Next Article in Journal
Mechanism by Which Heat Treatment Influences the Acoustic Vibration Characteristics of Bamboo
Previous Article in Journal
High-Frequency Plasma Electrolytic Oxidation of an Al–Si Alloy: Influence of Al2O3 and SiO2 Additives on Coating Microstructure and Tribological Performance
Previous Article in Special Issue
Purification and Preparation of Graphene-like Nanoplates from Natural Graphite of Canindé, CE, Northeast-Brazil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of Graphite Shells on Ferromanganese Oxide for Electromagnetic Wave Absorption

School of Materials and Chemistry, University of Shanghai for Science and Technology, No. 516 Jungong Road, Shanghai 200093, China
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(23), 5336; https://doi.org/10.3390/ma18235336
Submission received: 30 October 2025 / Revised: 20 November 2025 / Accepted: 24 November 2025 / Published: 26 November 2025
(This article belongs to the Special Issue Carbon Nanomaterials for Diverse Applications—Second Edition)

Abstract

Ferromanganese oxide (FMO), a by-product of steelmaking industry, was coated with polyacrylonitrile (PAN) to construct an electromagnetic wave absorber (FMO@C) with a core–shell structure. The effect of heat treatment from 600 to 1000 °C on the phase transformation of FMO and carbonization of PAN was studied. Upon the heat treatment at 1000 °C, the reflection loss and effective bandwidth of the FMO@C reached −18.20 dB and 3.08 GHz at a thickness of 1.6 mm, presenting a significant improvement over FMO which only exhibited a reflection loss of −2.31 dB at 10 mm. Boric acid was infiltrated into the PAN shells to catalyze the carbonization process and adjust the impedance matching, which further improved the reflection loss to a minimum value of −28.25 dB. Via varying the concentration of boric acid, the reflection loss of −22.01 dB with an effective bandwidth of 3.36 GHz at a thickness of 1.3 mm was achieved. The enhanced EMW absorption performance was attributed to multiple reflections and polarization caused by the core–shell structure, magnetic loss from the phase transformation of FMO, dielectric loss from carbon shells, as well as the tunable impedance matching by boron-catalyzed carbonization. The construction of the core–shell structure could be a promising downstream processing of FMO and could extend the application of the solid wastes.

1. Introduction

Electromagnetic waves (EMW) are essential to modern society, yet they pose risks such as harmful interference to electronic devices and potential health hazards from prolonged exposure [1,2]. EMW-absorbing materials are widely needed for human health, equipment shields, and information security [3].
Industrial wastes from steelmaking processes, including iron powders, iron oxide scales, and steel slag, possess excellent magnetic properties [4,5]. Research attempts have been made to utilize these industrial solid wastes as EMW absorbers. Steel slag was added in an EMW absorption mortar, which exhibited a minimum reflection loss (RL) of −11.5 dB at a thickness of 25 mm [6]. Cement-based composites were reinforced by waste iron powders and the minimum RL was −12.78 dB with an effective absorption bandwidth of 1.37 GHz at a thickness of 3.55 mm [7]. In addition, iron oxide scales were also used as EMW-shielding materials [4,8,9]. Ozturk et al. produced an EMW-shielding mortar containing 30 vol% of iron oxide scales, which exhibited a shielding effectiveness of approximately 40 dB at a thickness of 20 mm in the 11–18 GHz frequency range [4]. Adam Jakubas et al. prepared EMW-shielding materials by hot pressing recycled high-density polyethylene (HDPE) and iron oxide scales. The composites containing 70 wt% iron oxide scale in the HDPE matrix exhibited an electromagnetic shielding effectiveness of up to 52 dB in the 8.1–12.1 GHz range [8]. Nevertheless, the application of industrial solid wastes is limited due to their high content of impurity, which may result in unstable electromagnetic properties. Ferromanganese oxide (FMO) is a secondary processing product made from the recycled iron oxide scales, which opens an opportunity to produce low-cost EMW-absorbing materials in large quantity. Compared with iron oxide scales, the industrial-grade pre-sintered FMO powder, consisted of Mn2O3 and Fe2O3, have a well-defined composition and obvious magnetic characteristics. The addition of manganese oxide can generate interface inhomogeneity and interface polarization loss of EMW [10,11].
The structural design of carbon/magnetic composites is vital for optimizing their EMW absorption performance. Core–shell structures, skeleton structures, and hollow structures were studied, and all these structures enhanced EMW absorption through distinct mechanisms [12]. The skeleton structure employed a three-dimensional conductive network, which was lightweight and which promoted multiple scattering [12]. The hollow structure created internal cavities to enhance multiple reflections of EMW [13]. The core–shell structure was acknowledged as one of the most favourable structures for EMW absorption, as it can combine multi-function materials, protect the core materials, and facilitate interfacial polarization and multireflection [14,15,16]. Porous carbon-coated CoFe2O4 exhibited a minimum RL of −39.48 dB at 1.22 GHz with a 2 mm thickness [17]. BaFe12O19 was coated by glucose-based carbon, achieving an optimal RL of −73.42 dB at 17.84 GHz with a thickness of 1.40 mm [18]. MnO2-coated Fe3O4 microspheres showed an optimal RL of −48.5 dB at 11.2 GHz with a thickness of 2.5 mm [19]. Pyrolyzed carbon-coated Co had an optimal RL of −96.2 dB at 5.8 GHz with a thickness of 3.1 mm [20]. The high EMW absorption properties in these works are benefited from the core–shell structure with nano-sized or submicron-sized magnetic cores.
The doping of foreign elements on carbon materials can catalyze their graphitization process [21], improve their electrical conductivity, and facilitate the EMW absorption [22]. Boron has a similar atomic radius and less electronegativity than carbon, so it can enter the carbon lattice at substitutional position and function as a p-type dopant, which introduce additional defects, improves impedance matching, and enhances polarization loss, conductive loss and dielectric loss [22]. Immersion in boric acid, followed by heat treatment, was reported as a facile way to dope carbon fibres and catalyze their graphitization, which improved significantly the electrical and thermal conductivity of carbon fibres [21,23].
Aiming for an extended utilization of industrial solid wastes, a series of magnetic core@carbon shell (FMO@C) materials was constructed. Polyacrylonitrile (PAN) was pyrolyzed to produce the carbon shells. The effect of heat treatment was studied on the phase evolution of FMO and the carbonization of PAN. The boric acid was used to catalyze the carbonization and tune the impedance matching. The EMW-absorbing properties were analyzed based on the dielectric loss, magnetic loss, and the impedance matching characteristics. Eventually, the RL values of FMO@C reached −28.25 dB, exhibiting a much higher EMW absorption capability than FMO, owning to the controlled phase transformation of core materials and the catalyzed graphitization of carbon shells.

2. Materials and Methods

2.1. Fabrication of the FMO@C Powder

FMO was supplied by Shanghai Baosteel Magnetic Materials Co., Ltd. (Shanghai, China) PAN (average M.W. = 150,000, and was dissolved in N, N-dimethylformamide (DMF) at a concentration of 4.3%. FMO was then dispersed in DMF with a content of 10 wt% by sonicator for 10 min. After the evaporation of DMF at 120 °C, the mixture was calcined at 270 °C for 120 min for the cyclization of PAN (referred as FMO@PAN-270). Then, the products were carbonized in a tubular furnace under N2 atmosphere at different temperatures (600, 700, 800, 900, and 1000 °C) for 2 h. The FMO@C were denoted as FMO@C-600, FMO@C-700, FMO@C-800, FMO@C-900, and FMO@C-1000, according to the carbonization temperature. Alternatively, FMO@PAN-270 was dipped into a boron acid (H3BO3) solution of different concentrations (0.5%, 1%, and 2.5%) for 6 h at 80 °C before the carbonization process at 1000 °C. The samples were named as FMO@C-B1, FMO@C-B2, and FMO@C-B3, respectively.

2.2. Characterization

The morphological evolution and phase transition were tested by a scanning electron microscope (Thermo Apreo 2S HiVac, Thermo Fisher Scientific, Waltham, MA, USA), X-ray diffractometer (D8 ADVANCE, Bruker, Billerica, MA, USA), and Fourier transform infrared spectrometer (FT-IR, SPE CTRUM 100, PerkinElmer, Waltham, MA, USA). The graphitization degree was characterized by Raman spectroscopy (HORIBA, Kyoto, Japan) equipped with a 532 nm laser. The electrical conductivity of FMO@C was measured by the four-probe technique (SB120/2, China). The electromagnetic properties were evaluated using a vector network (Agilent N5244A, USA) across the 2–18 GHz frequency band. The FMO@C specimens were homogeneously blended with paraffin at a 1:1 weight ratio and compacted into a cylindrical mould (Φout = 7.00 mm, Φin = 3.00 mm).

3. Results and Discussion

3.1. The Phase Evolution of FMO@C

The FT-IR spectrum of PAN had a characteristic peak at 2245 cm−1, as shown in Figure 1a, which was attributable to -C≡N tensile vibration [24]. Upon cyclization at 270 °C, new peaks at 1603 cm−1, 1587 cm−1, and 1369 cm−1 were ascribed to -C=N, -C=C, and -C-H bonds, respectively, as a result of the typical transformation of PAN from a linear chain to a cyclized structure [25]. Upon carbonization at 1000 °C, the majority of the peaks were diminished, suggesting the pyrolysis process. A new peak at 1632 cm−1 was denoted to C=C or C=N conjugated bonds in aromatic rings [26].
The XRD spectra of FMO@C samples are presented in Figure 1b. The as-received FMO were composed of Fe2O3 and Mn2O3, with diffraction peaks corresponding to (JCPDS No. 33-0664) and (JCPDS No. 89-4836). Fe3O4 phase (JCPDS No. 19-0629) and MnO phase (JCPDS No. 07-0230) were formed in FMO@C-600 by reduction at 600 °C. XRD spectrum of the FMO@C-700 and FMO@C-800 samples showed mainly Fe upon further reduction. For FMO@C-900, the characteristic peaks of Fe1.8Mn1.2C (JCPDS No. 89-2546) appeared along with Fe and MnO phases. After the heat treatment at 1000 °C, the peaks corresponding to Fe and MnO disappeared, leaving only the Fe1.8Mn1.2C phase, which remained stable when the heating temperature was further increased to 1200 °C. The phase transformation played a fundamental role in regulating the final electromagnetic properties of the core–shell materials.
The graphitization degree of the shells can be analyzed by Raman spectra, as shown in Figure 1c. The G band around 1580 cm−1 arises from the stretching motions of ordered sp2 carbon bonds, whereas the D band near 1350 cm−1 comes from disordered breathing modes caused by structural imperfections [27]. The intensity of the ratio of the D to G bands (ID/IG) of the shells decreased continuously with increases in the heat treatment temperature, which suggests an increased degree of graphitization [28].
The as-received FMO particles had a diameter of 1–3 μm (Figure 2a). The PAN-wrapped FMO particles sticked together, forming large aggregates, as shown in Figure 2b. The polymetric morphology gradually disappeared upon the heat treatment (Figure 2c–e). Eventually, when the heating temperature reached 1000 °C, the aggregates were separated into core–shell particles because the carbonization of the PAN caused a volume shrinkage during the structural evolution of the shells, as shown in Figure 2f. With the higher magnification seen in Figure 2g, interconnected FMO@C-1000 particles were observed with a size of around 2 μm, similar with that of the as-received FMO. EDS Mn and Fe element mappings confirmed the wrapping of FMO in the particles as shown in Figure 2h,i. Along with the typical Raman spectra of carbon in Figure 1c, the core–shell structure can be proven.

3.2. EMW Absorption Properties of FMO@C

To investigate the EMW absorption properties of FMO@C, the RL values were calculated using the equations shown below [29]:
R L = 20 log | ( Z i n   Z 0 ) / ( Z i n   + Z 0 ) |
Z i n = Z 0 ( μ r / ε r ) 1 / 2 tanh   ( j ( 2 π f d / c ) ( μ r ε r ) 1 / 2
where Z i n and Z 0 are the input impedance of the absorber and the impedance of the free space. The f is the EMW frequency, h is the Planck constant, c is the light velocity, and d refers to the absorber thickness. The effective absorption bandwidth refers to the portion where RL is below −10 dB [30].
The RL values of FMO@C are shown in Figure 3. The pristine FMO had a virtually negligible capability of EMW absorption with a minimal RL of −2.31 dB at a thickness of 10.00 mm. As shown in Figure 3b–f, the FMO@C-600, FMO@C-900, and FMO@C-1000 showed a much higher EMW absorption capacity than FMO, FMO@C-700, and FMO@C-800. FMO@C-1000 had the minimum RL value of −18.20 dB with a thickness of 1.6 mm, and the effective bandwidth was 3.08 GHz. The minimum RL values of FMO@C-700 and FMO@C-800 were only −9.04 dB and −8.67 dB, respectively.
To acquire a more profound understanding of the mechanisms underlying EMW absorption properties, the electromagnetic parameters of FMO@C were investigated. The real and imaginary parts of the complex dielectric constant ( ε r = ε j ε ) and complex magnetic permeability ( μ r = μ j μ ) are shown in Figure 4a,b,d,e. The real parts (ε′ and μ′) represent the energy storage capacity within the EMW absorber, while the imaginary parts (ε″ and μ″) characterize its dissipative properties [31]. The ε′ and ε″ values of FMO were relatively low, suggesting that the dielectric storage and dissipation capabilities were weak. The ε′ and ε″ values of FMO@C-600 surpassed FMO, mainly due to the construction of the carbon shells. Moreover, the values of ε′ and ε″ of FMO@C-700 and FMO@C-800 significantly increased, caused by the transformation of Fe3O4 into Fe [32]. As the carbonization temperature was higher than 900 °C, the dielectric properties of FMO@C progressively decreased as the Fe and MnO converted to Fe1.8Mn1.2C.
The electrical conductivity of the samples are shown in Figure 5a. According to the equation derived from the free electron theory ( ε σ / 2 π ε 0 f , where σ represents the conductivity of materials), the conductivity is directly proportional to the value of ε″ [33]. The electrical conductivity of FMO and FMO@C-600 was lower than the testing limitation. The electrical conductivity of FMO@C-700 and FMO@C-800 was higher than that of FMO@C-900 and FMO@C-1000, exhibiting the same trend with ε″ values. The dielectric loss (tan δE) of the FMO@C was calculated by tan δ E = ε / ε and shown in Figure 4c. FMO@C-700 and FMO@C-800 possessed the highest tan δE values, indicating that they possessed the strongest dielectric loss capabilities.
Figure 4d–f show the μ′, μ″, and tan δM of FMO@C. The FMO@C-700 and FMO@C-800 exhibit higher tan δE and tan δM than other samples. However, their RL was not comparable to the other FMO@C specimens. Impedance matching ( Z = Z i n / Z 0 ) is another crucial parameter beyond the loss capability of materials. When the Z values of material are equal to 1, EMW will fully enter the material without reflection [34]. The Z values of FMO@C-900 and FMO@C-1000 were closer to 1, as shown in Figure 5, granting them higher RL values, even though their tan δE and tan δM values were not the highest. In contrast, FMO@C-700 and FMO@C-800 exhibited Z values significantly lower than one because the presence of elemental Fe resulted in their high dielectric constants and high electrical conductivity [34].

3.3. Effects of Boric Acid Catalyzation

To further enhance the EMW absorption properties of FMO@C, the graphitization process of the carbon shells was catalyzed by boric acid. At approximately 220 °C, boric acid underwent dehydration and decomposed into boron oxide (B2O3) [23]. As the temperature further increased, B2O3 reacted with carbon, playing a role in promoting graphitization [35].
Figure 6 shows the Raman spectra and electrical conductivity of samples. The ID/IG values of FMO@C-1000 were lower than that of pure PAN-based carbon after carbonization at 1000 °C, suggesting that FMO and its reduction products can facilitate the graphitization of carbon materials. The iron oxides or elemental Fe acted as catalysts that promoted the rearrangement of amorphous carbon into ordered graphitic layers through a dissolution–precipitation process [36,37]. The ID/IG of FMO@C gradually decreased as increasing the boric acid concentration, leading to the increasing of their electrical conductivity due to the enhanced graphitization degree of the carbon shells.
The FMO@C-B1 exhibited a RL value of −23.11 dB and effective bandwidth of 2.24 GHz with a thickness of 3.1 mm, as shown in Figure 7. The FMO@C-B2 achieved the best RL of −28.25 dB and effective bandwidth of 1.44 GHz with a thickness of 3.6 mm, while it had a RL value of −21.68 dB with an effective bandwidth of 2.92 GHz at a thickness of 2.0 mm. The RL of FMO@C-B3 was decreased to −22.01 dB, along with the widest effective bandwidth of 3.36 GHz at a thin thickness of 1.3 mm. All of the FMO@C-B specimens showed better EMW absorption properties than the specimens without the treatment by boric acid.
The ε′, ε″, and tan δE of FMO@C-B are shown in Figure 8a–c. The dielectric constant increased with the degree of graphitization. The dielectric constant of FMO@C-B was generally higher than those of FMO@C. The tan δE curves of all FMO@C-B exhibit multiple resonance peaks, indicating the presence of multiple polarization relaxations, which were beneficial for dielectric loss [20].
Figure 8d–f show the μ′, μ″, and tan δM of the FMO@C-B. The tan δE values of FMO@C were significantly higher than the tan δM values, indicating that dielectric loss played a primary role in the attenuation of EMW energy. All samples exhibited similar trends in tan δM curves and displayed resonance peaks at 9–12 GHz and 15–18 GHz, which may be related to natural resonance and exchange resonance [20].
The attenuation constant of FMO@C-1000 and FMO@C-B was calculated using the following equation [38]:
α = 2 c π f × ( μ ε μ ε ) + ( μ ε μ ε ) 2 + ( μ ε + μ ε ) 2
As shown in Figure 9, FMO@C-B2 and FMO@C-B3 exhibited higher attenuation constants than FMO@C-B1 and FMO@C-1000, indicating their stronger attenuation capabilities primarily attributed to their superior dielectric loss. Figure 10 illustrates the impedance matching of FMO@C-B specimens. It seems that the FMO@C-B2 had Z values nearer to one compared with the other samples. The attenuation ability and impedance matching of FMO@C-B2 were in agreement with its excellent RL value and explained the effects of boric acid on the EMW absorption properties.
The EMW absorption mechanisms of FMO@C-B are schematically presented in Figure 11. Firstly, the core–shell structure of the material facilitated multiple reflections of EMW at the interfaces. Meanwhile, the carbon shells of the particles were interconnected rather than isolated, which promoted the electron transportation between particles and enhanced the conductive loss of materials. Secondly, heterogeneous interfaces within the materials induced interfacial polarization, while the lattice defects and amorphous carbon structure in the carbon shells generated dipole polarization. These polarization relaxation processes amplified the dielectric loss of the materials. Furthermore, the magnetic cores generated multiple magnetic resonances, including natural resonance and exchange resonance. The dielectric and magnetic losses improved the material’s EMW absorption performance. Lastly, the impedance matching between magnetic cores and the carbon shells was optimized through the catalyzed graphitization of the carbon shells.

4. Conclusions

FMO@C with a core–shell structure was constructed using PAN-wrapped FMO as a precursor, aiming for the extended utilization of the industrial solid waste. When the heating temperature ranged from 600 to 1000 °C, FMO was transferred into different phases with varying dielectric and magnetic properties, and the graphitization degree of the carbon shells also changed. The FMO@C-1000 had a minimum RL of −18.20 dB and an effective bandwidth of 3.08 GHz at only a 1.6 mm thickness, which was superior to as-received FMO. Boric acid-catalyzed graphitization of carbon shells further enhanced the RL to a minimum value of −28.25 dB with an effective bandwidth of 1.44 GHz at a thickness of 3.6 mm because of the increased dielectric loss and tuned impedance matching. The RL of −22.01 dB with larger effective bandwidth of 3.36 GHz at smaller thickness of 1.3 mm was achieved by varying the concentration of boric acid. The FMO@C could be cost-efficient EMW absorbers and could be applied as fillers in composites for EMW-proof packaging or building. The core–shell construction, as a downstream processing of FMO, was simple and would be easy to upscale.

Author Contributions

Conceptualization, S.S. and J.L.; Methodology, Y.Z. and S.S.; Validation, Y.Z.; Investigation, Y.Z.; Resources, S.S. and J.L.; Writing—original draft, Y.Z.; Writing—review & editing, S.S. and J.L.; Supervision, J.L.; Project administration, S.S. and J.L.; Funding acquisition, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, D.; Peng, J.; Liu, S.; Huang, T.; Xiong, Z.; Huang, S.; Zhao, F.; Peng, Y.; Cao, M.; Liu, C. Prospects of porous-carbon-based electromagnetic wave absorbing materials. Carbon 2025, 244, 120650. [Google Scholar] [CrossRef]
  2. Chen, X.; Lan, D.; Zhou, L.; Liu, H.; Song, X.; Wang, S.; Zou, Z.; Wu, G. Review of recent advances in ferrite-based materials: From synthesis techniques to electromagnetic wave absorption performance. Int. J. Miner. Metall. Mater. 2025, 32, 591–608. [Google Scholar] [CrossRef]
  3. Zhao, R.; Liang, B.; Shi, Y.; Dong, Q.; Li, T.; Gu, J.; Ma, Y.; Zhang, J.; Melhi, S.; Alshammari, A.S.; et al. Recent progress of carbon-based magnetic fibers for electromagnetic wave absorption. Carbon 2024, 229, 119513. [Google Scholar] [CrossRef]
  4. Ozturk, M.; Depci, T.; Bahceci, E.; Karaaslan, M.; Akgol, O.; Sevim, U.K. Production of new electromagnetic wave shielder mortar using waste mill scales. Constr. Build. Mater. 2020, 242, 118028. [Google Scholar] [CrossRef]
  5. Zhang, X.; Zhao, J.; Liu, Y.; Li, J. Use of steel slag as carbonation material: A review of carbonation methods and evaluation, environmental factors and carbon conversion process. J. CO2 Util. 2024, 88, 102947. [Google Scholar] [CrossRef]
  6. Dai, Y.; Wu, J.; Wang, D.; Li, R.; Lu, C.; Xu, Z. Electromagnetic Wave-Absorbing Properties of Steel Slag. J. Mater. Eng. Perform. 2018, 28, 535–542. [Google Scholar] [CrossRef]
  7. Ma, C.; Wu, Z.; Xie, S.; Wang, Y.; Ji, Z.; Wang, J. Synergistic effects of carbon black and waste iron powder on electromagnetic wave absorbing and mechanical properties of cement mortar. J. Ind. Eng. Chem. 2025. [Google Scholar] [CrossRef]
  8. Jakubas, A.; Łada-Tondyra, E.; Makówka, M.; Suchecki, Ł. A Study on the Possibility of Using Iron Scale in the Construction of Electromagnetic Field Shields. Energies 2022, 15, 1332. [Google Scholar] [CrossRef]
  9. Alwaeli, M. The implementation of scale and steel chips waste as a replacement for raw sand in concrete manufacturing. J. Clean. Prod. 2016, 137, 1038–1044. [Google Scholar] [CrossRef]
  10. Lin, D.; Peng, J.; Guo, J.; Jiang, X. Unique frequency reversible conversion and bandwidth regulation of electromagnetic wave absorption performance for core-shell structured Fe3O4 and manganese oxide composites. Mater. Today Nano 2025, 29, 100568. [Google Scholar] [CrossRef]
  11. Meng, X.; Qiao, J.; Yang, Y.; Zhang, X.; Yang, Z.; Zheng, S.; Liu, J.; Wu, L.; Wang, Z.; Wang, F. Three-dimensional porous manganese oxide/nickel/carbon microspheres as high-performance electromagnetic wave absorbers with superb photothermal property. J. Colloid. Interface Sci. 2023, 629, 884–894. [Google Scholar] [CrossRef]
  12. Peng, H.; Zhang, D.; Xie, Z.; Lu, S.; Liu, Y.; Liang, F. Recent Advances in Structural Design of Carbon/Magnetic Composites and their Electromagnetic Wave Absorption Applications. Small 2025, 21, e2408570. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, J.; Guo, X.; Lan, D.; Wang, Y.; Huang, H.; Zhang, C.; Wu, G.; Zhang, S.; Jia, Z. Multifunctional electromagnetic wave absorbing materials: Research progress from component structural design to intelligent integration. Carbon 2025, 245, 120818. [Google Scholar] [CrossRef]
  14. Bhattacharjee, Y.; Bose, S. Core–Shell Nanomaterials for Microwave Absorption and Electromagnetic Interference Shielding: A Review. ACS Appl. Nano Mater. 2021, 4, 949–972. [Google Scholar] [CrossRef]
  15. Gai, L.; Zhao, H.; Wang, F.; Wang, P.; Liu, Y.; Han, X.; Du, Y. Advances in core—Shell engineering of carbon-based composites for electromagnetic wave absorption. Nano Res. 2022, 15, 9410–9439. [Google Scholar] [CrossRef]
  16. Wu, Z.; Cheng, H.W.; Jin, C.; Yang, B.; Xu, C.; Pei, K.; Zhang, H.; Yang, Z.; Che, R. Dimensional Design and Core-Shell Engineering of Nanomaterials for Electromagnetic Wave Absorption. Adv. Mater. 2022, 34, e2107538. [Google Scholar] [CrossRef]
  17. Ma, G.; Yin, P.; Zhang, L.; Huang, H.; Sun, X.; Zhang, Y.; Wang, J.; Feng, X. Biomass-derived porous carbon combined with CoFe2O4/CoFe@C for available low-frequency microwave dissipation. Powder Technol. 2023, 415, 118196. [Google Scholar] [CrossRef]
  18. Liu, Y.; Lin, Y.; Yang, H. Facile fabrication for core-shell BaFe12O19@C composites with excellent microwave absorption properties. J. Alloys Compd. 2019, 805, 130–137. [Google Scholar] [CrossRef]
  19. Qiao, M.; Lei, X.; Ma, Y.; Tian, L.; Su, K.; Zhang, Q. Dependency of tunable microwave absorption performance on morphology-controlled hierarchical shells for core-shell Fe3O4@MnO2 composite microspheres. Chem. Eng. J. 2016, 304, 552–562. [Google Scholar] [CrossRef]
  20. Wang, B.; Fu, Y.; Li, J.; Wu, Q.; Wang, X.; Liu, T. Construction of Co@C nanocapsules by one-step carbon reduction of single-crystal Co3O4 nanoparticles: Ultra-wideband microwave absorber verified via coaxial and arch methods. Chem. Eng. J. 2022, 445, 136863. [Google Scholar] [CrossRef]
  21. Lee, S.; Cho, S.Y.; Chung, Y.S.; Choi, Y.C.; Lee, S. High electrical and thermal conductivities of a PAN-based carbon fiber via boron-assisted catalytic graphitization. Carbon 2022, 199, 70–79. [Google Scholar] [CrossRef]
  22. Cui, A.; Wang, C.; Miao, Y.; Wang, X.; Wang, Y.; Lan, D.; Wu, S.; Song, G.; Wang, T.; Tian, Z.; et al. B─C Bonding Configuration Manipulation Strategy Toward Synergistic Optimization of Polarization Loss and Conductive Loss for Highly Efficient Electromagnetic Wave Absorption. Adv. Funct. Mater. 2024, 35, 2420292. [Google Scholar] [CrossRef]
  23. Huckstaedt, T.; Erdmann, J.; Lehmann, A.; Protz, R.; Ganster, J. Boric Acid as A Low-Temperature Graphitization Aid and Its Impact on Structure and Properties of Cellulose-Based Carbon Fibers. Polymers 2023, 15, 4310. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, W.; Sun, M.; Yin, J.; Abou-Hamad, E.; Schwingenschlogl, U.; Costa, P.; Alshareef, H.N. A Cyclized Polyacrylonitrile Anode for Alkali Metal Ion Batteries. Angew. Chem. Int. Ed. Engl. 2021, 60, 1355–1363. [Google Scholar] [CrossRef] [PubMed]
  25. Sun, Q.; Hu, G.; Peng, Z.; Cao, Y.; Zhu, F.; Zhang, Y.; Gao, H.; Du, K. Achieving a bifunctional conformal coating on nickel-rich cathode LiNi0.8Co0.1Mn0.1O2 with half-cyclized polyacrylonitrile. Electrochim. Acta 2021, 386, 138440. [Google Scholar] [CrossRef]
  26. Sayyar, S.; Moskowitz, J.; Fox, B.; Wiggins, J.; Wallace, G. Wet-spinning and carbonization of graphene/PAN-based fibers: Toward improving the properties of carbon fibers. J. Appl. Polym. Sci. 2019, 136, 47932. [Google Scholar] [CrossRef]
  27. Wang, T.; Wang, H.; Chi, X.; Li, R.; Wang, J. Synthesis and microwave absorption properties of Fe–C nanofibers by electrospinning with disperse Fe nanoparticles parceled by carbon. Carbon 2014, 74, 312–318. [Google Scholar] [CrossRef]
  28. Gai, L.; Zhao, H.; Li, X.; Wang, P.; Yu, S.; Chen, Y.; Wang, C.; Lan, D.; Han, F.; Du, Y. Shell engineering afforded dielectric polarization prevails and impedance amelioration toward electromagnetic wave absorption enhancement in nested-network carbon architecture. Chem. Eng. J. 2024, 501, 157556. [Google Scholar] [CrossRef]
  29. Xing, Y.; Fan, Y.; Yan, Z.; Zhao, B.; Huang, Y.; Pan, W. Core-shell LaOCl/LaFeO3 nanofibers with matched impedance for high-efficiency electromagnetic wave absorption. Sci. China Mater. 2022, 66, 1587–1596. [Google Scholar] [CrossRef]
  30. Liu, Q.; Cao, Q.; Bi, H.; Liang, C.; Yuan, K.; She, W.; Yang, Y.; Che, R. CoNi@SiO2 @TiO2 and CoNi@Air@TiO2 Microspheres with Strong Wideband Microwave Absorption. Adv. Mater. 2016, 28, 486–490. [Google Scholar] [CrossRef]
  31. Xu, X.; Wang, Y.; Yue, Y.; Wang, C.; Xu, Z.; Liu, D. Core-shell MXene/nitrogen-doped C heterostructure for wide-band electromagnetic wave absorption at thin thickness. Ceram. Int. 2022, 48, 30317–30324. [Google Scholar] [CrossRef]
  32. Yuan, S.; Wang, T.; Feng, T.; Kong, J. Electromagnetic wave absorption of fabricated Fe/Fe3O4/C hollow fibers derived from ceiba fiber templates. Mater. Sci. Eng. B 2024, 299, 117057. [Google Scholar] [CrossRef]
  33. Han, C.; Zhang, M.; Cao, W.-Q.; Cao, M.-S. Electrospinning and in-situ hierarchical thermal treatment to tailor C–NiCo2O4 nanofibers for tunable microwave absorption. Carbon 2021, 171, 953–962. [Google Scholar] [CrossRef]
  34. Wu, X.; Liu, K.; Ding, J.; Zheng, B.; Gao, F.; Qian, K.; Ma, Y.; Feng, Y.; Chen, L.; Zhang, P.; et al. Construction of Ni-based alloys decorated sucrose-derived carbon hybrid towards: Effective microwave absorption application. Adv. Compos. Hybrid Mater. 2022, 5, 2260–2270. [Google Scholar] [CrossRef]
  35. Mohammed, M.I.; Abud-Alnur, S.; Aliwi, S.M. Study of the mechanism of graphitization of phenolic resin carbon catalyzed by boron oxide. J. Polym. Res. 2021, 28, 174. [Google Scholar] [CrossRef]
  36. Du, Y.; Liu, W.; Qiang, R.; Wang, Y.; Han, X.; Ma, J.; Xu, P. Shell thickness-dependent microwave absorption of core-shell Fe3O4@C composites. ACS Appl. Mater. Interfaces 2014, 6, 12997–13006. [Google Scholar] [CrossRef]
  37. Krishnan, S.G.; White, C.E.; Zeng, K.; Kalarikkal, N.; Ok, Y.S.; Arnold, C.B.; Thomas, S.; Nzihou, A. Recent developments on multi- versus single-metallic catalytic graphitisation of biocarbon: A review. Fuel 2025, 396, 135330. [Google Scholar] [CrossRef]
  38. Liang, X.; Liu, S.; Zhong, S.; Zhang, S.; Meng, X.; Zhang, Y.; Yu, M.; Wang, C. A novel synthesis of Porous Fe(4)N/carbon hollow microspheres for thin and efficient electromagnetic wave absorbers. J. Colloid Interface Sci. 2023, 637, 123–133. [Google Scholar] [CrossRef]
Figure 1. (a) FT-IR curves of PAN, FMO@PAN-270 and FMO@C-1000, (b) XRD patterns of FMO and FMO@C, and (c) Raman spectrum of FMO@C.
Figure 1. (a) FT-IR curves of PAN, FMO@PAN-270 and FMO@C-1000, (b) XRD patterns of FMO and FMO@C, and (c) Raman spectrum of FMO@C.
Materials 18 05336 g001
Figure 2. The SEM images of (a) FMO, (b) FMO@PAN, (c) FMO@C-700, (d) FMO@C-800, (e) FMO@C-900, (f) FMO@C-1000 (low magnification), (g) FMO@C-1000 (high magnification), and the EDS (h) Mn and (i) Fe element mappings.
Figure 2. The SEM images of (a) FMO, (b) FMO@PAN, (c) FMO@C-700, (d) FMO@C-800, (e) FMO@C-900, (f) FMO@C-1000 (low magnification), (g) FMO@C-1000 (high magnification), and the EDS (h) Mn and (i) Fe element mappings.
Materials 18 05336 g002
Figure 3. The RL curves with different thicknesses in the frequency range of 2–18 GHz: (a) FMO, (b) FMO@C-600, (c) FMO@C-700, (d) FMO@C-800, (e) FMO@C-900, and (f) FMO@C-1000.
Figure 3. The RL curves with different thicknesses in the frequency range of 2–18 GHz: (a) FMO, (b) FMO@C-600, (c) FMO@C-700, (d) FMO@C-800, (e) FMO@C-900, and (f) FMO@C-1000.
Materials 18 05336 g003
Figure 4. Dielectric constant and permeability of FMO@C heated at different temperatures: (a) real part of permittivity, (b) imaginary part of permittivity, (c) the dielectric loss tangent, (d) real part of permeability, (e) imaginary part of permeability, and (f) the magnetic loss tangent.
Figure 4. Dielectric constant and permeability of FMO@C heated at different temperatures: (a) real part of permittivity, (b) imaginary part of permittivity, (c) the dielectric loss tangent, (d) real part of permeability, (e) imaginary part of permeability, and (f) the magnetic loss tangent.
Materials 18 05336 g004
Figure 5. (a) Electrical conductivity of FMO@C, the impedance matching of (b) FMO@C-600, (c) FMO@C-700, (d) FMO@C-FMO@C-800, (e) FMO@C-900, and (f) FMO@C-1000.
Figure 5. (a) Electrical conductivity of FMO@C, the impedance matching of (b) FMO@C-600, (c) FMO@C-700, (d) FMO@C-FMO@C-800, (e) FMO@C-900, and (f) FMO@C-1000.
Materials 18 05336 g005
Figure 6. (a) Raman spectrum and (b) electrical conductivity of PAN-based carbon, FMO@C-1000, and FMO@C-B.
Figure 6. (a) Raman spectrum and (b) electrical conductivity of PAN-based carbon, FMO@C-1000, and FMO@C-B.
Materials 18 05336 g006
Figure 7. The RL curves of (a) FMO@C-B1, (b) FMO@C-B2, and (c) FMO@C-B3.
Figure 7. The RL curves of (a) FMO@C-B1, (b) FMO@C-B2, and (c) FMO@C-B3.
Materials 18 05336 g007
Figure 8. Dielectric constant and permeability of FMO@C-B: (a) real part of permittivity, (b) imaginary part of permittivity, (c) the dielectric loss tangent, (d) real part of permeability, (e) imaginary part of permeability, and (f) the magnetic loss tangent.
Figure 8. Dielectric constant and permeability of FMO@C-B: (a) real part of permittivity, (b) imaginary part of permittivity, (c) the dielectric loss tangent, (d) real part of permeability, (e) imaginary part of permeability, and (f) the magnetic loss tangent.
Materials 18 05336 g008
Figure 9. The values of the attenuation constant for FMO@-1000 and FMO@C-B in the frequency range of 2.0–18.0 GHz.
Figure 9. The values of the attenuation constant for FMO@-1000 and FMO@C-B in the frequency range of 2.0–18.0 GHz.
Materials 18 05336 g009
Figure 10. The impedance matching of (a) FMO@C-B1, (b) FMO@C-B2, and (c) FMO@C-B3.
Figure 10. The impedance matching of (a) FMO@C-B1, (b) FMO@C-B2, and (c) FMO@C-B3.
Materials 18 05336 g010
Figure 11. Schematic illustration of EMW absorption mechanism of FMO@C-B.
Figure 11. Schematic illustration of EMW absorption mechanism of FMO@C-B.
Materials 18 05336 g011
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

Zhang, Y.; Shen, S.; Li, J. Construction of Graphite Shells on Ferromanganese Oxide for Electromagnetic Wave Absorption. Materials 2025, 18, 5336. https://doi.org/10.3390/ma18235336

AMA Style

Zhang Y, Shen S, Li J. Construction of Graphite Shells on Ferromanganese Oxide for Electromagnetic Wave Absorption. Materials. 2025; 18(23):5336. https://doi.org/10.3390/ma18235336

Chicago/Turabian Style

Zhang, Yuxiang, Shuling Shen, and Jing Li. 2025. "Construction of Graphite Shells on Ferromanganese Oxide for Electromagnetic Wave Absorption" Materials 18, no. 23: 5336. https://doi.org/10.3390/ma18235336

APA Style

Zhang, Y., Shen, S., & Li, J. (2025). Construction of Graphite Shells on Ferromanganese Oxide for Electromagnetic Wave Absorption. Materials, 18(23), 5336. https://doi.org/10.3390/ma18235336

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