Next Article in Journal
Combining Functional Units to Design Organic Materials with Dynamic Room-Temperature Phosphorescence under Continuous Ultraviolet Irradiation
Previous Article in Journal
Comparative Analysis of Structural and Functional Properties of Dietary Fiber from Four Grape Varieties
Previous Article in Special Issue
[FeIIICl(TMPPH2)][FeIIICl4]2: A Stand-Alone Molecular Nanomedicine That Induces High Cytotoxicity by Ferroptosis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Low-Density-Biomass-Carbon Composite Coated with Carpet-like and Dandelion-Shaped Rare-Earth-Doped Cobalt Ferrite for Enhanced Microwave Absorption

1
Department of Physics Science and Technology, Baotou Teacher’s College, Baotou 014030, China
2
Key Laboratory of Industrial Biocatalysis, Ministry of Education, Tsinghua University, Beijing 100084, China
3
Institute of Applied Chemistry, Department of Chemical Engineering, Tsinghua University, Beijing 100084, China
4
College of Science, Inner Mongolia University of Technology, Hohhot 010051, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(11), 2620; https://doi.org/10.3390/molecules29112620
Submission received: 15 April 2024 / Revised: 21 May 2024 / Accepted: 29 May 2024 / Published: 2 June 2024

Abstract

:
A novel low-density composite for the absorption of microwaves was prepared by loading La-doped spinel cobalt ferrite (La-CFO) onto biomass carbon (BC) derived from corn stalks using a hydrothermal method. This composite (La-CFO@BC) not only maintained the advantageous properties of low density and abundant porosity, but also exhibited a unique morphology, with La-CFO displaying a carpet-like structure interspersed with dandelion-shaped particles. The incorporation of La-CFO effectively tuned the electromagnetic parameters of the composite, thereby improving its impedance-matching attributes and its ability to absorb microwave radiation. At a frequency of 12.8 GHz for electromagnetic waves and with a thickness of 2.5 mm, La-CFO@BC demonstrated remarkable performance in microwave absorption, attaining a noteworthy minimum reflection (RLmin) of −53.2 dB and an effective absorption bandwidth (EAB) of 6.4 GHz. Furthermore, by varying the thickness of the La-CFO@BC within the range of 1.0 to 5.5 mm, the EAB could be broadened to 13.8 GHz, covering the entire X-band, the entire Ku-band, and a substantial portion of the C-band. This study demonstrated that La-CFO@BC was a promising alternative for electromagnetic wave attenuation, which offered superior performance in microwave absorption.

1. Introduction

Amidst the rapid evolution of contemporary communication technologies, especially following the rollout of the 5G network, electronic devices have become increasingly prevalent in enhancing human living standards. However, the abundant electromagnetic radiation has intensified the electromagnetic pollution in the environment, which not only provokes electromagnetic interference issues but also poses severe threats to telecommunications security, national defense security, and human health [1,2,3]. Given the complexity of these issues, the progress in the design and synthesis of electromagnetic wave (EM) absorbers has emerged as a pivotal area of research. Specifically, these materials must have features such as thinness, lightweight, broad bandwidth coverage, and high absorption capacity.
In recent years, due to their inherent lightweight nature and specific absorption capabilities, these carbon-based materials have gained increasing attention from researchers. Biomass carbon (BC) has unique pore structures and suitable electrical conductivity, and it is environmentally friendly, low cost, and easily available [4,5,6]. Additionally, the residual N, S, and Si elements in BC can create self-doping effects, making it a potential absorbing material. However, in the absorption of EW by BC the contribution of magnetic loss is relatively minor, and the main contribution arises from dielectric loss, thereby limiting its absorption capacity. Previous research has shown that loading magnetic particles onto the surface of carbon materials is an effective means to enhance magnetic loss in carbon materials [7,8,9,10]. Wang et al. [11] successfully prepared porous jute biochar (PJBC) composed of Fe3O4 nanoparticles using a chemical coprecipitation method. Research has shown that the minimum RL value of Fe3O4/PJBC composite material is −39.5 dB at a frequency of 6.4 GHz. Wang and co-workers [12] prepared a nano porous carbon (PC)@NiFe2O magnetic composite material using pomelo peels as raw material. The composite material demonstrated remarkable microwave-absorption capabilities, achieving the RLmin of −50.8 dB precisely at a frequency of 13.4 GHz. Furthermore, with a thickness of 2.5 mm, it exhibited an absorption bandwidth spanning 4.9 GHz. Liu et al. [13] conducted a significant study where they synthesized Fe3O4/CF (Fe3O4/carbon fiber) composite material. In this synthesis, bagasse waste was utilized as the primary raw material. The composite material exhibited a maximum RL value of −48.2 dB at 15.6 GHz, simultaneously achieving a wide EAB of 5.1 GHz at a thickness of 1.9 mm. Among various magnetic substances, cobalt ferrite (CoFe2O4) stands out for its stable chemical and excellent magnetic properties [14,15]. Additionally, it has been confirmed that the magnetic loss of iron oxide can be improved by doping it with rare-earth ions [15,16,17]. The slight addition of rare-earth ions modifies the coercivity, magnetic crystal anisotropy field, and diffusion activation energy of ferrite crystals, leading to an increase in magnetic-hysteresis loss. Rare-earth doping also boosts the natural resonance absorption peak and domain wall resonance absorption peak of ferrite, alters peak positions, and broadens absorption peaks, thereby enabling regulation of the material’s ability to absorb EW.
However, investigations into the composite of rare-earth-doped magnetic particles with BC and their synergistic effect on the material’s absorption performance have been rarely reported. Therefore, in this work, we have selected corn stalks, one of the most abundant agricultural waste biomass resources in China, as the raw material to use to synthesize BC. Subsequently, we employed a hydrothermal method to load the rare-earth-La-doped magnetic cobalt ferrite, CoLa0.12Fe1.88O4 (La-CFO), onto the BC, resulting in a novel composite microwave absorber, La-CFO@BC. The advantages of this composite material lay in the complementary interactions between cobalt ferrite doped with rare-earth elements and biomass-derived carbon, as well as with the unique carpet-like morphology exhibited by La-CFO, which together enhanced the impedance-matching conditions and attenuation capacity of the composite material. Consequently, La-CFO@BC exhibited exceptional EW attenuation characteristics. Notably, at the La-CFO@BC thickness of 2.5 mm, the composite achieved an RLmin of −53.2 dB for electromagnetic waves, accompanied by an EAB of 6.4 GHz. These remarkable properties rendered the La-CFO@BC composite a promising candidate for microwave-absorption applications.

2. Results and Discussion

2.1. Configuration Analysis

Using the PANalytical Empyrean for X-ray diffraction analysis (XRD), the crystalline configurations of BC, La-CFO@BC, and La-CFO composites were examined. From Figure 1a, it can be observed that biomass-char BC exhibited two broad diffraction peaks near 2θ of 25.2° and 44°. These observed reflections in the X-ray diffraction patterns matched the crystallographic orientations of the graphite’s (002) and (101) planes. The wider hump shape indicated that the corn stalk BC obtained in the experiment has a very low graphitization degree, and that the char was mainly present in the form of amorphous carbon. Additionally, seven distinct diffraction peaks were visible for La-CFO and La-CFO@BC samples, which corresponded to the (511), (440), (422), (400), (311), (220), and (111) crystal planes of the spinel ferrite face-centered cubic structure. These results indicated the CoLa0.12Fe1.88O4 and the La-CFO@BC contained structures present in the form of a spinel ferrite face-centered cubic crystal structure with good crystallinity. Furthermore, in the XRD pattern of La-CFO@BC, the absence of any distinctive peaks associated with La compounds suggested that La3+ had been successfully incorporated into the lattice of CoFe2O4.
Figure 1b displays the Raman spectroscopic signatures of BC and La-CFO@BC composites. In both samples, distinct peaks were evident at 1348 cm−1 (corresponding to the D-band) and 1584 cm−1 (associated with the G-band). The D-band is attributed to the vibrational modes stemming from the appearance of amorphous sp³ carbon bonds of disordered sp3 carbon configurations and structural defects within the graphitic. Alternatively, the G-band signifies the vibrational frequencies stemming from the stretching of sp2 carbon configurations in the ordered lattice framework. Consequently, ID/IG, which is the ratio of the intensities of two peaks, is a widely used indicator of the degree of graphitization and structural defects in the material [18,19]. Upon analysis, it has been determined that BC exhibited ID/IG values of 0.93, whereas the corresponding values for La-CFO@BC were found to be 1.05. The higher ID/IG ratio of La-CFO@BC indicated lower graphitization due to the increased defects and disrupted carbon-atom ordering during the activation process and during La-CFO loading. The introduced defects within the material can function as potent centers for electrode polarization within electromagnetic fields, thereby augmenting the electric dipole moments and enhancing the dielectric losses of the material towards electromagnetic waves [20,21]. Additionally, a pinning effect can be exerted on the magnetic moment reversal by these defects, making the reversal process more challenging, which consequently leads to an increase in magnetic hysteresis losses [22]. Furthermore, the presence of defects significantly impacts the conductivity of materials, as previously reported [23,24]. Specifically, defects within BC materials have the potential to increase the resistance of their inherent conductive networks, thereby enhancing the thermal effects associated with current flow and ultimately leading to increased conductivity losses. Collectively, these attributes contribute towards the enhanced electromagnetic wave-absorption capabilities of the La-CFO@BC.
The microscopic structures of BC, La-CFO and the composite La-CFO@BC were examined using the SEM as well as the TEM, as illustrated in Figure 2. According to Figure 2a,b, BC retained the 3D-network-porous-cell-wall structure with abundant pore-size distribution, mainly ranging from micro to sub-micro scales. The pore walls covered a wide range of thicknesses and, importantly, they contained numerous thin-layer curled and corrugated distributions. This special structure can enhance the specific surface area of the material, thereby providing a larger interface for potential applications. Additionally, the porous structure promoted the development of an interconnected three-dimensional (3D) conductive network. Meanwhile, it provided expanded regions for electromagnetic wave reflection and scattering. These characteristics were beneficial for improving the material’s microwave-absorption performance at different wavelengths. Figure 2b–d presents SEM images of La-CFO at different magnifications. It can be seen that La-CFO accumulated and was distributed in a granular manner, with particle sizes ranging from 700 nm to 1 μm. Additionally, the particle’s surface was covered with fine fluffy protrusions, giving the entire particle a dandelion-like morphology. As depicted in Figure 2e,f, the La-CFO@BC maintained a 3D-network-porous structure (the pore walls were rendered in green in the picture), and the surface of BC was densely coated with carpet-like, fluffy La-CFO interspersed with embedded dandelion-shaped La-CFO particles. Notably, the size of these embedded La-CFO particles was smaller compared to those in the pure La-CFO samples, with a distribution primarily ranging from 200 nm to 500 nm.
Figure 2j,k are the TEM images of BC, which clearly showed an abundance of microporous and mesoporous structures within its structure. Figure 2l,m are the TEM images of La-CFO@BC. It can be observed that there are also abundant microporous and mesoporous pore structures in La-CFO@BC. Among them, due to the relatively large thickness of the composite material, the pore structure was not obvious in the TEM image labeled (l). In addition, it can be observed that the surface of BC was covered with carpet-like La-CFO fluffy protrusions, and that the surface of dandelion-shaped La-CFO particles also exhibited abundant fluffy protrusions. The distinctive morphology of the material significantly contributed to increasing its specific surface area, which strengthened the surface polarization and improved the absorption of EW. Furthermore, this unique morphology played a pivotal role in modulating the surface impedance of BC, adding to its overall absorption functionality. The complete coverage of La-CFO on the BC surface effectively guided the electromagnetic waves towards the material’s interior, thus improving the internal absorption. This dual-faceted approach could significantly boost the material’s overall electromagnetic wave-absorption capabilities as discussed in the later sections.
To investigate the detailed surface area and porosity profiles of BC and La-CFO@BC, we conducted measurements using nitrogen (N₂) adsorption/desorption isotherms. As depicted in Figure 3a,b, both BC and La-CFO@BC exhibited combined features of Type II and Type IV isotherms, indicative of a coexistence of microporous and mesoporous architectures. Compared to BC, the adsorption capacity of La-CFO@BC was significantly increased. This was due to the presence of carpet-like fluffy and dandelion-shaped La-CFO, which strengthened the surface-adsorption capability of La-CFO@BC. Additionally, the gaps between BC and La-CFO further contributed to the enhanced pore-adsorption capacity of La-CFO@BC. Notably, La-CFO@BC exhibits a more pronounced hysteresis loop compared to BC. This observation could be attributed to the formation of capillary condensation between the micropores in BC and the gaps between BC and La-CFO, making desorption more difficult.
The pore-size distribution curve presented in Figure 3a,b was derived from the analysis of the adsorption data within the respective isotherms. It can be observed that both BC and La-CFO@BC had hierarchical pore structures composed of micropores, mesopores, and macropores at the microscopic level. Specifically, BC was primarily characterized by micropores, which were mainly distributed around 2 nm. La-CFO@BC was dominated by a combination of micropores and mesopores, exhibiting primary distributions centered at 2 nm and 28 nm, respectively. Overall, there was a significant increase in the total pore volume of La-CFO@BC compared to BC. Using the BET method, the specific surface areas of BC were found to be 753.5 m2/g, whereas for La-CFO@BC it was 989.6 m2/g. Such large specific surface areas were beneficial for enhancing the material’s electrode polarization and multiple reflections of EW, ultimately contributing to improved microwave-absorption properties of the materials.
The types, chemical states, and electronic structures of surface elements in the sample were studied using XPS spectroscopy. Figure 4a shows the survey XPS spectrum of La-CFO@BC, where eight elements, including C, O, N, S, Si, Co, Fe, and La could be identified. Figure 4b,c displays the C 1s and O 1s XPS spectra of La-CFO@BC. By using the Lorenz–Gaussian curve-fitting method, the C 1s spectral peak can be decomposed into three distinct sub-peaks, corresponding to the O-C=O, C-O/C-N, and C–C/C=C chemical bonds of the carbon element in BC, which are positioned at 288.71 eV, 285.62 eV, and 284.60 eV [19,21,25]. Similarly, the O 1s spectrum peak could be fitted into three sub-peaks: the one at 529.48 eV corresponding to the lattice O2− ions in La-CFO, and the other two sub-peaks at 533.01 eV and 531.12 eV were identified as representing the C=O and C-O bonds originated from carbonized residual oxygen elements in BC. Furthermore, it can be deduced that the residual elements of N, S, and Si observed in the spectra were remnants after the carbonization of BC. These residual elements in BC had high polarity and tended to form more polar functional groups [26,27,28]. The presence of polar functional groups has been favorably documented as enhancing the absorption proficiency of the material [29]. Consequently, BC may have good absorption properties. Furthermore, the presence of these residual elements has the potential to increase the surface activity of biomass carbon by generating additional surface-active groups and elevating its surface energy [30,31]. This increase subsequently facilitates the interaction of the biomass carbon with electromagnetic waves, thereby improving its wave-absorption capabilities.
Figure 4d is the detailed XPS spectrum of La-CFO@BC for Fe 2p. Due to spin-orbit splitting, two dominant peaks were observed at 725.70 eV and 711.74 eV. The spectral features align with the spin-orbit split components of Fe 2p1/2, and Fe 2p3/2, respectively. Additionally, the observation of secondary peaks situated at elevated binding energies, approximately 6 eV above the primary peaks, indicates the existence of Fe3+ ions in their oxidized state, as reported in references [32,33]. Furthermore, the Fe2p peak exhibits an asymmetric shape, indicating that Fe exists in multiple chemical states within the compound. By decomposing the Fe 2p1/2 as well as Fe 2p3/2 peaks, two sub-peaks were obtained, corresponding to Fe3+ ions occupying octahedral interstitial sites (Os) and tetrahedral interstitial sites (Ts) in spinel cobalt ferrite. Table 1 provides the binding-energy positions of the Fe3+ sub-peaks. To ascertain the proportional percentage composition of Fe3+ ions across the two distinct states, we analyzed the integral area ratio of the corresponding sub-peaks, yielding values of 71.96% for the Os state and 28.04% for the Ts state, respectively.
Figure 4e displays the detailed XPS spectrum of La-CFO@BC for Co 2p. Two dominant peaks were observed. Due to spin-orbit splitting, the one at 796.43 eV was attributed to the Co 2p1/2 spin-orbit component, whereas another one at 780.60 eV corresponds to the Co 2p3/2 spin-orbit component. The presence of strong high-energy side peaks at 802.10 eV and 785.75 eV indicates the oxidation state of Co as Co2+ [34,35]. By decomposing the Co 2p1/2 and Co 2p3/2 peaks, two sub-peaks were obtained, corresponding to Co2+ ions occupying Os and Ts. Table 1 provides the binding-energy positions of the Co2+ sub-peaks. The computed percentages of Co2+ ions in the two states were 42.08% for state Os and 56.92% for state Ts, respectively, indicating that La-CFO existed as a mixed phase of normal and inverse spinel structures.
Figure 4f shows the detailed XPS spectrum of La-CFO@BC for La 3d. The La 3d spectrum exhibited four peaks. The detection of two distinct XPS peaks, originating from the La 3d3/2 and La 3d5/2 energy levels, was observed at 851.46 eV and 834.71 eV, respectively. The results align well with previously reported positions of La3+ ions [36,37]. In addition, the other two peaks resulted from shake-up satellite features due to interactions between La 4f energy levels and O 2p [38]. These facts all indicated that La in La-CFO@BC exists as La3+ [15,39]. Decomposition of La 3d5/2 and La 3d3/2 peaks yielded four sub-peaks corresponding to La3+ ions occupying Os and Ts. Table 1 provided the binding-energy positions of the La3+ sub-peaks. The calculated occupancy ratios of La3+ in Os and Ts were 53.31% and 46.69%, respectively. In this way, a certain proportion of Fe3+, Co2+, and La3+ form a disordered distribution at the octahedral position, and electrons could quickly transfer between the oxidized states of metal ions, making La-CFO exhibit good conductivity. Within a certain range, as the conductivity increased, the dielectric constant of the material also increased accordingly, which can give an adjustment of the electromagnetic parameters to the material, so that its absorption performance is enhanced.

2.2. Magnetic Properties

The microwave-absorption performance of a sample is intricately intertwined with its magnetic characteristics. During the absorption of EW, the magnetic components in the material can effectively guide and absorb the electromagnetic waves [40]. Figure 5 presents the magnetic hysteresis loops of La-CFO and La-CFO@BC. Both samples exhibited an S-shaped magnetic hysteresis loop, indicating that they were typical ferromagnets. The coercivity (Hc) and magnetic saturation (Ms) values of La-CFO were 378.26 Oe and 49.38 emu/g, respectively. However, La-CFO@BC exhibited significantly reduced values of 264.88 Oe and 19.66 emu/g. This is because there is a non-magnetic component BC in the sample, which results in a relatively small contribution of La-CFO to the magnetic properties per unit mass of the sample. The decrease in coercivity and magnetic saturation of La-CFO@BC allowed microwave energy to be more easily absorbed and transformed into various energy forms, including thermal energy and others. This can improve energy conversion efficiency, which may enable La-CFO@BC to achieve better microwave-absorption performance. This finding highlighted the potential application of La-CFO@BC in materials science and engineering due to its unique chemical and physical properties.

2.3. Microwave-Absorption Characteristics

The material’s EW absorption capability is commonly quantified through reflection loss (RL) [7,41]. According to the theory of reflection transmission, RL refers to the loss of energy that is partially reflected back but not absorbed by EW entering a material. Ideally, for efficient absorption, the incident wave should penetrate maximally into the absorber, where it can be effectively dissipated rather than reflected. Consequently, the magnitude of RL directly correlates with the absorber’s efficacy in absorbing electromagnetic waves. A larger absolute RL value signifies less reflection and a greater proportion of the wave penetrating into the absorber, indicating superior absorption performance [15,42]. The RL value could be accurately determined by application of a specified mathematical formulation, grounded in the principles of transmission line theory [43]:
RL(dB) = 20 log|(Z0Zin)/(Z0 + Zin)|,
where Z0 represents the fundamental impedance for surrounding vacuum or air, whereas Zin signifies the resistive property encountered by an incoming wave at the interface of the microwave-absorption material. Typically, the input impedance (Zin) of the microwave-absorption material can be derived using the mathematical expression outlined below:
Zin = Z0 (μr/εr)1/2 tanh[j(εrμr1/2 (2πdf/c)]
where the thickness of the absorption layer, denoted by d, along with the velocity of EW in a vacuum, c, and the frequency of the incoming EW, f, play vital roles in determining the performance of the absorber. Additionally, the complex permittivity (εr) and complex magnetic permeability (μr) of the absorber material are crucial parameters in characterizing its electromagnetic properties. Therefore, for a given incident frequency of electromagnetic waves, optimal impedance-matching and absorption effects could be attained through adjusting the absorbing material’s d.
The electromagnetic parameters of the samples are tested by the coaxial method using a vector network analyzer. The paraffin wax and the sample to be tested were mixed in a weight ratio of 7:3. Subsequently, the mixed material was compressed into a circular ring-shaped specimen, possessing an outer diameter measuring 7 mm, an inner diameter of 3 mm, and a thickness precisely set at 2 mm. The electromagnetic parameters ε′, ε″, μ′, and μ″ of the material are tested for the sample within the frequency range of 2–18 GHz. Figure 6 shows the actual photos of the VAM instrument connection, the ring-shaped tested sample, and the coaxial fixture for the tested sample.
With the values of ε′, ε″, μ′, and μ″ obtained through experiments, the reflection loss RL of each sample was derived through Equations (1) and (2). The RL curves, along with their corresponding 3D distribution maps and 2D RL contour plots were given in Figure 7. Firstly, it could be observed that the RL of all samples reached its peak at a certain frequency, regardless of their thickness. And with an increase in thickness, the corresponding minimum peak frequency of the RL spectrum gradually decreases. The quarter-wavelength resonance model, described by the given Equation (3), can explain this observed phenomenon [25,44]:
dm = nc/4fm (|μr||εr|)1/2
In this equation, the parameters dm and fm designate the respective thickness and frequency of the incident EW, at which the absorbing material attains its RL peak. Therefore, dm and fm exhibit a negative correlation, indicating that as dm increases, fm decreases.
As depicted in Figure 7, it is evident that the RL peak of each sample attains a minimum value RLmin. For example, at a frequency of 14.02 GHz, with a 2 mm thickness, sample BC achieved an RLmin of −19.31 dB. This result was similar to that recently reported on absorption effects of biomass carbon (the RLmin of carbonized waste coffee grounds was −20.89 dB at 13.92 GHz, 2.0 mm) [20]. It is noteworthy to mention that the EAB is defined as the frequency range wherein the RL value diminishes to less than −10 dB, signifying that a substantial fraction, exceeding 90% of the electromagnetic wave energy is dissipated within that range. A wider EAB indicated a better frequency absorption range and more practical electromagnetic-wave-absorption capabilities. As shown in the Figure 7a–c, sample BC had an EAB of 4.41 GHz, which was also similar to previously reported results (the RLmin of waste coffee grounds was −20.89 dB at 13.92 GHz, 2.0 mm) [20]. According to Figure 7d–f, at a frequency of 14.02 GHz and with a thickness of 2 mm, La-CFO achieved a minimum RL of −30.32 dB and possessed an EAB of 4.94 GHz. This was a significant improvement compared to pure CoFe2O4 particles (at 13.4 GHz and 2.5 mm, RLmin of P-CFO was about −9.91 dB) [15].
As shown in the Figure 7g–i, after loading La-CFO onto BC, the sample La-CFO@BC exhibited a remarkable microwave-absorbing performance, attaining a RLmin value of −53.19 dB at a frequency of 12.80 GHz and a thickness of 2.5 mm, while demonstrating an EAB spanning 6.49 GHz. Table 2 compares some recently reported biomass-carbon-absorbing materials, showing that the synthesized La-CFO@BC magnetic biomass material in this work had certain advantages in terms of absorption amount and EAB. It indicated that La-CFO@BC was a highly promising absorbing material for practical applications.
Overall, the BC provided effective EW absorption in the high frequency range, whereas the magnetic particles played a role in the low and medium frequency ranges. This synergistic effect enabled hybrid materials to effectively absorb electromagnetic waves over a wider frequency range.
The complex permittivity εr = ε′ − ″ and complex permeability μr = μ′ − ″ of a material are closely related to its wave-absorbing performance. The real components ε′ as well as μ′ represent the capacity of the material to store the EW energy. Meanwhile, the imaginary components ε″ and μ″ denote the dissipation capacity of the EW energy, within the material [54,55]. Secondly, when the EW is incident to the material surface, results from the disparity in impedance between the absorber and the surrounding medium, a segment of the EW is reflected to free space but cannot enter the absorbing body [56,57]. The impedance behaviors of the material are dependent on its εr and μr. To analyze the absorption mechanism of each sample, Figure 8 gives the frequency-dependent εr and μr of BC, La-CFO and La-CFO@BC using a vector analyzer.
Figure 8a,b illustrates the dielectric parameters ε′ and ε″ of BC, which were both higher than those of La-CFO. This means that the BC had a stronger dielectric-loss capability compared to that of La-CFO. Moreover, the ε′ value together with the ε″ value of La-CFO@BC were lower than those of BC alone. This could be ascribed to the introduction of La-CFO with relatively lower ε′ and ε″ values. It was also noticed that ε″ exhibited a decrease as the frequency rose, a trend that can be traced back to the reduction in space-charge polarization and the consequent decrease in dielectric loss as frequency increases. Additionally, as shown in Figure 8c,d, BC exhibited very low values of μ′ and μ″ compared to those of La-CFO. This was because BC was a non-magnetic material exhibiting minimal responses to magnetic fields. This observation indicated that BC had a weaker magnetic-loss capability compared to that of La-CFO. The relatively high values of μ′ and μ″ in La-CFO contributed to the moderate μ′ and μ″ values in the composite material La-CFO@BC. This not only enhanced the magnetic-loss capability of La-CFO@BC but also facilitated the impedance matching.
Figure 9a,b depicts the relationships between frequency and the tanδe = ε″/ε′ as well as tanδm = μ″/μ′ of BC, La-CFO, and La-CFO@BC, which are usually used to evaluate dielectric and magnetic losses, respectively [44]. As shown in Figure 9a, La-CFO@BC exhibits the highest tanδe value, indicating its high dielectric-loss capability. The observed phenomenon can be ascribed to the thin, layered, and corrugated carbon pore walls within BC, which were serving as conduits that facilitate the formation of a conductive network and were ultimately bolstering the electrical conductivity of La-CFO@BC. Additionally, the pore walls of BC were densely covered with pores of various sizes, ranging from micro to macro, offering an extensive surface area that permitted BC to carry a large amount of La-CFO. After loading, La-CFO@BC possessed a large interface between BC and La-CFO, providing a strong interfacial-polarization capability. In addition, the La-CFO deposited on the surface of BC exhibited both carpet-like and dandelion-shaped morphologies, leading to a significant expansion of the overall surface area of the material. This, in turn, increased the interface between the absorber and its surroundings. As a result, a significant presence of dangling bonds and coordination-unsaturated atoms emerged on the absorber surface, thereby intensifying electronic, ionic, and interfacial-polarization mechanisms within the material. Consequently, La-CFO@BC exhibited a higher electrical polarization rate. Typically, high polarization rates and electrical conductivities result in high dielectric losses [58,59], leading to the higher tanδe = ε″/ε′ value observed for La-CFO@BC. In addition, reference [60] presents a phenomenological hypothesis regarding the EW absorption capacity of conductive composites, suggesting the existence of a voltage-controlled Fowler-Nordheim tunneling effect. The tunneling phenomenon exerts an influence on the composite material’s effective conductivity via the Drude damping factor, thereby generating alternative current flow routes within the non-conductive resin matrix. Consequently, these extra current paths effectively transform the resin into a virtually conductive medium, thereby altering the effective dielectric constant of the two-phase composite material. In our composite material, La-CFO@BC, numerous gaps exist during the combination of BC and La-CFO. It is possible that a tunneling effect occurs within these gaps, influencing the composite’s effective conductivity through the Drude damping factor and modifying its effective dielectric constant. This, in turn, has the potential to enhance the material’s absorption of electromagnetic wave energy.
It can also be observed in Figure 9b that La-CFO exhibited the highest tanδm value. This phenomenon stemmed from the intrinsic properties of CFO ferrite materials, including significant eddy current loss, magnetic aftereffect loss, magnetic hysteresis loss, domain wall resonance loss, and natural resonance-loss capabilities. Moreover, the incorporation of the element La increased the coercivity and magnetic anisotropy field of the CFO ferrite crystal, thereby enhancing its magnetic hysteresis loss. Furthermore, compared to Fe and Co ions, the larger radius of La ions exerted a pinning effect on domain wall motion, enhancing the natural resonance absorption and the domain wall resonance absorption of CFO ferrite. Additionally, the rare-earth element La had a numerous atomic energy level. When it was doped into the ferrite crystal, it increased the density of material energy levels and also enhanced the opportunity for energy level transitions. This increased energy transfer between ions and improved the multi-band energy absorption capability of the absorbing material. Compared to BC and La-CFO, the sample La-CFO@BC exhibited an intermediate tanδm value while maintaining a similar trend to La-CFO. It indicated that La-CFO@BC possessed a certain magnetic loss capability and exhibited magnetic-loss patterns similar to those of La-CFO. By comparing the tanδm plot of La-CFO@BC (Figure 9b) with its reflection loss plot (Figure 7g), it could be observed that the tanδm of the sample peaked at a frequency of 13.20 GHz, which was very close to the frequency corresponding to its minimum loss peak at 12.80 GHz. Therefore, it could be concluded that the improvement in microwave-absorption capability of La-CFO@BC was predominantly attributed to the augmented magnetic loss.
In addition, the Debye relaxation model provides the basis for Equation (4) [61],
ε ε + ε s / 2 2 + ε 2 = ε ε s / 2 2
where εs represents the stationary dielectric permittivity, whereas ε denotes the limiting relative dielectric constant at high frequencies. Figure 10 presents the Cole–Cole plot of ε″ as a function of ε′. Typically, the presence of a semicircle on the Cole–Cole plot signifies the occurrence of a polarization-relaxation mechanism, whereas the emergence of multiple semicircles denotes the concurrent occurrence of numerous polarization-relaxation processes. As can be seen from Figure 10, there are numerous Cole–Cole semicircles in all three samples, with La-CFO and La-CFO@BC showing more semicircles, indicating that multiple polarization-relaxation processes existed in all samples, with La-CFO and La-CFO@BC having more polarization-relaxation processes. Additionally, it could be observed that the Cole–Cole curves of BC and La-CFO@BC exhibited a linear long tail, indicating the presence of strong conductive losses [19,62]. This phenomenon could potentially originate from the conductive losses associated with the conductive network established by BC.
According to the theory of transmission lines, when incident EW in free space reaches the absorbing body, they may be partially reflected due to impedance mismatch. To facilitate the efficient absorption of EW energy by the absorbing body, it is imperative to minimize the reflection of EW and maximize its penetration into the interior of the absorbing material. This requires the establishment of an optimal impedance match between the absorber and the surrounding free space. The effect of impedance matching is usually evaluated using the parameter Z (Equal to 1 in ideal matching), which can be obtained by converting Equation (5) from Equation (2).
Z = Z in Z 0 = ( μ r ε r ) 1 / 2   tanh [ j ( ε r μ r )   1 / 2   ( 2 π d f c ) ]
The thicknesses of BC, La-CFO, and La-CFO@BC were fixed as 2.5 mm, so that their Z values were calculated based on Equation (5). Figure 11a shows the Z-f curve of each sample. One can observe that, compared to the other two samples, the Z value of La-CFO@BC was closer to 1, and the spectral band corresponding to the Z value close to 1 was broader. Therefore, EM waves more easily penetrated into the interior of La-CFO@BC absorbers than into the other materials. This result indicated that the EM parameters of the hybrids could be adjusted by rational composite mixing of BC with La-CFO to achieve superior impedance-matching characteristics [63]. Another significant characteristic parameter that characterizes the performance of the absorber is its ability to absorb electromagnetic waves internally, denoted as α. This parameter α can be expressed mathematically using Equation (6) [64]:
α = 2 π f / c × μ ε μ ε + ( ε μ + μ ε ) 2 + ( ε μ ε μ ) 2 1 / 2
Referring to Figure 11b, the frequency-dependent α-value curves of the three samples, which were computed based on electromagnetic parameters, are depicted. It is observable that the α-value of La-CFO@BC was generally higher than that of BC and La-CFO. Based on the previous analysis, the high α-value of La-CFO@BC could be attributed to the strong interface polarization between BC and La-CFO, in conjunction with the collaborative actions of dipole polarization, electron transport, and magnetic resonance, ultimately enhancing the EW absorption capacity of La-CFO@BC.
Based on the above experimental data and on discussions considering the unique morphology of La-CFO@BC, the potential absorption process of the La-CFO@BC was illustrated in Figure 12. The thin, corrugated carbon pore walls of La-CFO@BC form a 3D=-interconnected conductive network. The presence of numerous micropores and mesopores on these pore walls, along with defects introduced by the loading of La-CFO, provide appropriate resistance and capacitance to the conductive network. Under the influence of incident electromagnetic waves, this 3D conductive network can act as a resistive-inductive–capacitive coupling circuit, generating time-varying electromagnetic induction currents. These complex, long-distance induction currents decay as they propagate through the 3D network, converting electromagnetic energy into thermal energy, thus contributing to electrical-conduction loss. Moreover, La-CFO@BC exhibits a diverse porosity distribution, allowing EW to undergo numerous reflections in the material’s micropores, mesopores, and macropore channels, leading to multiple reflection phenomena. This increases the depth and area of electromagnetic wave propagation within the absorbing material, enabling the absorption of as much electromagnetic energy as possible. Furthermore, the integration of BC and La-CFO enhances the composite’s overall impedance-matching profile, balancing its magnetic-loss capacity and dielectric-loss capacity. From another perspective, the rich polarization enhances absorber’s dielectric-loss ability. Firstly, the formation of numerous defects during the loading of La-CFO onto BC promotes the generation of internal dipoles, resulting in rich dipole polarization. Secondly, the numerous contact interfaces between BC and La-CFO provide abundant interfacial polarization for the material. Finally, the carpet-like and dandelion-like surface textures of La-CFO deposited on the material provide significant surface polarization. Therefore, the doping with the rare-earth element La enhances the magnetic-loss capacity of the composite material, broadening its effective absorption frequency bandwidth. The unique carpet-like and dandelion-like morphologies of La-CFO also contribute to enhancing the magnetic-loss capacity of the material, thus playing important roles in absorbing magnetic field energy. Based on these characteristics, La-CFO@BC exhibits excellent absorption properties.

3. Materials and Methods

3.1. Preparation of the Composites

Firstly, the treated corn stalks were heated at a rate of 10 °C per minute in a vacuum tube furnace under N2 atmosphere to the constant temperature of 450 °C, and then calcined for 3 h. Once the temperature had gradually and naturally subsided to ambient room temperature, the carbonized corn stalks (BC) were taken out. The BC was then activated and treated. Subsequently, 1.5 g of the treated BC, 1.3 g of NaAc (Phygene, Fuzhou, China), and 1 g of polyethylene glycol (Aladdin, Shanghai, China) were added into 70 mL of ethylene glycol (EG, Aladdin, Shanghai, China) and stirred for 10 min. Then, according to the stoichiometric ratio of CoLa0.2Fe1.8O4, a total amount of 2.0 g of Fe(NO3)3·9H2O, La(NO3)3·6H2Oand Co(NO3)3·6H2O (Aladdin, Shanghai, China) were introduced to the solution. The mixed solution was magnetically stirred for another 10 min and sonicated for 30 min, then poured into a stainless-steel autoclave liner, which was kept at 140 °C for 24 h, subsequently reaching ambient temperature naturally. The mixed solution underwent filtration to yield a black substance; after undergoing numerous ethanol rinses, the material was then dehydrated in an oven kept at a temperature of 80 °C, ultimately resulting in La-CFO@BC (the weight ratio of the CoLa0.2Fe1.8O4 in the La-CFO@BC was 20%). In addition, for comparative experiments, a similar procedure was employed to synthesize pure CoLa0.2Fe1.8O4 without loading on BC, which was labeled as La-CFO.

3.2. Characterization

The synthesized samples were subjected to various analytical techniques to measure their physical and chemical properties. X-ray diffraction methodology (XRD, PANalytical Empyrean, Almelo, The Netherlands) was applied to determine the phase structure of the product, with a scan range set from 10° to 80° in 2θ. Raman spectrometer spectroscopy (RAM, Thermo Fisher, Waltham, MA, USA) was used to analyze the molecular structure information of BC. To examine the physical appearance and microscopic characteristics of the product outcomes, scanning electron microscopy (SEM, Thermo Scientific Apreo 2C, Waltham, MA, USA) as well as transmission electron microscopy (TEM, FEI Tecnai F20, Hillsboro, OR, USA) were utilized. The chemical configurations of the product were examined through the utilization of X-ray photoelectron spectroscopy (XPS, Thermo Fisher Escalab 250 xi, Waltham, MA, USA). The XPS spectra were calibrated with reference to the C 1s signal at 284.6 eV. The N2 adsorption–desorption isotherm analysis was conducted utilizing a specialized surface area and pore-size analyzer at −196 °C (BET, Micromeritics Asap 2020, Norcross, GA, USA). An assessment of the magnetic characteristics of the specimens was conducted using a physical property measurement system (PPMS, Quantum Design Inc., San Diego, CA, USA). Using a vector network analyzer (VNA, Agilent E5071c, Santa Clara, CA, USA), the electromagnetic parameters of the sample were measured utilizing the coaxial method within the frequency range of 2–18 GHz. For the purpose of conducting the tests, the paraffin wax and the sample were blended together in a ratio of 7 parts wax to 3 parts sample, by weight. The blended mixture was then compressed into circular rings, featuring an internal diameter of 3 mm, an external diameter of 7 mm, and a height of 2 mm.

4. Conclusions

Biomass carbon (BC) was prepared from corn stalks, and rare-earth-doped spinel ferrite La-CFO was directly loaded on BC via a hydrothermal method, resulting in the successful synthesis of a low-density composite for microwave absorption, La-CFO@BC. The composite not only retained the characteristics of low density and abundant porosity of typical biomass-carbon materials but also exhibited a unique morphology. Specifically, La-CFO covering the BC surface displayed a carpet-like structure with dandelion-shaped La-CFO particles embedded within it. The loading of La-CFO modulated the electromagnetic parameters of the composite, addressing the weakness of poor magnetic loss in pure BC materials. As a result, La-CFO@BC exhibited superior impedance-matching characteristics and displayed outstanding microwave-absorption capability. Under the conditions of an incident EW frequency of 12.8 GHz, the RLmin value of La-CFO@BC reached −53.2 dB at a thickness of 2.5 mm, with an EAB of 6.4 GHz. By varying the thickness of the composite between 1.0 and 5.5 mm, the effective absorption bandwidth of La-CFO@BC could be extended to 13.8 GHz, encompassing the entire X-band, the entire Ku-band, and most of the C-band electromagnetic waves. The findings of this work may thus provide a substantial advancement in the field, offering a promising alternative to conventional microwave absorbers.

Author Contributions

Conceptualization, T.S.; supervision, T.S.; writing—original draft, T.S.; investigation, X.Z.; methodology X.Z.; visualization, H.Z.; validation H.Z.; software, Y.S.; resources, R.W.; formal analysis, R.W.; writing—review and editing, T.S. and X.Z.; funding acquisition, T.S. and X.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, (No. 52362013, 22178197), the Natural Science Foundation of The Inner Mongolia Autonomous Region (No. 2021MS01015), and the High-level Talent Introduction Startup Fund of Baotou Teacher’s College (No. BTTCRCQD2019).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data of the compounds are available from the authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ding, C.; Shao, C.; Songsong, W.; Ma, Y.; Liu, Y.; Ma, S.; Hu, X.; Cao, Z.; Ren, X.; Zhong, B. A review of 1D carbon-based materials assembly design for lightweight microwave absorption. Carbon 2023, 213, 118279. [Google Scholar] [CrossRef]
  2. Zhao, Z.; Qing, Y.; Kong, L.; Xu, H.; Fan, X.; Yun, J.; Zhang, L.; Wu, H. Advancements in microwave absorption motivated by interdisciplinary research. Adv. Mater. 2023, 36, 2304182. [Google Scholar] [CrossRef]
  3. Xiang, L.; Qi, X.; Rao, Y.; Wang, L.; Gong, X.; Chen, Y.; Peng, Q.; Zhong, W. A simple strategy to develop heterostructured carbon paper/Co nanoparticles composites with lightweight, tunable and broadband microwave absorption. Mater. Today Phys. 2023, 34, 101030. [Google Scholar] [CrossRef]
  4. Guan, H.; Wang, Q.; Wu, X.; Pang, J.; Jiang, Z.; Chen, G.; Dong, C.; Wang, L.; Gong, C. Biomass derived porous carbon (BPC) and their composites as lightweight and efficient microwave absorption materials. Compos. Part B 2021, 207, 108562. [Google Scholar] [CrossRef]
  5. Shi, Q.; Zhao, Y.; Li, M.; Li, B.; Hu, Z. 3D lamellar skeletal network of porous carbon derived from hull of water chestnut with excellent microwave absorption properties. J. Colloid Interface Sci. 2023, 641, 449–458. [Google Scholar] [CrossRef]
  6. Zhang, R.; Qiao, J.; Zhang, X.; Yang, Y.; Zheng, S.; Li, B.; Liu, W.; Liu, J.; Zeng, Z. Biomass-derived porous carbon for microwave absorption. Mater. Chem. Phys. 2022, 289, 126437. [Google Scholar] [CrossRef]
  7. Song, W.; Zhao, Q.; Wang, Z. Magnetic biomass porous carbon@Co/CoO nanocomposite for highly efficient microwave absorption. Mater. Res. Bull. 2023, 167, 112371. [Google Scholar] [CrossRef]
  8. Huang, X.; Wang, Y.; Lou, Z.; Chen, Y.; Li, Y.; Lv, H. Porous, magnetic carbon derived from bamboo for microwave absorption. Carbon 2023, 209, 118005. [Google Scholar] [CrossRef]
  9. Chang, Q.; Xie, Z.; Long, C.; Feng, X.; Shi, B.; Wu, H. Magnetic hierarchically porous biomass carbon for realizing broadband and strong absorption of electromagnetic wave. Ceram. Int. 2023, 49, 27015–27023. [Google Scholar] [CrossRef]
  10. Shang, T.; Lu, Q.; Zhao, J.; Chao, L.; Qin, Y.; Ren, N.; Yun, Y.; Yun, G. Novel three-dimensional graphene-like networks loaded with Fe3O4 nanoparticles for efficient microwave absorption. Nanomaterials 2021, 11, 1444. [Google Scholar] [CrossRef]
  11. Wang, L.; Guan, H.; Hu, J.; Huang, Q.; Dong, C.; Qian, W.; Wang, Y. Jute-based porous biomass carbon composited by Fe3O4 nanoparticles as an excellent microwave absorber. J. Alloys Compd. 2019, 803, 1119–1126. [Google Scholar] [CrossRef]
  12. Wang, Y.; Gao, X.; Zhou, H.; Wu, X.; Zhang, W.; Wang, Q.; Luo, C. Fabrication of biomass-derived carbon decorated with NiFe2O4 particles for broadband and strong microwave absorption. Powder Technol. 2019, 345, 370–378. [Google Scholar] [CrossRef]
  13. Liu, Y.; Chen, Z.; Xie, W.; Song, S.; Zhang, Y.; Dong, L. In-situ growth and graphitization synthesis of porous Fe3O4/carbon fiber composites derived from biomass as lightweight microwave absorber. ACS Sustain. Chem. Eng. 2019, 7, 5318–5328. [Google Scholar] [CrossRef]
  14. Khanahmadi, S.; Masoudpanah, S. Preparation and microwave absorption properties of CoFe2O4/NiCo2O4 composite powders. Ceram. Int. 2023, 50, 9779–9788. [Google Scholar] [CrossRef]
  15. Shang, T.; Lu, Q.; Chao, L.; Qin, Y.; Yun, Y.; Yun, G. Effects of ordered mesoporous structure and La-doping on the microwave absorbing properties of CoFe2O4. Appl. Surf. Sci. 2018, 434, 234–242. [Google Scholar] [CrossRef]
  16. Sahu, S.; Gurrala, R.C.; Dobbidi, P. Effect of rare earth substitution on electromagnetic and microwave absorption properties of nickel-cobalt nano ferrite. Ceram. Int. 2023, 49, 40875–40893. [Google Scholar] [CrossRef]
  17. Routray, K.L.; Saha, S.; Sanyal, D.; Behera, D. Role of rare-earth (Nd3+) ions on structural, dielectric, magnetic and Mossbauer properties of nano-sized CoFe2O4: Useful for high frequency application. Mater. Res. Express 2018, 6, 026107. [Google Scholar] [CrossRef]
  18. Raji, S.; Sharma, G.K.; Aranya, B.; Prabhakaran, K. Carbon composite foams from the wasted banana leaf for EMI shielding and thermal insulation. Carbon 2023, 213, 118259. [Google Scholar] [CrossRef]
  19. Ning, T.; Li, Q.; Ren, Q.; Wang, J.; Sun, Y.; Zhang, P.; Ma, W.; Yan, J.; Zhai, C.; Zhao, W. Kapok fibers-derived carbon microtubes as efficient electromagnetic wave absorption materials. Ceram. Int. 2023, 49, 29339–29347. [Google Scholar] [CrossRef]
  20. Wen, X.; Li, C.; Liu, H.; Fan, G.; Tang, Y.; Hao, C.; Ma, L.; Song, P. Green carbonization of waste coffee grounds into porous C/Fe hybrids for broadband and high-efficiency microwave absorption. J. Mater. Sci. Technol. 2024, 170, 1–10. [Google Scholar] [CrossRef]
  21. Che, J.; Zheng, H.; Lu, Z.; Yang, Z.; Gao, Y.; Wang, Y. Bimetallic sulfides embedded into porous carbon composites with tunable magneto-dielectric properties for lightweight biomass-reinforced microwave absorber. Ceram. Int. 2023, 49, 27094–27106. [Google Scholar] [CrossRef]
  22. Cao, A.; Shen, B.; Lin, Q.; Chen, S.; Huang, Z.; Ji, Z.; Zhang, Z. Influence of Stone-Wales defect on graphene friction: Pinning effect and wrinkle modification. Comput. Mater. Sci. 2020, 173, 109423. [Google Scholar] [CrossRef]
  23. Harrington, G.F.; Sun, L.; Yildiz, B.; Sasaki, K.; Perry, N.H.; Tuller, H.L. The interplay and impact of strain and defect association on the conductivity of rare-earth substituted ceria. Acta Mater. 2019, 166, 447–458. [Google Scholar] [CrossRef]
  24. Nepal, K.; Ugwumadu, C.; Gautam, A.; Kappagantula, K.; Drabold, D.A. Electronic conductivity in metal-graphene composites: The role of disordered carbon structures, defects, and impurities. J. Phys. Mater. 2024, 7, 025003. [Google Scholar] [CrossRef]
  25. Zhao, Y.; Li, M.; Shi, Q.; Li, B.; Hu, Z.; Li, J.; Wan, X.; Yu, H. Co/C composites generated from biomass exhibit outstanding electromagnetic wave absorption. Diamond Relat. Mater. 2023, 138, 110191. [Google Scholar] [CrossRef]
  26. Leng, L.; Liu, R.; Xu, S.; Mohamed, B.A.; Yang, Z.; Hu, Y.; Chen, J.; Zhao, S.; Wu, Z.; Peng, H. An overview of sulfur-functional groups in biochar from pyrolysis of biomass. J. Environ. Chem. Eng. 2022, 10, 107185. [Google Scholar] [CrossRef]
  27. Chun, Y.; Sheng, G.; Chiou, C.T.; Xing, B. Compositions and sorptive properties of crop residue-derived chars. Environ. Sci. Technol. 2004, 38, 4649–4655. [Google Scholar] [CrossRef] [PubMed]
  28. Hu, B.; Wang, K.; Wu, L.; Yu, S.H.; Antonietti, M.; Titirici, M.M. Engineering carbon materials from the hydrothermal carbonization process of biomass. Adv. Mater. 2010, 22, 813–828. [Google Scholar] [CrossRef] [PubMed]
  29. Jin, H.; Zhou, J.; Tao, J.; Gu, Y.; Yu, C.; Chen, P.; Yao, Z. Commonly neglected ester groups enhanced microwave absorption. Small 2023, 19, 2304536. [Google Scholar] [CrossRef]
  30. Zeng, S.; Li, K.; Xu, X.; Zhang, J.; Xue, Y. Efficiently catalytic degradation of tetracycline via persulfate activation with plant-based biochars: Insight into endogenous mineral self-template effect and pyrolysis catalysis. Chemosphere 2023, 337, 139309. [Google Scholar] [CrossRef]
  31. Janu, R.; Mrlik, V.; Ribitsch, D.; Hofman, J.; Sedláček, P.; Bielská, L.; Soja, G. Biochar surface functional groups as affected by biomass feedstock, biochar composition and pyrolysis temperature. Carbon Resour. Convers. 2021, 4, 36–46. [Google Scholar] [CrossRef]
  32. Yadav, R.S.; Havlica, J.; Masilko, J.; Kalina, L.; Wasserbauer, J.; Hajdúchová, M.; Enev, V.; Kuřitka, I.; Kožáková, Z. Impact of Nd3+ in CoFe2O4 spinel ferrite nanoparticles on cation distribution, structural and magnetic properties. J. Magn. Magn. Mater. 2016, 399, 109–117. [Google Scholar] [CrossRef]
  33. Wang, W.P.; Yang, H.; Xian, T.; Jiang, J.L. XPS and magnetic properties of CoFe2O4 nanoparticles synthesized by a polyacrylamide gel route. Mater. Trans. 2012, 53, 1586–1589. [Google Scholar] [CrossRef]
  34. Chagas, C.A.; de Souza, E.F.; de Carvalho, M.C.; Martins, R.L.; Schmal, M. Cobalt ferrite nanoparticles for the preferential oxidation of CO. Appl. Catal. A 2016, 519, 139–145. [Google Scholar] [CrossRef]
  35. Benedet, M.; Rizzi, G.A.; Barreca, D.; Gasparotto, A.; Maccato, C. XPS analysis of graphitic carbon nitride functionalized with CoO and CoFe2O4. Surf. Sci. Spectra 2023, 30, 014004. [Google Scholar] [CrossRef]
  36. Siddiqui, H.; Shrivastava, M.; Parra, M.R.; Pandey, P.; Ayaz, S.; Qureshi, M. The effect of La3+ ion doping on the crystallographic, optical and electronic properties of CuO nanorods. Mater. Lett. 2018, 229, 225–228. [Google Scholar] [CrossRef]
  37. Zeng, J.; Zhong, J.B.; Li, J.Z.; Wang, S.H.; Hu, W. Improved photocatalytic activity of La3+-doped Bi2O3. Adv. Mater. Res. 2011, 239, 86–89. [Google Scholar] [CrossRef]
  38. Van Der Heide, P. Systematic x-ray photoelectron spectroscopic study of La1-xSrx-based perovskite-type oxides. Surf. Interface Anal. 2002, 33, 414–425. [Google Scholar] [CrossRef]
  39. Dong, S.; Wang, Y.; Zhao, Y.; Zhou, X.; Zheng, H. La3+/La (OH)3 loaded magnetic cationic hydrogel composites for phosphate removal: Effect of lanthanum species and mechanistic study. Water Res. 2017, 126, 433–441. [Google Scholar] [CrossRef]
  40. Qin, M.; Zhang, L.; Wu, H. Dielectric loss mechanism in electromagnetic wave absorbing materials. Adv. Sci. 2022, 9, 2105553. [Google Scholar] [CrossRef]
  41. Chang, Q.; Li, C.; Sui, J.; Waterhouse, G.I.; Zhang, Z.-M.; Yu, L.-m. Ni/Ni3ZnC0.7 modified alginate-derived carbon composites with porous structures for electromagnetic wave absorption. Carbon 2022, 200, 166–177. [Google Scholar] [CrossRef]
  42. Shen, Z.; Zu, Y.; Chen, Y.; Gong, J.; Sun, C. Microwave absorption performance of porous carbon particles modified by nickel with different morphologies. J. Mater. Sci. Technol. 2023, 137, 79–90. [Google Scholar] [CrossRef]
  43. Zhang, Z.; Chen, Q.; Huang, L.; Yang, X.; Yuan, Y. Recycling Polyethylene into High-Value Porous Carbon Composites for Microwave Absorption. Adv. Eng. Mater. 2023, 25, 2300366. [Google Scholar] [CrossRef]
  44. Jiang, H.; Wang, C.; Cui, B.; Xu, X.; Li, M.; Xu, Z.; Tan, H.; Wang, C.; Wang, Y. Lotus leaf derived NiS/carbon nanofibers/porous carbon heterogeneous structures for strong and broadband microwave absorption. Small 2023, 19, 2304918. [Google Scholar] [CrossRef] [PubMed]
  45. Yue, J.; Yu, J.; Jiang, S.; Chen, Y. Biomass carbon materials with porous array structures derived from soybean dregs for effective electromagnetic wave absorption. Diamond Relat. Mater. 2022, 126, 109054. [Google Scholar] [CrossRef]
  46. Ren, L.; Wang, Y.; Chai, L.; Zhou, S.; He, Q.; Du, H.; Wu, G. Hierarchical porous carbon prepared using swelling-induced biomass structure-controllable method with excellent microwave absorption performance. Mater. Chem. Phys. 2022, 279, 125739. [Google Scholar] [CrossRef]
  47. Fang, J.; Li, P.; Liu, Y.; Min, Y. Cobalt magnetic particles and carbon composite microtubes as high-performance electromagnetic wave absorbers. J. Mater. Chem. C 2021, 9, 2474–2482. [Google Scholar] [CrossRef]
  48. Wu, Z.; Yao, C.; Meng, Z.; Deng, Y.; Wang, Y.; Liu, J.; Wang, Y.; Zhou, H. Biomass-Derived Crocodile Skin-Like Porous Carbon for High-Performance Microwave Absorption. Adv. Sustain. Syst. 2022, 6, 2100454. [Google Scholar] [CrossRef]
  49. Lou, Z.; Wang, Q.; Kara, U.I.; Mamtani, R.S.; Zhou, X.; Bian, H.; Yang, Z.; Li, Y.; Lv, H.; Adera, S. Biomass-derived carbon heterostructures enable environmentally adaptive wideband electromagnetic wave absorbers. Nano-Micro Lett. 2022, 14, 11. [Google Scholar] [CrossRef]
  50. Zhao, H.; Jin, C.; Lu, P.; Xiao, Z.; Cheng, Y. Anchoring well-dispersed magnetic nanoparticles on biomass-derived 2D porous carbon nanosheets for lightweight and efficient microwave absorption. Composites Part A 2022, 154, 106773. [Google Scholar] [CrossRef]
  51. Liang, N.; Yin, Z.; Guo, J.; Fang, W.; Wang, Q.; Tian, G.; Zhang, D.; Yue, H.; Feng, S. Yolk-shell structure synergistic defect engineering for boosting electromagnetic wave absorption in Co9S8@Humins-derived carbon. J. Mater. Sci. Technol. 2023, 164, 140–149. [Google Scholar] [CrossRef]
  52. Mao, F.; Long, L.; Zeng, G.; Chen, H.; Li, Y.; Zhou, W. Achieving excellent electromagnetic wave absorption property by constructing VO2 coated biomass carbon heterostructures. Diamond Relat. Mater. 2022, 130, 109422. [Google Scholar] [CrossRef]
  53. Cheng, Y.; Cao, J.; Li, Y.; Li, Z.; Zhao, H.; Ji, G.; Du, Y. The outside-in approach to construct Fe3O4 nanocrystals/mesoporous carbon hollow spheres core-shell hybrids toward microwave absorption. ACS Sustain. Chem. Eng. 2018, 6, 1427–1435. [Google Scholar] [CrossRef]
  54. Xie, A.; Sheng, D.; Liu, W.; Chen, Y.; Cheng, S. Enhancing electromagnetic absorption performance of Molybdate@Carbon by metal ion substitution. J. Mater. Sci. Technol. 2023, 163, 92–100. [Google Scholar] [CrossRef]
  55. Zhang, M.; Zhao, L.; Zhao, W.; Wang, T.; Yuan, L.; Guo, Y.; Xie, Y.; Cheng, T.; Meng, A.; Li, Z. Boosted electromagnetic wave absorption performance from synergistic induced polarization of SiCNWs@MnO2@PPy heterostructures. Nano Res. 2023, 16, 3558–3569. [Google Scholar] [CrossRef]
  56. Jia, T.; Hao, Y.; Qi, X.; Rao, Y.; Wang, L.; Ding, J.; Qu, Y.; Zhong, W. Interface engineering and impedance matching strategy to develop core@shell urchin-like NiO/Ni@carbon nanotubes nanocomposites for microwave absorption. J. Mater. Sci. Technol. 2024, 176, 1–12. [Google Scholar] [CrossRef]
  57. He, X.; Xiong, Z.; Lei, C.; Shen, Z.; Ni, A.; Xie, Y.; Liu, C. Excellent microwave absorption performance of LaFeO3/Fe3O4/C perovskite composites with optimized structure and impedance matching. Carbon 2023, 213, 118200. [Google Scholar] [CrossRef]
  58. Zeng, X.; Cheng, X.; Yu, R.; Stucky, G.D. Electromagnetic microwave absorption theory and recent achievements in microwave absorbers. Carbon 2020, 168, 606–623. [Google Scholar] [CrossRef]
  59. Wu, Y.; Chen, L.; Han, Y.; Liu, P.; Xu, H.; Yu, G.; Wang, Y.; Wen, T.; Ju, W.; Gu, J. Hierarchical construction of CNT networks in aramid papers for high-efficiency microwave absorption. Nano Res. 2023, 16, 7801–7809. [Google Scholar] [CrossRef]
  60. Micheli, D.; Pastore, R.; Gradoni, G.; Marchetti, M. Tunable nanostructured composite with built-in metallic wire-grid electrode. AIP Adv. 2013, 3, 112132. [Google Scholar] [CrossRef]
  61. Li, Z.; Liang, J.; Wei, Z.; Cao, X.; Shan, J.; Li, C.; Chen, X.; Zhou, D.; Xing, R.; Luo, C. Lightweight foam-like nitrogen-doped carbon nanotube complex achieving highly efficient electromagnetic wave absorption. J. Mater. Sci. Technol. 2024, 168, 114–123. [Google Scholar] [CrossRef]
  62. Qiu, Y.; Yang, H.; Cheng, Y.; Bai, X.; Wen, B.; Lin, Y. Constructing a nitrogen-doped carbon and nickel composite derived from a mixed ligand nickel-based a metal–organic framework toward adjustable microwave absorption. Nanoscale 2021, 13, 9204–9216. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, X.; Xing, X.; Zhu, H.; Li, J.; Liu, T. State of the art and prospects of Fe3O4/carbon microwave absorbing composites from the dimension and structure perspective. Adv. Colloid Interface Sci. 2023, 318, 102960. [Google Scholar] [CrossRef]
  64. Yu, L.; Zhu, Q.; Guo, Z.; Cheng, Y.; Jia, Z.; Wu, G. Unique electromagnetic wave absorber for three-dimensional framework engineering with copious heterostructures. J. Mater. Sci. Technol. 2024, 170, 129–139. [Google Scholar] [CrossRef]
Figure 1. The XRD and the Raman spectroscopic signatures of La-CFO@BC composite. (a) X-ray diffraction patterns; and (b) Raman spectroscopic signatures.
Figure 1. The XRD and the Raman spectroscopic signatures of La-CFO@BC composite. (a) X-ray diffraction patterns; and (b) Raman spectroscopic signatures.
Molecules 29 02620 g001
Figure 2. The images from SEM of the BC (a,b) La-CFO (ce) and La-CFO@BC (fi) particles, and the TEM images of the BC (j,k) and La-CFO (l,m) particles.
Figure 2. The images from SEM of the BC (a,b) La-CFO (ce) and La-CFO@BC (fi) particles, and the TEM images of the BC (j,k) and La-CFO (l,m) particles.
Molecules 29 02620 g002
Figure 3. N2 isotherm adsorption/desorption curve and pore-size distribution curve of BC (a) and La-CFO@BC (b).
Figure 3. N2 isotherm adsorption/desorption curve and pore-size distribution curve of BC (a) and La-CFO@BC (b).
Molecules 29 02620 g003
Figure 4. XPS survey spectra of La-CFO@BC. (a) The full spectrum of XPS; (b) C 1s; (c) O 1s; (d) Fe 2p; (e) Co 2p; and (f) La 3d.
Figure 4. XPS survey spectra of La-CFO@BC. (a) The full spectrum of XPS; (b) C 1s; (c) O 1s; (d) Fe 2p; (e) Co 2p; and (f) La 3d.
Molecules 29 02620 g004
Figure 5. Magnetization curves of La-CFO and La-CFO@BC.
Figure 5. Magnetization curves of La-CFO and La-CFO@BC.
Molecules 29 02620 g005
Figure 6. The actual photos of the VAM instrument connection (a), the ring-shaped tested sample (b), and the coaxial fixture for the tested sample (c).
Figure 6. The actual photos of the VAM instrument connection (a), the ring-shaped tested sample (b), and the coaxial fixture for the tested sample (c).
Molecules 29 02620 g006
Figure 7. The RL curves, RL 3D distribution maps and 2D RL contour plots of BC (ac); La-CFO (df); La-CFO@BC (gi).
Figure 7. The RL curves, RL 3D distribution maps and 2D RL contour plots of BC (ac); La-CFO (df); La-CFO@BC (gi).
Molecules 29 02620 g007
Figure 8. The relationships between frequency and the ε′ and ε″ (a,b) as well as μ′ and μ″ (c,d) for BC, La-CFO, and La-CFO@BC.
Figure 8. The relationships between frequency and the ε′ and ε″ (a,b) as well as μ′ and μ″ (c,d) for BC, La-CFO, and La-CFO@BC.
Molecules 29 02620 g008
Figure 9. The relationships between frequency and tanδe (a) as well as tanδm (b) for BC, La-CFO, and La-CFO@BC.
Figure 9. The relationships between frequency and tanδe (a) as well as tanδm (b) for BC, La-CFO, and La-CFO@BC.
Molecules 29 02620 g009
Figure 10. Cole–Cole curves of (a) BC, (b) La-CFO, and (c) La-CFO@BC.
Figure 10. Cole–Cole curves of (a) BC, (b) La-CFO, and (c) La-CFO@BC.
Molecules 29 02620 g010
Figure 11. Z (a) and α (b) in relation to frequency for BC, La-CFO, and La-CFO@BC.
Figure 11. Z (a) and α (b) in relation to frequency for BC, La-CFO, and La-CFO@BC.
Molecules 29 02620 g011
Figure 12. The schematic for the potential EW-loss process of La-CFO@BC.
Figure 12. The schematic for the potential EW-loss process of La-CFO@BC.
Molecules 29 02620 g012
Table 1. The XPS sub-peak positions of Fe 2p, Co 2p, and La 3d for the La-CFO@BC.
Table 1. The XPS sub-peak positions of Fe 2p, Co 2p, and La 3d for the La-CFO@BC.
Energy Position (eV)
Spectrum Fe 2 p1/2Fe 2p3/2Co 2p1/2Co 2p3/2La 3d3/2La 3d5/2
AssignmentOs724.81710.85796.28780.33850.86834.11
Ts726.20712.21798.16782.21852.12835.37
Table 2. Comparative analysis of microwave-absorbing performance for different biomass-carbon materials.
Table 2. Comparative analysis of microwave-absorbing performance for different biomass-carbon materials.
FeedstockSamplesRLmin Value
(dB)
EAB
(GHz)
Thickness at EAB (mm)Ref.
Soybean dregsFK-SDC−18.504.803.50[45]
Shaddock peelG-800−29.505.801.70[46]
KapokfibersCMT-900−30.756.782.06[19]
Cotton fibersPCMT/Co−36.86.71.4[47]
Wheat flourPCP@Ni-chain−38.425.22.0[42]
AlginateNi/Ni3ZnC0.7/C-0.1−40.25.42.0[41]
Rice huskBHPC−47.463.42.8[48]
BambooGC-8−51.004.21.66[49]
Shaddock peelsFe3O4/PCS-2−50.34.11.6[50]
HuminsCS@C-700−51.45.21.84[51]
Coffee groundsWCG-20–750−52.866.403.0[20]
Corn stalksLa-CFO@BC−53.196.492.5This work
Chestnut needlesVO2–50−54.02.52.87[52]
Pine needlesPNC−56.303.441.4[53]
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

Shang, T.; Zhu, H.; Shang, Y.; Wu, R.; Zhao, X. A Novel Low-Density-Biomass-Carbon Composite Coated with Carpet-like and Dandelion-Shaped Rare-Earth-Doped Cobalt Ferrite for Enhanced Microwave Absorption. Molecules 2024, 29, 2620. https://doi.org/10.3390/molecules29112620

AMA Style

Shang T, Zhu H, Shang Y, Wu R, Zhao X. A Novel Low-Density-Biomass-Carbon Composite Coated with Carpet-like and Dandelion-Shaped Rare-Earth-Doped Cobalt Ferrite for Enhanced Microwave Absorption. Molecules. 2024; 29(11):2620. https://doi.org/10.3390/molecules29112620

Chicago/Turabian Style

Shang, Tao, Hongwei Zhu, Yichun Shang, Ruixia Wu, and Xuebing Zhao. 2024. "A Novel Low-Density-Biomass-Carbon Composite Coated with Carpet-like and Dandelion-Shaped Rare-Earth-Doped Cobalt Ferrite for Enhanced Microwave Absorption" Molecules 29, no. 11: 2620. https://doi.org/10.3390/molecules29112620

Article Metrics

Back to TopTop