Next Article in Journal
Synthesis, Characterization, Magnetic Properties, and Applications of Carbon Dots as Diamagnetic Chemical Exchange Saturation Transfer Magnetic Resonance Imaging Contrast Agents: A Review
Previous Article in Journal
Precisely Targeted Nanoparticles for CRISPR-Cas9 Delivery in Clinical Applications
Previous Article in Special Issue
Research Progress on High-Temperature-Resistant Electromagnetic Wave Absorbers Based on Ceramic Materials: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

New Electromagnetic Interference Shielding Materials: Biochars, Scaffolds, Rare Earth, and Ferrite-Based Materials

1
Vinča Institute of Nuclear Sciences, National Institute of the Republic of Serbia, University of Belgrade, P.O. Box 522, 11001 Belgrade, Serbia
2
University of Lille, CNRS, Centrale Lille, University Polytechnique Hauts-de-France, UMR 8520-IEMN-Institut d’Electronique de Microélectronique et de Nanotechnologie–Lille, 59650 Villeneuve-d’Ascq, France
3
Department of Computer Science, Division of Microrobotics and Control Engineering, University of Oldenburg, Ammerländer Heerstraße 114-118, 26129 Oldenburg, Germany
4
Faculty of Polymer Technology, Ozare 19, 2380 Slovenj Gradec, Slovenia
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(7), 541; https://doi.org/10.3390/nano15070541
Submission received: 7 March 2025 / Revised: 30 March 2025 / Accepted: 31 March 2025 / Published: 2 April 2025

Abstract

:
In this review, a comprehensive systematic study of the research background, developments, classification, trends, and advances over the past few years in research on new electromagnetic interference (EMI) shielding materials will be described. The following groups of new materials for EMI shielding will be discussed: biochars, scaffolds, rare earth, and ferrite-based materials. We selected two novel, organic, lightweight materials (biochars and scaffolds) and compared their shielding effectiveness to inorganic materials (ferrite and rare earth materials). This article will broadly discuss the EMI shielding performance, the basic principles of EMI shielding, the preparation methods of selected materials, and their application prospects. Biochars are promising, eco-friendly, sustainable, and renewable materials that can be potentially used as a filter in polymer composites for EMI shielding, along with scaffolds. Scaffolds are new-generation, easy-to-manufacture materials with excellent EMI shielding performance. Rare earth (RE) plays an important role in developing high-performance electromagnetic wave absorption materials due to the unique electronic shell configurations and higher ionic radii of RE elements. Ferrite-based materials are often combined with other components to achieve enhanced EMI shielding, mechanical strength, and electrical and thermal conductivity. Finally, the current challenges and future outlook of new EMI shielding materials will be highlighted in the hope of obtaining guidelines for their future development and application.

Graphical Abstract

1. Introduction

The human population in urban and rural regions is exposed to non-ionizing electromagnetic fields (EMFs) due to the rapid development of telecommunication technology and digital systems for information transfer [1,2]. The association of the fifth generation (5G) and upcoming sixth generation (6G) will lead to the emergence of innovative applications (Internet of Things and self-driving vehicles) [3,4]. Wearable and portable electronics will soon become more commonly used [4]. The COVID-19 pandemic triggered the implementation of technology trends such as digital payments, telemedicine, and robotics, which use radiofrequency (RF) electromagnetic waves (EMWs) in the 100 kHz–300 GHz frequency range [5]. These technological advances caused an increase in artificial sources of EMF and resulted in the chronic exposure of people and the environment, creating electromagnetic pollution. Electromagnetic pollution is the continuous and uncontrolled exposure to electromagnetic fields from any emitting source of the EMF spectrum [5]. From 1950 to 2010, the levels of exposure to RF-EMF have increased from extremely low natural levels by about 1018 times [6]. Many studies have shown the harmful effects of EMF on human health [7,8,9,10,11] and the environment [12,13,14,15]. In addition to the potential health risks, the interference of EMW worsens heat accumulation, significantly shortening equipment lifespan [16]. Thus, new electromagnetic interference (EMI) shielding materials are urgently needed.
The conventional EMI shielding materials are metals, carbons, ceramics, cement, conductive polymers, and associated composites. Metals and carbons are the most functional materials due to their high conductivity, while ceramics, cement, and (non-conductive) polymers are less efficient unless combined with functional components/materials [17,18]. Shielding barriers should meet the required dimensions and shapes at the lowest possible cost for various applications.
Among the various new materials, organic materials such as graphene, carbon black, and carbon nanotubes are highly conducting, lightweight, flexible, and highly efficient EMI shielding materials. Still, their high price and the complexity of production and purification lead to high cost and limit their use. Inorganic materials, such as metal powders and metal oxides, are highly efficient EMI shielding materials due to their high electrical conductivity accompanied by magnetic permeability. Challenges in their application are associated with their high prices, chemical instability, corrosion, and heavy weight. Thus, in this review paper, we are focusing on new, suitable, low-cost organic materials—biochars and scaffolds. These organic EMI shielding materials are compared to novel inorganic shielding materials—rare earth and ferrite-based composites. By reviewing their production methods, yields, physico-chemical properties, and EMI shielding effectiveness, the future direction in the development of EMI shielding materials is estimated.
According to the reviewed literature, biochars are promising, eco-friendly, sustainable, and renewable materials that can potentially be used as filters in polymer composites for EMI shielding applications in electronic devices and construction materials. Previous research showed that scaffolds are new-generation, easy-to-manufacture materials with excellent EMI shielding performance.
Recently, great attention has been paid to traditional magnetic materials/alloys with excellent absorption of EMWs, but they are often limited by thickness and range of use. However, according to Snoek’s limit, where the direct current (DC) susceptibility and the cut-off frequency are constant, the permeability value decreases in the high-frequency range, which leads to an apparent decrease in magnetic loss capability [19,20,21]. Improving the permeability and magnetic loss of magnetic materials/alloys at high frequencies is crucial for obtaining good electromagnetic/microwave absorption properties [22]. It was shown that the electronic structure and well-chosen methods for designing magnetic materials/alloys can improve magnetism at high frequencies [23].
Therefore, magnetic metal oxide composites or perovskite oxides doped with rare earth element ions have been in development for the last few years due to their specific structures, interfacial polarization, multiple reflections, and excellent conducting and EMI shielding properties [24].
Rare earth materials are often characterized by various names, such as rare earth elements (REEs), rare earth metals (REMs), rare earth oxides, or yttrium-based rare earth material. This group represents a set of 17 chemical elements, as listed in the periodic table, from lanthanum (La) to lutetium (Lu), including 15 lanthanides and scandium (Sc) and yttrium (Y) with a tendency to occur in the same core deposits as lanthanides and exhibiting identical chemical properties [25]. In the crystal structure, REEs are most often found in the trivalent state (RE3+), although the divalent state (RE2+) is also possible. REEs with strong inter-electron interactions with their localized 3d and 4f electrons as well as strong coupling with magnetic materials have emerged as outstanding material dopants [26]. REEs based on neodymium (Nd), dysprosium (Dy), and samarium (Sm), due to their relatively large number of unpaired electrons in the atomic structure, possess a high remanent magnetization and coercivity value determining the stability of the remanent state. The orbital electron structure of these elements contains many unpaired electrons, which allows them to store large amounts of magnetic energy. The total magnetic moment originates from these unpaired electrons [27,28]. The reviewed literature revealed the growing interest in rare-earth-based materials as materials for EMI shielding applications [29,30]. These materials can be classified into several groups: RE-doped ferrites (RE-Fs), F2O3/Fe3O4, RE–transition metal intermetallics (RETMIs), RE oxides (REOs), RE-based alloys, RE spinels, RE-doped MoS2, and RE polymers. The above-mentioned groups of rare-earth-based materials will be discussed in detail throughout this manuscript. The main advantages of rare-earth-based materials over traditional EMI shielding materials are the magneto-dielectric effects and tunable dielectric properties in an electric or magnetic field. These properties make them attractive for exploring their EMI shielding applications.
New nanocomposite-based foams and heterogeneous layered structures have shown outstanding EMI shielding properties. Namely, thermoplastic polyether-block-amide elastomer beads coated with Ti3C2Tx showed an EMI shielding efficiency of 44 dB at 8.5–12.5 GHz [29], microcellular aramid nanofiber/Ti3C2Tx MXene foams reached a value of 64.9 dB [30], and foams based on the conductive polymer polypyrrole with Fe3O4 achieved a value of 41.1 dB [31]. Combining MXenes with carbon nanotubes and silver nanowires in a layer-by-layer architecture, Zhang et al. achieved an EMI shielding efficiency of 53.1 dB in the X-band [32].
The development, classification, trends, and advancements of research on novel shielding materials in electromagnetic irradiation during the past few years will be thoroughly and methodically reviewed herein. The recent progress in the development of new-generation shielding materials based on biochars, 3D network structures (referred to as scaffolds in the following), materials with rare earth (referred to as rare-earth-based materials in the following), and ferrite materials will be described (Figure 1). This paper will comprehensively discuss the EMI shielding performance, basic principles and mechanisms, preparation methods, structure, morphology, and application prospects of these materials.

2. New EMI Shielding Materials: Biochars, Scaffolds, Rare Earth, and Ferrite-Based Materials

2.1. Biochar as EMI Shielding Material

Among conventional EMI shielding materials, metals such as Cu, Ag, Al, etc., are highly efficient EMI shielding materials due to their electrical conductivity and ability to reflect incident EMWs [16]. Regardless, the use of these metals is limited owing to their high density, inconvenient processing, and poor resistance to corrosion [16]. Recently, shielding materials have been investigated in the form of composites with plastic or metals as substrates [34]. As alternative solutions, conductive polymer-based composites, porous conductive foams, and polymers with micro- and nanoscale filters such as carbon material (graphite, graphene, reduced graphene oxide, carbon nanotubes), MXenes, metal nanoparticles, and nanowires have been explored due to their low mass densities, good flexibility, and stretchability [4,35].
One of the latest approaches in developing new, sustainable, and eco-friendly solutions for EMI shielding is the integration of natural, renewable carbon materials—biochars—into polymers [4]. Biochars originate from various biomass sources, such as bamboo, sugarcane, and cork. To create biochars from biomass, biomass resources are converted to carbon materials by pyrolytic processes in an oxygen-limited environment. A schematic presentation of biochar starting materials and production steps is displayed in Figure 2. After high-temperature carbonization, biochars show extraordinary hardness, excellent thermal stability, and high electrical conductivity, making them suitable fillers for polymers [36]. By combining biochars and polymers, the mechanical, physical, and electrical properties of both materials can be upgraded successfully.
The electrical conductivity of biochar is closely related to the source properties and carbonization conditions [37]. Nevertheless, some researchers have disregarded the electrical properties of biochars, so their potential application in EMI shielding has rarely been examined [38,39,40].
In 2018, Li et al. conducted an initial research study [39]. Namely, they used commercially available bamboo charcoal as the biochar filler and/or an ultra-high-molecular-weight polyethylene (UHMWPE)/linear low-density polyethylene (LLDPE) blend as a potential EMI shielding composite. The composite with 80 wt.% of biochar was fabricated using the mass-producing extrusion and hot compression methods. Through the carbonization of bamboo charcoal at 1100 °C, a graphitic-like structure with good electrical conductivity and a high specific surface area was obtained. Namely, an electrical conductivity of 107.6 S/m and EMI shielding effectivity (SE) of 48.7 dB were measured at 1.5 GHz. The specific EMI SE of the composite was 39.0 dB cm3/g. This value is four times higher than the specific EMI SE of copper (10.0 dB cm3/g) [41].
Tolvanen et al. applied the same approach to prepare a biodegradable two- and three-phase composite of graphite, biochar derived from pine chips, and polylactic acid (PLA) [4]. The EMI shielding performances were tested in the K-band (18–26.5 GHz). The composite showed an outstanding EMI SE of >32 dB when the materials were prepared in the form of a film with a thickness of 25 μm and a high specific shielding effectiveness (SSE/t) of >890 dB cm2/g. It can be seen that these composites are convenient for the application of wearable/portable and stationary devices.
Shortly after, Akgül et al. presented a biochar–iron (BC-Fe) material obtained by the pyrolysis of industrial tea waste biomass and the encapsulation of BC-Fe into polymethyl methacrylate (PMMA) polymer [42]. During synthesis, Fe(III) ions promoted the graphitization of the amorphous carbon in biomass and contributed to biochar surface stabilization. The results of EMI SE measurements showed that the reflection component of the EMI SE of pure PMMA was near −10 dB at frequencies of 7.8 and 10.1 GHz, while it was reduced by ~60% when the content of BC-Fe in the polymer was 40 wt.% in the frequency range of 8.5–12.9 GHz.
Liang et al. successfully developed a mortar/brick structure using wood-derived porous carbon (WPC) as a skeleton and conductive 3D MXene aerogel [1]. The MXene aerogel/WPC composite was composed of highly ordered honeycomb cells inside WPC as a microreactor. Herein, a high graphitization level of natural wood was achieved by applying a high carbonization temperature of 1500 °C. The MXene aerogel/WPC composite showed an excellent EMI SE value of 71.3 dB in a frequency range of between 8.5 GHz and 12.5 GHz, while the sample density was only 0.197 g/cm3. In the study, a wall-like mortar brick structure (WPC as mortar and MXene aerogel as brick) solved the instability of the MXene aerogel network as well as prolonged the transmission path of EM waves, dissipating the incident EM waves in the form of heat and electrical energy. The material’s superior EMI shielding performance was achieved thanks to a specific design.
The shielding efficiency of a gypsum–biochar drywall-like composite was investigated by Natalio et al. [43]. They used wooden chips and eucalyptus biochar and combined them with gypsum. An enhancement in the shielding efficiency was recorded with an increased biochar content in the composites. Namely, the EMI SE values of the drywall-like plates with 10%, 20%, and 40% w/w biochar contents were 11.65 ± 1.6, 19.2 ± 5.7, and 19.25 ± 1.8 dB at a frequency of 6 GHz. This investigation contributes to expanding innovative bio-based sustainable materials with EM shielding properties in the microwave region.
Savi et al. used sewage sludge to obtain biochar and investigated its EMI-shielding features [44,45]. The biochar was mixed with epoxy polymer (20 wt%) and cast into a film 4 mm thick. The composite showed a promising EMI SE of −10 dB [44]. Savi et al. reported an outstanding electrical conductivity of 300 S/m [44]. The same group compared the EMI SE properties of biochar produced from carbonized sludge with graphene nanoplatelets [46]. Herein, the authors produced polymer composites with polyvinylidene fluoride (PVDF), while fillers (biochar and graphene) were added at 90 wt%. The biochar–PVDF composite showed transmission scattering of less than −30 dB.
Savi et al. also studied commercial biochar produced from wood biomass to coat several layers of common building components, such as drywall panels [47]. They recorded EMI SE values of 17 dB at 1 GHz and 25 dB at 18 GHz. Drywall panels coated with several layers of biochar are easy to fabricate and a low-cost solution to realize a protected surrounding for healthcare applications (chemotherapy and tomography) to minimize the intensity of the EM field close to electronic equipment. The same research group produced a cement-based composite with commercial lignin-based biochar and polyvinyl chloride (PVC) [48]. They introduced biochar powder into cement paste to improve its shielding properties. PVC, which was used as a filler, was obtained by decommissioning old electrical cables that would have ended up in a landfill. A combination of 10 wt.% biochar and 6 wt.% PVC revealed the best shielding capacity, around 16 dB in the 5.4–8 GHz frequency range. This investigation shows great importance in the EMI shielding application and from the aspect of the circular economy.
Miao et al. fabricated a conductive EBC@CNF@MWCNT composite aerogel by the freeze-drying process, mixing ferrite chloride with electrically conductive bamboo charcoal (EBC), with cellulose nanofibrils (CNFs) as a skeleton and multi-walled carbon nanotubes (MWCNTs) as a conductive enhancer in the freeze-drying process [36]. Afterward, the aerogel was soaked with polydimethylsiloxane and hot-pressed into the membrane. In the composite, EBC and MWCNTs were arranged uniformly in the CNF skeleton to form a 3D conducting network. The composite displayed exceptional electrical conductivity (47 S/m) and a high EMI shielding effectiveness of 39.5 dB, with an adsorption loss of ~75%.
Recently, Nikolopoulus et al. used biochar prepared from olive tree pruning to fabricate composite samples with carbon black and polytetrafluoroethylene as binders [49]. The EMI shielding capacity was measured in the 1–3 GHz frequency range. The results demonstrated that the raw pure biochar had a low EMI SE, between 1.5 and 4 dB, which was enhanced with an increase in the thickness from 0.1 to 0.5 mm. On the contrary, the composite significantly improved the EMI SE, reaching a value of 39 dB. This work indicates that biochar could be used as a basis for developing composites with high EMI SE values.
Milenkovic et al. recently reported that agricultural biowaste collected after apple and quince processing could be converted into EMI shielding material [50]. The starting materials were chopped, homogenized, dried, and mixed with an equal mass of KOH. The mixtures were carbonized at 850 °C, under nitrogen flow. The resulting biochars were mixed with sodium silicate at 40 wt.%. The film of only 0.2 mm thickness showed an EMI SE of 15.5 dB in the 8–12 GHz frequency range. Although moderate EMI SE values were found, the study demonstrates the outstanding potential of selected biowaste in the production of sustainable EMI shielding materials.
Another study by Perumal et al. proposed a new organic waste for biochar production [37]. Within the study, the composite built from Ricinus communis outer shell-based biochar and epoxy was fabricated by slow pyrolysis at 400–700 °C. The results showed that the biochar pyrolysis at 700 °C led to a maximal electrical conductivity of 95 S/m due to the presence of graphitic carbon. A maximum EMI SE of 26.5 dB was found in the X-band frequency range, at 40 wt.% biochar to epoxy matrix.
According to the reviewed literature, biochar is a promising, eco-friendly, sustainable, and renewable material that could be used as a filter for EMI shielding applications in electronic devices and construction materials. The biochar shielding effectiveness is associated with the electrically conductive graphene region in biochars and porous structure [50]. Namely, EMW attenuation is the result of wave reflection from the graphene region and multiple reflections inside the pores. Biochar features rely on two main factors:
(1)
The experimental conditions (temperature, duration of heating, addition of chemical agent for pore formation or activitvation of graphitization, selected gas);
(2)
The starting material’s chemical composition [39].
With an increased temperature, the porosity, specific free space, and carbon content increase too, while the selection of source materials affects both the properties and yields of the resulting biochar [39].

2.2. Scaffolds as EMI Shielding Material

Many structural patterns, including highly porous materials and conductive structures such as multilayer composites obtained by stacking cellulose nanofibers/reduced graphene oxide (rGO), rGO films, and multistage composite foams, are recognized as excellent EMI shielding materials. These structures are obtained using vacuum-assisted filtration, freeze-drying techniques, or electroless deposition [51]. The mentioned procedures allow for a reduction in the reflection coefficient, but they are complex and often not reliable. The innovative manufacturing technologies imply 3D network structures (scaffolds) that can be decorated with metal nanoparticles [52], coated by polymers, encapsulated with paraffin [53], etc. These approaches enable design freedom, flexibility, precise control over the shape and size, and the connectivity of porous structures.
Cellulose scaffolds (CSs) are recognized as a promising material for EMI shielding applications due to the ability of cellulose nanofibers to improve the thermal and mechanical features of nanocomposites. Tran et al. represented a 3D network structure of cellulose scaffolds decorated with silver nanoparticles (AgNPs) [52]. The scaffolds were filtered at the epoxy matrix and cured at 40 °C to fix the nanoparticles. The obtained composite possesses a thermal conductivity of 2.52 W/m/K, which is over 11-fold the thermal conductivity of pure epoxy. The extremely high electrical conductivity of 53.691 S/m caused a remarkable SE value of 69.1 dB. He et al. designed a scaffold connected with metal nanoparticles [53]. In this work, the shape-stable composite was composed of a paraffin impregnated in a biological porous carbon scaffold, followed by a coating with polyurethane and Fe3O4 nanoparticles. The biological porous carbon was obtained from a loofah sponge by immersion in a phenolic resin solution followed by carbonization. Scanning electron microscopy (SEM) results revealed that the system kept its original shape with a 3D and honeycomb-like porous structure of single fibers after carbonization. The polyurethane coating provided adequate mechanical support and effective leak-proof performance. The produced framework shows strong EMI shielding performance (up to 32 dB). Another study based on the encapsulation of Fe3O4 nanoparticles in carbon scaffolds was conducted by Wei et al. [54]. Herein, laminar cellulose-paper-based scaffolds with bidirectional gradient distributed Fe3O4 nanoparticles were constructed via immersion, drying, and carbonization processes. The resulting carbon scaffolds exhibited high in-plane electrical conductivity (96.3 S/m) and high shielding efficiency (1805.9 dB/cm2g).
A 3D cellulose scaffold was combined with a CNT/MXene composite [55]. Nanosheets of the composite were inserted into a cellulose scaffold by vacuum impregnation. Finally, a hydrophobic and multifunctional 3D system was produced by wrapping it with poly (dimethylsiloxane). The composite had a high compressive strength of 1.53 MPa, a maximum strain at fracture of 74.1%, and an outstanding SE (29.3 dB).
Hu et al. reported a dual-ice templating assembly strategy to prepare a dual-interpenetrated scaffold [56]. The scaffold was linked with a high-quality graphene array and porous MXene-Co aerogel. It showed an absorption-dominated EMI SE of 72.86 dB.
Three-dimensional-printing technology was applied for scaffold fabrication based on a triply periodic minimal surface and 70% porosity [42]. The monolithic layered dipping method was used to regulate the gradient distribution of carbon nanotubes on the 3D-printed scaffold surface and ensure integrity. This system blocked 99.9% of EM waves. The value of the SE in this case was 35.9 dB.
Previous research showed that scaffolds are new-generation, easy-to-manufacture materials with excellent EMI shielding performance. Due to their exceptional three-dimensional construction and inner network, they show anisotropic thermal conductivity and allow heat dissipation, making them efficient EMI shielding materials [52,56]. The results from the literature based on biochar and scaffolds in shielding applications are listed in Table 1.

2.3. Rare-Earth-Based Materials as EMI Shielding Material

Next-generation microwave-absorbing compounds and hybrid materials could potentially outperform traditional materials in terms of their ability to absorb microwaves, their persistence, their stability at high temperatures, weight reduction, their corrosion resistance, and the need to expand the frequency spectrum over which materials can efficiently absorb microwaves. Research and development in the last few years have promoted magnetic metal oxide composites or perovskite oxides doped with REEs as innovative materials for good electromagnetic wave absorbance due to their specific structures, interfacial polarization, and multiple reflections, showing excellent conductive properties and magnetic losses [60]. By doping rare earth elements, the dielectric loss and dipole polarization can be increased, and the wave absorption performance of ferrite can be further improved. Nikzad et al. [61] prepared a Nd3+-substituted Gd–Co ferrite composite and found this to be a promising EW absorption material with a wide effective absorption bandwidth in the X-band. On the other hand, it was observed that appropriate La doping enhances the dielectric and conductive mismatch levels of the material. This enhancement leads to a stronger interface polarization capability [62].
Different microstructures and morphologies manifest different EMI shielding properties. Addressing the challenge of creating absorption-dominant EMI shielding requires significant attention to detailed structural design [63]. This strategy aims to produce structures that minimize the reflection of undesirable electromagnetic noise, capturing electromagnetic waves (EMWs) through various attenuation mechanisms.
The rare earths (REs) for electromagnetic wave absorption applications can be classified into several groups: RE-doped ferrites (RE-Fs), RE–transition metal intermetallics (RETMIs), RE oxides (REOs), and other categories [64]. Graphene-based materials, metal–organic frameworks (MOFs), ferrites, molybdenum disulfide, MoS2, and other innovative materials can significantly enhance EMW absorption properties by doping or substitution with REs due to the magnetic moment originating from their unpaired electrons [65]. The high-frequency permittivity of rare-earth Er-doped MoS2 films was studied [66]. The size of nanocrystals depends on the reaction time; the reactant concentration and composition control the morphology [67]. With a lowered concentration, the resulting nanoparticles exhibited a spherical morphology but gradually changed to flower-shaped with increasing concentration.
Pure MoS2 has a relatively homogeneous loss mechanism, limited impedance matching, and a limited ability to absorb EW. Doping is an effective method to improve the electromagnetic properties of materials [68]. For instance, palladium-doped few-layer MoS2 nanosheets grown on the surface of multi-walled CNTs and the 10% Pd nanohybrid possess a higher EMI shielding effectiveness (SE ∼26.50 dB) than that of SE ∼21.55 dB in the undoped MoS2/CNT at the same thickness of 1 mm [69].
Laser-cladded novel FeCo-based alloys with different RROs, such as La2O3, Y2O3, Nd2O3, and Gd, possess enhanced electromagnetic shielding performance with a shielding effectiveness (SE) of up to ~96 dB at 23.9 GHz and an ability to absorb 99.9999% of EMWs [70,71].
Li et al. [72] discovered that REE substitution in Fe4N could modulate magnetic moments and magnetic exchange coupling interactions and the change in spin polarizability. Zhang et al. [73] studied the microwave absorption properties of high-entropy hexaborides containing rare earth, including Ce, Y, Sm, Er, and Yb, prepared by a boron carbide reduction reaction, which showed good absorption properties with an effective bandwidth of 4.3 GHz, while Chen et al. [74] studied the absorption properties of high-entropy rare earth silicide carbides and high-entropy rare earth oxides based on different ratios of Tm, Y, Dy, Gd, Tb, Pr, and Tb ions, prepared by a solid-state reaction method, which showed good absorption properties with an effective bandwidth of 4.5 GHz. Absorbent, magnetic cores made of alloys based on Fe and Nb are suitable for EMI reduction signals with a 6–7 MHz frequency. However, Zamborszky et al. constructed a broadband coaxial probe head allowing the measurement of Fe-based nanocrystalline cores containing Nb up to 1 GHz using a reflectometry method with a vector network analyzer (VNA) [75].
The effects of the 4f-3d interaction on electronic locality, magnetism, and charge migration during hybridization between Er and Fe/Co were studied [22,76]. Due to enhanced EMI shielding properties, La, Ce, Pr, Gd texaphyrin, Tb, and Er were studied and selected as potential new shielding materials, as well as Co–Mg–La ferrite/graphene composites [77,78,79]. Gd3+-doped MnFe2O4 synthesized by the sol–gel method expressing very low dielectric loss, saturation, and minimum coercivity can be a potential material with applications in storage and microwave absorption devices [80]. Gd-doped Fe3O4 oxide (GdxFe3−xO4) samples with different percentages of Gd contents were prepared by the hydrothermal method [81].
The substitution of smaller Fe ions with RE ions leads to significant changes in physical properties such as the structural distortion of crystal units, the formation of a new phase, and changes in the morphological, optical, electrical, and dielectric properties due to the different ionic radii between ions.
The partial substitution of RE ions, such as with Sm, Ce, Er, Dy, La, and Nd for Fe3+, leads to a structural bend in the spinel structure which induces strain and considerably modifies the electrical and dielectric properties [82].
All ferrites exhibit a cubic spinel structure with the Fd-3m space group. The ionic radius of RE3+ ions in this environment (Gd3+ = 0.94 Å, Eu3+ = 0.95 Å, Sm3+ = 0.96 Å, Nd3+ = 0.96 Å, Pr3+ = 0.99 Å) is larger than that of Fe2+ ions (0.65 Å) [83]. The substitution of Fe2+ by RE3+ ions limits solubility in the lattice and inhibits grain growth. The shifting of the positions of diffraction peaks to smaller two-theta angles with an increase in the RE3+ concentration results from the increased lattice constant [84,85,86].
RE substitution has been reported to immensely affect cobalt ferrite’s structural, morphological, and physical properties [87,88]. It induces strain, causes the distortion of the crystal lattice, promotes the creation of vacancies, reduces crystallite sizes, increases grain–grain boundary interfaces, and promotes nanoscale magnetic effects such as spin canting and the core–shell effect. The insertion of RE in place of Fe promotes 3d–4f coupling, leading to variation in the magnetic, dielectric, and conduction mechanisms, and an enhancement in the SE can be obtained.
RE substitution can also decrease the grain size, which is an important factor in low-noise media. Cheng et al. obtained an increase in the polar Kerr rotation for Er- and Tm-doped cobalt ferrite films, while no significant changes in the magneto-optical response were observed for the Ho, Yb, and Lu ion substitutions [89].
For example, different Gd3+ doping concentrations in the ferrite structure led to noticeable variations in the magnetic saturation values. This can be explained by the finite size effects of nanoparticles, indirect interactions of Fe-Gd and Co-Gd and relatively weaker Fe-Fe interactions, or surface dipole disorder induced by lattice distortions upon dopant substitution with different ionic radii [90]. These interactions can significantly influence the electromagnetic properties of spinel ferrites [91]. A low-density polyethylene/MWCNT/graphene/LaFe2O3 composite and La3+ ions substituted into CoLaxFe2−xO4 and NiLaxFe2−xO4 resulted in enhanced microwave absorption at an X-band frequency range of 8–12 GHz [92,93,94].
The nanocomposite Gd-doped MoS2/reduced graphene oxide Gd-MoS2/(rGO) for high-performance EMI shielding applications was prepared using a hydrothermal method with a varying percentage of Gd doping. The MoS2/rGO nanocomposite without dopant showed a low total SE, ~16.18 dB, and the one with 20% Gd-doped MoS2/rGO nanocomposite showed a higher total SE, ~20.47 dB, in the frequency range of 8.0–12.0 GHz due to enhanced electrical conductivity, defect dipole polarization, and interfacial polarization [95]. A 5% Gd-doped MoS2 electrode compound synthesized using cost-effective one-step hydrothermal methods can find applications in high-performance energy storage systems [96]. The importance and application of MoS2-based microwave-absorbing materials are in current social needs, including military radar stealth and civil electronic communication [97].
Flower-shaped Gd-doped FeNi3 and a novel raspberry-like absorbent based on a biomimetic design with high electromagnetic wave (EMW) absorption performance for an ultra-wide bandwidth of 12.24 GHz were designed by thermal catalysis [98]. These materials were investigated in the 2–18 GHz frequency range, while EMW attenuation was assigned to dipole polarization, multiple wave scattering, and reflections. According to free electron theory, closely related to electrical conductivity, the hierarchical flower-like and NiO/rGO composites were designed using the hydrothermal method. The NiO/rGO composites with a thickness of 2 mm showed an improved reflection loss of −60 dB at 9.8 GHz in comparison with a reflection loss of −40 dB at 10.8 GHz for NiO [99]. On the other hand, Pr3+-doped Co0.5Zn0.5Bi0.4−xPr0.1Fe1.5+xO4 spinel ferrite series with different contents of Bi synthesized by the micro-emulsion method and Pr-doped (Bi0.5Na0.5)1−xPrxTiO3 ceramics obtained by using the conventional solid-state method have remarkable potential for application in EMI shielding in the X-band frequency range of 8.2–12.4 GHz [100,101].
Neodymium (Nd)-doped spinel ferrites induce an increase in the polarization intensity, improving their saturation magnetization, electrical resistivity, thermal stability, and ability to absorb microwave radiation, making them suitable for advanced electronics, telecommunications, and power systems [102,103,104,105]. Classes of Ce-doped W-type barium ferrites, Ba1−xCexNi2Fe15.4O27, with varying concentrations of Ce were successfully prepared by a sol–gel approach, whereby a Ce mole fraction of 0.3 gave the minimum reflection loss (RL) value of −52 dB, while a ferrite with Ce mole fraction of 0.4 had an RL value of −38 dB at 1.36 GHz frequency, suggesting the Ce enhancing EMW absorption of ferrites [106].
RE-metal-based MOFs have large specific surface areas. Their composites exhibit good dielectric loss, making them efficient EMW absorbers. The minimum RL value of the Ce(IV)-MOF is −32.12 dB at a frequency of 4.56 GHz at a thickness of 5.5 mm [107,108]. Excellent EMW absorption performance and mechanisms have been unveiled for the first time for Gd2O2S/rGO composites with an absorption capacity of −65 dB and an absorption bandwidth of 5.6 GHz. Gd2O2S nanosheets with a 1 nm thickness were produced via a facile hot injection method and mixed with rGO [109]. The EMW absorption performances (minimum reflection loss, RL, effective absorption bandwidth, EAB, thickness, and frequency range used) of various RE–ferrite composites, RE–transition metal intermetallics, REO composites, RE-MOFs, and other RE compounds are presented in [64].

2.3.1. Synthesis, Morphology, and EMI Shielding Efficiency of RE-Based Materials

Synthesis of GdxFe3−xO4, x = 0, 0.02, 0.04, 0.06, 0.08, and 0.1

A series of GdxFe3−xO4 materials with different Gd contents, x = 0, 0.02, 0.04, 0.06, 0.08, and 0.1, was obtained by using the hydrothermal method at a temperature of T = 180 °C and high vapor pressure. Gd(NO3)·6H2O, FeCl3·6H2O, and FeCl2·4H2O, as precursors of Gd3+, Fe3+, and Fe2+ ions, were dissolved in water separately and then mixed in the corresponding ratio, and the pH was adjusted up to the value of 12.8 using 10 M NaOH. The mixture was transferred to Teflon-lined stainless steel sealed containers and heated for 2 h at a temperature of 180 °C. After cooling, the obtained samples were washed several times with deionized water. For specific absorption rate (SAR) measurement, the obtained samples were functionalized by stirring overnight in a PEG solution [81].

Synthesis of CoFe2−xDyxO4 Nanoparticles

The first series of CoFe2−xDyxO4 nanoparticles with a stoichiometric ratio of Co(NO3)2·6H2O, Fe(NO3)3·9H2O, and Dy(NO3)3·H2O precursor solutions in distilled water was synthesized by the co-precipitation approach. Oleic acid, followed by a liquid ammonia solution, was added dropwise at pH values of between 10 and 12. The obtained precipitates were washed and sintered at 500 °C for 2 h to form a fine powder. The nanocomposites of polypyrrole and 10 wt% CoFe2−xDyxO4 were obtained via in situ chemical oxidative polymerization, and the synthesis is presented in detail in Figure 3 [110]. The obtained nanocomposites are promising candidates for applications in the 12.4 GHz to 18 GHz range, particularly in satellite communication systems, due to their impressive shielding performance, with an SE of approximately 16.3 dB at a shield thickness of 1.5 mm.

Synthesis of Gd- and Er-Doped α-MnO2 Nanorods

Gd- and Er-doped α-MnO2 with a homogenously distributed rod-shaped morphology, as can be seen in Figure 4, was prepared by a modified chemical-route-assisted hydrothermal method. Er(CH3COO)2, Gd(CH3COO)2, and Mn(CH3COO)2, as precursors of Er, Gd, and Mn ions, were dissolved in deionized water separately, according to the stoichiometry ratio of 7 wt.% to PVA; KMnO4 was used as a reaction catalyst. The PVA was dissolved in hot (T = 80 °C), double-distilled water with continuous stirring. When it was dissolved in water, it became transparent. When the PVA became a transparent solution and cooled down to room temperature, the solution of Er/Gd and Mn ions precursors was added to it very slowly with constant stirring, followed by the addition of KMnO4. The pH value was adjusted to ~10 by adding KOH solution drop by drop, and the mixture of solutions was left for aging. The brownish-black precipitate was centrifuged and washed several times with distilled water, put into the hydrothermal reactor, and heated at 160 °C for 8 h. The brownish-black precipitate was again centrifuged, washed, dried additionally at 80 °C for 18 h, mortared, and annealed at 450 °C for 2 h to acquire a fine powder of Gd- and Er-doped α-MnO2 samples with good crystallinity. The EMI SE (presented in Figure 5a–d, showing total absorption and reflection and skin depth variation in the frequency range of 8 to 18 GHz) achieved a maximum of −43 dB at 15.3 GHz and −46 dB at 15.3 GHz for thin-layer Gd- and Er-doped α-MnO2 at a of thickness ~600 μm, respectively. A schematic representation of the impact of hazardous EM waves radiated from various electronic sources on livelihood and a demonstration of the most possible EMI shielding mechanism to prevent it are presented in Figure 5e [111].
Rare earth elements exhibit high magneto-dielectric effects, and their dielectric properties can be tuned by applying an electric or magnetic field. This property makes them attractive for use in electromagnetic shielding applications where the magnetic field can be used to control the dielectric constant of the material. For example, the doping of MnO2 with Er and Gd can enhance its dielectric properties, which make it a promising material for high-performance electronic devices. The incorporation of Er ions into the MnO2 matrix can potentially modify the material’s electrical conductivity, magnetic properties, and structural stability, thereby influencing its effectiveness in attenuating electromagnetic waves [111].

2.4. Iron Oxide-Based Materials as EMI Shielding Material

Iron oxides are inorganic compounds of iron and oxygen that exhibit very different stoichiometries due to the difference in the Fe oxidation state. Briefly, they can be classified into three groups: oxides of Fe2+ (FeO, wüstite), oxides of Fe3+ (various forms of Fe2O3, hematite, maghemite), and combined Fe2+/Fe3+ oxides (Fe3O4, magnetite, Fe4O5, Fe5O6, etc.). Apart from the difference in the Fe oxidation state, many existing crystal structures also enhance the diversity of iron oxides depending on the synthetic routes. For example, a variety of polymorphs of iron (III) oxide has been reported to date [112]. α-Fe2O3 is the most common form of iron (III) oxide and occurs as the mineral hematite. It has a rhombohedral structure and shows antiferromagnetic behavior below ~260 K and weak ferromagnetism at temperatures of between 260 and 950 K [113]. Common methods to prepare α-Fe2O3 are thermal decomposition and precipitation in the liquid phase. The β-phase of Fe2O3 is metastable, and at temperatures above 773 K, it converts to the α-phase. It can be obtained by the thermal decomposition of iron (III) salts. Similar to the β-phase, γ-Fe2O3 is also metastable and converts to the α-phase at high temperatures [114]. It has a cubic structure, and in the bulk form, it is ferromagnetic, but γ-Fe2O3 nanoparticles less than 10 nm in size show superparamagnetic behavior. The methods of γ-Fe2O3 preparation involve the thermal dehydration of gamma iron (III) oxide-hydroxide, the modest oxidation of Fe3O4, or the thermal decomposition of suitable iron salts [115]. The ε-phase is rhombic with properties between those of the alpha and gamma phases. It is metastable with a tendency to transform to the alpha phase at temperatures of between 773 and 1000 K. The synthesis of pure ε-Fe2O3 is challenging and involves the oxidation of iron in an electric arc or sol–gel precipitation from iron (III) nitrate [116]. In addition, an amorphous form of Fe2O3 also exists, which can be prepared under high pressure.
Oxides of iron have become interesting candidates for electromagnetic interference (EMI) shielding due to several factors. Fe2O3 is a magnetic material, while Fe3O4 is a ferromagnetic. Magnetic materials can interact with and absorb electromagnetic waves, especially at lower frequencies (like radio waves), thus reducing EMI. Iron oxides exhibit a certain level of electrical conductivity, allowing them to contribute to EMI shielding by reflecting or absorbing the electrical components of electromagnetic waves. In addition, they can contribute to the absorption of electromagnetic energy through dielectric loss, a mechanism that involves converting electromagnetic energy into heat. This makes it effective in dissipating energy and preventing the transmission of EMI. Finally, iron oxides are abundant, inexpensive, and easy to combine in composite materials, such as polymers or carbon-based materials, and can maintain their properties over time.
For EMI shielding applications, iron oxides are often combined with other materials, such as various polymers, graphene oxide, carbon nanotubes, or similar, to achieve desirable EMI shielding, mechanical strength, and electrical and thermal conductivity. Using an in situ polymerization technique, Azadmanjiri et al. created iron oxide and polypyrrole nanocomposites and examined their EMI shielding properties in the 0.1–18 GHz frequency range [117]. The composites’ intimate contact between the conducting and magnetic phases increased absorption by 10.10 dB at the instrument’s highest frequency limit (17–18 GHz), while iron oxide nanoparticles only increased the absorption by 2.6 dB. As the possible cause of this improvement, they proposed a better match between dielectric loss, magnetic loss, and improved dispersion of the magnetic/conductive nanocomposites in the matrix. Gupta et al. [118] reported that the microwave shielding properties were affected by the various morphologies of iron oxides. Using a two-step sol–gel process, they produced various ferrite structures, including cubes, rods, and flakes, covered with multilayer rGO. Their EMI shielding capabilities were evaluated in the Ku-band frequency range. In comparison to the flake- and cube-shaped iron oxides, the rod-shaped iron oxide covered with rGO sheets had the highest shielding efficiency value of ~33.30 dB (>99.9% attenuation). This resulted from the combined effect of magnetic and dielectric losses. Anisotropy energy in the composites, eddy current effects, and natural resonances were the sources of magnetic loss. The nanoferrite particle content of the composite was the main cause of eddy currents in the microwave ranges. The surface anisotropic field would result in higher anisotropy energy for the small materials due to the small-size effect. The higher anisotropy energy also contributed to greater microwave absorption. Moreover, the magnetic iron oxide was coated with an rGO layer, which increased the interfaces and surface polarization charges. Interfacial polarization is a significant polarization process, and the corresponding relaxation will lead to a loss mechanism. One of the main causes of dielectric loss could be interfacial polarization. It is well known that some heating-induced absorption losses are caused by the interaction of surface-formed molecular dipoles with the microwave field.
Dhawan et al. created a conducting ferrimagnetic PANI nanocomposite implanted with titanium dioxide (70–90 nm) and γ-Fe2O3 (9–12 nm) nanoparticles using micro-emulsion polymerization [119]. They discovered that the high shielding effectiveness value of −45 dB owing to absorption (SEA) was caused by dielectric and magnetic losses that resulted from the combined action of γ-Fe2O3 and TiO2. In contrast, the SEA of PANI-TiO2 was around 22.4 dB, but that of PANI-γ-Fe2O3 was approximately 8.8 dB. In another work, PANI tubes made of rGO coated with γ-Fe2O3 nanoparticles were synthesized and characterized by Singh et al. [120]. The intercalated iron oxide nanoparticles were produced by thermally breaking down ferric acetyl acetonate in a reducing atmosphere. The β-naphthalene sulphonic acid-induced oxidative polymerization of aniline, which produced the core–shell shape, was also used to enclose those nanoparticles. At a thickness of 2.5 mm, the presence of rGO-γ-Fe2O3 in the PANI core structures increased the composite’s interfacial polarization and effective anisotropy energy, which increased scattering and produced a high shielding efficiency of about 51 dB.
Iron oxides are frequently combined with different carbon nanomaterials. The composite material can effectively absorb, reflect, and dissipate electromagnetic radiation over a wider frequency range by fusing the electrical conductivity of carbon nanotubes with the magnetic qualities of iron oxides. Fe3O4-nanoparticle-loaded functionalized multi-walled carbon nanotubes were produced by Bhaskara Rao et al. [121]. They discovered a high overall specific shielding efficiency of around 49.56 dB/(g cm−3), along with improved absorption (15.85 dB) and reflection (9.43 dB). Liu et al. [122] developed trilayer-type laminated nanocomposites with a matching layer of 15 wt.% Fe3O4, an absorbent layer of 5 wt.% MWCNTs, and a reflecting layer of 10 wt.% MWCNTs. Their results showed that such trilayer-type laminated nanocomposites have an excellent ability to absorb microwaves up to 40 dB in the 13 GHz to 40 GHz frequency range. Ferroferric oxide (Fe3O4) and MWCNTs were integrated by Li et al. into a core–shell system made of high-density polyethylene (HDPE), polyvinylidene fluoride (PVDF), and polystyrene (PS) [123]. The composite with MWCNTs in the PS shell and Fe3O4 in the PVDF matrix had the highest SE, measuring 25 dB at 9.5 GHz with 1 vol.% Fe3O4 and 1 vol.% MWCNTs. The SE was over 20 dB throughout the tested frequency range (X-band). Prasad et al. [124] developed a facile two-step hydrothermal process for the synthesis of a MoS2–reduced graphene oxide/Fe3O4 (MoS2-rGO/Fe3O4) nanocomposite and its application as an enhanced shielding material against electromagnetic interference. The Fe3O4 nanoparticles were spherical and evenly distributed throughout the MoS2-rGO composite. The MoS2-rGO/Fe3O4 nanocomposite was found to be an extremely effective electromagnetic shielding material in the 8.0–12.0 GHz X-band, according to an examination of its electromagnetic shielding efficiency. The MoS2-rGO composite showed low shielding performance (SET ~3.81 dB) compared to the MoS2-rGO/Fe3O4 nanocomposite (SET ~8.27 dB). This was caused by interfacial polarization in the presence of an electromagnetic field.
By the chemical oxidative polymerization of pyrrole, Sambyal et al. created a conducting polymer-based composite encapsulated with barium strontium titanate (BST), rGO, and Fe3O4 nanoparticles [125]. Filler components in the conducting polymer matrix produced an absorption-dominated shielding efficiency value of 48 dB in the 8.2–12.4 GHz (X-band) frequency range. Furthermore, the chemical and thermal stability of the composite material was enhanced by the use of magnetic and dielectric fillers. By using a self-assembly process, rGO was also added to latex coupled with magnetic iron oxide (Fe3O4) and flexible natural rubber [126]. Compared to NRG composites, Fe3O4 enhanced the electromagnetic interference shielding effectiveness (EMI SE) of natural rubber/reduced graphene oxide (NRG) composites. The EMI SE value of the NRMG composite with 10 parts per 100 parts of rubber rGO is 1.4 times higher than that of the NRG composite with the same rGO content in the 8.2–12.4 GHz frequency band. With a specific EMI SE of 26.4 dB mm−1, the NRMG composite outperforms the previously reported polymer/Fe3O4@rGO composites with a low rGO concentration. Remarkably, the NRMG composite’s EMI SE only decreases by 3.5% after 2000 bending–release cycles, indicating that it may find application in flexible shielding materials.
In order to create 3D network porous graphene nanoplatelet (GNP)/Fe3O4/epoxy nanocomposites with a low density of 0.34–0.73 g/cm3, Liu et al. suggested a new and simple method called epoxy–water–inorganic filler suspended emulsion polymerization [127]. The porous nanocomposite that resulted from loading 7 wt.% graphene nanoplatelets and 7 wt.% Fe3O4 nanoparticles showed a satisfactory specific electromagnetic interference (EMI) shielding effectiveness of about 37.03 dB/(g/cm3), which was much higher than that of the solid equivalents (28.30 dB/(g/cm3)). In another study, Fe3O4/thermally annealed graphene aerogel (Fe3O4/TAGA) was created by first thermally annealing ethylenediamine-functionalized Fe3O4 (NH2-Fe3O4) nanoparticles with graphene oxide (GO), followed by the addition of l-ascorbic acid [128]. Then, the Fe3O4/TAGA/epoxy nanocomposites were created using a template-casting technique. The resulting Fe3O4/TAGA/epoxy nanocomposites achieved the highest electromagnetic interference shielding effectiveness (EMI SE of 35 dB in the X-band) when the mass ratio of GO to NH2-Fe3O4 was 2:1 and the total mass fraction of Fe3O4/TAGA was 2.7 wt.% (comprising 1.5/1.2 wt.% Fe3O4/TAGA). This was significantly better than epoxy nanocomposites with the same Fe3O4/thermally annealed graphene oxide (Fe3O4/TAGO) loading, which only showed an EMI SE of 10 dB.
Liu et al. created magnetic reduced graphene oxide rGO@Fe3O4 nanoplatelets (NPs), which were used as fillers, by co-precipitation and electrostatic self-assembly [129]. The nanocomposites were made by applying external magnetic fields to align rGO@Fe3O4 NPs during epoxy curing. Because of the anisotropic properties of rGO@Fe3O4 NPs and external magnetic fields, the nanocomposite containing aligned rGO@Fe3O4 NPs showed anisotropic thermal conductivity. The produced sample exhibited exceptional thermal stability and 13.45 dB EMI shielding at 8.2 GHz. Overall, the rGO@Fe3O4 NPs’ in-plane interaction was enhanced by aligning them under a magnetic field, which promoted the growth of horizontal thermal conductive networks. Reflection is the main EMI shielding mechanism. Three-dimensional Fe3O4-decorated carbon nanotube/reduced graphene oxide foam/epoxy (3D Fe3O4-CNT/rGF/EP) nanocomposites with highly aligned three-dimensional structures were created by Liang et al. using a simple template approach [130]. The obtained 3D Fe3O4-CNT/rGF/EP nanocomposites with 0.24 wt.% rGF and 2.76 wt.% Fe3O4-CNTs showed a remarkable electrical conductivity of 15.3 S/m and an EMI SE of 36 dB within the X-band range, which was a nearly 482% improvement when compared to the EMI SE value of physically blended Fe3O4-CNT/EP nanocomposites without a three-dimensional structure (~6 dB). Using a supercritical carbon dioxide (Sc-CO2) foaming technique, lightweight and flexible methyl vinyl silicone rubber (VMQ)/multi-walled carbon nanotube (MWCNT)/ferriferrous oxide (Fe3O4) nanocomposite foams with superior EMI shielding capabilities were created [131]. The addition of a cellular structure and magnetic Fe3O4 nanoparticles greatly improved the VMQ/MWCNT/Fe3O4 foams’ microwave-absorbing capacity by successfully lowering secondary electromagnetic wave interference brought on by reflection. In the 8.2–12.4 GHz frequency range, these nanocomposite foams, which have an approximate density of 0.48 g/cm3, showed an EMI shielding efficiency (SE) of 27.5 dB and an average absorption ratio of up to 64%. The foams had a specific EMI SE of about 72 dB g−1 cm3 and a high conductivity of about 14.6 S/m with a filler loading of 1.78 vol.%.
Shu et al. [132] used rice-husk-based activated carbon (AC) and produced compo-sites with acicular or octahedral Fe3O4 nanoparticles. A significant effect on the EMI shielding effectiveness of the composites was assigned to the morphology of the metallic nanoparticles and the layered structure of the C component of the composites. They achieved an EMI SE of −52.14 dB.
Record values of EMI shielding effectiveness were achieved for MXene by incorporating magnetic nanoparticles as intercalators between layers [133] and 3D N-doped GO with silver nanowires (−79.99 dB) at a thickness of 2.660 mm [134].

3. Conclusions

This review has provided a comprehensive overview of the developments, classification, trends, and advances in new electromagnetic shielding materials. It gives a thorough literature review of the EMI shielding properties of new EMI shielding materials such as biochars, scaffolds, rare earth, and ferrite-based materials. A detailed discussion is given regarding the preparation methods, structure, EMI shielding performance, EMI shielding mechanisms, and application perspectives of these materials.
Looking ahead, the prospects for and research on electromagnetic shielding materials should focus on scalability, improving performance, cost-effectiveness, sustainability, versatility, and production methods to advance new materials like biochars, scaffolds, rare earth, and ferrite-based materials as novel EMI shielding solutions. Biochars are promising eco-friendly, sustainable, and renewable materials that can be potentially used as EMI shielding materials in electronic devices and construction materials, similar to scaffolds, new-generation, easy-to-manufacture materials with excellent EMI shielding performance. Although MXene, carbon nanotubes, and graphene-based composites are highly efficient EMI shielding materials due to their large free surface area, high aspect ratio, chemical stability, and lightweight nature, the large number of synthetic phases leads to their high price. Carbon-based nanomaterials derived from biomass are being developed in a quest for new, cost-efficient, sustainable, and affordable materials.
The current disadvantages of these materials are the thickness of the shielding barrier and lack of transparency, which limit some applications.
Materials doped with rare earth elements have been developing over the last few years. They have specific structures, interfacial polarization, multiple reflections, and excellent conductive and EMI shielding properties. Combining the magnetic properties of ferrite-based materials with the electrical conductivity of carbon nanomaterials, composite materials can efficiently absorb, reflect, and dissipate electromagnetic energy across a broader frequency range. Although the materials showed outstanding EMI shielding performance, the price of REEs will limit the application of these composites to more sophisticated and sensitive fields such as medicine.
Their EMI shielding effectiveness, good magnetic and dielectric properties, excellent thermal stability, and high electrical conductivity and mechanical strength make biochars, scaffolds, rare earth, and ferrite-based materials ecological and sustainable solutions as new products for blocking EMWs.

Author Contributions

Conceptualization, S.J. and D.M.; validation, S.D., M.Y., K.H., B.N. and D.K.; resources, S.J. and M.Y.; writing—original draft preparation, S.J., S.D., D.K., M.Y., K.H., B.N. and D.M.; writing—review and editing, S.J., M.Y. and D.M.; supervision, S.J., M.Y., K.H., B.N. and D.M.; project administration, S.J.; funding acquisition, S.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the European Union’s Horizon Europe Coordination and Support Actions Programme under grant agreement No 101079151-GrInShield. S.J., S.D., D.K., and D.M. thank the Ministry of Science, Technological Development and Innovation of the Republic of Serbia (grant number 451-03-136/2025-03/200017).

Data Availability Statement

This review article does not contain any original data. All data referenced in this article are publicly available from the sources cited in the references. No new datasets were generated or analyzed in this work.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Liang, C.; Qiu, H.; Song, P.; Shi, X.; Kong, J.; Gu, J. Ultra-light MXene aerogel/wood-derived porous carbon composites with wall-like “mortar/brick” structures for electromagnetic interference shielding. Sci. Bull. 2020, 65, 616–622. [Google Scholar] [CrossRef]
  2. Li, J.; Cui, M.; Wen, J.; Chen, Y.; Shi, B.; Fan, H.; Xiang, J. Leather-like hierarchical porous composites with outstanding electromagnetic interference shielding effectiveness and durability. Compos. B Eng. 2021, 225, 109272. [Google Scholar] [CrossRef]
  3. Manassas, A.; Apostolidis, C.; Iakovidis, S.; Babas, D.; Samaras, T. A study of the long term changes in the electromagnetic environment using data from continuous monitoring sensors in Greece. Sci. Rep. 2023, 13, 13784. [Google Scholar] [CrossRef]
  4. Tolvanen, J.; Hannu, J.; Hietala, M.; Kordas, K.; Jantunen, H. Biodegradable multiphase poly(lactic acid)/biochar/graphite composites for electromagnetic interference shielding. Compos. Sci. Technol. 2019, 181, 107704. [Google Scholar] [CrossRef]
  5. Calvente, I.; Núñez, M.I. Is the sustainability of exposure to non-ionizing electromagnetic radiation possible? Med. Clín. 2024, 162, 387–393. [Google Scholar] [CrossRef]
  6. Bandara, P.; Carpenter, D.O. Planetary electromagnetic pollution: It is time to assess its impact. Lancet Planet. Health 2018, 2, e512–e514. [Google Scholar] [CrossRef]
  7. Karimi, A.; Ghadiri Moghaddam, F.; Valipour, M. Insights in the biology of extremely low-frequency magnetic fields exposure on human health. Mol. Biol. Rep. 2020, 47, 5621–5633. [Google Scholar] [CrossRef] [PubMed]
  8. Lin, Y.; Gao, P.; Guo, Y.; Chen, Q.; Lang, H.; Guo, Q.; Miao, X.; Li, J.; Zeng, L.; Guo, G. Effects of long-term exposure to L-band high-power microwave on the brain function of male mice. Biomed Res. Int. 2021, 2021, 2237370. [Google Scholar] [CrossRef]
  9. Touitou, Y.; Selmaoui, B.; Lambrozo, J. Assessment of cortisol secretory pattern in workers chronically exposed to ELF-EMF generated by high voltage transmission lines and substations. Environ. Int. 2022, 161, 107103. [Google Scholar] [CrossRef]
  10. Alshammary, R.N.; Mohammed Zaki, Z.D.; Al-Haaik, A.G. Effect of mobile frequencies exposure on histology of retina and cornea in pregnant albino mice. Iraqi J. Vet. Sci. 2022, 36, 245–249. [Google Scholar] [CrossRef]
  11. Tuhanioğlu, B.; Erkan, S.O.; Gürgen, S.G.; Özdaş, T.; Görgülü, O.; Çiçek, F.; Günay, İ. The effect of very low dose pulsed magnetic waves on cochlea. Braz. J. Otorhinolaryngol. 2019, 85, 282–289. [Google Scholar] [CrossRef] [PubMed]
  12. Dunham, A.; Pegg, J.R.; Carolsfeld, W.; Davies, S.; Murfitt, I.; Boutillier, J. Effects of submarine power transmission cables on a glass sponge reef and associated megafaunal community. Mar. Environ. Res. 2015, 107, 50–60. [Google Scholar] [CrossRef] [PubMed]
  13. Harsanyi, P.; Scott, K.; Easton, B.A.A.; de la Cruz Ortiz, G.; Chapman, E.C.N.; Piper, A.J.R.; Rochas, C.M.V.; Lyndon, A.R. The effects of anthropogenic electromagnetic fields (EMF) on the early development of two commercially important crustaceans, European lobster, Homarus gammarus (L.) and Edible crab, Cancer pagurus (L.). J. Mar. Sci. Eng. 2022, 10, 564. [Google Scholar] [CrossRef]
  14. Piccinetti, C.C.; De Leo, A.; Cosoli, G.; Scalise, L.; Randazzo, B.; Cerri, G.; Olivotto, I. Measurement of the 100 MHz EMF radiation in vivo effects on zebrafish D. rerio embryonic development: A multidisciplinary study. Ecotoxicol. Environ. Saf. 2018, 154, 268–279. [Google Scholar] [CrossRef]
  15. Boga, A.; Emre, M.; Sertdemir, Y.; Uncu, İ.; Binokay, S.; Demirhan, O. Effects of GSM-like radiofrequency irradiation during the oogenesis and spermiogenesis of Xenopus laevis. Ecotoxicol. Environ. Saf. 2016, 129, 137–144. [Google Scholar] [CrossRef]
  16. Chen, L.; Mai, T.; Ji, X.-X.; Wang, P.-L.; Qi, M.-Y.; Liu, Q.; Ding, Y.; Ma, M.-G. 3D printing of customizable and lightweight multilayer MXene/nanocellulose architectures for tunable electromagnetic interference shielding via direct ink writing. Chem. Eng. J. 2023, 476, 146652. [Google Scholar] [CrossRef]
  17. Chung, D.D.L. Materials for electromagnetic interference shielding. Mater. Chem. Phys. 2020, 255, 123587. [Google Scholar] [CrossRef]
  18. Orasugh, J.T.; Ray, S.S. Functional and structural facts of effecttive electromagnetic interference shielding materials: A review. ACS Omega 2023, 8, 8134–8158. [Google Scholar] [CrossRef]
  19. Li, Y.; Wu, H.; Deng, K.; Jiang, J.; Yang, Z.; Zhang, R. Preparation of graphite/ferrite/resin-based composite high-efficiency wave-absorbing materials by selective laser sintering. Adv. Eng. Mater. 2023, 25, 2301080. [Google Scholar] [CrossRef]
  20. Ji, Z.; Wang, Q.; Wang, Z.; Duan, Y.; Dong, C.; Liaw, P.K. Electromagnetic wave-absorbing behavior of soft-magnetic medium entropy alloys with BCC/L21 coherent microstructure. Mater. Des. 2022, 222, 111054. [Google Scholar] [CrossRef]
  21. Zhao, Y.; Lin, Z.; Huang, L.; Meng, Z.; Yu, H.; Kou, X.; Zou, Z.; Huang, P.; Wang, Y.; Xi, D.; et al. Simultaneous optimization of conduction and polarization losses in CNT@NiCo compounds for superior electromagnetic wave absorption. J. Mater. Sci. Technol. 2023, 166, 34–46. [Google Scholar] [CrossRef]
  22. Zhou, Y.; Zhu, Y.; Chen, P.; Li, X. Effect of rare earth-transition metal electronic interaction on magnetism in FeCoEr alloys. J. Rare Earths 2024, in press. [CrossRef]
  23. Liu, H.; Yang, S.; Wang, G.; Liu, H.; Peng, Y.; Sun, C.; Li, J.; Chen, J. Strong electronic orbit coupling between cobalt and single-atom praseodymium for boosted nitrous oxide decomposition on Co3O4 catalyst. Environ. Sci. Technol. 2022, 56, 16325–16335. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, H.; Zhang, H.; Zhao, K.; Nie, A.; Alharthi, S.; Amin, M.A.; El-Bahy, Z.M.; Li, H.; Chen, L.; Xu, B.B.; et al. Research progress on electromagnetic wave absorption based on magnetic metal oxides and their composites. Adv. Compos. Hybrid Mater. 2023, 6, 120. [Google Scholar] [CrossRef]
  25. Jha, A.R. Rare Earth Materials: Properties and Applications; CRC Press Taylor and Francis Group: Boca Raton, FL, USA, 2014. [Google Scholar]
  26. Wen, H.; Zhao, W.; Han, X. Constructing Co3O4/La2Ti2O7 p-n heterojunction for the enhancement of photocatalytic hydrogen evolution. Nanomaterials 2022, 12, 1695. [Google Scholar] [CrossRef] [PubMed]
  27. Legvold, S. Chapter 3: Rare earth metals and alloys in book. Handb. Ferromagn. Mater. 1980, 1, 183–295. [Google Scholar] [CrossRef]
  28. Wang, Y.; Wu, X.; Zhang, W.; Chen, W. Synthesis and electromagnetic properties of La-doped Ni–Zn ferrites. J. Magn. Magn. Mater. 2016, 398, 90–95. [Google Scholar] [CrossRef]
  29. Kumar, P.; Pathak, S.; Singh, A.; Verma, R.; Khanduri, H.; Jain, K.; Tawale, J.; Wang, L.; Pant, R.P. Augmented magnetic nanoparticle assimilation in rGO sheets for tailored static and dynamic magnetic properties in surface functionalized Co0.8Zn0.2Fe2O4 nanoferrite–rGO hybrid structures. J. Mater. Chem. C 2024, 12, 18036–18047. [Google Scholar] [CrossRef]
  30. Pathak, S.; Verma, R.; Singhal, S.; Chaturvedi, R.; Kumar, P.; Sharma, P.; Pant, R.P.; Wang, X. Spin dynamics investigations of multifunctional ambient scalable Fe3O4 surface decorated ZnO magnetic nanocomposite using FMR. Sci. Rep. 2021, 11, 3799. [Google Scholar] [CrossRef]
  31. Ma, Z.; Jiang, R.; Jing, J.; Kang, S.; Ma, L.; Zhang, K.; Li, J.; Zhang, Y.; Qin, J.; Yun, S.; et al. Lightweight Dual-Functional Segregated Nanocomposite Foams for Integrated Infrared Stealth and Absorption-Dominant Electromagnetic Interference Shielding. Nano-Micro Lett. 2024, 16, 223. [Google Scholar] [CrossRef]
  32. Ma, Y.; Jiang, R.; Zhang, Y.; Ma, L.; Bai, Y.; Zhang, K.; Zuo, X.; Zuo, Y.; Jing, H.; Qin, J.; et al. Lightweight and mechanically strong MXene-Based microcellular nanocomposite foams for integrated electromagnetic interference shielding and thermal management. Compos. Sci. Technol. 2025, 260, 110988. [Google Scholar] [CrossRef]
  33. Kruželák, J.; Kvasničáková, A.; Hložeková, K.; Dosoudil, R.; Gořalík, M.; Hudec, I. Electromagnetic Interference Shielding and Physical-Mechanical Characteristics of Rubber Com-posites Filled with Manganese-Zinc Ferrite and Carbon Black. Polymers 2021, 13, 616. [Google Scholar] [CrossRef] [PubMed]
  34. Chu, W.; Li, J.; Lin, J.; Li, W.; Xin, J.; Liu, F.; He, X.; Ma, Z.; Zhao, Q. Honeycomb-like Polyimide/Fe3O4@PPy foam for electromagnetic wave shielding with excellent absorption characteristics. Compos. Sci. Technol. 2024, 249, 110489. [Google Scholar] [CrossRef]
  35. Zhou, Z.; Song, Q.; Huang, B.; Feng, S.; Lu, C. Facile Fabrication of Densely Packed Ti3C2 MXene/Nanocellulose Composite Films for Enhancing Electromagnetic Interference Shielding and Electro-/Photothermal Performance. ACS Nano 2021, 15, 12405–12417. [Google Scholar] [CrossRef]
  36. Miao, Y.; Lin, J.; Wang, E.; Liang, Y.; Li, W.; Dai, C.; Huang, J.; Zhang, W. Electrically conductive bamboo charcoal@cellulose nanofibrils based composite membranes designed for electromagnetic interference shielding and flame retardant. Ind. Crop. Prod. 2023, 206, 117713. [Google Scholar] [CrossRef]
  37. Perumal, R.S.; Muralidharan, B. Valorization of Ricinus communis outer shell biomass to biochar: Impact of thermal decomposition temperature on physicochemical properties and EMI shielding performance. Results Eng. 2024, 24, 103097. [Google Scholar] [CrossRef]
  38. Tomczyk, A.; Sokołowska, Z.; Boguta, P. Biochar physicochemical properties: Pyrolysis temperature and feedstock kind effects. Rev. Environ. Sci. Biotechnol. 2020, 19, 191–215. [Google Scholar] [CrossRef]
  39. Li, S.; Huang, A.; Chen, Y.-J.; Li, D.; Turng, L.-S. Highly filled biochar/ultra-high molecular weight polyethylene/linear low density polyethylene composites for high-performance electromagnetic interference shielding. Compos. B Eng. 2018, 153, 277–284. [Google Scholar] [CrossRef]
  40. Gabhi, R.S.; Kirk, D.W.; Jia, C.Q. Preliminary investigation of electrical conductivity of monolithic biochar. Carbon 2017, 116, 435–442. [Google Scholar] [CrossRef]
  41. Gokce, E.C.; Calisir, M.D.; Selcuk, S.; Gungor, M.; Acma, M.E. Electromagnetic interference shielding using biomass-derived carbon materials. Mater. Chem. Phys. 2024, 317, 129165. [Google Scholar] [CrossRef]
  42. Akgül, G.; Demir, B.; Gündoğdu, A.; Türk, A.S.; Sözer, S. Biochar-iron composites as electromagnetic interference shielding material. Mater. Res. Express 2020, 7, 015604. [Google Scholar] [CrossRef]
  43. Natalio, F.; Corrales, T.P.; Feldman, Y.; Lew, B.; Graber, E.R. Sustainable lightweight biochar-based composites with electromagnetic shielding properties. ACS Omega 2020, 5, 32490–32497. [Google Scholar] [CrossRef]
  44. Savi, P.; Yasir, M. Waveguide measurements of biochar derived from sewage sludge. Electron. Lett. 2020, 56, 335–337. [Google Scholar] [CrossRef]
  45. Savi, P.; Yasir, M.; Bartoli, M.; Giorcelli, M.; Longo, M. Electrical and Microwave Characterization of Thermal Annealed Sewage Sludge Derived Biochar Composites. Appl. Sci. 2020, 10, 1334. [Google Scholar] [CrossRef]
  46. Yasir, M.; Zaccagnini, P.; Palmara, G.; Frascella, F.; Paccotti, N.; Savi, P. Morphological Characterization and Lumped Element Model of Graphene and Biochar Thick Films. C 2021, 7, 36. [Google Scholar] [CrossRef]
  47. Savi, P.; Ruscica, G.; di Summa, D.; Natali Sora, I. Shielding effectiveness measurements of drywall panel coated with biochar layers. Electronics 2022, 11, 2312. [Google Scholar] [CrossRef]
  48. Ruscica, G.; Peinetti, F.; Natali Sora, I.; Savi, P. Analysis of electromagnetic shielding properties of cement-based composites with biochar and PVC as fillers. C 2024, 10, 21. [Google Scholar] [CrossRef]
  49. Nikolopoulos, C.D.; Baklezos, A.T.; Kapetanakis, T.N.; Vardiambasis, I.O.; Tsubota, T.; Kalderis, D. Characterization of the electromagnetic shielding effectiveness of biochar-based materials. IEEE Access 2023, 11, 6413–6420. [Google Scholar] [CrossRef]
  50. Milenkovic, M.; Saeed, W.; Yasir, M.; Nassar, K.E.S.; Syrgiannis, Z.; Milivojevic, D.; Spanopoulos, I.; Azmy, A.; Bajuk-Bogdanovic, D.; Maletić, S.; et al. Carbonized apples and quinces stillage for electromagnetic shielding. Nanomaterials 2024, 14, 1882. [Google Scholar] [CrossRef]
  51. Wang, C.; Wang, Y.; Zou, F.; Fang, B.; Zhao, J.; Zhang, H.; Guo, J.; Jia, L.; Yan, D. Construction of lightweight, high-energy absorption 3D-printed scaffold for electromagnetic interference shielding with low reflection. Compos. B Eng. 2025, 291, 112043. [Google Scholar] [CrossRef]
  52. Tran, T.T.V.; Vo, D.-V.N.; Nguyen, S.T.; Luu, S.D.N.; Mofijur, M.; Vu, C.M. In situ sintered silver decorated 3D structure of cellulose scaffold for highly thermoconductive electromagnetic interference shielding epoxy nanocomposites. J. Appl. Polym. Sci. 2021, 138, 51193. [Google Scholar] [CrossRef]
  53. He, H.; Wang, Y.; Zhao, Z.; Wang, Q.; Wei, Q.; Cai, Y. Dual-encapsulated multifunctional phase change composites based on biological porous carbon for efficient energy storage and conversion, thermal management, and electromagnetic interference shielding. Energy Storage Mater. 2022, 55, 105358. [Google Scholar] [CrossRef]
  54. Wei, Z.; Cheng, Y.; Hu, X.; Meng, Y.; Zhan, Y.; Li, Y.; Xia, H.; Jiang, X.; Chen, Z. Cellulose–derived carbon scaffolds with bidirectional gradient Fe3O4 distribution: Integration of green EMI shielding and thermal management. Int. J. Biol. Macromol. 2024, 275, 133724. [Google Scholar] [CrossRef]
  55. Wang, Z.-X.; Han, X.-S.; Zhou, Z.-J.; Meng, W.-Y.; Han, X.-W.; Wang, S.-J.; Pu, J.-W. Lightweight and elastic wood-derived composites for pressure sensing and electromagnetic interference shielding. Compos. Sci. Technol. 2021, 213, 108931. [Google Scholar] [CrossRef]
  56. Hu, B.; Guo, H.; Cui, Y.; Li, J.; Cao, M.; Qi, W.; Cao, X.; Li, B. Engineering multifunctional phase change composites enabled by dual-interpenetrating hybrid scaffold for excellent thermal conductivity and electromagnetic absorption. J. Chem. Eng. 2024, 492, 152259. [Google Scholar] [CrossRef]
  57. Yasir, M.; di Summa, D.; Ruscica, G.; Natali Sora, I.; Savi, P. Shielding properties of cement composites filled with commercial biochar. Electronics 2020, 9, 819. [Google Scholar] [CrossRef]
  58. Babu, M.; Bapu, B.R.T.; Muruganantham, P.; Anita, R.; Nagaraju, V.; P. J. Kumar, S. Role of cashew shell biochar on EMI shielding behaviour of carbon fibre–epoxy nanocomposites in E, F, I and J band–microwave frequencies. Biomass Convers. Biorefinery 2023, 13, 375–382. [Google Scholar] [CrossRef]
  59. Suresh, N.; Sivakumar, P.; Malathi, A.C.J.; Balamurugan, K.S. Effect of temperature on EMI shielding behavior of jack fruit rags biochar and waste silk fiber-reinforced vinyl ester composite. J. Mater. Sci. Mater. Electron. 2025, 36, 100. [Google Scholar] [CrossRef]
  60. Sharma, S.; Parne, S.R.; Srihari, S.; Panda, S.; Gandi, S. Progress in microwave absorbing materials: A critical review. Adv. Colloid Interface Sci. 2024, 327, 103143. [Google Scholar] [CrossRef]
  61. Nikzad, A.; Parvizi, R. Presence of neodymium and gadolinium in cobalt ferrite lattice: Structural, magnetic and microwave features for electromagnetic wave absorbing. J. Rare Earth 2020, 38, 411–417. [Google Scholar] [CrossRef]
  62. Zhang, C.; Wang, L.; Ji, X.; Chen, L.; Yan, H.; Xing, Z.; Wang, B. Preparation and electrorheological behavior of rare-earth La ion doping MIL-125 nanoparticles. Adv. Eng. Mater. 2024, 26, 2401323. [Google Scholar] [CrossRef]
  63. Isari, A.A.; Ghaffarkhah, A.; Hashemi, S.A.; Wuttke, S.; Arjmand, M. Structural design for EMI shielding: From underlying mechanisms to common pitfalls. Adv. Mater. 2024, 36, 2310683. [Google Scholar] [CrossRef]
  64. Qiao, J.; Li, L.; Liu, J.; Wu, N.; Liu, W.; Wu, F.; Zeng, Z. The vital application of rare earth for future high-performance electromagnetic wave absorption materials: A review. J. Mater. Sci. Technol. 2024, 176, 188–203. [Google Scholar] [CrossRef]
  65. Deng, Y.; Wang, L.; Liu, W.; Wu, N.; Liu, J.; Pan, F.; Zeng, Z. Research progress on controllable absorption properties of rare earth element doped electromagnetic wave absorbing materials. Chin. J. Chem. 2024, 42, 2536–2549. [Google Scholar] [CrossRef]
  66. Mohapatra, P.P.; Ghosh, S.; Jain, A.; Aich, S.; Dobbidi, P. Rare earth substituted lithium ferrite/carbon black ceramic composites for shielding electromagnetic radiation. J. Magn. Magn. Mater. 2023, 573, 170678. [Google Scholar] [CrossRef]
  67. Zhu, X.; Zhang, J.; Liu, J.; Zhang, Y. Recent progress of rare-earth doped upconversion nanoparticles: Synthesis, optimization, and applications. Adv. Sci. 2019, 6, 1901358. [Google Scholar] [CrossRef]
  68. Liu, J.; Pan, Y.; Yu, L.; Gao, Z.; Zhang, S.; Lan, D.; Jia, Z.; Wu, G. MoS2-based composites for microwave absorption mechanism-oriented structural optimization and design perspectives. Carbon 2025, 238, 120233. [Google Scholar] [CrossRef]
  69. Prasad, J.; Singh, A.K.; Gahlot, A.P.S.; Tomar, M.; Gupta, V.; Singh, K. Electromagnetic interference shielding properties of hierarchical core-shell palladium-doped MoS2/CNT nanohybrid materials. Ceram. Int. 2021, 47, 27586–27597. [Google Scholar]
  70. Zhang, H.; Ma, G.; Wu, H.; Yuan, M.; Liu, X.; Huang, Z. Enhanced mechanical and wave-absorption properties of SiC–Si3N4–FeSi porous ceramics by introducing Al. J. Am. Ceram. Soc. 2024, 107, 6821–6832. [Google Scholar] [CrossRef]
  71. Wang, L.; Chen, S.; Zhu, X.; Chen, J.; Liang, J.; Wang, M. Influence of various rare earth elements co-integration on the properties of super multifunctional metal-based materials fabricated by laser cladding. Appl. Mater. Today 2024, 37, 102093. [Google Scholar] [CrossRef]
  72. Li, Z.; Mi, W.; Bai, H. The role of rare–earth dopants in tailoring the magnetism and magnetic anisotropy in Fe4N. J. Phys. Chem. Solids 2018, 116, 7–14. [Google Scholar] [CrossRef]
  73. Zhang, W.; Zhao, B.; Xiang, H.; Dai, F.-Z.; Wu, S.; Yanchun, Z. One-step synthesis and electromagnetic absorption properties of high entropy rare earth hexaborides (HEREB6) and high entropy rare earth hexaborides/borates (HEREB6/HE REBO3) composite powders. J. Adv. Ceram. 2021, 10, 62–77. [Google Scholar] [CrossRef]
  74. Chen, H.; Zhao, B.; Zhao, Z.; Xiang, H.; Dai, F.-Z.; Liu, J.; Zhou, Y. Achieving strong microwave absorption capability and wide absorption bandwidth through a combination of high entropy rare earth silicide carbides/rare earth oxides. J. Mater. Sci. Technol. 2020, 47, 216–222. [Google Scholar] [CrossRef]
  75. Zámborszky, F.; Gyüre Garami, B.; Jánosi, B.; Vajtai, L.; Hegyessy, L.; Gresits, I.; Simon, F. High-frequency characterization of Fe-based nanocrystalline cores. J. Magn. Magn. Mater. 2022, 564, 170027. [Google Scholar] [CrossRef]
  76. Azmat, M.; Yang, J.; Li, Q.; Zhang, J.; Haibo, J.; Kashif, N.M.; Li, J. Role of 4f electrons and 3d-4f hybridization in metal-insulator transition in RE (La, Nd, Sm, Eu, Dy and Er)-doped vanadium dioxide for thermochromic applications. Ceram. Int. 2024, 50, 11119–11128. [Google Scholar] [CrossRef]
  77. Nagaraj, N.; Manjunatha, H.C.; Vidya, Y.S.; Seenappa, L.; Sridhar, K.N.; Gupta, P.S.D. Investigations on lanthanide polymers for radiation shielding purpose. Radiat. Phys. Chem. 2022, 199, 110310. [Google Scholar] [CrossRef]
  78. Hasan, M.S.; Khan, M.I.; Mandal, G.; Awais, M.; Farhat, L.B.; Liu, J. Integrating the structural, electro-optical, dielectric, and magnetic features of Co–Mg–La ferrites/graphene composites. J. Am. Ceram. Soc. 2025, 108, e20329. [Google Scholar] [CrossRef]
  79. Jia, Z.; Gao, Z.; Kou, K.; Feng, A.; Zhang, C.; Xu, B.; Wu, G. Facile synthesis of hierarchical A-site cation deficiency perovskite LaxFeO3−y/RGO for high efficiency microwave absorption. Compos. Commun. 2020, 20, 100344. [Google Scholar] [CrossRef]
  80. Serrano, A.; García-Martín, E.; Granados-Miralles, C.; Gorni, G.; López-Sánchez, J.; Ruiz-Gómez, S.; Pérez, L.; Quesada, A.; Fernández, J.F. Hexaferrite-based permanent magnets with upper magnetic properties by cold sintering process via a non-aqueous solvent. Acta Mater. 2021, 219, 117262. [Google Scholar] [CrossRef]
  81. Kahil, H.; Faramawy, A.; El-Sayed, H.; Abdel-Sattar, A. Magnetic properties and SAR for gadolinium-doped iron oxide nanoparticles prepared by hydrothermal method. Crystals 2021, 11, 1153. [Google Scholar] [CrossRef]
  82. Choudhary, N.; Verma, M.K.; Sharma, N.D.; Sharma, S.; Singh, D. Correlation between magnetic and transport properties of rare earth doped perovskite manganites La0.6R0.1Ca0.3MnO3 (R = La, Nd, Sm, Gd, and Dy) synthesized by Pechini process. Mater. Chem. Phys. 2020, 242, 122482. [Google Scholar] [CrossRef]
  83. Available online: http://abulafia.mt.ic.ac.uk/shannon/ptable.php (accessed on 30 March 2025).
  84. Rajeshwari, A.; Punithavathy, I.K.; Jeyakumar, S.J.; Jothibas, M. Dependance of gadolinium ions on structural, magnetic and dielectric properties of manganese nanoferrites. Mater. Chem. Phys. 2023, 296, 127195. [Google Scholar] [CrossRef]
  85. Kadam, A.B.; Mande, V.K.; Kadam, S.B.; Kadam, R.H.; Shirsath, S.E.; Borade, R.B. Influence of gadolinium (Gd3+) ion substitution on structural, magnetic and electrical properties of cobalt ferrites. J. Alloys Compd. 2020, 840, 155669. [Google Scholar] [CrossRef]
  86. Geetha, P.; Taddesse, P.; Murali, N.; Narayana, P.V.L. Impact of Gd3+ and Nd3+ ions substitution on structural and magnetic properties of Co0.5Ni0.5Fe2O4 ferrite system. J. Indian Chem. Soc. 2022, 99, 100255. [Google Scholar] [CrossRef]
  87. Manner, O.; Sarmah, S.; Patra, K.P.; Maji, D.; Ravi, S.; Bora, T. Effect of rare earth (Ho and Er) co-substitution on the magnetic and dielectric properties of nanocrystalline cobalt ferrites. Ceram. Int. 2024, 50, 30912–30930. [Google Scholar] [CrossRef]
  88. Mohapatra, P.P.; Dobbidi, P. Development of spinel ferrite-based composites for efficient EMI shielding. Mater. Chem. Phys. 2023, 301, 127581. [Google Scholar] [CrossRef]
  89. Cheng, F.; Jia, J.; Xu, Z.; Zhou, B.; Liao, C.; Yan, C.-H.; Chen, L.-Y.; Zhao, H.-B. Microstructure, magnetic and magneto-optical properties of chemical synthe sized Co–RE (RE, Ho, Er, Tm, Yb, Lu) ferrite nanocrystalline films. J. Appl. Phys. 1999, 86, 2727–2732. [Google Scholar] [CrossRef]
  90. Bansal, M.; Ahlawat, S.D.; Singh, A.; Rathee, S.P.; Kumar, V.; Maan, A.; Singh, M. Sol-gel synthesized CoFe2−xGdxO4:SiO2 nanocomposites for structural, thermal and magnetic investigation. Next Mater. 2025, 6, 100470. [Google Scholar] [CrossRef]
  91. Srinivas, C.; Praveen, K.N.; Kumar, E.R.; Singh, S.; Meena, S.S.; Bhatt, P.; Rao, C.T.V.; Sarkar, D.; Arun, B.; Raju, K.C.J.; et al. Microwave absorption properties of rare earth (RE) ions doped Mn–Ni–Zn nanoferrites (RE = Dy, Sm, Ce, Er) to shield electromagnetic interference (EMI) in X-band frequency. Ceram. Int. 2022, 48, 33891–33900. [Google Scholar] [CrossRef]
  92. Kumar, T.M.M.; Kini, H.J.; Praveen, M.; Kumar, M. Electromagnetic interference shielding performance of lanthanum ferrite with MWCNT and graphene in the polyethylene polymer matrix in X-band frequency. Diam. Relat. Mater. 2024, 141, 110701. [Google Scholar] [CrossRef]
  93. Mashadi, Y.; Winatapura, D.S.; Setiawan, J.; Mulyawan, A.; Fakhrudin, M.; Taryana, Y.; Sudrajat, N.; Adi, W.A.; Gunanto, Y.E. Microwave absorbing material of cobalt lanthanum ferrite: The contributions of intrinsic and extrinsic factors on the microwave absorption properties. Appl. Phys. A 2024, 130, 921. [Google Scholar] [CrossRef]
  94. Yunasfi; Dewi, S.H.; Mashadi; Winatapura, D.S.; Setiawan, J.; Mulyawan, A.; Edi Gunanto, Y.; Ari Adi, W. Exploring the structural and magnetic properties of La-doped nickel ferrite for microwave absorbing application. J. Magn. Magn. Mater. 2024, 603, 172267. [Google Scholar] [CrossRef]
  95. Prasad, J.; Singh, K.A.; Haldar, K.K.; Gupta, V.; Singh, K. Electromagnetic interference shielding effectiveness in 3D flower-like MoS2-rGO/gadolinium-doped nanocomposites. J. Alloys Compd. 2019, 788, 861–872. [Google Scholar] [CrossRef]
  96. Das, S.; Banerjee, A.; Pal, P.; Rudra, S.; Nandi, U.; Ghosh, A. Hydrothermally synthesized gadolinium doped molybdenum disulfide for electrochemical supercapacitor applications. J. Energy Storage 2024, 99, 113268. [Google Scholar] [CrossRef]
  97. Li, J.; Li, X.; He, L.; Guo, H.; Xia, W.; Sun, B.; Cao, C.; Sha, L.; Zhou, D. MoS2-based nanocomposites for microwave absorption: A review. ACS Appl. Nano Mater. 2024, 7, 5761–5775. [Google Scholar] [CrossRef]
  98. Hu, R.; He, X.; Luo, Y.; Liu, C.; Liu, S.; Lv, X.; Yan, J.; Peng, Y.; Yuan, M.; Che, R. Biomimetic multi-interface design of raspberry-like absorbent: Gd-doped FeNi3@covalent organic framework derivatives for efficient electromagnetic attenuation. Small Methods 2024, in press. [CrossRef]
  99. Phani, P.S.D.R.; Sahu, S.; Gurrala, R.C.; Dobbidi, P.; Raidongia, K.; Latha, B.S.; Babu, B.K.; Annapurna, N. Leveraging synergistic interfaces in NiO and NiO/rGO heterostructures for enhanced microwave absorption. Surf. Interface 2025, 56, 105617. [Google Scholar] [CrossRef]
  100. Akbar, R.; Shifa, M.S.; Saleem, A.; Zaib, A.; Raheem, F.; Khaliq, M.W. Tuning the properties of praseodymium cobalt-zinc ferrites by substitution of bismuth. J. Mat. Phy. Sci. 2022, 3, 1–13. [Google Scholar] [CrossRef]
  101. Das, D.; Dikshit, A.P.; Samal, R.R.; Parashar, K.; Parashar, S.K.S. A-site Pr-doped BNT ceramics for absorption-dominated EMI shielding in X-band. J. Mater. Sci. Mater. Electron. 2025, 36, 139. [Google Scholar] [CrossRef]
  102. Seethalakshmi, K.; Sakthipandi, K.; Sethuraman, B.; Alhashmi, B.; Venkatesan, K.; Rajkumar, G.; Alqarni, A.S.; Ansari, I.A.; Raghavan, M.S. Investigation of electromagnetic shielding effectiveness and magnetic phase transitions of neodymium-doped Cu0.25Ni0.5Zn0.25Fe2−xNdxO4 nanoferrites. J. Rare Earth, 2025; in press. [Google Scholar] [CrossRef]
  103. Majeed, M.; Akhtar, M.; Khatoon, R.; Amin, N.; Morley, N.; Tung, L.D.; Amami, M.; Abbas, W.; Siddeeg, S.M.; Thanh, N.T.K.; et al. Dielectric and magnetic response of Cu-Co-Sm ferrite impregnated with graphene nanoplatelets for high-frequency device applications. J. Alloys Compd. 2024, 986, 173770. [Google Scholar] [CrossRef]
  104. Kanna, R.R.; Jaisiva, S.; Dhineshbabu, N.R.; Kesavan, M.P.; Lenin, N.; Prasad, L.G.; Kumar, P.R. Microwave-absorbing behavior of rare-earth-ion-doped copper manganese nanoferrites in X-band frequency. Ceram. Int. 2024, 50, 39763–39774. [Google Scholar] [CrossRef]
  105. Kumar, P.V.P.; Suryanarayana, B.; Vemula, V.L.; Rao, D.J.; Uppugalla, S.; Ramakrishna, Y. Effect of rare earth ions (RE = La3+, Sm3+, Nd3+, and Gd3+) substitution on structural, magnetic properties, and dc electrical resistivity of Co0.5Ni0.5Fe2O4 ferrite. Appl. Phys. A 2023, 129, 496. [Google Scholar] [CrossRef]
  106. Deng, Y.; Li, L.; Wang, L.; Wu, N.; Jin, H.; Gao, F.; Zeng, Z. Rare earth Ce-doped W-type barium ferrites for tunable electromagnetic waves absorption performance. Mater. Res. Bull. 2024, 176, 112792. [Google Scholar] [CrossRef]
  107. Yang, Y.; Jin, P.; Li, Y.; Li, S.; Zhu, L. The fabrication of Ce-MOFs with the effective electromagnetic wave absorption performance. Inorg. Chem. Commun. 2024, 168, 112950. [Google Scholar] [CrossRef]
  108. Deng, Y.; Guo, H.; Zhao, C.; Huang, J.; Song, F.; Zeng, X. Rare-earth metal-based CuLa@NC nanorods anchoring Mo-MXene layers for electromagnetic wave absorption. Colloids Surf. A Physicochem. Eng. Asp. 2025, 708, 135964. [Google Scholar] [CrossRef]
  109. Li, N.; Wen, B.; Li, X.; Zuo, A.; Yang, S.; Ding, S.; Yang, G. High-quality ultrathin Gd2O2S nanosheets with oxygen vacancy-decorated rGO for enhanced electromagnetic wave absorption. ACS Appl. Mater. Interfaces 2023, 15, 53891–53901. [Google Scholar] [CrossRef]
  110. Gahlawat, R.; Shukla, R. Enhanced EMI shielding performance of CoFe2−xDyxO4/polypyrrole nanocomposites in Ku-band. Surf. Interface 2025, 56, 105629. [Google Scholar] [CrossRef]
  111. Mondal, D.; Bhattacharya, D.; Mondal, T.; Kundu, M.; Sarkar, S.; Mandal, T.K.; Paul, B.K.; Das, S. Rare earth ion-infused α-MnO2 nano-rods for excellent EMI shielding efficiency: Experimental and theoretical insights. Sustain. Mater. Technol. 2023, 38, e00772. [Google Scholar] [CrossRef]
  112. Zboril, R.; Mashlan, M.; Petridis, D. Iron(III) oxides from thermal processes synthesis, structural and magnetic properties, Mössbauer spectroscopy characterization, and applications. Chem. Mater. 2002, 14, 969–982. [Google Scholar] [CrossRef]
  113. Raming, T.P.; Winnubst, A.J.A.; van Kats, C.M.; Philipse, A.P. The synthesis and magnetic properties of nanosized hematite (α-Fe2O3) particles. J. Colloid Interface Sci. 2002, 249, 346–350. [Google Scholar] [CrossRef]
  114. Morrish, A.H. Morphology and Physical Properties of Gamma Iron Oxide. In Growth and Properties; Freyhardt, H.C., Ed.; Springer: Berlin/Heidelberg, Germany, 1980; pp. 171–197. [Google Scholar] [CrossRef]
  115. Uyama, T.; Mukai, K.; Yamada, I. Facile and low-temperature synthesis of γ-Fe2O3 nanoparticles with thermally stable ferrimagnetism for use in magnetic recording tapes. ACS Appl. Nano Mater. 2020, 3, 10678–10690. [Google Scholar] [CrossRef]
  116. Kelm, K.; Mader, W. Synthesis and structural analysis of ϵ-Fe2O3. Z. Anorg. Allg. Chem. 2005, 631, 2383–2389. [Google Scholar] [CrossRef]
  117. Azadmanjiri, J.; Hojati-Talemi, P.; Simon, G.P.; Suzuki, K.; Selomulya, C. Synthesis and electromagnetic interference shielding properties of iron oxide/polypyrrole nanocomposites. Polym. Sci. Eng. 2011, 51, 247–253. [Google Scholar] [CrossRef]
  118. Gupta, A.; Singh, A.P.; Varshney, S.; Agrawal, N.; Sambyal, P.; Pandey, Y.; Singh, B.P.; Singh, V.N.; Gupta, B.K.; Dhawan, S.K. New insight into the shape-controlled synthesis and microwave shielding properties of iron oxide covered with reduced graphene oxide. RSC Adv. 2014, 4, 62413–62422. [Google Scholar] [CrossRef]
  119. Dhawan, S.K.; Singh, K.; Bakhshi, A.K.; Ohlan, A. Conducting polymer embedded with nanoferrite and titanium dioxide nanoparticles for microwave absorption. Synth. Met. 2009, 159, 2259–2262. [Google Scholar] [CrossRef]
  120. Singh, A.P.; Mishra, M.; Sambyal, P.; Gupta, B.K.; Singh, B.P.; Chandra, A.; Dhawan, S.K. Encapsulation of γ-Fe2O3 decorated reduced graphene oxide in polyaniline core–shell tubes as an exceptional tracker for electromagnetic environmental pollution. J. Mater. Chem. A 2014, 2, 3581–3593. [Google Scholar] [CrossRef]
  121. Rao, B.V.B.; Chengappa, M.; Kale, S.N. Lightweight, flexible and thin Fe3O4-loaded, functionalized multi walled carbon nanotube buckypapers for enhanced X-band electromagnetic interference shielding. Mater. Res. Express 2017, 4, 045012. [Google Scholar] [CrossRef]
  122. Liu, Y.; Song, D.; Wu, C.; Leng, J. EMI shielding performance of nanocomposites with MWCNTs, nanosized Fe3O4 and Fe. Compos. B Eng. 2014, 63, 34–40. [Google Scholar] [CrossRef]
  123. Li, L.-Y.; Li, S.-L.; Shao, Y.; Dou, R.; Yin, B.; Yang, M.-B. PVDF/PS/HDPE/MWCNTs/Fe3O4 nanocomposites: Effective and lightweight electromagnetic interference shielding material through the synergetic effect of MWCNTs and Fe3O4 nanoparticles. Curr. Appl. Phys. 2018, 18, 388–396. [Google Scholar] [CrossRef]
  124. Prasad, J.; Singh, A.K.; Shah, J.; Kotnala, R.K.; Singh, K. Synthesis of MoS2-reduced graphene oxide/Fe3O4 nanocomposite for enhanced electromagnetic interference shielding effectiveness. Mater. Res. Express 2018, 5, 055028. [Google Scholar] [CrossRef]
  125. Sambyal, P.; Dhawan, S.K.; Gairola, P.; Chauhan, S.S.; Gairola, S.P. Synergistic effect of polypyrrole/BST/rGO/Fe3O4 composite for enhanced microwave absorption and EMI shielding in X-Band. Curr. Appl. Phys. 2018, 18, 611–618. [Google Scholar] [CrossRef]
  126. Zhan, Y.; Wang, J.; Zhang, K.; Li, Y.; Meng, Y.; Yan, N.; Wei, W.; Peng, F.; Xia, H. Fabrication of a flexible electromagnetic interference shielding Fe3O4@reduced graphene oxide/natural rubber composite with segregated network. J. Chem. Eng. 2018, 344, 184–193. [Google Scholar] [CrossRef]
  127. Liu, H.; Liang, C.; Chen, J.; Huang, Y.; Cheng, F.; Wen, F.; Xu, B.; Wang, B. Novel 3D network porous graphene nanoplatelets/Fe3O4/epoxy nanocomposites with enhanced electromagnetic interference shielding efficiency. Compos. Sci. Technol. 2019, 169, 103–109. [Google Scholar] [CrossRef]
  128. Huangfu, Y.; Liang, C.; Han, Y.; Qiu, H.; Song, P.; Wang, L.; Kong, J.; Gu, J. Fabrication and investigation on the Fe3O4/thermally annealed graphene aerogel/epoxy electromagnetic interference shielding nanocomposites. Compos. Sci. Technol. 2019, 169, 70–75. [Google Scholar] [CrossRef]
  129. Liu, Y.; Lu, M.; Wu, K.; Yao, S.; Du, X.; Chen, G.; Zhang, Q.; Liang, L.; Lu, M. Anisotropic thermal conductivity and electromagnetic interference shielding of epoxy nanocomposites based on magnetic driving reduced graphene oxide@Fe3O4. Compos. Sci. Technol. 2019, 174, 1–10. [Google Scholar] [CrossRef]
  130. Liang, C.; Song, P.; Ma, A.; Shi, X.; Gu, H.; Wang, L.; Qiu, H.; Kong, J.; Gu, J. Highly oriented three-dimensional structures of Fe3O4 decorated CNTs/reduced graphene oxide foam/epoxy nanocomposites against electromagnetic pollution. Compos. Sci. Technol. 2019, 181, 107683. [Google Scholar] [CrossRef]
  131. Yang, J.; Liao, X.; Li, J.; He, G.; Zhang, Y.; Tang, W.; Wang, G.; Li, G. Light-weight and flexible silicone rubber/MWCNTs/Fe3O4 nanocomposite foams for efficient electromagnetic interference shielding and microwave absorption. Compos. Sci. Technol. 2019, 181, 107670. [Google Scholar] [CrossRef]
  132. Shu, X.; Fang, B.; Wu, W.; Song, Y.; Zhao, Z. Acicular or octahedral Fe3O4/rice husk-based activated carbon composites through graphitization synthesis as superior electromagnetic wave absorbers. Compos.-A Appl. Sci. 2021, 151, 106635. [Google Scholar] [CrossRef]
  133. Shu, X.; Cheng, J.; Fang, B.; Wang, J.; Song, Y.; Lu, W.; Zhao, Z. Morphology-dependent magnetic role of ZIFs in nitrogen-doped MXene as metallic conductor microwave absorber. J. Chem. Eng. 2023, 474, 145817. [Google Scholar] [CrossRef]
  134. Shu, X.; Yan, S.C.; Fang, B.; Song, Y.; Zhao, Z. A 3D multifunctional nitrogen-doped RGO-based aerogel with silver nanowires assisted self-supporting networks for enhanced electromagnetic wave absorption. J. Chem. Eng. 2023, 451, 138825. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of new shielding materials based on biochars, scaffolds, rare earth, and ferrite materials discussed in this review. Part of Figure 1 (lower, left part) is adapted from reference [33].
Figure 1. Schematic representation of new shielding materials based on biochars, scaffolds, rare earth, and ferrite materials discussed in this review. Part of Figure 1 (lower, left part) is adapted from reference [33].
Nanomaterials 15 00541 g001
Figure 2. Different starting materials (Ricinus communis outer shell, bamboo, wood, apple, and quince stillage, olive tree pruning, eucalyptus) in biochar production, at various temperatures, usually in an oxygen-controlled atmosphere, lead to the production of porous materials with domains with graphene and amorphous carbon.
Figure 2. Different starting materials (Ricinus communis outer shell, bamboo, wood, apple, and quince stillage, olive tree pruning, eucalyptus) in biochar production, at various temperatures, usually in an oxygen-controlled atmosphere, lead to the production of porous materials with domains with graphene and amorphous carbon.
Nanomaterials 15 00541 g002
Figure 3. Schematic representation of in situ chemical oxidative polymerization technique. The figure is adapted from reference [110]. Copyright 2025, Elsevier B.V.
Figure 3. Schematic representation of in situ chemical oxidative polymerization technique. The figure is adapted from reference [110]. Copyright 2025, Elsevier B.V.
Nanomaterials 15 00541 g003
Figure 4. Morphology and particle size analysis by FESEM images at different resolutions of (a,b) Gd- and (c,d) Er-doped α-MnO2 samples. The figure is adapted from reference [111]. Copyright 2023, Elsevier B.V.
Figure 4. Morphology and particle size analysis by FESEM images at different resolutions of (a,b) Gd- and (c,d) Er-doped α-MnO2 samples. The figure is adapted from reference [111]. Copyright 2023, Elsevier B.V.
Nanomaterials 15 00541 g004
Figure 5. The EMI shielding efficiency of the Er-doped and Gd-doped α-MnO2 samples according to the (a) total, (b) absorption, (c) reflection, and (d) skin depth variation in the frequency range of 8 to 18 GHz. Schematic representation of (e) impact of hazardous EM waves radiated from various electronic sources on livelihood and demonstration of most possible EMI shielding mechanism to prevent it. The figure is adapted from reference [111]. Copyright 2023, Elsevier B.V.
Figure 5. The EMI shielding efficiency of the Er-doped and Gd-doped α-MnO2 samples according to the (a) total, (b) absorption, (c) reflection, and (d) skin depth variation in the frequency range of 8 to 18 GHz. Schematic representation of (e) impact of hazardous EM waves radiated from various electronic sources on livelihood and demonstration of most possible EMI shielding mechanism to prevent it. The figure is adapted from reference [111]. Copyright 2023, Elsevier B.V.
Nanomaterials 15 00541 g005
Table 1. Summarized results for biochar and scaffolds as EMI shielding materials.
Table 1. Summarized results for biochar and scaffolds as EMI shielding materials.
MaterialReinforcement and Matrix ThicknessEMI Shielding EffectivenessFrequency
Bamboo charcoal/HMWPE/LLDPE composite3 mm, 140 mm48.7 dB1500 MHz[39]
Pine chip/PLA composite0.25 mm>32 dB18–26.5 GHz[4]
MXene aerogel/WPC composite3 mm71.3 dB8.5–12.5 GHz[1]
Gypsum–biochar drywall-like composite>2 mm11.65 ± 1.6 dB, 19.2 ± 5.7 dB, and 19.25 ± 1.8 dB for 10, 20, and 40% w/w biochar contents6 GHz[43]
Drywall panels coated with commercial wood biochar10 mm17 dB
25 dB
1 GHz
18 GHz
[47]
Cement-based/commercial lignin-derived biochar/PVC composite4 mm16 dB5.4–8 GHz[48]
EBC@CNF@MWCNT composite~8 mm>32 dB8–12 GHz[36]
Olive tree-derived biochar/polytetrafluoroethylene composite0.1–0.5 mm39 dB1–3 GHz[49]
Apple and quince biowaste-based biochars0.2 mm15.5 dB8–12 GHz[50]
Ricinus communis outer shell-based biochar/epoxy composite0.15 mm26.5 dB8–12 GHz[37]
18 wt.% lignin-based biochar/cementitious composite4 mm15 dB10 GHz[57]
Cashew shell biochar/carbon fiber-reinforced epoxy resin compositenot reported−48.6 dB18 GHz[58]
Jackfruit rag biochar/waste silk fiber-reinforced vinyl ester compositenot reported31.5 dB
47.25 dB
63 dB
68.25 dB
8 GHz
12 GHz
16 GHz
18 GHz
[59]
Cellulose scaffold/AgNP composite1 mm69.1 dB8.2–12.4 GHz[52]
Carbon scaffold/polyurethane/Fe3O4 NP composite~8 mm32 dB8.2–12.4 GHz[53]
Cellulose-paper-based scaffold/Fe3O4 NP composite1.3 mm1805.9 dB/cm2 g10.3 GHz[54]
3D cellulose scaffold/CNT/MXene composite0.25 mm29.3 dB18–26.5 GHz[55]
Hybrid scaffold coupled with high-quality graphene array/MXene-Co aerogel6 mm72.86 dB8.2–12.4 GHz[56]
3D-printed scaffold/CNTs10 mm35.9 dB8.2–12.4 GHz[51]
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

Marinković, D.; Dorontić, S.; Kepić, D.; Haddadi, K.; Yasir, M.; Nardin, B.; Jovanović, S. New Electromagnetic Interference Shielding Materials: Biochars, Scaffolds, Rare Earth, and Ferrite-Based Materials. Nanomaterials 2025, 15, 541. https://doi.org/10.3390/nano15070541

AMA Style

Marinković D, Dorontić S, Kepić D, Haddadi K, Yasir M, Nardin B, Jovanović S. New Electromagnetic Interference Shielding Materials: Biochars, Scaffolds, Rare Earth, and Ferrite-Based Materials. Nanomaterials. 2025; 15(7):541. https://doi.org/10.3390/nano15070541

Chicago/Turabian Style

Marinković, Dragana, Slađana Dorontić, Dejan Kepić, Kamel Haddadi, Muhammad Yasir, Blaž Nardin, and Svetlana Jovanović. 2025. "New Electromagnetic Interference Shielding Materials: Biochars, Scaffolds, Rare Earth, and Ferrite-Based Materials" Nanomaterials 15, no. 7: 541. https://doi.org/10.3390/nano15070541

APA Style

Marinković, D., Dorontić, S., Kepić, D., Haddadi, K., Yasir, M., Nardin, B., & Jovanović, S. (2025). New Electromagnetic Interference Shielding Materials: Biochars, Scaffolds, Rare Earth, and Ferrite-Based Materials. Nanomaterials, 15(7), 541. https://doi.org/10.3390/nano15070541

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