Open Access Article
R. Suresha
a,
H. K. Sachidanandab,
B. Shivamurthy*c,
Gibin Georged and
Sampath Parasurame
aSchool of Engineering & Information Technology, Department of Electrical & Electronics Engineering, Manipal Academy of Higher Education, Dubai, United Arab Emirates. E-mail: suresha@manipaldubai.com
bSchool of Engineering & Information Technology, Department of Mechanical Engineering, Manipal Academy of Higher Education, Dubai, United Arab Emirates. E-mail: sachidananda@manipaldubai.com
cDepartment of Mechanical & Industrial Engineering, Manipal Institute of Technology, Manipal Academy of Higher Education, Manipal-576104, India. E-mail: shiva.b@manipal.edu
dDepartment of Mechanical Engineering, SCMS School of Engineering and Technology, Karukutty, Kerala, India. E-mail: gibingeorge@scmsgroup.org
eDepartment of Materials Engineering, Indian Institute of Science, Bangalore, India. E-mail: parasurams@iisc.ac.in
First published on 22nd October 2025
This study investigates the structural and electromagnetic interference (EMI) shielding performance of bidirectional carbon fabric/epoxy composites reinforced with Fe2O3 nanoparticles for aerospace applications. Composites with varying Fe2O3 loadings (1–3 wt%) were fabricated using a hand lay-up method, followed by mechanical and electromagnetic shielding characterisation in the X-band frequency range (8.2–12.4 GHz). Tensile testing revealed that 1–2 wt% Fe2O3 loading enhanced the stiffness and tensile strength of the composite due to improved fibre–matrix interfacial bonding, while 3 wt% caused agglomeration and reduced strength. EMI shielding measurements showed absorption-dominated performance across all configurations, with multiple layers significantly improving total shielding effectiveness (SET). The highest SET (25.3 dB) was achieved for a two-layer laminate with 3 wt% Fe2O3, attributed to synergistic dielectric and magnetic losses and enhanced internal reflections. The results demonstrate that optimised Fe2O3 content and laminate layering can deliver lightweight, structurally robust composites with effective EMI shielding, making them suitable for advanced aerospace structures requiring mechanical integrity and electromagnetic compatibility. These findings highlight that optimized nanoparticle loading, and laminate architecture can yield lightweight composites that unite mechanical robustness with effective EMI shielding, offering strong potential for aerospace structures demanding both structural performance and electromagnetic compatibility.
In addition to onboard sources, aircraft encounter external EMI from thunderstorms, electrostatic discharges, and solar activity, making shielding essential for flight safety.2 At the same time, the drive for lightweight, high-performance structures has accelerated the use of carbon fibre-reinforced epoxy (CF/E) composites, which combine high specific strength, fatigue resistance, thermal stability, and corrosion resistance, positioning them as attractive alternatives to traditional metals in aerospace applications.3
Although carbon fiber/epoxy composites provide excellent structural performance, their semiconducting nature limits intrinsic EMI shielding. Consequently, research increasingly targets enhancing their effectiveness in the 8–12 GHz X-band frequency, a range vital for radar, satellite communication, navigation, and wireless networks, where reliable EMI protection is essential for uninterrupted operation.4
Bidirectional carbon fabric/epoxy composites offer superior EMI shielding compared to unidirectional ones, but excessive nanofiller addition can cause agglomeration and degrade mechanical strength, underscoring the need for optimized loading.5 Recent studies emphasize strategies such as incorporating magnetic nanofillers, forming conductive networks, and designing multilayer structures to enhance shielding effectiveness in the X-band (8.2–12.4 GHz).
Ahmad et al.6 showed that alternating carbon fiber plies reinforced with MWCNTs and Fe2O3 enhanced EMI shielding via combined dielectric and magnetic losses. Waseem et al.7 developed a hybrid layered composite using ZrB2-SiC, Fe3O4-loaded carbon fabric, and carbonized cotton fibre, achieving lightweight, thermally stable, absorption-dominated shielding. Shao et al.8 reported a Janus-structured CNT/Fe3O4/MXene membrane (∼84.9 μm) with 44.56 dB shielding in the X-band, along with excellent thermal conductivity, flexibility, and low weight. Likewise, Yu et al.9 created a nacre-inspired Fe3O4/CNT composite with alternating conductive and magnetic layers, delivering durable, absorption-dominant shielding that retained 96% efficiency after physical damage.
Zhang et al.10 engineered ultralight rGO–CNT–epoxy aerogels with Fe3O4 nanoparticles, achieving a reflection loss of −58.13 dB at 12.08 GHz through a 3D porous structure that enhanced scattering and impedance matching. Wilson et al.11 reported TPU laminates with graphite and CoFe2O4, where the G/F/G configuration reached approximately 54 dB shielding from synergistic dielectric and magnetic losses. Durmaz et al.12 developed bio-based PA11/PLA composites with 30 wt% carbon fibres, yielding 28 dB SE at 10 GHz through reflection-dominated shielding. Ahmad et al.13 demonstrated graphene-Fe2O3 polymer composites with broadband EMI attenuation from combined dielectric reflection and magnetic absorption. Duan et al.14 fabricated CF/GO/Fe3O4/epoxy multilayers that improved impedance matching and promoted absorption-based shielding. Fallah et al.15 achieved 36.6 dB SE at 8.2 GHz in epoxy nanocomposites with Fe2O3 and carbon black, showing tunable frequency response. Bhaskaran et al.16 developed Fe3O4@GNP hybrids with ∼9.6 dB SE at 1 mm, outperforming individual fillers through synergistic effects. Xu et al.17 produced epoxy-cotton fibre scaffolds with GNPs and Fe3O4, delivering 33.1 dB SET and absorption-dominant shielding with structural stability for defence and electronics.
Veeramani et al.18 studied carbon fiber/epoxy composites fabricated via vacuum-assisted resin infusion, incorporating polyphenylene ether (PPE) as a toughening agent. PPE dispersed in the interlaminar region enhanced toughness, with DMA results showing a notable increase in glass transition temperature. In another work, Veeramani et al.19 developed self-healing composites by embedding urea-formaldehyde-encapsulated dicyclopentadiene (DCPD) microcapsules with epoxy, chopped CF, and CNTs. These systems enabled self-detection and prevention of microcrack propagation, with potential applications in aerospace, wind energy, and marine structures. Eken et al.20 investigated two-layer CFRP composites reinforced with hematite (Fe2O3) and goethite (FeO(OH)) produced by hand lay-up, evaluating EMI shielding across 700–6000 MHz and impact resistance. They reported overall improvements in both shielding effectiveness and impact strength compared to unreinforced samples.
Fe2O3 was selected in this study due to its lower cost, environmental benignity, chemical stability, and strong magnetic loss response in the X-band compared to fillers like Fe3O4 and CoFe2O4. Unlike earlier works that broadly combined carbon-based conductive and magnetic fillers, the novelty of this research lies in optimizing Fe2O3 loading within carbon fabric/epoxy laminates to simultaneously balance structural performance and absorption-dominated EMI shielding. The exclusive focus on the X-band is justified by its critical importance in aerospace applications, particularly radar, satellite communication, and navigation systems, where reliable EMI protection is essential for operational safety.
Previous studies on conductive and magnetic nanofillers have advanced EMI shielding but often overlooked structural robustness. Next-generation aerospace, defence, and wearable systems require materials that are both lightweight and multifunctional. This study addresses that gap by investigating Fe2O3-reinforced carbon fabric/epoxy composites, linking shielding effectiveness with mechanical performance. By integrating structural strength with absorption-dominated shielding, the work proposes a multifunctional composite system that meets aerospace demands for both mechanical integrity and electromagnetic compatibility, thereby contributing to the design of advanced materials capable of balancing performance requirements often neglected in earlier research.
:
20 resin-to-hardener weight ratio, following an amine addition curing mechanism. Ethanol was chosen as the dispersant for Fe2O3 nanoparticles due to its low toxicity, high volatility, and good compatibility with both the epoxy system and nanoparticle surfaces, which facilitated uniform dispersion while minimizing health and environmental hazards compared to DMF or acetone. To ensure complete removal of residual ethanol prior to curing, the nanoparticle–resin mixtures were subjected to continuous magnetic stirring and mild heating at 60 °C, followed by vacuum drying, thereby eliminating solvent traces that could otherwise affect curing or composite integrity.
The preparation of the Fe2O3-filled epoxy resin is illustrated in Fig. 2. First, equal amounts of ethanol and epoxy were mixed with a magnetic stirrer for 30 minutes and termed as solution A. In a separate container, Fe2O3 nanoparticle was dispersed in ethanol via a probe-type ultrasonicator for 2 hours, and this mixture was termed solution B. After a uniform dispersion of Fe2O3 in ethanol was achieved, solution B was mixed with solution A and ultrasonic dispersion was performed for an additional 2 hours. This mixture was termed solution C. Solution C was mixed with hardener and mechanically stirred for 5 minutes to prepare the hybrid matrix for composite laminate fabrication. This procedure for matrix preparation was repeated for different Fe2O3 nanoparticle concentrations of 1 wt%, 2 wt%, and 3 wt%. The material compositions of the hybrid matrix prepared using the previously described method are summarised in Table 1.
| Sl no. | Sample designation | Fe2O3 (wt%) | Epoxy (vol%) | Hardener (vol%) |
|---|---|---|---|---|
| 1 | E | 0 | 100 | 20 |
| 2 | 1FE/E | 1 | 100 | 20 |
| 3 | 2FE/E | 2 | 100 | 20 |
| 4 | 3FE/E | 3 | 100 | 20 |
The cleaned carbon fabric was then placed on a glass plate, and the prepared hybrid epoxy matrix was spread over the surface of the carbon fabric using a roller and brush. Similarly, fabrics were stacked one above the other, with the matrix applied between the fabric layers and then consolidated by pressing with a roller. Before placing the carbon fabric on the glass plate, a releasing film was used to facilitate easy removal of the laminate. Likewise, after completing the layup, a releasing film was placed on the top along with the glass plate. The entire stack, impregnated with resin, was kept at room temperature for curing, which took about 24 hours, resulting in the production of the laminates. Table 2 presents the material composition and designation of the composite sample laminates.
| Sl. no. | Fe2O3 (wt%) | Number of layers | Epoxy : hardener (vol, %) |
Sample designation |
|---|---|---|---|---|
| 1 | 0 | 1 | 100 : 20 |
CF/E-1Layer |
| 2 | 2 | CF/E-2Layer | ||
| 3 | 4 | CF/E-4Layer | ||
| 4 | 1 | 1 | CF/1FE/E-1Layer | |
| 5 | 2 | CF/1FE/E-2Layer | ||
| 6 | 4 | CF/1FE/E-4Layer | ||
| 7 | 2 | 1 | CF/2FE/E-1Layer | |
| 8 | 2 | CF/2FE/E-2Layer | ||
| 9 | 4 | CF/2FE/E-4Layer | ||
| 10 | 3 | 1 | CF/3FE/E-1Layer | |
| 11 | 2 | CF/3FE/E-2Layer | ||
| 12 | 4 | CF/3FE/E-4Layer |
Among the measured parameters, S11 (reflection coefficient at port 1) quantifies the portion of incident power reflected back from the material, while S21 (transmission coefficient from port 1 to port 2) indicates the amount of electromagnetic power transmitted through the specimen. Parameters S12 and S22, typically associated with reverse transmission and reflection at port 2, were considered less critical for this application and thus excluded from the analysis. The S21 parameter is of primary importance in determining shielding effectiveness, as it directly correlates with the transmitted power and, therefore, the attenuation capability of the shielding material.
The specimens for EMI shielding (22.86 mm × 10.16 mm) were cut with an abrasive water jet machine from the prepared CF/E and CF/FE/E laminates with strict dimensional tolerances and high edge quality. The samples were mounted in a WR-90 rectangular waveguide sample holder, where the scattering parameters (S-parameters) were measured at frequencies between 8 and 12 GHz. Each specimen was inserted into the measurement system after performing a total two-port calibration and recording the S-parameters from a baseline for each specimen to minimise systematic error.
As shown in Fig. 3, the block diagram depicts the experimental setup for evaluating EMI shielding effectiveness (SE) in the X-band frequency range (8.2–12.4 GHz) using a rectangular waveguide. The vector network analyzer (VNA) generates and receives microwave signals using two coaxial ports. The signals are generated on port 1 and transmitted via a coaxial cable into the waveguide input, where the test specimen is mounted in a custom holder. The attenuated or transmitted signals exit through the waveguide output and are then directed to port 2 of the VNA. In this setup, it is now possible to measure the scattering parameters (S11 and S21) to evaluate a specimen's shielding performance. Total shielding effectiveness was computed from the measured forward transmission as given in eqn (1) and (2) of the manuscript consistent with waveguide SE formulations reported for planar samples. Because the WR-90 fixture with a flat, isotropic specimen is a passive reciprocal two-port, S21 = S12 by reciprocity and the diagonal terms are dominated by the same port reflections; thus, reporting S11 and S21 suffices and is standard practice in SE studies using waveguides/VNAs. We now cite representative sources and general standards to support these choices.
All EMI shielding measurements were conducted in a controlled laboratory environment at 23 ± 2 °C and 50 ± 5% relative humidity, with specimens conditioned under the same conditions for 48 h prior to testing. The vector network analyzer (VNA) was calibrated using a thru-reflect-line (TRL) standard provided with the WR-90 waveguide kit to eliminate systematic errors from cables, connectors, and fixtures. Calibration was repeated before each measurement series to ensure stability. By reporting these environmental controls and calibration procedures, we provide greater transparency and confidence in the accuracy of the shielding effectiveness data.
The shielding effectiveness due to absorption (SEA) and reflection (SER) was determined from the measured S-parameters using eqn (1) and (2), respectively:
![]() | (1) |
![]() | (2) |
The total shielding effectiveness (SET) was then computed by summing the contributions from absorption and reflection, as shown in eqn (3).
| SET = SEA + SER | (3) |
000× and Fig. 4(b) at 100
000× magnification. It is found that the Fe2O3 nanofillers used in this study are a mixture of spherical and irregular morphologies.
At 250
00× magnification (refer Fig. 4(a)), the Fe2O3 nanoparticles exhibit a highly agglomerated and irregular cluster-like structure. The particles appear to form dense aggregates, likely due to strong van der Waals forces and magnetic interactions characteristic of iron oxide nanoparticles. The surface is rough, with visible granularity indicating the polycrystalline nature of Fe2O3. At 100
000× magnification (refer Fig. 4(b)), the fine nanostructures within the agglomerates become clearer. The particles appear to have a roughly spherical to polyhedral shape, with an average size likely in the 20–80 nm range (based on visual approximation). The higher magnification reveals that these clusters are formed from much smaller primary nanoparticles.
The primary particle size (in Fig. 4(b)) seems to be in the nanometer range, but the secondary clusters (visible in Fig. 4(a)) are significantly larger, spanning micrometre scales. This hierarchical structure such as small primary nanoparticles forming larger agglomerates is typical for Fe2O3 due to high surface energy and magnetic dipole–dipole interactions. The particle distribution appears non-uniform, with some densely packed regions and others showing voids or gaps, which may affect dispersion if used as a filler in composites.
c), with lattice parameters of a = b = 5.034 Å and c = 13.748 Å (JCPDS no. 033-0664). As no additional diffraction peaks were identified, this also shows that the Fe2O3 nanoparticles are of high purity, and this agrees with the stated 99% purity.
![]() | ||
| Fig. 6 FTIR spectra of Fe2O3 nanoparticles, epoxy, carbon fabric and carbon fabric/Fe2O3/epoxy (CF/3FE/F) composite. | ||
| Wavenumber (cm−1) | Functional group/vibrationation mode | Origin |
|---|---|---|
| 3400 | O–H stretching | Fe2O3, CF, epoxy, composite |
| 2912, 2845 | C–H stretching | CF, epoxy, composite |
| 1793 | C O stretching |
CF |
| 1679–1604 | O–H bending/aromatic | Fe2O3, epoxy |
| 1582 | C C stretching |
CF, composite |
| 1484–1459 | Aromatic C–H, ring | Epoxy, composite |
| 1247–1244, 1215 | C–O–C stretching (epoxy) | Epoxy, composite |
| 1110–1056 | C–O stretching | Epoxy, composite |
| 917, 835 | Fe–O vibration/aromatic ring | Fe2O3, composite |
| 751–533 | Fe–O vibration/aromatic | Fe2O3, epoxy, composite |
The observed bands in the 470–580 cm−1 region corroborate Fe–O vibrations of α-Fe2O3, distinguishing physical adsorption from chemical bonding at the polymer–filler interface requires thermal or surface-energetic evidence. We therefore now frame the FTIR discussion as indicative rather than conclusive and note that DSC/TGA and contact-angle/XPS measurements would be appropriate to substantiate interfacial mechanisms in future work.
Neat epoxy is highly flaw-sensitive under tensile loading, small variations in cure, micro voids, or surface defects act as critical flaws that trigger failure, producing wide specimen-to-specimen scatter. Also, processing related microstructure often differs between neat and filled systems that directly affect tensile strengths. Small amounts of nanoparticles can reduce effective void content by occupying free volume will increase if dispersion is poor. In these cases, both voids and partial cure reduce tensile strength and increase test-to-test variability. Similarly, well dispersed nanoparticles can stabilize the microstructure and narrow the scatter. Fe2O3 additions can create competing mechanisms that make the filled curves look “trendless” unless dispersion is tightly controlled. At very low loadings, well-dispersed Fe2O3 can increase tensile strength via crack deflection/pinning and improved interfacial polarization-induced energy dissipation, at higher loadings, agglomerates behave as stress concentrators that depress strength and re-introduce scatter. Therefore, the trend depends upon the balance between dispersion-assisted toughening and agglomeration-induced embrittlement.
Referring to the tensile failure sample of CF/E composites previously reported by our group, which is shown in Fig. 8(a),23 the tensile failure of CF/FE/E composite samples in this study is displayed in Fig. 8(b), (c), and (d). It is observed that the CF/E and CF/FE/E samples exhibited similar behaviour, with less delamination and good integrity of filler, fibre, and matrix, as shown in Fig. 8(a) and (b). The sample with 2 wt% Fe2O3 filler showed matrix cracking with delamination, as shown in Fig. 8(c), and this failure was more severe in the 3 wt% filler-loaded sample, as shown in Fig. 8(d). This indicates that at higher Fe2O3 concentrations, the material failed due to brittle fracture of the matrix and increased delamination. In the current study, the failure analysis presented in Fig. 8 is qualitative, showing features such as river markings, fibre pull-out, and matrix cracking that are indicative of interfacial adhesion and filler–matrix interactions. While this provides useful insight, we acknowledge the limitation that no quantitative fractographic metrics (e.g., crack propagation length, void density, or fibre-matrix debonding area) were extracted.
![]() | ||
| Fig. 8 Fractured tensile specimen of (a) CF/E composite, (b) CF/1FE/E composite, (c) CF/2FE/E composite, (d) CF/3FE/E composite. | ||
From the tensile stress versus strain plot, the Young's modulus of the neat CF/E composite is computed and found to be 26.83 GPa and a maximum tensile strength of 531 MPa. Incorporation of Fe2O3 significantly influenced stiffness, with a maximum modulus of 35.20 GPa observed for CF/2FE/E, representing a ∼31% increase over the neat composite shown in Fig. 9. This enhancement is attributed to improved stress transfer due to strong interfacial adhesion between the Fe2O3-modified epoxy matrix and the CF reinforcement, consistent with findings by Shukla (2019) and Ahmad et al. (2021), where optimal nanofiller content improved modulus via synergistic reinforcement mechanisms.4,6 In terms of tensile strength, the highest value (575 MPa) was recorded for CF/1FE/E, marginally surpassing the neat composite. However, strength declined at higher loadings, with CF/3FE/E showing the lowest value (413 MPa). This reduction is attributed to nanoparticle agglomeration, void formation, and reduced fibre–matrix interfacial integrity at excessive filler content, as similarly reported in studies on nanoparticle-reinforced polymer composites.8,24 Overall, the results indicate that moderate filler addition (1–2 wt%) optimizes stiffness, while excessive loading compromises tensile strength despite maintaining high modulus. This trend aligns with literature highlighting the balance between dispersion quality, interfacial bonding, and filler content in determining mechanical performance.24,25
From previous studies on CF/E composites in the X-band, we determined that the SEA of the single- and double-layer samples remained relatively stable throughout the frequency range, while the four-layer samples exhibited serious frequency-dependent variability in SEA due to the skin effect. The skin depth, defined as the extent to which an electromagnetic wave will penetrate a material, is inversely proportional to the square root of frequency (f) which is displayed in (4). Note that δ represents the skin depth, μ is magnetic permeability (H m−1), σ is electrical conductivity (S m−1), and f is the frequency of the impinging electromagnetic field.
![]() | (4) |
According to eqn (4), as the skin depth δ decreases as frequency increases, means that, at higher frequencies, EM waves tend to travel along the surface rather than penetrate the material. At lower frequencies, larger δ allows for deeper penetration and greater absorption, which explains the superior low-frequency performance of four-layer samples. Increasing the number of layers enhances absorption-dominated shielding, as multiple carbon fabric layers better attenuate EM waves through successive absorption and reduced reflection.
Fig. 10(a–c) presents the SEA, SER, and SET of CF/1FE/E composites with single-, double-, and four-layer structures in the X-band region. As depicted in Fig. 10(d), the SEA contribution increases from 61.2% to 67.6% with the transition from a single to a two-layer configuration. A comparable trend in SEA and SER is observed for the two- and four-layer samples. Notably, the SER of the CF/1FE/E composite is lower than that of the CF/E system, which can be attributed to the enhanced absorption of EM radiation by the Fe2O3 fillers. Therefore, incorporation of Fe2O3 promotes absorption-dominated shielding effectiveness in the composite.
SEA, SER, and SET of CF/2GNP/E composites with one-, two-, and four-layer structures in the X-band are shown in Fig. 11(a–c), and the percentage SE is illustrated in Fig. 11(d).
CF/2FE/E composites demonstrate absorption-dominated EMI shielding in all configurations, with SET and SEA values increasing as the number of layers increases. The four-layer laminate achieves the highest shielding effectiveness (∼22 dB), confirming the benefit of layer stacking in enhancing performance. SEA contributes more than 66% to the overall shielding across all samples, indicating that absorption is the primary shielding mechanism in the X-band region.
In the X-band region, the SEA, SER, and SET values of the CF/3FE/E composites with single, double-, and four-layer configurations are represented in Fig. 12(a–c) and the corresponding percentage SE is given in Fig. 12(d). The presence of 3 wt% Fe2O3 as a filler increased the absorption-related shielding effectiveness due to a more efficient conductive network formed in the epoxy; in addition, multilayered structures also augmented absorption-dominated shielding. The two-layer CF/3FE/E demonstrably the best overall shielding effectiveness at 25.3 dB among the tested samples, which indicates significant potential for practical use in the X-band frequency range.
The shielding effectiveness and mechanisms in case of GNP-based epoxy composites is that graphene nanoplatelets contribute primarily through high electrical conductivity and the formation of percolated conductive networks. In this case the dominant shielding mechanism is reflection due to impedance mismatch and is supplemented by some absorption via multiple scattering inside the conductive network. So, these systems achieve relatively high SE values at moderate loadings. In case of carbon fabric with Fe2O3 nanoparticle–epoxy composites, carbon fibers provide a long range continuous conductive pathways and mechanical reinforcement. Also, Fe2O3 nanoparticles introduce magnetic loss (natural resonance, eddy current loss) and enhance interfacial polarization. Thus, shielding arises from a synergistic combination of reflection from conductive carbon fabric, absorption from Fe2O3-induced magnetic/dielectric losses, and multiple scattering due to heterogeneous interfaces. So, compared with GNP alone, this dual-action usually yields higher absorption-dominated shielding, which is desirable in applications where weight reduction, structural integrity and EMI performance must be simultaneously achieved.
Fig. 13 illustrates the comparative analysis of CF/1FE/E, CF/2FE/E, and CF/3FE/E composites across single-, double-, and four-layer configurations, revealing a clear trend of increasing total shielding effectiveness (SET) with higher filler content and additional layers. Among all, CF/3FE/E-2Layer exhibits the maximum SET of 25.3 dB, closely followed by CF/3FE/E-4Layer (23.4 dB). SEA consistently dominates over SER in all samples, indicating absorption as the primary EMI shielding mechanism, enhanced by Fe2O3-induced magnetic losses and carbon fabric conductivity. Stacking layers boosts SEA via repeated internal reflections; however, the higher Fe2O3 content strengthens dielectric and magnetic losses, resulting in excellent X-band shielding performance. In aerospace and defence sectors, EMI shielding materials are generally expected to achieve ≥30 dB attenuation, which corresponds to ∼99.9% reduction of incident electromagnetic radiation across the relevant frequency ranges (typically 8–12 GHz or even broader). For critical avionics enclosures or military-grade applications, the requirement is even stricter, often ≥60–80 dB to protect sensitive electronic systems and ensure electromagnetic compatibility (EMC). Civil aircraft structural composites are usually benchmarked against the 30 dB threshold, balancing performance with weight reduction and cost. The 25.3 dB SET value reported here corresponds to ∼99.7% attenuation. While this is a strong result for a lab-developed epoxy composite with Fe2O3 nanoparticles and carbon fabric, it falls slightly below the minimum 30 dB level generally recommended for aerospace-grade shielding panels. This study achieves near-threshold performance with a structural composite (carbon fabric reinforced epoxy + Fe2O3) is significant because it demonstrates dual functionality of mechanical integrity and EMI shielding which is a crucial step toward lightweight multifunctional aerospace materials. Table 4 compares the EMI-SE of various carbon fabric composites reported previously and in the present study. The present study reports a maximum SE of 25.3 dB, this value should be interpreted relative to filler loading, laminate thickness, and electrical conductivity. Unlike several literature examples where high SE values (>50 dB) are achieved at substantially higher filler concentrations (e.g., 40–60 wt% Fe3O4, CNT/graphite systems, or dual nanofillers) and thicker laminates (>20–100 layers), our design intentionally maintains low filler content (1–4 wt% Fe2O3) and relatively thin laminates (1–3 layers, ∼2–3 mm thickness) to balance mechanical integrity with processability. Consequently, the lower conductivity of our composites compared to highly conductive matrices (e.g., MXene- or CNT-based systems) inherently limits SE, as predicted by classical shielding theory where SE ∝ σ0.5 × t. However, the present study demonstrates that even with minimal filler loading and simpler hand lay-up processing, SE values in the 13–25 dB range are achievable, which is significant for lightweight EMI shielding applications. Furthermore, our results are consistent with other low-filler, low-thickness Fe2O3-based systems, which typically yield SE values in the 20–30 dB range, underscoring the material's comparative efficiency.
| Matrix, primary reinforcement & secondary reinforcement | Process | N | t (mm) | σ (S cm−1) | SE | Ref. | ||||
|---|---|---|---|---|---|---|---|---|---|---|
| Fibre/fabric/matrix | Filler | Concentration (wt%) | SEA (dB) | SER (dB) | SET (dB) | |||||
| T800-CF | MWCNT + Fe2O3 | 10 | Prepreg + compression molding | 20 | 3.0 | 35.0 | ∼50 | ∼10 | 59.0 | 6 |
| C fabric-plain weave | Fe3O4 | 60 | Solution casting + hot pressing | 1 | 1.54 | 1.17 | ∼47.1 | ∼3.8 | ∼58.93 | 7 |
| PTFE + CNT dual NF | Fe3O4 + MXene | 50 + 70 | Shear-induced in situ fibrillation + vacuum filtration | 2 | 0.0849 | 4.0278 | ∼24.8 | ∼19.8 | 44.56 | 8 |
| CFAN-artificial nacre | Fe2O3 | <10 | Matrix-directed mineralisation + lamination | ∼100 | <0.001 | — | ∼21–25 | ∼4–7 | ∼27–29 | 9 |
| rGCE-aerogel | Fe3O4 | 10 | In situ chemical precipitation | — | 2.5 | — | — | — | −58.13 | 10 |
| TPU | Graphite + CoFe2O4 | 40 + 40 | Melt mixing + compression molding | 3 (G/F/G) | 2.4 | — | — | — | 54 | 11 |
| PA11/PA + SCF | — | 30 | Extrusion + compression molding | 1 | 1 | ∼2.28 × 10−5 | ∼22.7 | ∼5.3 | 28 | 12 |
| UDCF/E prepreg (T800, 12 K) | Gp + Fe2O3 | 4 + 6 | Hand lay-up followed by compression molding) | 16 [0/90]8 | 2.5 | G-4 | ∼90.5 | ∼5 | 95.57 | 13 |
| F-6 | ||||||||||
| C fabric/E | GO + Fe3O4 | 8 | Layer-by-layer deposition + hot press curing | 8 | ∼1.35 | ∼3.6 | 22.8 | 10.7 | 32.9 | 14 |
| Epoxy resin | Fe3O4 + CB | 15 + 50 | Mixing + casting | — | 0.7 | — | 11.85 | −17.6 | 36.6 | 15 |
| Epoxy resin | GNP + Fe3O4 | 12 + 4 | Melt blending | 1 | 1 | — | — | — | ∼7 | 16 |
| Fe3O4 decorated GNP | 12 + 4 | 9 | ||||||||
| Cotton fibre/epoxy | Fe3O4 + GNP | 5 + 30 | Epoxy impregnation + hot press | 2 | 2.68 | — | ∼28.1 | ∼5.0 | 33.1 | 17 |
| C fabric-plain weave | Hand lay-up method | 1 | 7.8 | 6.0 | 13.9 | 23 | ||||
| 2 | 10.2 | 8.8 | 19 | |||||||
| 4 | 12.6 | 8.0 | 20.6 | |||||||
| Present study | Fe2O3/CF | 1 | Hand lay-up method | 1 | — | 8.15 | 5.14 | 13.3 | ||
| 2 | 8.79 | 4.25 | 13 | |||||||
| 4 | 10.4 | 4.37 | 14.8 | |||||||
| 2 | 1 | — | 12.1 | 4.43 | 16.5 | |||||
| 2 | 10.1 | 5.05 | 15.1 | |||||||
| 4 | 16.9 | 5.18 | 22.1 | |||||||
| 3 | 1 | — | 8.03 | 6.63 | 14.6 | |||||
| 2 | 17.3 | 8.03 | 25.3 | |||||||
| 4 | 17.3 | 6.04 | 23.4 | |||||||
The improvement in tensile strength observed at moderate Fe2O3 loadings reflects effective stress transfer at the fiber–matrix interface, which simultaneously promotes the formation of continuous conductive or semi-conductive networks essential for EMI attenuation. Conversely, the reduction in mechanical properties at higher filler contents can be attributed to agglomeration and interfacial defects, which not only act as stress concentrators but also disrupt the uniformity of the shielding pathways, thereby limiting shielding effectiveness despite increased filler volume. Quantitatively, this trade-off can be rationalized by considering that both load transfer efficiency (σ ∝ Vf × τi, where τi is interfacial shear strength) and shielding effectiveness (SE ∝ σ0.5 × t, where σ is conductivity and t is thickness) are maximized under conditions of uniform dispersion and strong interfacial bonding. By correlating tensile strength retention with SE trends across different filler loadings, we demonstrate that optimal performance arises not simply from increasing filler concentration, but from achieving synergistic dispersion and interfacial compatibility. This mechanistic perspective situates our findings within broader composite design principles and highlights pathways for future optimization (e.g., surface functionalization of Fe2O3, hybrid filler strategies, or controlled laminate stacking) to simultaneously advance mechanical robustness and EMI shielding efficiency.
The carbon fabric/Fe2O3 nanoparticle/epoxy (CF/3FE/E-Layer) composite developed in this study shows multifunctional characteristics as good mechanical integrity from carbon fabric, magnetic and dielectric loss from Fe2O3, and structural reliability from epoxy. While the aerospace sector is an obvious target, several other industries could benefit from such composites. Which included automotive and electric vehicles (EV) with increasing use of sensors, radar, and advanced driver-assistance systems (ADAS), EMI shielding is crucial for safety and reliability. Also, the composites can replace heavier metallic shielding, lower vehicle weight and improve fuel efficiency. These composites can also be used in EMI-safe enclosures for EV batteries and control units, reducing electromagnetic crosstalk. The carbon fabric/Fe2O3 nanoparticle/epoxy (CF/3FE/E-Layer) composite developed in this study also can be used in structural panels for antenna housings, base stations, and smart towers where both strength and EMI shielding are needed. Also, it can be used in laptops, tablets, and high-frequency devices that benefit from lightweight shielding to prevent interference.
To further improve the shielding effectiveness (SET) beyond the current 25.3 dB, future work should focus on multilayer or gradient architectures that enhance multiple reflections, hybridization of Fe2O3 with conductive 2D fillers such as graphene or MXenes to combine magnetic and conductive losses, and surface functionalization of Fe2O3 to improve dispersion and interfacial polarization. In addition, thickness optimization guided by skin-depth calculations and the introduction of porous or foam-like sublayers can significantly increase absorption without adding excessive weight. Together, these strategies are expected to raise SET above the aerospace benchmark of 30 dB and broaden the applicability of the composites to advanced structural and electronic shielding applications.
| This journal is © The Royal Society of Chemistry 2025 |