Chinmay Chandan
Parhi
a,
Avanish Babu
Thirumalasetty
a,
Ajit Raymond
James
b,
Yogesh Kumar
Choukiker
c and
Madhuri
Wuppulluri
*d
aDepartment of Physics, School of Advanced Sciences, Vellore Institute of Technology, Vellore, 632014, Tamil Nadu, India
bCeramic Composite Group, Defense Metallurgical Research Laboratory, Hyderabad, 500058, Telengana, India
cSchool of Electronics Engineering, Vellore Institute of Technology, Vellore, 632014, Tamil Nadu, India
dCenter for Functional Materials, Vellore Institute of Technology, Vellore, 632014, Tamil Nadu, India. E-mail: madhuriw12@hotmail.com
First published on 19th December 2023
The flourishing electronic industries and advanced technologies have created an emerging issue by escalating the level of electromagnetic radiation in the environment. In the current investigation, a sophisticated, cutting-edge magnetoelectric composite material is designed and developed in contrast to its counterparts, and exhibits excellent electromagnetic interference shielding performance. A novel multiferroic material composed of ferroelectric BaZr0.2Ti0.8O3 (BZT) and ferromagnetic Co0.75Ni0.25Fe2O4 (CNF) with graphene oxide decoration is developed through a novel microwave sintering technique. The relative permittivity, AC conductivity, magnetic coercivity, and polarization of the composites are all found to be enhanced by the incorporation of graphene oxide into the composite. A high relative permittivity of more than 50k at 10 kHz is observed in the magnetoelectric composite along with relaxor behaviour at 1.5 wt% graphene oxide reinforcement. The magnetoelectric composite's energy storage capacity has been found to be improved by the addition of GO reinforcement. Impedance studies revealed the Debye relaxation behavior for the BZT–CNF composite and composite with 0.5 wt% GO. The performance of the shielding against electromagnetic interference is evaluated in the X-band frequency range. Furthermore, the BZT–CNF composite has the lowest reflection loss (RL) of −54.9 dB with a maximum total shielding efficiency of 66.8 dB, whereas the magnetoelectric composite with 1.5 wt% GO exhibits improved microwave attenuation with the highest specific shielding efficiency of 75.03 dB cm2 g−1 and the maximum total shielding efficiency of 95.44 dB. The discovered material has excellent microwave absorption and shielding properties, making it very useful for radar and stealth technologies in the military, as well as in the electronics, medical, and electronic industries.
Several nations are developing high-energy electromagnetic weapons that pose a danger to the planet by immensely effecting communication satellites, computer networks, and electrical power systems. Only EMI shielding can mitigate this new kind of pollution, which is a global issue. Thus, electromagnetic interference (EMI) shielding materials, especially the novel ones that can absorb incoming EM signals, are growing rapidly to meet strategic requirements and prevent electromagnetic leakage to protect electronic communication and the environment.8,9 These materials reflect, absorb, or inhibit EM radiation to protect electronic equipment from EM waves for better functioning and durability.10 The emergence of innovative functional materials with enhanced properties boosted the development of microwave absorption and EMI shielding solutions.11–15 EMI shielding materials are essential to electronic devices and equipment, yet difficulties with performance exist. Conventionally, metal shrouds are used for EM shielding, but their weight and corrosion resistance made them unsuitable for many applications.16 In the past, polymer composites, carbon compounds, ferrites, ceramics, and 2D materials have been researched to achieve high shielding effectiveness (SE).17–20 Since 1950, several microwave shielding materials have been investigated and still search for high performance materials is going on.21
It is known that impedance mismatch between interfaces induced by uniform or simple structures makes electromagnetic wave (EMW) dissipation problematic. The shielding effective (SE) of a material relies on its dielectric strength, magnetic permeability, shield thickness, and radiation frequency.22 Magnetic and dielectric losses are the two main absorption mechanisms of EMW absorbers. Dielectric loss needs strong electrical conductivity so that eddy current heats the conductor to attenuate electromagnetic waves in metals, carbon materials, composites, etc.23,24 Magnetic loss needs excellent magnetism; the abrupt change in electromagnetic wave impedance at the interface attenuates electromagnetic waves, in ferrite and transition metals.23 However, microwave-absorbing ferrite nanoparticles alone have a limited bandwidth, low absorptivity, and low impedance matching. High magnetic loss with poor impedance matching reduces microwave absorption of ceramic soft ferrites alone.25,26 Nevertheless, due to its well-known Snoek limit and weak dielectric loss capacity, ferrite's fundamental restriction is its abrupt microwave absorption characteristic deterioration in the gigahertz-frequency range.27 In addition to exceeding the Snoek limit, efforts have been made to hybridize ferrites with dielectrics such as BaTiO3 28 and SiO2.29 Benefiting from the synergistic effect of magnetic and dielectric losses, the high-frequency microwave absorption and shielding properties can be enhanced. Recent studies have shown that absorption-dominated EMI shielding can be achieved through a reasonable design of magnetic/electric gradient structures owing to enhanced multiple absorption/reflection and interfacial polarization losses.22
Magnetoelectric composites provide excellent microwave absorption and shielding due to structural properties, physical properties, ferroelectricity, ferromagnetism, and magnetoelectric coupling.28 BaTiO3 and its derivatives exhibit high permittivity, microwave absorption, and low toxicity, as reported by Saini et al. and Shi et al.30–32 Furthermore, to improve microwave shielding, it has been reported that constructing a large number of interfaces to enhance the interface polarization effect is more conducive to realizing the optimization of EMI SE. Introducing a carbon filler may effectively strengthen the impedance matching characteristics, enhance the interfacial polarization due to its defects and further improve the EMI shielding performance of composites.17 From previous reports, it is established that graphene-based fillers are most often used for microwave shielding owing to their conductivity, high aspect ratios, and skin effects, which result in a high-frequency bandwidth. EM wave absorbing materials, which can be used as fillers to carbon materials, shield the devices more efficiently by converting the incident EM wave energy into thermal energy with the help of electric and/or magnetic losses and dissipating it through the surface.10,16 Graphene-based fillers are predominantly considered as potential candidates for designing microwave shielding due to their conductivity, high aspect ratios, and skin effects, leading to absorption of high-frequency bandwidths. Graphene oxide (GO) due to its superior mechanical, thermal, electrical, chemical, and optical properties with the presence of defect sites and attached functional groups can serve as a low-density, high mechanical strength, and efficient shielding material.10
In the current investigation, a novel multifunctional magnetoelectric composite comprising Co0.75Ni0.25Fe2O4 (CNF) and BaZr0.2Ti0.8O3 (BZT) is developed through hybrid microwave sintering. For the first time, graphene oxide is reinforced into the magnetoelectric composite and investigated the effect of GO on the properties of the magnetoelectric composite. Here, graphene oxide is reinforced into magnetoelectric composites like 0.5BZT + 0.5CNF + $% GO ($ = 0.5, 1, 1.5) to improve microwave absorption and shielding in the X-band frequency range. This composite exhibit multiple noteworthy findings. The composite with GO reinforcement exhibits reduced band gap energy. The study shows that GO decoration to the composite results in improving electromagnetic interference shielding capabilities. The relative permittivity, impedance and conductivity were found to be improved through graphene oxide reinforcement. The polarizability, magnetic anisotropy and magnetic coercivity of the magnetoelectric composite along with capacity to store energy are found to be enhanced via GO decoration. Overall, it is believed that the current investigation could lead the way for the design and development of multifunctional microwave shielding materials suitable for future defence, medical, and communication applications.
Fig. 1 Schematic representation of the synthesis procedure of the GO decorated 0.5CNF–0.5BZT magnetoelectric composite through the novel microwave sintering technique. |
Fig. 2(b) shows the low angle XRD pattern of synthesized pristine graphene oxide, which indicates the characteristic graphene oxide peaks of (001), (002), and (100) at 2θ positions of 12.1°, 24.8° and 42.52°.39,40 In order to understand the evolution of the structural nature of the composite with the increase in graphene oxide concentrations, the XRD profile in the Bragg's angle range from 25–70° is illustrated in Fig. 2(b). Although no observable structural transformation has taken place, the intensity of both phases increases with the introduction of GO. It has been observed that with an increase in GO concentration, the peak intensity of the ferroelectric and ferrite phases slightly decreases. It is found that the diffraction peaks of GO-reinforced CNF–BZT become sharp and undergo peak broadening with an increase in GO concentration in the magnetoelectric composite. This is due to the creation of homogeneous particles scattered along the graphene oxide membrane, and the degree of agglomeration reduces as the concentration of GO in the materials increases. The existence of GO content is assumed to have supplied more surface area for the growth of tiny magneto-electric particles, resulting in a broadening of the XRD peaks.39 However, the XRD pattern of the GO-decorated magnetoelectric composite indicates little shifting of peaks with GO concentration due to the effect of the semiconducting 2D graphene oxide layer over crystalline structures.
The density of 0.5BZT–0.5CNF–$% GO ($ = 0, 0.5, 1, 1.5) is calculated through Archimedes' principle in xylene. The porosity in the composites is estimated with the help of the relative density of the composites.41 It is found that the increase of GO content in the magnetoelectric composite enhances the densification and reduces the porosity in the composites. The average crystallite size, dislocation density, and structural features of GO-reinforced BZT–CNF magnetoelectric composites are estimated by applying the Scherer approach. The physical characteristics of the GO-decorated magnetoelectric composite are summarized in Table 1. It has been found that adding GO to the magnetoelectric matrix induces the crystallites to grow in size. The composite with $ = 0.5 has the largest crystallite size, measuring 54.8 nm. A larger crystallite size after GO reinforcement suggests that crystal growth may be stimulated by the incorporation of GO into the composite. It is hypothesized that the GO with a large surface area serves as a crystal nucleus, stimulating the crystallization process.42 The results demonstrated that the lattice characteristics and structure of the samples were not significantly altered by the addition of GO, suggesting that the GO still exists between the grains, serving as a connecting bridge.
Avg. crystallite size (nm) | Dislocation density (m−2) × 10−9 | Avg. grain size (μm) | Lattice parameters (Å) | Phase (%) | Exp. density (g cm−3) | Relative density (%) | |||
---|---|---|---|---|---|---|---|---|---|
Ferromagnetic phase | Ferroelectric phase | BZT | CNF | ||||||
CNF | 25.5 | 1.53 × 10−3 | — | a = 8.35 | — | — | 100 | 5.65 | 93.03 |
BZT | 32.07 | 9.53 × 10−4 | — | — | a = b = 3.98, c = 3.99 | 100 | — | 5.48 | 89.7 |
$ = 0 | 27.51 | 1.32 × 10−3 | 5.44 | a = 8.37 | a = b = 4.05, c = 4.06 | 50.7 | 49.5 | 4.87 | 95.52 |
$ = 0.5 | 54.87 | 3.32 × 10−4 | 3.45 | a = 8.36 | a = b = 4.03, c = 4.02 | 50.4 | 48.9 | 5.34 | 97.78 |
$ = 1 | 45.14 | 4.90 × 10−4 | 1.11 | a = 8.36 | a = b = 4.03, c = 4.02 | 50.8 | 49.6 | 5.22 | 98.46 |
$ = 1.5 | 42.15 | 5.62 × 10−4 | 1.26 | a = 8.36 | a = b = 4.03, c = 4.02 | 50.6 | 49.2 | 5.30 | 98.79 |
It is observed that as the concentration of GO reinforcement increases, the intensity of the peak decreases due to the layer structure of graphene oxide. The IR spectra of GO-modified magneto-electric composites exhibit a redshift with the increase in GO concentration may be because of interface formation between the magnetoelectric particles and graphene oxide sheet.39 FT-IR analysis indicates the formation of metal oxide composites and vibrational bands exhibited by the composites, indicating the presence of both the ferroelectric (FE) BZT phase and ferromagnetic (FM) CNF phase inside the composite, which was also confirmed by XRD analysis.
(αhν) = B(hν − Eg)n | (1) |
The band gap energy (Eg) for 0.5CNF–0.5BZT is found to be 1.44 eV. It is observed that the band gap energy value of the composite is in between the band gap energies of constituent ferroelectric BZT (3.1 eV) and ferromagnetic CNF (1.2 eV). The reinforcement of GO in the composite results in a decrease in the optical band gap energy value. The reduction in band gap energy is due to the introduction of nano-2D GO into the composite, which has higher excitation energy. This is in good agreement with Kubo's theory.47 With the increase of GO concentration, the band gap energy increases. This is due to the increase in surface charge between the magneto-electric composite and graphene oxide. The band gap value for 0.5CNF–0.5BZT–1% GO is 1.09 eV and then decreases to 0.88 eV for 0.5CNF–0.5BZT–1.5% GO.
Apart from BZT and CNF phases, there is no presence of additional phases due to the absence of diffusion between individual magnetic and electric phases as concluded by the XRD analysis. Here the synthesis and sintering technique of the composite plays an important role in the formation of a highly dense magneto-electric composite. In conventional sintering, it is difficult to get a homogeneous and dense ceramic composite due to the difference in the thermal coefficient and sintering temperature of BZT and CNF.37,48 The grain size of composites is estimated with the help of ImageJ software and represented through histograms illustrated in Fig. 3(a)–(d). The average size of the grains of the composites is found to decrease from 5.4 microns at $ = 0 to 1.11 microns at $ = 1 upon incorporating GO into the magnetoelectric composite. The addition of GO in the composite improves its scattering cross section, which subsequently influences the reduction in grain size.
The hetero-junction interfaces between BZT–CNF and GO might be regarded as a magneto-dielectric junction between magnetoelectric particles and a dielectric GO layer; consequently, it could be considered as a GO membrane coating BZT–CNF particles. The special hetero-junction interfaces, with features including a more extensive array of magneto-dielectric layers, have been considered to be responsible for the wide-band absorption properties and GHz frequency absorption detailed in the following sections.
Dielectric properties at 1 MHz | Conductivity | Impedance (ohm) | Ferroelectric properties | ||||||||
---|---|---|---|---|---|---|---|---|---|---|---|
T m (°C) | ε r | Tanδ | Tanδ at 1 kHz | σ ac (S cm−1) × 10−4 at 1 MHz | A | n | |Z| | P max (μC cm−2) at 8.2 kV cm−1 | P r (μC cm−2) at 8.2 kV cm−1 | E c (kV cm−1) | |
$ = 0 | 621 | 1324 | 0.09 | 2.25 | 0.277 | 4.06 × 10−12 | 1.00 | 6138.35 | 0.96 | 0.81 | 5.86 |
$ = 0.5 | 460 | 1769 | 0.18 | 0.50 | 0.376 | 4.61 × 10−12 | 1.01 | 4009.79 | 1.29 | 0.97 | 5.18 |
$ = 1 | 440 | 2633 | 0.19 | 0.51 | 2.42 | 2.12 × 10−10 | 0.88 | 1150.51 | 3.06 | 2.20 | 4.11 |
$ = 1.5 | 431 | 2751 | 0.23 | 0.51 | 4.51 | 8.5 × 10−10 | 0.83 | 505.41 | 10.46 | 4.45 | 6.46 |
It has been observed at lower frequencies and higher temperatures that improved values of the dielectric constant are obtained. Two different explanations can account for this phenomenon. In the first place, this is due to the space charge polarization effect within the ferroelectric and ferrite phases.50 A charge accumulation at the grain boundaries is caused by the existence of a potential barrier driven by space charge polarization at the grain boundaries, which develops at higher dielectric constant values.51 The ferrite's hopping conduction mechanism is another justification for the enhanced dielectric constant at low frequency and high temperature.52 The mechanism of hopping conduction is triggered by the temperature. Elevated temperatures promote increased electron hopping between Fe3+ and Fe2+. The high value of permittivity is due primarily to the fact that being subjected to an external electric field, the dipoles Fe3+ → Fe2+ align their axes parallel to the electric field. The polarization value is boosted even further by the fact that ferrites have their intrinsic ionic polarization.53 The lower the frequency and the higher the temperature, the greater the effectiveness of these mechanisms. As a result, the dielectric constant steadily rises along with temperature. Once a certain temperature is reached, ε is found to decrease. Higher temperatures cause electrons and ions to vibrate more randomly, which causes the dielectric constant to drop.
The reinforcement of GO in the ceramic composite results in an increment in the permittivity/dielectric constant. However, the Tc or temperature of maximum permittivity (Tm) for composites decreases gradually with an increase in GO concentration. The increase in relative permittivity on the addition of GO to the composite is due to an increase in the average electric field in the ferroelectric matrix and the formation of micro-capacitor structures. The GO sheets operate as conducting electrodes, while the magneto-electric material between two GO sheets acts as an insulator, and the application of a large number of such micro-capacitors results in high capacitance. When an external field is introduced, additional oxygen-containing groups on GO contribute to orientation polarization. This improves interfacial polarization, which increases as GO concentration increases. This leads to an increase in the relative permittivity of the synthesized composites.54 Also, the increase in the density of Fe2+–O dipoles enhances the dipolar polarization in composites and the addition of GO further increases the MWS (Maxwell–Wigner–Sillars polarization). Because the ferroelectric matrix, ferromagnetic materials, and GO have different dielectric constants, a lot of charge carriers build up on these interfaces and strong MWS polarization occurs, which also play a crucial role in the improvement of the relative permittivity of the GO-reinforced magneto-electric composite.55 Introducing GO at 0.5 wt% into the composite influences the composite to exhibit two dielectric relaxations at 1 kHz. For 0.5BZT–0.5CNF–0.5% GO, the degree of dielectric relaxation decreases with an increase in frequency and shows diffused phase transition at a frequency above 1 kHz. The diffused phase transition (Tc) temperature for 0.5BZT–0.5CNF–0.5% GO is in the range of 450 °C to 540 °C. The ε′ vs. T plot of all the composites indicates that the GO concentration in the composites reduces the dielectric relaxation.
It is observed from Fig. 4(a)–(d) that the relaxor behavior exhibited by the magneto-electric ceramics is influenced by increasing graphene oxide (GO) concentration in the composite. The composites CNF–BZT–1% GO and CNF–BZT–1.5% GO have a strong frequency dispersion of the dielectric constant with a peak below and above the maximum temperature (Tm). The Tm for $ = 1 and 1.5 shifts towards higher temperatures across 370 °C to 450 °C and 325 °C to 430 °C respectively with an increasing frequency from 1 kHz to 1 MHz. This suggests that the relaxor dielectric nature of the composite 0.5BZT–0.5CNF–$% GO ($ = 0, 0.5, 1, 1.5) exhibits a large dielectric constant. However, the composite $ = 1.5% exhibits a higher permittivity value in a lower frequency region compared to other composites. This could be due to the presence of excess microcapacitor formation. Hence $ = 1% can be considered as a percolation limit of GO in the 0.5CNF–0.5BZT composite. It has been accepted that the dynamics of the polar nanoregions (PNRs) govern the dielectric features of relaxor ferroelectrics.56 The dynamics of these PNRs are potentially affected by the sample's high concentration of grain boundaries and the related internal stresses. As a result of domain refinement and long-range polar ordering in magnetoelectric composite, the residual crystalline phase and reduction in grain size contribute to improved relaxor behaviour and dielectric characteristics.57
Fig. 4(a)–(d) illustrate the relationship between relative permittivity and temperature at different frequencies. The reduced values with increasing frequency show the polar character of the dielectric materials in all composites. The dielectric constant decreases significantly in the low-frequency region (<10 kHz) and approaches a constant value in the high-frequency region. As a result of the heterogeneity of the CNF–BZT–GO interface in the composites, the permittivity is enhanced in the low-frequency region. Koop's model, based on Maxwell–Wegner interfacial polarization, explains this phenomenon.58 A rise in the dielectric constant is predicted by this model due to the conductivity discrepancy among grains and grain boundaries. Several charge hopping events, in which charges jump from one valence state of the cations found in BZT, CNF, and GO to another, are caused by the conductivity inconsistency between grains and grain boundaries in the composites. Permittivity remains constant over a range of high frequencies because dipoles cannot change their orientation in response to an external electric field.
To understand the nature of the inherent dissipation of electromagnetic energy of the synthesized composites, analysis of dielectric loss/tanδ is performed as a function of temperature at different frequencies. As the frequency increases from 1 kHz to 1 MHz, the tanδ peaks relocate in the direction of greater temperatures, demonstrating that the relaxation mechanism is engaged by heat.59 The composites have a strong frequency dispersion of dielectric loss similar to the dielectric constant. The dielectric loss decreases with an increase in frequency owing to ion mobility and direct current conduction in the composites. At lower frequencies, the space charge effect as well as the frictional force between space charges and the interfaces is attributed to the high dielectric loss.38 It is observed that up to 1 kHz frequency, the dielectric loss decreases with an increase in the percentage of GO in the composite owing to the activity of the Maxwell–Wegner relaxation at a lower frequency. At 1 MHz, the dielectric loss/tanδ increases with an increase in graphene oxide concentration in the magneto-electric composites. This observation can be explained based on the influence of Debye relaxation from the orientation polarization of the dipoles at higher frequencies. In general, dipole relaxation induces dielectric loss; however, in this study, the creation of a conducting network plays a major role in the increase of dielectric loss with relaxation. The high conductivity of GO sheets improves charge mobility inside the composite and allows charge flow through it when an electric field is applied. The flow of charge transforms electrical energy to thermal energy, and as a result of local thermal runaway, composite loss increases.60 This is the cause of the rise in dielectric loss when GO concentration increases in the synthesized composite. The high dielectric constant and enhanced dielectric loss along with relaxor behaviour prove the suitability of the synthesized composites for various applications including EMI shielding.
The ferroelectric examination demonstrates that increasing graphene oxide content improves the polarization in magnetoelectric composites. The high polarizability dipoles produced by the GO's numerous defects and surface functional groups may contribute to the composite's enhanced polarization.65 The addition of GO into the magnetoelectric composite results in an enhancement in maximum polarization (Pm) and a decline in the coercive field (Ec). The efficiency of energy dissipation rises in tandem with the fraction of graphene oxide used. Energy dissipation causes the separation of the charge and voltage signals, which results in loops with a substantial area of the curve.66 The charge storage capability of a material is defined by the area of the curve; the greater the area of the curve, the greater the charge storage capability of the material.67 The rise in remnant polarization (Pr) induced by the inclusion of the graphene oxide filler is due to hetero-polarization caused by the interaction of conducting fillers and ceramic composites. The Pr increment with GO content can be due to the larger number of switching dipoles in the composite. As the GO content increases, more charge density accumulates on the surface of GO, resulting in a larger number of switching dipoles. The decrease in the coercive field up to $ = 1% of GO as an additive in the composite is due to the space charge effect at the interface.68 It is observed that at $ = 1.5, the magneto-electric composite exhibit a large jump in maximum polarization and also exhibits improvement in the coercive field (Ec) as compared to other GO decorated magneto-electric composites. This can be attributed to the percolation limit of the 0.5BZT–0.5CNF–$% GO composite; $ = 1 as explained earlier. These results indicate that graphene oxide fillers are effective in enhancing the polarizing properties of the magneto-electric composite.
The energy storage capacity of the magnetoelectric composite is estimated through the polarization as a function of applied electric field curve. Although the composite is designed and developed for high frequency applications such as EMC/EMI shielding, the effect of graphene oxide (GO) reinforcement on energy storage capacity is worth studying. The energy loss (Wloss) in composites is represented through the closed area inside P–E curve where the recoverable energy (Wrec) can be determined through the area between Pmax to Pr and Y-axis. The recoverable energy and its storage efficiency (η) in the magnetoelectric composite are estimated through eqn (2) and (3) (ref. 33) at an applied electric field of 7.4 kV cm−1. The magneto-electric 0.5BZT–0.5CNF composite's recovered energy value is shown to grow with the addition of GO, and its storage efficiency also improves, rising from 5.1 percent for $ = 0 to 9.77 percent for $ = 0.5 and 11.11% for $ = 1. The increase in storage capacity can be attributed to the formation of multiple micro-capacitors and a magneto-dielectric interface upon GO reinforcement. For the composite with GO decoration above $ = 1, the efficiency of energy storage was found to be decrease to 4.84%. This type of disparity may be due to the formation of a conducting percolative pathway across the $ = 1.5 composite. Table S1 in the ESI† lists the energy storage parameters for the GO reinforced magnetoelectric composites. Owing to its ferroelectric attributes, the magnetoelectric composite can be used for harnessing the energy storage capabilities. The efficiency of the energy storage of the composite can be improved by tuning the FM% in the FE matrix of the BZT–CNF composite. The storage efficiency can be enhanced further by GO reinforcement up to $ = 1.
(2) |
(3) |
The leakage current density characteristics have been measured for CNF–BZT magnetoelectric ceramics incorporated with GO. Fig. 5(e) depicts the measured current density (J) against the electric field (E) for 0.5BZT–0.5CNF–$% GO at room temperature and in a field of 0 to ±5 kV cm−1. The current density attributes of the composites are examined under both forward and reverse bias conditions and found to exhibit symmetry under both positive and negative electric field conditions. The leakage current of 0.5CNF–0.5BZT composites can be significantly influenced by the incorporation of GO. Fig. 5(e) demonstrates that as GO content increases in magnetoelectric composites, the leakage current density rises from 3.59 × 10−5 A cm−2 at $ = 0 to 2.67 × 10−4 A cm−2 at $ = 1.5. From the observation, it can be hypothesized that the leakage current density was shown to increase with GO concentration in composites provided that the added amount of GO was greater than or equivalent to 1 wt%. The inclusion of GO as a connecting bridge weakens the insulation between the grains and the dielectric breakdown strength of the magnetoelectric composite, resulting in an increase in leakage current.
The electric conduction mechanism of the ceramic has been represented by plotting the logarithmic relationship between the leakage current J and the electric field E as shown in Fig. 5(f). Previous efforts to explain the conduction mechanism in ceramics have presented two distinct models: the ohmic conduction model and the space charge limited current model.33 It is established that the current density follows a power law distribution i.e., J ∝ En. In the case of n = 1, conduction unfolds through the ohmic conduction mechanism; nevertheless, with a value of n = 2, the mechanism transitions to the space charge limited current.69 Fitted n values are indicated in Fig. 5(f). Slope n with values of 0.97, 0.97, 1.03, and 1.11 for 0.5BZT–0.5CNF–$% GO ($ = 0, 0.5, 1, and 1.5) reveals that ohmic conduction is the only mechanism responsible for transport and leakage in these composites. For the ohmic conduction process, it is well known that current density J = eμNeE, where e is the electron's charge, μ is a free electron's mobility, Ne denotes the density of electrons, and E is the applied electric field.70 Since the leakage current density is proportional to the concentration of free electron–hole pairs (Ne), decreasing the optical band gap energy dramatically increases Ne.71 As was discussed in the prior section, the optical bandgap energy of 0.5CNF–0.5BZT was found to decrease with increasing GO concentration. Thus an increase in GO content leads to an increase in leakage current density. Both intrinsic and extrinsic sources contribute to the overall leakage current density of the composites.72 The increasing GO content in the composite alters the intrinsic properties like the band gap of the composite resulting in an increment in leakage current density. The increasing n value with an increase in GO content is due to extrinsic properties like defects in the composites. High leakage current will influence the improvement of the EMI shielding capabilities of the magnetoelectric composites.
(4) |
The AC conductivity values are obtained from the relationship, σac = ωε′ε0tanδ, where ω is the angular frequency, ε′ is the real component of the dielectric constant, ε0 is the permittivity of vacuum, and tanδ is the dielectric loss. The frequency-dependent variation for composites with different GO concentrations is depicted in Fig. 6(a). The conductivity of the sample is increased by the incorporation of GO, with the increase being modest up to 0.5% GO in the BZT–CNF composite. Nevertheless, it was discovered that the conductivity of magnetoelectric composites significantly improved at 1.5% GO. The BZT–CNF–GO composite's conductive properties come from a combination of ohmic and tunneling conduction. Enhanced conductivity above 1% can be attributed to the same phenomenon of ohmic conduction occurring owing to direct contact of GO sheets when concentration is above the criterion. Below the threshold, when an external field is applied, charges tunnel from one GO sheet to another across the magneto-electric composites.
The composite's electrical conductivity is found to increase with frequency at room temperature. Fig. 6(a) depicts two distinct features of AC conductivity: (i) low-frequency upland and (ii) high-frequency diffusion with a gradual slope distribution. The term “hopping frequency” is commonly used to describe the varying frequency slope. Funke uses the jump-relaxation model73 to explain the spread of conductivity with frequency. This model predicts that at low frequencies, electrons engage in a long sequence of successful hopping motions, which ultimately results in DC conductivity that is independent of frequency. As per Funke, backward hopping coexists with successful ones, just as they do in the field; the dissipative conductivity at high frequencies is a function of the ratio of these two types of hopping, as well as the relaxation of adjacent charged particles. Vault charge particles, which are distinct from free-charged particles, and confinement of charge particles in defects can be linked to the frequency-dependent rise in conductivity. At low frequencies, the hopping process is suppressed because fewer conducting grain boundaries are sandwiched between more conducting grains, as predicted using the Maxwell–Wagner bi-layer model. The hopping mechanism is strengthened by the increased activity of grains at higher frequencies. So the conductivity of ac power rises as the frequency rises.74Eqn (5) represents the universal power law, which provides insight into the subservience of conductivity.75
σtot = Aωn + σ0 | (5) |
A complex impedance study is performed to understand the charge transport properties of 0.5BZT–0.5CNF–$% GO ($ = 0, 0.5, 1 and 1.5) composites. Fig. 6(f) illustrates the Bode plots and real and imaginary parts of impedance as a function of frequency at room temperature. At room temperature, the magnitude of the impedance Z′ is maximum in the low-frequency range and least in the high-frequency range. Due to the reduction in the barrier properties of composites caused by the release of space charge, they combine at higher frequencies. Increases in conductivity due to immobile charges at low frequency and defects at high frequency likely cause a steady drop in the impedance Z′ with increasing frequency. Furthermore, it demonstrates that the impedance Z′ of magnetoelectric composite materials at low frequencies reduces as the concentration of GO increases. Since a smaller Z′ indicates less grain clustering and a smaller potential barrier between grains, less resistance is encountered by charge carriers.68 It can be shown in Fig. 6(b) that at lower frequencies the imaginary impedance Z′′ has different values due to differing compositions, while at higher frequencies it converges, possibly because of a lack of the space-charge effect. It also demonstrates that the impedance Z′′ reaches its maximum value, which contradicts the system's tendency to relax. As relaxation time and vacancies diminish, the frequency at which the impedance Z′′ reaches its maximum value shifts upwards. Increases in frequency cause a decrease in polarization due to the accumulation of space charges in the material, which does not need a longer time to relax at higher frequencies.51 Additionally, the peak in the imaginary part of complex impedance (Bode plots) for the composite $ = 0, 0.5 suggests an activation-mediated conduction mechanism.79
Fig. 6(g) represents the Nyquist plot for the magneto-electric composites at room temperature. The advantage of the Nyquist plot is that it reveals potential mechanisms influencing phenomena by highlighting the distinctive impedance arc and the form of the curve produced by an activation-controlled process with a measurable time constant.80 Graphene oxide's (GO) electrical transport capabilities are largely determined by its oxidation state and chemical composition, both of which can be modified by removing or adding oxygen groups to control the ratio of sp2 to sp3 carbon.81 It is well known that the semiconducting properties of GO are primarily observed in the bulk form, although behaviour may depend on the level of oxidation of GO sheets.82 From the Nyquist plot, it is found that composites with $ = 0 and $ = 0.5 exhibit two semicircles in the impedance plot indicating the semiconducting nature of the grains and the insulating nature of the grain boundary. It is evident that for $ = 0 and $ = 0.5, the grain resistance (Rg) increases and the grain boundary resistance (Rgb) decreases with an increase in GO concentration in the magnetoelectric composite. These composites exhibit Debye relaxation behaviour across the measured frequency range, where two semicircular arcs have been noticed, one smaller at lower frequencies and one larger at higher frequencies, demonstrating the involvement of two distinct relaxation processes in the composite. The magnetoelectric composite $ = 0 and $ = 0.5's Nyquist plot and impedance analysis supported the composite's feasibility for electrochemical applications.
The composite with $ = 1 and $ = 1.5 displayed one arc at a lower frequency, indicating that just one relaxation process is operational. In the composite with $ = 1 and 1.5, the appearance of a semicircle that is both asymmetric and depressed suggests the existence of non-Debye relaxation in the composites, as the relaxation duration are dispersed across the composite. A reduction in resistance as a consequence of increasing GO concentration accounts for the presence of a single semicircular arc in $ = 1 and 1.5. It's believed that the rapid increase in resistance owing to grains observed at $ = 1 is caused by the percolation limit of GO in the 0.5CNF–0.5BZT composite. A possible explanation for the occurrence of non-Debye relaxation at $ = 1 and $ = 1.5 is the large grain size of the composite. This may be linked to the increased electron scattering caused by the incorporation of larger grains into the GO matrix.79 Increased composite resistance points to 0.5BZT–0.5CNF particles embedded between or on the surface of graphene oxide (GO) sheets. As resistance drops and conductivity rises, the semicircle's radius decreases, which is be evident from the Nyquist plot of composites. This further validates the role that grain and grain boundary resistance play in the composites' intrinsic conduction process.
For all the composites, the loop is complete and well saturated with a small opening of the hysteresis curve, which signifies the soft magnetic behavior of the synthesized material. The soft magnetic nature of the composites makes them suitable for multiple industrial applications. To understand the magnetic attributes and impact of GO addition to the composites, the magnetic parameters like the anisotropy constant (K), experimental magnetic moment (ηB) value, and squareness value (R) are estimated and tabulated in Table 3. The squareness values (R) of the composites are measured via the relationship83 and the R-value in the 0.5BZT–0.5CNF–$% GO ($ = 0, 0.5, 1, 1.5) composite is less than 0.5, which signifies the presence of multi-phase domains in the composite.84 It is noticed that the K value initially decreases with the decoration of GO at 0.5 wt% in 0.5BZT–0.5CNF. However, a further increase in GO concentration in the composite leads to an increment in the magnetic anisotropy constant and the diamagnetic behaviour of GO reduced the value for experimental magnetic moment per unit formula in the Bohr magnetron of the composite as shown in Table 3.
Magnetic properties | Electromagnetic shielding (dB) | |||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|
M s (emu g−1) | M r (emu g−1) | H c (Oe) | M r/Ms | K (emu Oe g−1) × 102 | η B | RLmin | SEmax | SEr max | SEabs max | SEtotal max | SSEmax | |
$ = 0 | 35.82 | 9.86 | 268.08 | 0.27 | 100.05 | 1.52 | −54.9 | 38.50 | 20.55 | 64.75 | 77.01 | 66.84 |
$ = 0.5 | 33.19 | 9.45 | 272.81 | 0.28 | 94.32 | 1.41 | −46.9 | 43 | 17.3 | 84.43 | 85.99 | 67.60 |
$ = 1 | 33.89 | 10.95 | 345.75 | 0.32 | 122.08 | 1.44 | −42.1 | 41.97 | 17.31 | 80.94 | 83.95 | 67.26 |
$ = 1.5 | 33.46 | 10.70 | 351.26 | 0.31 | 122.44 | 1.42 | −33.2 | 47.72 | 24.54 | 95.09 | 95.44 | 75.03 |
The prepared ceramic composite BZT–CNF is a mixture of the multiferroic FE phase of BZT and the FM phase of CNF. Furthermore, the coupling of two order parameters enables multiferroic composites to demonstrate magnetism well within domain wall ordering when in the FE phase.85 However, the magnetic behaviour of the composites is dominated and mostly contributed by the magnetic ordering of the CNF phase. It is observed that upon the introduction of the FM phase into the FE matrix in a particulate manner, the saturated magnetization decreased whereas the coercive field increased. This resulted due to the pinning of magnetic spins by the ferroelectric domains in the magneto-electric composites. For magnetic spins, the ferroelectric domain boundaries act as pins, and the spin density fluctuates in the area around ferroelectric walls.86 This signifies the magnetic loss in the prepared ceramic composite, which can be exploited for electromagnetic interference shielding applications.
Fig. 6(i) displays the variation of saturation magnetization and the coercive field as a function of GO content in the 0.5BZT–0.5CNF composite. The introduction of GO into the magneto-electric composite results in further decreases in saturation magnetization. The difference in saturation magnetization (Ms) in GO-reinforced composites is due to the paramagnetic nature of GO. Since GO functions as a composite membrane, it is evident that when GO concentration rises, the composite's magnetic moment will also decline.87 The anonymous little increase in Ms for the composite with 1% GO may be due to the effect of critical concentration of GO reinforcement on the composite. The coercivity (Hc) of the composite increases with GO concentration. With 1% GO in the composite, there is a large change in the Hc value, which corresponds to the critical concentration of GO, which is also observed and explained in the previous characterization. However, there are no significant changes in the coercive field for the GO-reinforced magneto-electric composite, thus indicating the soft magnetic nature of microwave-sintered magnetoelectric composites. The GO concentration in the BZT–CNF composite increases the magnetic loss, thus making this prepared ceramic composite a suitable candidate for EMI/EMC applications.
RL = 10log10(R) | (6) |
SE = 10log10(T) | (7) |
SER = −10log10(1 − R) | (8) |
(9) |
The intrinsic shielding performance of a material can be determined through the specific shielding efficiency (SSE) per unit thickness of the material, which is provided as16
(10) |
To evaluate the efficacy of 0.5BZT–0.5CNF–$% GO ($ = 0, 0.5, 1, 1.5) magneto-electric composites to attenuate electromagnetic interference, the S parameters are measured in the X band range (8.2–12.4 GHz) at a sample thickness of 2.4 mm. Fig. 7(a) illustrates the reflection loss and shielding efficiency of GO-reinforced magnetoelectric composites in the X-band range. For $ = 0, two RL peaks are obtained at 9.68 GHz and 11.6 GHz and the lowest value of reflection loss, −54.9 dB, occurs at a frequency of 9.68 GHz. For 0.5BZT–0.5CNF, the maximum total shielding efficiency due to absorption and reflection is found to be 77.01 dB at 8.54 GHz frequency and an SSE value of 66.84 dB cm2 g−1 is achieved. Table 3 depicts the shielding parameter of 0.5BZT–0.5CNF–$% GO ($ = 0, 0.5, 1, 1.5) composites. The microwave absorption capabilities of the ferrite-ferroelectric composite are improved by combining the magnetic properties of CNF with high-permittivity BZT compared to microwave attenuation of individual phases. An apparent dielectric and magnetic interface can generate the magneto-dielectric hetero-junction, which could serve as a focal point for microwave attenuation.89
To be an effective shielding layer, a material ought to exhibit the impedance matching condition, necessitating that the magnetic permeability value of the shielding material is close to its electric permittivity value. Impedance mismatch can be mitigated and EM radiation absorption can be enhanced by incorporating a suitable filler (graphene oxide for the present investigation) for the ferromagnetic and ferroelectric matrix into the magnetoelectric composites.90Table 3 and Fig. 7(d) and (e) shows that the $ = 1.5 composite exhibited the highest SET, SEabs, and SSE among all the composites. Incorporating GO into the composites enhanced the microwave absorption attributes, evidenced in the table and the figure. Similar observations were reported previously by Sumit et al.16 and Samadi et al.91 The maximum RL for the 0.5CNF–0.5BZT composite moves to a higher frequency range of 11.5 GHz after GO is introduced, from a frequency of 9.68 GHz for the magnetoelectric composite. As discussed in the microstructural study, the composite's dispersibility improves with GO reinforcement, resulting in smaller particle sizes equally dispersed on the GO layers. As a result of this enhanced dispersibility, it is hypothesized that the magneto-dielectric heterostructure interfaces between CNF and BZT, which serve as the microwave absorption center, occur more frequently.39 Microwave absorption and the fractional bandwidth are improved by increasing the surface area and compactness brought about by the creation of hetero-structural connections, which in turn result from an increase in the number of interfaces.
The improvement in shielding efficiency is attributed to the creation of a hybrid conducting network inside the composite materials, which is a necessity for effective EMI shielding. As was previously explained, the AC conductivity of magneto-electric composites was enhanced by the inclusion of graphene oxide. It, therefore, indicates that the hybrid composite's electrical conductivity can be improved by the incorporation of GO. Notably, two-dimensional graphene oxide has interwoven to create a continuous percolated network, which in turn built hierarchical structures to increase the number of channels that electrons can travel through with high conductivity.92 For electromagnetic interference shielding the introduction of GO increases the number of free electrons, functional groups, and defects available to interact with the incident EM radiation, leading to the formation of electric dipoles that dissipate the energy through relaxation losses. Subsequently, the addition of GO generates a rise in interfacial relaxation losses due to an increase in electron hopping across the conductive network of GO, thereby turning more EM energy into heat.
In the magnetoelectric composite, the ferrite particles serve as magnetization centers, while conductive graphene oxide (GO) networks and the ferroelectric BZT matrix serve as polarization centers, allowing the hybrid structure to trap maximum incident electromagnetic radiations within the multiple heterogeneous interfacial regions. The magnetoelectric particles act as electron migration/hoping bridges, causing tiny current networks to form and increasing the conduction losses within the material.93 Micro-currents are generated as a result of the interaction of electromagnetic waves in these network structures, and the resulting heat dissipates the electromagnetic energy. Within the arranged, interconnected network structure of composite with graphene oxide, electromagnetic radiations undergo absorption and multiple internal scattering.
The results reflect that the GO-reinforced composite shield has increased electric and magnetic losses, leading to a high absorption efficiency for incident EM waves. The best total shielding efficiency of 95.44 dB at 11.26 GHz is exhibited in magnetoelectric composites with $ = 1.5, and the highest specific shielding efficiency of 75.03 dB cm2 g−1 is evident in these materials as well. The efficiency of magnetoelectric composite materials as an EW shield is shown in Table 4, together with existing results given in the literature; these results show that the hybrid composite networks defined in this study improve shielding effectiveness. In conclusion, the EMI shielding material suggested by the composite samples proves to be highly effective and chemically stable.
Synthesis conditions | σ ac (S cm−1) | M s (emu g−1) | Frequency (GHz) | SE (dB) | Ref. | |
---|---|---|---|---|---|---|
NiFe2O4–BaTiO3–AP | High energy milling | — | 15.2 | 16.8 | RLmin = −39.8 | 28 |
BaFe12O19–CQD@BaTiO3 | Sol–gel auto-combustion | — | — | 10.21 | RLmin = −25.3 | 94 |
BaTiO3–CoFe2O4–LDPE | Co-precipitation and melt mixing | — | — | 9.4 | SEt = 7.9 | 95 |
NiCoZnFe2O4–BaTiO3 fiber | Electrospinning with sol–gel | — | 34.5 | 8.2 | RLmin = −30.7 | 96 |
Y–Ag–BiFeO3 | Sol–gel auto-combustion | — | 0.98 | 8.36 | SE = 23.05 | 97 |
SrFe12O19/CoFe2O4/NiO | Solvothermal | 240 S m−1 at 12 GHz | 24.6 | 10.7 | RLmin = −36 | 98 |
PVDF–MWCNT | Film casting-compression molding | 75.7 S m−1 | — | 34 | SEt = 99.3 | 99 |
Ni–MOF–rGO aerogel | Facile solvothermal | — | 36 | 17.52 | RLmin = −51.9 | 100 |
CFF/Ag/PDMS | Electroless plating | 540.9 | — | 8.2–12.4 | SE = 105 | 101 |
NiFe2O4–BaTiO3 | Solid state reaction | — | — | 8–12 | SE > 34 | 102 |
TOCNF/Ti3C2Tx/AgNW | Vacuum assisted filtration | 1.2 × 107 S m−1 | — | 8.2–12.4 | SE = 45.57 | 92 |
MXene/polyaramid | Hydrogen-bonding induced self-assembly | — | — | 8.2–12.4 | SE = 38.9 | 103 |
MXene/Co2–Ba hexaferrite/PVDF | Electrospinning-microwave treatment | — | — | 11.9 | RLmin = −32 | 104 |
Phosphate-bonded CFO–BTO | Phosphate bonded ceramic approach | — | — | 45 | SE = 40 | 105 |
Nb2O5–N doped CNF | Electrospinning | 1.47 | — | 8.2–26.5 | SE = 67 | 106 |
ZnO/PVDF/MXene | Hydrothermal-electrospinning | — | — | 8–12 | SE = 61 | 21 |
CoNiFe2O4–BaZrTiO3–GO | Microwave sintering | 4.51 × 10−4 at 1 MHz | 35.8–33.46 | 11.24 | $ = 1.5; SEt max = 95.44 | Present work |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ta06292a |
This journal is © The Royal Society of Chemistry 2024 |