li, T.; Liu, J.; et al. Temperature-stable dielectric properties of Nb2O5-doped BaTiO3 based ceramics with fine grain via controlling grain growth. Ceram. Int. 2025, 51, 26791–26798. [Google Scholar] [CrossRef]
  102. O’Reilly, T.; Holsgrove, K.; Gholinia, A.; Woodruff, D.; Bell, A.; Huber, J.; Arredondo, M. Exploring domain continuity across BaTiO3 grain boundaries: Theory meets experiment. Acta Mater. 2022, 235, 118096. [Google Scholar] [CrossRef]
  103. Rana, S.; Karimunnesa, S.; Alam, F.; Das, B.C.; Khan, F.A. Strain induced structural phase transition and compositional dependent magnetic phase transition in Ti doped Bi0.80Ba0.20FeO3 ceramics. Heliyon 2022, 8, e12530. [Google Scholar] [CrossRef]
  104. Reddy, T.; Akepati, S.R.; Nagalakshmi, V.; Rao, D.J.; Madaka, R. Effect of rare earth (Sm3+) substitution in mixed Ni-Zn-Co ferrites: Structural, magnetic, and DC electrical resistivity studies. In. Chem. Com. 2024, 165, 112480. [Google Scholar] [CrossRef]
  105. Alikin, D.; Turygin, A.; Ushakov, A.; Kosobokov, M.; Alikin, Y.; Hu, Q.; Liu, X.; Xu, Z.; Wei, X.; Shur, V. Competition between ferroelectric and ferroelastic domain wall dynamics during local switching in rhombohedral PMN-PT single crystals. Nanomaterials 2022, 12, 3912. [Google Scholar] [CrossRef]
  106. Meng, Z.; Zhang, Y.; Liu, X.; Wang, H. Advances in polymer dielectrics with high energy storage performance by designing electric charge trap structures. Adv. Mater. 2024, 36, 2310272. [Google Scholar] [CrossRef]
  107. Anaele Opara, F.; Ibekwe, S.; Okoye, R.; Uzoegwu, D.; Madueke, I. Progress in polymer-based composites as efficient materials for electromagnetic interference shielding applications: A review. Curr. Mater. Sci. 2023, 16, 235–261. [Google Scholar] [CrossRef]
  108. González, J.; Singh, E.; Salas, C.; Kumar, R.; Bhatia, R. Advanced cellulose–nanocarbon composite films for high-performance triboelectric and piezoelectric nanogenerators. Nanomaterials 2023, 13, 1206. [Google Scholar] [CrossRef]
  109. You, L.; Chen, Z.; Wang, Y.; Zhao, J.; Zhang, D. Energy storage performance of polymer-based dielectric composites with two-dimensional fillers. Nanomaterials 2023, 13, 2842. [Google Scholar] [CrossRef]
  110. Mishra, R.K. A Study of Control Mechanisms in Micro and Nano System-Enhanced Polymer Nanocomposites Under Mechanical and Electrical Stimuli: An Experimental and Computational Investigation. Ph.D. Thesis, Cranfield University, Bedford, UK, 2023. [Google Scholar]
  111. Ghule, B.; Laad, M. Polymer composites with improved dielectric properties: A review. Ukr. J. Phys. 2021, 66, 166. [Google Scholar] [CrossRef]
  112. Meng, X.; Zhou, W.; Chen, X.; Kong, F.; Zhao, J.; Li, W.; Zhang, Y.; Wang, F.; Yuan, M. Synergistic regulation of intra-particle and inter-particle polarizations in BaTiO3@ Al2O3/PVDF nanocomposites towards boosted overall dielectric properties. Mat. Today Chem. 2025, 43, 102492. [Google Scholar] [CrossRef]
  113. Feng, Y.; Liu, L.; Chen, P.; Bo, M.; Xie, J.; Deng, Q. Enhancing surface polarization and reducing bandgap of BaTiO3 nanofiller for preparing dielectric traits-improved composites via its hybridization with layered g-C3N4. Sur. Int. 2022, 31, 102060. [Google Scholar] [CrossRef]
  114. Pan, D. Lead zirconate titanate (PZT) piezoelectric ceramics: Applications and prospects in human motion monitoring. Ceram.-Silik. 2024, 68, 444–458. [Google Scholar] [CrossRef]
  115. Song, S.; Zhang, Y.; Li, J.; Chen, R. Research advances in rare-earth-based solid electrolytes for all-solid-state batteries. Small 2025, 21, 2502008. [Google Scholar] [CrossRef]
  116. Hernández Lara, J.P.; Estrada, W.; Reyes-López, J.C.; Ramírez-Bon, R. Structural evolution and electrical properties of BaTiO3 doped with Gd3+. Mater. Res. 2017, 20, 538–542. [Google Scholar] [CrossRef]
  117. Ianculescu, A.C.; Tiseanu, C.; Voicu, F.; Preda, S.; Gartner, M.; Ianculescu, M. Formation mechanism and characteristics of lanthanum-doped BaTiO3 powders and ceramics prepared by the sol–gel process. Mater. Charact. 2015, 106, 195–207. [Google Scholar] [CrossRef]
  118. Xiao, Y.; Zhang, H.; Liu, Y.; Wang, F.; Zhang, Y. Cold-Sintered ZnO Ceramic Composites Co-Doped with Polytetrafluoroethylene and Oxides. Molecules 2023, 29, 129. [Google Scholar] [CrossRef]
  119. Yang, L.; Ditta, A.; Feng, B.; Zhang, Y.; Xie, Z. Study of the comparative effect of sintering methods and sintering additives on the microstructure and performance of Si3N4 ceramic. Materials 2019, 12, 2142. [Google Scholar] [CrossRef]
  120. Chinelatto, A.S.A.; Santos, C.M.A.; Zaghete, M.A.; Gimenes, R.; Longo, E. Mechanisms of microstructure control in conventional sintering. In Sintering of Ceramics—New Emerging Techniques; Lakshmanan, A., Ed.; InTech: Rijeka, Croatia, 2012; pp. 401–422. [Google Scholar]
  121. Bram, M.; Zapf, M.; Buchkremer, H.P.; Stöver, D. Application of electric current-assisted sintering techniques for the processing of advanced materials. Adv. Eng. Mater. 2020, 22, 2000051. [Google Scholar] [CrossRef]
  122. Li, X.Y.; Zhang, H.L.; Sun, J.L.; Liu, X.Y.; Li, W.Z. The development and application of spark plasma sintering technique in advanced metal structure materials: A review. Powder Metall. Met. Ceram. 2021, 60, 410–438. [Google Scholar] [CrossRef]
  123. Radingoana, P.M. Densification and Microstructural Characterization of ZnO-Based Ceramics Obtained by SPS Sintering for Thermoelectric Application. Ph.D. Thesis, Université Paul Sabatier-Toulouse III, Toulouse, France, 2019. [Google Scholar]
  124. Colomban, P. Chemical preparation routes and lowering the sintering temperature of ceramics. Ceramics 2020, 3, 312–339. [Google Scholar] [CrossRef]
  125. Almeida, R.M.; Gonçalves, M.C. Sol–Gel process and products. In Encyclopedia of Glass Science, Technology, History, and Culture; Richet, P., Conradt, R., Takada, A., Dyon, J., Eds.; Wiley: Hoboken, NJ, USA, 2021; Volume 2, pp. 969–979. [Google Scholar]
  126. Polinger, V.; Bersuker, I.B. Origin of polar nanoregions and relaxor properties of ferroelectrics. Phys. Rev. B 2018, 98, 214102. [Google Scholar] [CrossRef]
  127. Zhang, X.L.; Zhu, J.J.; Zhang, J.Z.; Xu, G.S.; Hu, Z.G.; Chu, J.H. Photoluminescence study on polar nanoregions and structural variations in Pb(Mg1/3Nb2/3)O3–PbTiO3 single crystals. Opt. Express 2014, 22, 21903–21911. [Google Scholar] [CrossRef]
  128. Su, R.; Zhang, X.; Li, W.; Li, H.; Wang, Q. High energy density performance of polymer nanocomposites induced by designed formation of BaTiO3@sheet-like TiO2 hybrid nanofillers. J. Phys. Chem. C 2016, 120, 11769–11776. [Google Scholar] [CrossRef]
  129. Pattipaka, S.; Manukonda, S.R.; Gattu, K.P.; Gangavaram, V.; Chukka, S.; Kumar, C.M. Enhanced energy storage performance and efficiency in Bi0.5(Na0.8K0.2)0.5TiO3–Bi0.2Sr0.7TiO3 relaxor ferroelectric ceramics via domain engineering. Materials 2023, 16, 4912. [Google Scholar] [CrossRef]
  130. Khaliq, A.; Sheeraz, M.; Ullah, A.; Seog, H.J.; Ahn, C.W.; Kim, T.H.; Kim, I.W. Ferroelectric seeds-induced phase evolution and large electrostrain under reduced poling field in bismuth-based composites. Ceram. Int. 2018, 44, 13278–13285. [Google Scholar] [CrossRef]
  131. Sheeraz, M.; Khaliq, A.; Ullah, A.; Han, H.S.; Khan, A.; Ullah, A.; Ahn, C.W. Stress driven high electrostrain at low field in incipient piezoelectrics. J. Eur. Ceram. Soc. 2019, 39, 4688–4696. [Google Scholar] [CrossRef]
  132. Matizamhuka, W.R. Fabrication of fine-grained functional ceramics by two-step sintering or Spark Plasma Sintering (SPS). In Design and Manufacturing; IntechOpen: Rijeka, Croatia, 2019. [Google Scholar]
  133. Bordia, R.K.; Kang, S.-J.L.; Olevsky, E.A. Current understanding and future research directions at the onset of the next century of sintering science and technology. J. Am. Ceram. Soc. 2017, 100, 2314–2352. [Google Scholar] [CrossRef]
  134. Tayari, F.; Nassar, K.I.; Carvalho, J.P.; Teixeira, S.S.; Hammami, I.; Gavinho, S.R.; Graça, M.P.; Valente, M.A. Sol–Gel Synthesis and Comprehensive Study of Structural, Electrical, and Magnetic Properties of BiBaO3 Perovskite. Gels 2025, 11, 450. [Google Scholar] [CrossRef]
  135. Alrebdi, T.A.; Alkallas, F.H.; Amin, L.G.; Abdel-Aty, A.H.; Alshehri, K.; Alofi, A.; Gomaa, H.M. Influence of calcium substitution on the structural, thermal, and electrical properties of PbTiO3 perovskite. J. Mater. Sci. Mater. Electron. 2025, 36, 1235. [Google Scholar] [CrossRef]
  136. Daniel, T.O.; Okafor, L.C.; Arikpo, J.U.; Udeh, E.; Shaibu, O.D.; Edaogbogun, G.O.; Olawale, P.E. Synthesis and Characterization of Calcium and Tin Co-Doped Barium Titanate (Ba0.91Ca0.09Sn0.01Ti0.99O3) Ceramic Using Solid State Synthesis for Energy Storage Application in Ceramic Capacitor. J. Chem. Soc. Pak. 2025, 47. [Google Scholar]
  137. Freeman, C.J. Use of Sol Chemistry and Fine Grained Precursors in the Production of Controlled Microstructural Polycrystalline Continuous Oxide Fibres. Ph.D. Thesis, University of Wolverhampton, Wolverhampton, UK, 2005. [Google Scholar]
  138. Prompa, K.; Swatsitang, E.; Putjuso, T. Very low loss tangent and giant dielectric properties of CaCu3Ti4O12 ceramics prepared by the sol–gel process. J. Mater. Sci. Mater. Electron. 2017, 28, 15033–15042. [Google Scholar] [CrossRef]
  139. Sharma, N.; Alam, S.N.; Ray, B.C. Fundamentals of spark plasma sintering (SPS): An ideal processing technique for fabrication of metal matrix nanocomposites. In Spark Plasma Sintering of Materials: Advances in Processing and Applications; Springer: Cham, Switzerland, 2019; pp. 21–59. [Google Scholar]
  140. Akinyemi, S.A.; Ogundele, D.A.A.; Akinyemi, A.M. Cefixime and Aspirin alters antioxidant response, biochemical composition, and primary productivity of phytoplankton: A Mesocosm approach. Afr. J. Pure Appl. Sci. 2025, 6, 37–45. [Google Scholar]
  141. Bobylev, G.S.; Prikhodko, Y.M.; Pikalov, M.Y.; Besshaposhnikov, A.V.; Melnichenko, A.V.; Bazhin, V.Y. Full-scale experimental test production of high-basicity sinter in the Enakievo metallurgical plant sinter shop. Metallurgist 2020, 64, 741–749. [Google Scholar] [CrossRef]
  142. Kim, C.W.; Park, J.H.; Han, J.; Kim, K.S.; Kim, Y.D.; Kim, S.J. A selective morphosynthetic approach for single crystalline hematite through morphology evolution via microwave-assisted hydrothermal synthesis. J. Ind. Eng. Chem. 2017, 53, 341–347. [Google Scholar] [CrossRef]
  143. Canu, G.; Buscaglia, V. Hydrothermal synthesis of strontium titanate: Thermodynamic considerations, morphology control and crystallisation mechanisms. CrystEngComm 2017, 19, 3867–3891. [Google Scholar] [CrossRef]
  144. Walton, R.I. Perovskite oxides prepared by hydrothermal and solvothermal synthesis: A review of crystallisation, chemistry, and compositions. Chem.–Eur. J. 2020, 26, 9041–9069. [Google Scholar] [CrossRef]
  145. Tayari, F.; Attaf, A.; Attaf, N.; Al-Douri, Y.; Omri, K.; Touati, F. Progress and developments in the fabrication and characterization of metal halide perovskites for photovoltaic applications. Nanomaterials 2025, 15, 613. [Google Scholar] [CrossRef]
  146. Teixeira, S.S.; Graça, M.P.F.; Essid, M.; Nassar, K.I. Investigating structural, dielectric, and electrical characteristics of sol–gel synthesized perovskite ceramic Bi0.7Ba0.3(FeTi)0.5O3. J. Sol-Gel Sci. Technol. 2024, 112, 601–613. [Google Scholar] [CrossRef]
  147. Benamara, M.; Iben Nassar, K.; Rivero-Antúnez, P.; Essid, M.; Soreto Teixeira, S.; Zhao, S.; Serrà, A.; Esquivias, L. Study of electrical and dielectric behaviors of copper-doped zinc oxide ceramic prepared by spark plasma sintering for electronic device applications. Nanomaterials 2024, 14, 402. [Google Scholar] [CrossRef]
  148. Alanazi, A.Z.; Alhazzani, K.; El-Wekil, M.M.; Ali, A.M.B.H.; Darweesh, M.; Ibrahim, H. Sol–gel derived ceramic nanocomposite CNFs anchored with a nanostructured CeO2 modified graphite electrode for monitoring the interaction of a selective tyrosine kinase inhibitor capmatinib with dsDNA. RSC Adv. 2024, 14, 34448–34456. [Google Scholar] [CrossRef]
  149. Li, Y. Joining Technology and Application of Advanced Materials; Springer: Cham, Switzerland, 2023. [Google Scholar]
  150. Wang, S.W.; Chen, L.D.; Hirai, T. Densification of Al2O3 powder using spark plasma sintering. J. Mater. Res. 2000, 15, 982–987. [Google Scholar] [CrossRef]
  151. Somasundaram, M.; Uttamchand, N.K.; Annamalai, A.R.; Jen, C.P. Insights on spark plasma sintering of magnesium composites: A review. Nanomaterials 2022, 12, 2178. [Google Scholar] [CrossRef] [PubMed]
  152. Nie, B.; Chang, W.; Chen, X.; Fu, K. Cold sintering-enabled interface engineering of composites for solid-state batteries. Front. Energy Res. 2023, 11, 1149103. [Google Scholar] [CrossRef]
  153. Jabr, A.; Jambhulkar, S.; Masadeh, A.; Li, J.F.; Randall, C.A. Scaling up the cold sintering process of ceramics. J. Eur. Ceram. Soc. 2023, 43, 5319–5329. [Google Scholar] [CrossRef]
  154. Bartoletti, A.; Populoh, S.; Sata, N.; Battaglia, C. Exploring the potential of cold sintering for proton-conducting ceramics: A review. Materials 2024, 17, 5116. [Google Scholar] [CrossRef] [PubMed]
  155. Kalyani, V.; Palani, I.A.; Liu, Y.; Hu, Z. Hydrothermal synthesis of SrTiO3 mesocrystals: Single crystal to mesocrystal transformation induced by topochemical reactions. Cryst. Growth Des. 2012, 12, 4450–4456. [Google Scholar] [CrossRef]
  156. Tazim, T.Q.; Kumar, A.; Thakur, N.S.; Majumdar, D.; Singh, S. Hydrothermal synthesis of nano-metal oxides for structural modification: A review. Next Nanotechnol. 2025, 7, 100167. [Google Scholar] [CrossRef]
  157. Almutairi, H.H.; Parveen, N.; Ansari, S.A. Hydrothermal synthesis of multifunctional bimetallic Ag-CuO nanohybrids and their antimicrobial, antibiofilm and antiproliferative potential. Nanomaterials 2022, 12, 4167. [Google Scholar] [CrossRef]
  158. Cao, W.; Wang, Z.; Wang, J.; Zeng, H.; Tan, X.; Zhang, X. Interfacial-polarization engineering in BNT-based bulk ceramics for ultrahigh energy-storage density. Adv. Sci. 2024, 11, 2409113. [Google Scholar] [CrossRef]
  159. Thakur, Y.; Ghosh, S.; Katiyar, R.S.; Guha, S. Enhancement of the dielectric response in polymer nanocomposites with low dielectric constant fillers. Nanoscale 2017, 9, 10992–10997. [Google Scholar] [CrossRef] [PubMed]
  160. Cao, W.; Wu, Y.; Yang, X.; Guan, D.; Huang, X.; Li, F.; Guo, Y.; Wang, C.; Ge, B.; Hou, X.; et al. Breaking polarization-breakdown strength paradox for ultrahigh energy storage density in NBT-based ceramics. Nat. Commun. 2025, 16, 6228. [Google Scholar] [CrossRef] [PubMed]
  161. Guillemet, S.; Estournès, C.; Laurent, C.; Peigney, A.; Rousset, A. Colossal permittivity in ultrafine grain size BaTiO3–x and Ba0.95La0.05TiO3–x materials. Adv. Mater. 2008, 20, 551–555. [Google Scholar] [CrossRef]
  162. Al-Momin, A. Effect of Ferromagnetic and Ferroelectric Phases on the Magnetic and Transport Properties of xLi0.1Ni0.2Mn0.6Fe2.1O4+(1−x) Bi1−yRyFeO3 Multiferroic Composites. Ph.D. Thesis, Bangladesh University of Engineering and Technology (BUET), Dhaka, Bangladesh, 2020. [Google Scholar]
  163. Naveed-Ul-Haq, M. Magnetoelectric Effect in Lead-Free Multiferroic Composites and Thin Films; Universität Duisburg-Essen: Essen, Germany, 2018. [Google Scholar]
  164. Valiulina, L.I.; Cherepanov, V.N.; Valiev, R.R. Relationship between the electric polarizability and aromaticity of metallocene-containing macrocycles. Eurasian J. Chem. 2023, 28, 3. [Google Scholar] [CrossRef]
  165. Kuziemski, A.; Łączkowski, K.Z.; Baranowska-Łączkowska, A. Theoretical investigation of electric polarizability in porphyrin–zinc and porphyrin–zinc–thiazole complexes using small property-oriented basis sets. Int. J. Mol. Sci. 2024, 25, 11044. [Google Scholar] [CrossRef]
  166. Yan, X.J.; Tang, Y.; Zhang, X.; Li, Z.; Gao, H. Dependence of microwave dielectric properties on ceramic connectivity in ultralow-εr Al2O3 porous ceramics. Microstructures 2025, 5, 2025065. [Google Scholar] [CrossRef]
  167. Koval, V.; Jastrabik, L.; Kovacova, M.; Bobak, A.; Fetea, D.; Buryy, O. Dielectric relaxation and conductivity phenomena in ferroelectric ceramics at high temperatures. J. Eur. Ceram. Soc. 2024, 44, 2886–2902. [Google Scholar] [CrossRef]
  168. Zhang, Y. Optical Properties and Modulated Luminescence of Metal Ion Doped Phosphors. Ph.D. Thesis, Hong Kong Polytechnic University, Hong Kong, China, 2012. [Google Scholar]
  169. Zaman, T. Study of Structural, Dielectric and Ferroelectric Properties of Lead-Free BaTiO3–K0.5Na0.5NbO3 Ceramics. Master’s Thesis, Rajshahi University of Engineering & Technology (RUET), Rajshahi, Bangladesh, 2016. [Google Scholar]
  170. Sun, Y.; Zhang, L.; Chen, H.; Xiong, R.; Zhao, H. Defect engineering in perovskite oxide thin films. Chem. Commun. 2021, 57, 8402–8420. [Google Scholar] [CrossRef]
  171. Patsidis, A.C.; Papathanassiou, A.N.; Sakellis, E. Dielectric response of ZnO/PMMA nanocomposites with atmospheric pressure plasma-modified surfaces. Materials 2024, 17, 4063. [Google Scholar] [CrossRef]
  172. Anoufa, M.; Kiat, J.M.; Bogicevic, C. Ultrahigh permittivity in core-shell ferroelectric ceramics: Theoretical approach and practical conclusions. arXiv 2017, arXiv:1712.04197. [Google Scholar] [CrossRef]
  173. Liu, X.; Wang, T.; Zhang, Y.; Chen, L.; Li, Y. A BaTiO3-based flexible ferroelectric capacitor for non-volatile memories. J. Mater. 2025, 11, 100870. [Google Scholar] [CrossRef]
  174. Moise, T.; Summerfelt, S.; Rodriguez, J. A framework for embedded non-volatile memory development: Examples from Pb(ZrxTi1−x)O3 ferroelectric memory development at Texas Instruments. Electronics 2025, 14, 818. [Google Scholar] [CrossRef]
  175. Goel, R.; Gautam, D.; Singh, R.; Ghosh, S.; Pandey, D. Magnetoelectric coupling susceptibility in novel lead-free 0–3 type multiferroic particulate composites of (1−x)Na0.5Bi0.5TiO3–(x)CoCr0.4Fe1.6O4. Mater. Chem. Phys. 2022, 282, 126004. [Google Scholar] [CrossRef]
  176. Sato, K.; Asakura, N. Domain switching dynamics in relaxor ferroelectric Pb(Mg1/3Nb2/3)O3–PbTiO3 revealed by time-resolved high-voltage electron microscopy. J. Appl. Phys. 2021, 130, 164101. [Google Scholar] [CrossRef]
  177. Tennery, V.J.; Hang, K.W.; Novak, R.E. Ferroelectric and structural properties of the Pb(Sc1/2Nb1/2)1−xTixO3 system. J. Am. Ceram. Soc. 1968, 51, 671–674. [Google Scholar] [CrossRef]
  178. Mao, P.; Liu, H.; Luo, D.; Guo, C.; Li, Q.; Wu, K.; Zhuang, Y. Tunable dielectric polarization and breakdown behavior for high energy storage capability in P(VDF–TrFE–CFE)/PVDF polymer blended composite films. Phys. Chem. Chem. Phys. 2020, 22, 13143–13153. [Google Scholar] [CrossRef]
  179. Hou, X.; Li, W.; Liu, Y.; Li, Y.; Wang, L.; Liu, Y. Polymer nanocomposite dielectric based on P(VDF-TrFE)/PMMA/BaTiO3 for TIPs-pentacene OFETs. Org. Electron. 2015, 17, 247–252. [Google Scholar] [CrossRef]
  180. Rehman, M.U.; Zhang, J.; He, W.; Liu, Y.; Zhang, J.; Fan, J. The effects of SiO2 addition on the phase, microstructure, dielectric, and energy storage properties of BaTiO3-based ceramics. Mater. Sci. Eng. B 2023, 288, 116190. [Google Scholar] [CrossRef]
  181. He, D.; Wang, Y.; Chen, X.; Deng, Y. Core–shell structured BaTiO3@Al2O3 nanoparticles in polymer composites for dielectric loss suppression and breakdown strength enhancement. Compos. Part A Appl. Sci. Manuf. 2017, 93, 137–143. [Google Scholar] [CrossRef]
  182. Selvi, N.; Sankar, S.; Dinakaran, K. Effect of shell ZnO on the structure and optical property of TiO2 core@shell hybrid nanoparticles. J. Mater. Sci. Mater. Electron. 2015, 26, 2271–2277. [Google Scholar] [CrossRef]
  183. Eliseev, E.A.; Morozovska, A.N.; Kalinin, S.V.; Svechnikov, S.V.; Maksymovych, P. Strain-induced polarization enhancement in BaTiO3 core-shell nanoparticles. Phys. Rev. B 2024, 109, 014104. [Google Scholar] [CrossRef]
  184. Mohamed, M.M.; Reda, S.M.; Amer, A.A. Enhanced performance of BiFeO3@nitrogen doped TiO2 core–shell structured nanocomposites: Synergistic effect towards solar cell amplification. Arab. J. Chem. 2020, 13, 2611–2619. [Google Scholar] [CrossRef]
  185. Kumari, M.L.A.; Shanker, R.; Chatterjee, S.; Ramana, C.V. Mechanochemical synthesis of ternary heterojunctions TiO2(A)/TiO2(R)/ZnO and TiO2(A)/TiO2(R)/SnO2 for effective charge separation in semiconductor photocatalysis: A comparative study. Environ. Res. 2022, 203, 111841. [Google Scholar] [CrossRef]
  186. Upadhyay, M.; Bandyopadhyay, S.; Nath, T.K.; Pal, A. Thickness dependence of dielectric properties in sub-nanometric Al2O3/ZnO laminates. Solid-State Electron. 2021, 186, 108070. [Google Scholar] [CrossRef]
  187. Grabowska, E.; Zaleska-Medynska, A.; Marchelek, M.; Hupka, J.; Gazda, M.; Zasada, F. TiO2/SrTiO3 and SrTiO3 microspheres decorated with Rh, Ru or Pt nanoparticles: Highly UV–vis responsible photoactivity and mechanism. J. Catal. 2017, 350, 159–173. [Google Scholar] [CrossRef]
  188. Mukherjee, S.; Mukherjee, A.; Das, S.; Das, P.; Banerjee, A. Structure and electronic effects from Mn and Nb co-doping for low band gap BaTiO3 ferroelectrics. J. Phys. Chem. C 2021, 125, 14910–14923. [Google Scholar] [CrossRef]
  189. Wang, J.; Li, C.; Wu, W.; Sun, X.; Xu, L.; Liu, Y. Ferroelectric BaTiO3@ZnO core–shell heterojunction triboelectric nanogenerators for electrochemical degradation of MO. Ceram. Int. 2024, 50, 4841–4850. [Google Scholar] [CrossRef]
  190. Voora, V.M.; Venkateswara Rao, A.; Ramachandran, R. Electrical properties of ZnO–BaTiO3–ZnO heterostructures with asymmetric interface charge distribution. Appl. Phys. Lett. 2009, 95, 082902. [Google Scholar] [CrossRef]
  191. Desmond, M.; Mavrogiannis, N.; Gagnon, Z. Maxwell-Wagner Polarization and Frequency-Dependent Injection at Aqueous Electrical Interfaces. Phys. Rev. Lett. 2012, 109, 187602. [Google Scholar] [CrossRef] [PubMed]
  192. Kuriakose, M.; Longuemart, S.; Depriester, M.; Delenclos, S.; Sahraoui, A.H. Maxwell-Wagner-Sillars effects on the thermal-transport properties of polymer-dispersed liquid crystals. Phys. Rev. E 2014, 89, 022511. [Google Scholar] [CrossRef]
  193. Bauer, G.E.W.; Tang, P.; Iguchi, R.; Xiao, J.; Shen, K.; Zhong, Z.; Yu, T.; Rezende, S.M.; Heremans, J.P.; Uchida, K. Polarization transport in ferroelectrics. Phys. Rev. Appl. 2023, 20, 050501. [Google Scholar] [CrossRef]
  194. Cowley, R.A.; Gvasaliya, S.N.; Lushnikov, S.G.; Roessli, B.; Rotaru, G.M. Relaxing with relaxors: A review of relaxor ferroelectrics. Adv. Phys. 2011, 60, 229–327. [Google Scholar] [CrossRef]
  195. Sun, J.; Li, Y. Research on improving energy storage density and efficiency of dielectric ceramic ferroelectric materials based on BaTiO3 doping with multiple elements. J. Compos. Sci. 2023, 7, 233. [Google Scholar] [CrossRef]
  196. Wu, W.; Liu, X.; Qiang, Z.; Yang, J.; Liu, Y.; Huai, K.; Zhang, B.; Jin, S.; Xia, Y.; Fu, K.K.; et al. Inserting insulating barriers into conductive particle channels: A new paradigm for fabricating polymer composites with high dielectric permittivity and low dielectric loss. Compos. Sci. Technol. 2021, 216, 109070. [Google Scholar] [CrossRef]
  197. Hsu, H.-J.; Yang, K.H.; Hsiang, H.-I. Enhanced dielectric properties and thermal stability of BaTiO3 ceramics sintered in a reducing atmosphere through MnNb2O6 doping. Ceram. Int. 2025; in press. [Google Scholar] [CrossRef]
  198. Jiang, X.; Chen, W. Sol-gel method preparation and first-principles calculations of CaCu3Ti4O12. J. Phys. Conf. Ser. 2025, 2941, 012026. [Google Scholar] [CrossRef]
  199. Wang, E.; Yang, H.; Guo, H.; Li, H.; Zhang, H.; Li, J.; Gu, M.; Yang, T.; Zhang, Y. Enhancement of Energy Storage Performance in NaNbO3-Modified BNT-ST Ceramics. Coatings 2025, 15, 504. [Google Scholar] [CrossRef]
  200. Riaz, M.; Ali, B.; Ali, S.M.; Khan, M.I.; Sahar, M.S.U.; Shahid, M.; Alam, M. Stress-induced transformation on the cubic perovskite RbTaO3 for high-temperature applications: A DFT approach. J. Comput. Electron. 2024, 23, 483–497. [Google Scholar] [CrossRef]
  201. Hickman, J.; Mishin, Y. Thermal conductivity and its relation to atomic structure for symmetrical tilt grain boundaries in silicon. Phys. Rev. Mater. 2020, 4, 033405. [Google Scholar] [CrossRef]
  202. Mahesh, M.L.V.; Bhanu Prasad, V.V.; James, A.R. Effect of sintering temperature on the microstructure and electrical properties of zirconium doped barium titanate ceramics. J. Mater. Sci. Mater. Electron. 2013, 24, 4684–4692. [Google Scholar] [CrossRef]
  203. Tuichai, W.; Danwittayakul, S.; Manyam, J.; Chanlek, N.; Takesada, M.; Thongbai, P. Giant dielectric properties of Ga3+–Nb5+ Co-doped TiO2 ceramics driven by the internal barrier layer capacitor effect. Materialia 2021, 18, 101175. [Google Scholar] [CrossRef]
  204. Elsehsah, K.A.A.A.; Noorden, Z.A.; Mat Saman, N. Current insights and future prospects of graphene aerogel-enhanced supercapacitors: A systematic review. Heliyon 2024, 10, 17. [Google Scholar] [CrossRef] [PubMed]
  205. Buscaglia, V.; Randall, C.A. Size and scaling effects in barium titanate. An overview. J. Eur. Ceram. Soc. 2020, 40, 3744–3758. [Google Scholar] [CrossRef]
  206. Jiang, B.; Iocozzia, J.; Zhao, L.; Zhang, H.; Harn, Y.W.; Chen, Y.; Lin, Z. Barium titanate at the nanoscale: Controlled synthesis and dielectric and ferroelectric properties. Chem. Soc. Rev. 2019, 48, 1194–1228. [Google Scholar] [CrossRef]
  207. Su, X.; Riggs, B.C.; Tomozawa, M.; Nelson, J.K.; Chrisey, D.B. Preparation of BaTiO3/low melting glass core–shell nanoparticles for energy storage capacitor applications. J. Mater. Chem. A 2014, 2, 18087–18096. [Google Scholar] [CrossRef]
  208. Pillai, V.V.; Ramasubramanian, B.; Sequerth, O.; Pilla, S.; Wang, T.; Mohanty, A.K.; Govindaraj, P.; Alhassan, S.M.; Salim, N.; Kingshott, P.; et al. Nanomaterial advanced smart coatings: Emerging trends shaping the future. Appl. Mater. Today 2025, 42, 102574. [Google Scholar] [CrossRef]
  209. Chamankar, N.; Khajavi, R.; Yousefi, A.A.; Rashidi, A.; Golestanifard, F. A flexible piezoelectric pressure sensor based on PVDF nanocomposite fibers doped with PZT particles for energy harvesting applications. Ceram. Int. 2020, 46, 19669–19681. [Google Scholar] [CrossRef]
  210. Luo, C.; Hu, S.; Xia, M.; Li, P.; Hu, J.; Li, G.; Jiang, H.; Zhang, W. A flexible lead-free BaTiO3/PDMS/C composite nanogenerator as a piezoelectric energy harvester. Energy Technol. 2018, 6, 922–927. [Google Scholar] [CrossRef]
  211. Li, C.; Huang, L.; Li, T.; Lü, W.; Qiu, X.; Huang, Z.; Liu, Z.; Zeng, S.; Guo, R.; Zhao, Y.; et al. Ultrathin BaTiO3-based ferroelectric tunnel junctions through interface engineering. Nano Lett. 2015, 15, 2568–2573. [Google Scholar] [CrossRef]
  212. Vélu, G.; Blary, K.; Burgnies, L.; Carru, J.C.; Delos, E.; Marteau, A.; Lippens, D. A 310°/3.6-dB K-band phaseshifter using paraelectric BST thin films. IEEE Microw. Wirel. Compon. Lett. 2006, 16, 87–89. [Google Scholar] [CrossRef]
  213. Vélu, G.; Blary, K.; Burgnies, L.; Marteau, A.; Houzet, G.; Lippens, D.; Carru, J.C. A 360° BST Phase Shifter With Moderate Bias Voltage at 30 GHz. IEEE Trans. Microw. Theory Tech. 2007, 55, 438–444. [Google Scholar] [CrossRef]
  214. Zidani, J.; Tajounte, L.; Benzaouak, A.; Touach, N.; Duong, A.; Zannen, M.; Lahmar, A. Advances in Lead-Free Flexible Piezoelectric Materials for Energy and Evolving Applications. Adv. Ind. Eng. Polym. Res. 2025; in press. [Google Scholar] [CrossRef]
  215. McLellan, B.C.; Corder, G.D.; Ali, S.H. Sustainability of rare earths—An overview of the state of knowledge. Minerals 2013, 3, 304–317. [Google Scholar] [CrossRef]
  216. Guo, X.; Yang, H.; Zhu, X.; Zhang, L. Preparation and properties of nano-SiC-based ceramic composites containing nano-TiN. Scr. Mater. 2013, 68, 281–284. [Google Scholar] [CrossRef]
Figure 1. Dielectric permittivity of GO-Ag-PVA nanocomposites as a function of (a) GO loading, (b) ionic liquid, and (c) temperature effect on 3GO-Ag-PVA (IL) [96].
Figure 1. Dielectric permittivity of GO-Ag-PVA nanocomposites as a function of (a) GO loading, (b) ionic liquid, and (c) temperature effect on 3GO-Ag-PVA (IL) [96].
Nanomaterials 15 01329 g001
Figure 2. Frequency and Temperature Dependence of Dielectric Permittivity in Bi0.75Ba0.25(FeMn)0.5O3 Ceramic [97].
Figure 2. Frequency and Temperature Dependence of Dielectric Permittivity in Bi0.75Ba0.25(FeMn)0.5O3 Ceramic [97].
Nanomaterials 15 01329 g002
Figure 3. Frequency dependence of (a) ε′ and (b) tan δ of (Sr0.75Ag0.25)(NiNb)0.5O3 ceramic for various temperatures [98].
Figure 3. Frequency dependence of (a) ε′ and (b) tan δ of (Sr0.75Ag0.25)(NiNb)0.5O3 ceramic for various temperatures [98].
Nanomaterials 15 01329 g003
Figure 4. (a) Dielectric constant as a function of frequency at different temperature for the sample x = 0.2. (b) ε′ as a function of frequency of the prepared samples at room temperature. (c) Frequency dependence of dielectric loss tg at various temperatures for Ba0.67Ni0.33Mn1−xFxO3 (x = 0.0 and 0.2) perovskite samples [99].
Figure 4. (a) Dielectric constant as a function of frequency at different temperature for the sample x = 0.2. (b) ε′ as a function of frequency of the prepared samples at room temperature. (c) Frequency dependence of dielectric loss tg at various temperatures for Ba0.67Ni0.33Mn1−xFxO3 (x = 0.0 and 0.2) perovskite samples [99].
Nanomaterials 15 01329 g004
Figure 5. Humidity-induced transition from ferroelastic to ferroelectric domain switching in PMN-PT single crystals [105].
Figure 5. Humidity-induced transition from ferroelastic to ferroelectric domain switching in PMN-PT single crystals [105].
Nanomaterials 15 01329 g005
Figure 6. The permittivity (a) and dielectric losses (b) of samples at room temperature [118].
Figure 6. The permittivity (a) and dielectric losses (b) of samples at room temperature [118].
Nanomaterials 15 01329 g006
Figure 7. The schematic of the preparation of ZnO-based composites with PTFE and metal oxides using CSP [118].
Figure 7. The schematic of the preparation of ZnO-based composites with PTFE and metal oxides using CSP [118].
Nanomaterials 15 01329 g007
Figure 8. Schematic diagram of (a) recoverable energy density (blue region, Wrec) and hysteresis loss (pink region, Wloss) from the P–E hysteresis loop of a dielectric material. (b) Domain evolution and formation of FE to RFE transition with the substitution of BST into BNKT (indicated by the green arrow), where the red arrows represent domain orientations. This incorporation results in enhanced Wrec and efficiency (η) [129].
Figure 8. Schematic diagram of (a) recoverable energy density (blue region, Wrec) and hysteresis loss (pink region, Wloss) from the P–E hysteresis loop of a dielectric material. (b) Domain evolution and formation of FE to RFE transition with the substitution of BST into BNKT (indicated by the green arrow), where the red arrows represent domain orientations. This incorporation results in enhanced Wrec and efficiency (η) [129].
Nanomaterials 15 01329 g008
Figure 9. Schematic illustration of the fabrication process of CeNPs@CNF-CF modified pencil graphite electrodes (CeNPs@CNF-CF/PGRs) [148].
Figure 9. Schematic illustration of the fabrication process of CeNPs@CNF-CF modified pencil graphite electrodes (CeNPs@CNF-CF/PGRs) [148].
Nanomaterials 15 01329 g009
Figure 10. Schematic representation of (a) spark plasma sintering; (b) comparison between SPS and conventional sintering; and (c) D.C. pulse current between particles (adapted) [151].
Figure 10. Schematic representation of (a) spark plasma sintering; (b) comparison between SPS and conventional sintering; and (c) D.C. pulse current between particles (adapted) [151].
Nanomaterials 15 01329 g010
Figure 11. Schematic representation of the cold sintering process (a) and the thermocompression apparatus (b) [154].
Figure 11. Schematic representation of the cold sintering process (a) and the thermocompression apparatus (b) [154].
Nanomaterials 15 01329 g011
Figure 12. Systematic steps involved in the synthesis of the Ag-CuO nanohybrids [157].
Figure 12. Systematic steps involved in the synthesis of the Ag-CuO nanohybrids [157].
Nanomaterials 15 01329 g012
Figure 13. Emerging applications of ceramic nanocomposites in advanced electronics, energy devices, sensors, and biomedical platforms.
Figure 13. Emerging applications of ceramic nanocomposites in advanced electronics, energy devices, sensors, and biomedical platforms.
Nanomaterials 15 01329 g013
Figure 14. Roadmap of Challenges and Strategic Solutions for Future Ceramic Nanocomposite Applications.
Figure 14. Roadmap of Challenges and Strategic Solutions for Future Ceramic Nanocomposite Applications.
Nanomaterials 15 01329 g014
Table 1. Summary of polarization mechanisms in dielectric nanocomposites, including their physical origin, active frequency range, and typical influence on dielectric behavior [88,89,90,91,92,93,94,95].
Table 1. Summary of polarization mechanisms in dielectric nanocomposites, including their physical origin, active frequency range, and typical influence on dielectric behavior [88,89,90,91,92,93,94,95].
MechanismOriginFrequency RangeTypical Role
ElectronicDisplacement of electron clouds in atomsOptical (1015 Hz)Minor, fast response in high-frequency fields
IonicRelative motion of positive and negative ionsInfrared (1013 Hz)Moderate contributor to polar ceramics
Dipolar (Orientational)Rotation of permanent dipoles (e.g., in polymers)Microwave (109–1011 Hz)Strong effect on polymer-based composites
Interfacial (MWS)Charge accumulation at interfaces in heterogeneous materialsLow frequency (<106 Hz)Dominant in nanocomposites with mixed phases
Table 2. Interfacial polarization in selected ceramic nanocomposites and its effect on performance [109,110,111,112,113].
Table 2. Interfacial polarization in selected ceramic nanocomposites and its effect on performance [109,110,111,112,113].
Composite SystemFiller TypeInterfacial Polarization EffectPerformance Impact
BaTiO3–CNT1D conductiveHigh MWS polarization; may increase leakage at high loading↑ εr ↓ breakdown strength if CNTs percolate
PVDF–BaTiO3Ceramic–polymerStrong dipolar + interfacial effects↑ flexibility and dielectric constant
BaTiO3–Graphene Oxide2D conductiveEnhanced interface area; tailored permittivity↑ εr with controlled loss tangent
KNN–SiO2Ceramic boundarySpace charge modification via SiO2 additives↑ breakdown strength; ↓ dielectric loss
Note: ↑ indicates an increase; ↓ indicates a decrease.
Table 3. Effect of Sintering Techniques on Ceramic Microstructure [119,120,121,122,123,124,125].
Table 3. Effect of Sintering Techniques on Ceramic Microstructure [119,120,121,122,123,124,125].
Sintering TechniqueTemp (°C)Grain ControlKey BenefitsChallenges
Conventional Sintering1200–1400Coarse grain growthSimple, widely usedHigh porosity, energy intensive
Spark Plasma Sintering800–1100Excellent grain retentionFast densification, fine grainsExpensive equipment
Cold Sintering<300Nanograin preservationEco-friendly, low cost, fastLimited to specific material systems
Sol–gel + Annealing600–900Moderate grain controlMolecular-level mixing and uniform compositionAgglomeration if not well controlled
Table 4. Comparative summary of processing techniques for selected dielectric and ferroelectric nanocomposites, including typical compositions, processing temperatures, grain sizes, dielectric constants, and key advantages [134,135,136,137,138,139,140,141,142,143,144,145].
Table 4. Comparative summary of processing techniques for selected dielectric and ferroelectric nanocomposites, including typical compositions, processing temperatures, grain sizes, dielectric constants, and key advantages [134,135,136,137,138,139,140,141,142,143,144,145].
MethodTypical Material ExamplesTemp (°C)Grain Size (nm)Dielectric Constant (εr)Key Advantages
Solid-State SinteringBaTiO3, K0.5Na0.5NbO3 (KNN), Pb(Zr,Ti)O3 (PZT)1300–1500~4000~1200Simple, scalable, suitable for bulk fabrication
Sol–GelBaTiO3, BiFeO3, PbTiO3600–900~50~1800Fine grains, low porosity, precise stoichiometry control
Spark Plasma SinteringBaTiO3–CNT, Ba(Zr,Ti)O3 (BZT), CaCu3Ti4O12 (CCTO)900–1100~300~2500High density, rapid sintering, nanoscale grain retention
Cold SinteringZnO–PTFE, BaTiO3–PVDF<300~200~1400Low-energy, compatible with polymers, eco-friendly
HydrothermalBaTiO3, SrTiO3, TiO2 nanostructures100–250~70~1600Morphology control, selective crystallization, high purity
Table 5. Comparative Summary of Core–Shell and Heterojunction Interfaces in Ceramic Nanocomposites.
Table 5. Comparative Summary of Core–Shell and Heterojunction Interfaces in Ceramic Nanocomposites.
AspectCore–Shell StructuresHeterojunction Interfaces
Interface TypeConformal, typical uniform around coreDiscontinuous or planar, between distinct phases
Polarization MechanismEnhanced interfacial and ferroelectric polarizationDominantly Maxwell–Wagner interfacial polarization
Permittivity BehaviorHigh dielectric constant with improved breakdown strengthStrong low-frequency dielectric dispersion
Thermal StabilityHigh, due to encapsulation and grain growth suppressionModerate; depends on phase interaction
Energy StorageHigh energy density in dense, well-aligned systemsModerate; often limited by interface defects
ExamplesBaTiO3@SiO2, BiFeO3@TiO2, ZnO@Al2O3BaTiO3–PVDF, TiO2–SnO2, BaTiO3–Nb:SrTiO3
ApplicationsCapacitors, energy harvesters, high-temperature dielectricsFlexible electronics, tunable dielectrics, EMI shielding
Table 6. Key challenges in ceramic nanocomposites and corresponding existing or proposed solutions, based on current literature and emerging research directions.
Table 6. Key challenges in ceramic nanocomposites and corresponding existing or proposed solutions, based on current literature and emerging research directions.
Key ChallengeProposed Solution
Scalability of processing techniquesAdoption of low-temperature scalable sintering (e.g., SPS, cold sintering)
Interfacial engineering limitationsAdvanced core–shell and surface functionalization strategies
Long-term stability and reliabilityImproved encapsulation and barrier coatings
Eco-toxicity of additives and solventsDevelopment of green synthesis and biodegradable matrices
Integration with flexible substratesEngineering polymer–ceramic hybrids with elastic interfaces
Limited AI-guided material discoveryUse of machine learning for structure–property predictions
Cost and energy consumptionOptimization of processing-energy tradeoffs using LCA tools
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

Ahmed Althumairi, N.; Hjiri, M.; Aldukhayel, A.M.; Jbeli, A.; Nassar, K.I. Recent Advances in Dielectric and Ferroelectric Behavior of Ceramic Nanocomposites: Structure Property Relationships and Processing Strategies. Nanomaterials 2025, 15, 1329. https://doi.org/10.3390/nano15171329

AMA Style

Ahmed Althumairi N, Hjiri M, Aldukhayel AM, Jbeli A, Nassar KI. Recent Advances in Dielectric and Ferroelectric Behavior of Ceramic Nanocomposites: Structure Property Relationships and Processing Strategies. Nanomaterials. 2025; 15(17):1329. https://doi.org/10.3390/nano15171329

Chicago/Turabian Style

Ahmed Althumairi, Nouf, Mokhtar Hjiri, Abdullah M. Aldukhayel, Anouar Jbeli, and Kais Iben Nassar. 2025. "Recent Advances in Dielectric and Ferroelectric Behavior of Ceramic Nanocomposites: Structure Property Relationships and Processing Strategies" Nanomaterials 15, no. 17: 1329. https://doi.org/10.3390/nano15171329

APA Style

Ahmed Althumairi, N., Hjiri, M., Aldukhayel, A. M., Jbeli, A., & Nassar, K. I. (2025). Recent Advances in Dielectric and Ferroelectric Behavior of Ceramic Nanocomposites: Structure Property Relationships and Processing Strategies. Nanomaterials, 15(17), 1329. https://doi.org/10.3390/nano15171329

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop