Next Article in Journal
Improving Thermal Conductivity Coefficient in Oriented Strand Lumber (OSL) Using Sepiolite
Previous Article in Journal
Combating Antibiotic Resistance through the Synergistic Effects of Mesoporous Silica-Based Hierarchical Nanocomposites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ultrafine FeNi3 Nanocrystals Embedded in 3D Honeycomb-Like Carbon Matrix for High-Performance Microwave Absorption

1
School of Material and Energy, Guangdong University of Technology, Guangzhou 510006, China
2
School of Computer Science and Technology, Guangdong University of Technology, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2020, 10(4), 598; https://doi.org/10.3390/nano10040598
Submission received: 18 February 2020 / Revised: 6 March 2020 / Accepted: 16 March 2020 / Published: 25 March 2020

Abstract

:
The reasonable design of magnetic carbon-based composites is of great significance to improving the microwave absorption (MA) performance of the absorber. In this work, ultrafine FeNi3 nanocrystals (5–7 nm) embedded in a 3D honeycomb-like carbon matrix (FeNi3@C) were synthesized via a facile strategy that included a drying and carbonization process. Because of the soft magnetic property of the FeNi3 nanocrystals and their unique 3D honeycomb-like structure, the FeNi3@C composites exhibit excellent MA abilities. When the filler loading ratio of FeNi3@C/paraffin composites is only 30 wt%, the maximum reflection loss (RL) value is −40.6 dB at 10.04 GHz. Meanwhile, an ultra-wide absorption frequency bandwidth of 13.0 GHz (5.0–18.0 GHz over −10 dB) can be obtained in the thickness range of 2.0–4.5 mm, and this means that the absorber can consume 90% of the incident waves. It benefits from the dual loss components, multiple polarizations, and multiple reflections for improving MA performances of FeNi3@C composites. These observations suggest that the 3D honeycomb-like FeNi3@C composites have broad application prospects in exploring new MA materials that have a wide frequency bandwidth and strong absorption.

1. Introduction

In recent years, a wide usage of electronic equipment has caused more and more electromagnetic pollution [1,2,3]. Long-term exposure to this pollution will cause specific harm to human health, and will also interfere with the operation of sensitive electronic equipment in civil and military fields [4].Thus, microwave absorption (MA) materials are urgently needed to attenuate superfluous electromagnetic (EM) energies [5,6]. The MA materials effectively convert incident EM microwaves energy into thermal energy or dissipate them via interference [4,7,8]. To date, a range of traditional MA materials (including dielectric materials, such as graphene, carbon nanotubes, carbon fiber, and ZnO, and magnetic materials, such as Fe, Co, Ni, Fe2O3, and Fe3O4) have been widely studied for their MA performance [4,9,10,11]. However, it is difficult to achieve excellent MA properties for a single loss component, because these traditional absorbers have the drawbacks of thick thickness, weak absorption, and narrow absorption bandwidth. Therefore, exploring novel microwave absorbers is highly needed. More and more researchers think highly of nanoscale soft magnetic materials and their alloys because they own low coercive forces, high saturation magnetization, high Curie temperatures, and a high Snoek’s limit at high frequency bands. FeNi3 alloy, as a typical soft magnetic material, has aroused great interest from researchers, due to its superior magnetic properties.
Accordingly, a simple and effective way to improve MA performance is to combine dual loss (magnetic and dielectric loss) components into composite materials using a well-designed method [12]. Che et al. [13] filled Fe into carbon nanotubes (CNTs) via a simple catalytic pyrolysis routine, and the composite material showed excellent microwave absorbability. Cao et al. [14] introduced multi-walled carbon nanotubes (MWCNTs)/Fe3O4 nanocrystals composites for EM microwaves absorbers via a co-precipitation way. MA properties of the composites are better than those of pure MWCNTs and other Fe3O4 hybrid materials. Thus, many composites such as Ni/C porous nanofibers [15], Co@C spheres [16], FeCo-C core-shell nanoparticles [17], Fe3O4/CNTs [18], and C@NiCo2O4@Fe3O4 [19], have been explored for microwave absorption. Normally, both the compounds in the composites as well as the microstructure affect the MA properties. As a result, numerous morphologies have been successfully prepared to attain strong microwave attenuation ability. Liu and coworkers [20] designed flower-like Co20Ni80 alloy spheres with an absorption bandwidth of 5.5 GHz and an urchin-like Co20Ni80 alloy with a reflection loss (RL) value of −33.5 dB. Shen et al. [21] prepared Co-C nanofibers via electrospinning, and the nanofibers had an optimal RL value of −33.1 dB at 14.1 GHz. Zhao et al. [22] investigated the MA properties of 2D sponge-like Ni/derivative heterostructures, and they found that the value of maximum RL was −37.3 dB at 7.1 GHz when the thickness was 3.3 mm. As above, substantial improvements to the MA performance were achieved with various morphological structures. However, there is still a long way to go to optimize MA composites with a consideration of their microstructure and composition at the same time.
In our work, we successfully synthesized 3D honeycomb-like FeNi3@C composites using a facile strategy. As-prepared FeNi3 nanocrystals are uniformly surrounded by a carbon matrix that protects the FeNi3 nanocrystals from agglomeration. The FeNi3@C composites benefit from the soft magnetic property of the FeNi3 nanoparticles and unique 3D honeycomb-like structure, and thus they exhibit excellent MA abilities. The synergy effects, concluding dielectric/electrical losses and magnetic loss, efficiently improve the MA performance. The 3D honeycomb-like FeNi3@C composites obtain the maximum RL value of −40.6 dB and a broad frequency bandwidth of 13.0 GHz. This work provides a facile approach for the synthesis of a magnetic alloys/carbon composites absorber with wide bandwidth and strong absorption.

2. Materials and Methods

2.1. Materials

Sodium lauryl sulfate (SDS), styrene (C8H8), divinylbenzene (DVB), potassium persulfate (K2S2O8), tetrahydrofuran (THF), polyvinylidene fluoride (PVDF), bis (η5-2, 4-cyclopentadien-1-yl) nickel (C10H10Ni), and ferrocene (C10H10Fe) were acquired from Aladdin Industrial Co, Shanghai. All chemical reagents used in reaction are analytical grade and without further treatment.

2.2. Preparation of Polystyrene (PS) Spheres Emulsion Solution

An emulsion solution of PS spheres was prepared via a typical emulsion polymerization method. The specific procedure was described as follows: 240 mg SDS was completely dissolved in 1200 mL distilled water with a 30-minute stirring. Then, 30 mL C8H8 and 3.6 mL DVB were added under continuous stirring. Subsequently, the solution was heated to 75 °C under a N2 atmosphere, with 1.2 g K2S2O8 adding. After 3 h, 3.6 mL DVB was added to the mixed solution system again. Finally, the PS emulsion solution was formed and stored for later use.

2.3. Synthesis of FeNi3@C Composites

The as-papered PS emulsion solution was dried overnight at 80 °C under vacuum to obtain a powder of the PS spheres. Ultrasonication was used to disperse 0.15 g of the powder of PS spheres in 20 mL of THF. Then, 1.00 g C10H10Ni, 0.33 g C10H10Fe and x g of PVDF (x = 0.3, 0.15, and 0.075) were added to the mixed solution with a 2-h stirring. The mixtures were then transferred into a homemade oven to de-solvate overnight under heat at 80 °C. After removing the solvent (THF), the remaining precursors were calcined at 600 °C for 2 h under an Ar/H2 atmosphere; the heating rate was 5 °C min−1. The obtained FeNi3@C composites were labeled as S1 (0.30 g PVDF), S2 (0.15 g PVDF), and S3 (0.075 g PVDF).

2.4. Preparation of the Coaxial Sample

The FeNi3@C composites and paraffin were mixed in a certain ratio and heated at 80 °C for 10 min. Then the system was cooled to room temperature and pressed into a ring-shaped sample ( φ i n = 3.04   mm , φ o u t   = 7.00   mm ) under a pressure of 1.5 MPa. The photos of pristine powder and ring-shaped sample can be seen in Figure 1.

2.5. Characterizations

The phase and component information of the composites were characterized using an X-ray diffraction instrument (XRD, Germany, D8 ADVANCE). Both field emission scanning electron microscopy (FESEM, Hitachi, SU8220) and high-magnification transmission electron microscopy (HRTEM, FEI, TALOS F200S) were performed to observe the morphology of the product. Energy dispersive spectrometry (EDS), as an accessory of FESEM, was used to detect the distribution elements. The operation voltage was 20 KV for FESEM and 200 KV for HRTEM. Thermogravimetric analysis (TGA) was performed from room temperature to 900 °C with a heating rate of 10 °C min−1 on a TGA instrument. The degree of graphitization of the products was analyzed using Raman spectroscopy with an excitation wavelength of 532 nm (HORIBA Jobin Yvon, France, LabRAM HR Evolution). X-ray photoelectron spectroscopy (XPS) was used to investigate the chemical states and compositions of the composites. Vibrating sample magnetometry (VSM) was carried out to survey the magnetic properties of the FeNi3@C composites. An AV3618 network analyzer was used for coaxial measurements of the electromagnetic parameters in the frequency range of 2–18 GHz.

3. Results and Discussion

Figure 2 is a schematic illustration of the whole preparation process of the 3D honeycomb-like FeNi3@C composites. The PS spheres were synthesized via an emulsion polymerization method, and then a precursor template was used. The PS spheres, C10H10Fe, C10H10Ni, and PVDF were added to the THF solvent; the mixture was then sequentially dried and carbonized. PVDF serves as a carbon source and a binder to connect the spheres during the drying process, and this causes the metal cations to be evenly distributed in the space of spheres. During the final carbonization process, PS spheres were removed to produce nanoscale pores. Simultaneously, PVDF decomposed and formed a carbon matrix. With prolonged heating, the Fe/Ni-based oxides that were produced during the early stage were reduced by H2 to the corresponding metals [23,24]. The reaction of intermetallic FeNi3 can be simply described as follows [25]: Fe + 3Ni→FeNi3.
Scanning electron microscopy (SEM) and HRTEM were used to assess the microstructure of the as-prepared FeNi3@C product. Notably, all the samples showed a 3D honeycomb-like structure with no obvious differences (Figure S1a–d). Therefore, taking S2 sample as an example, the microstructure analysis is discussed. As shown in Figure 3a–c, nanopores are distributed uniformly and in an orderly manner in the carbon matrix under different magnifications. The average size of the pores is about 130–150 nm, with mainly controlling from the PS spheres. There is no doubt that this unique porous structure will enhance the scattering of EM waves, thus facilitating the enhancement of MA properties, which have been proved [10]. Figure 3d,e represent typical TEM images of sample S2, in which the FeNi3 nanocrystals are well coated by a honeycomb-like carbon layer. The ultrafine nanocrystals are sphere-like and the average size of particles is 5–7 nm. In Figure 3f the periodic fringe spacing is 0.02 nm, which can be assigned to the (111) plane of FeNi3. Figure 3g–i shows EDS mapping of S2. The results indicate that Fe, Ni, and C were well distributed, and this is in good agreement with the XPS analysis.
Figure 4a shows the XRD patterns of all of samples. All samples are highly crystalized with three diffraction peaks. The peaks located at 44.28°, 51.53°, and 75.87°, respectively, correspond to the (111), (200), and (220) planes of FeNi3 (PDF#38-0419) with face-centered cubic (FCC) structures. Our products are free of the oxidation disruption because no peaks of Fe/Ni-based oxides can be found in the composite. Considering the absence of graphitization peaks, the honeycomb carbon matrix is considered as an amorphous structure, which is consistent with the results of EDS and TEM. Raman spectra (Figure 4b) were further studied to examine the presence and crystallinity of carbon matrix. There are two prominent peaks at 1333 cm−1 and 1590 cm−1, and these respectively correspond to the D and G bands. Generally speaking, the D band of the A1g vibrational mode is considered to represent sp3 carbon atoms in disordered graphite, and the G band of the E2g vibrational mode of the C-C bond stretching represents sp2 carbon atoms in a 2D hexagonal lattice [26,27]. The value of the ID/IG ratio can be used to evaluate the degree of disorder in the FeNi3@C composites and indicates the presence of structural defects. Furthermore, these defects facilitate the polarization of dipole and thus improve the microwave absorption performance, as reported by Ding [26]. The ID/IG values of S1, S2, and S3 are 0.8570, 0.8640, and 0.8685, respectively. Obviously, the ID/IG values of S1, S2, and S3 show an increasing trend, and this can be attributed to a decrease in the amount of added PVDF, which has a lower degree of graphitization.
As shown in Figure 4c, the thermal gravimetric curves of all of the FeNi3@C samples present two weight loss stages. One stage (below 200 °C) is caused by the removal of surface-absorbed water. The second stage (between 320 °C and 530 °C) corresponds to the combustion of the carbon layer. The final products are FeNi3 nanocrystals with Fe/Ni-based oxides and the values of the residual weight of S1, S2, and S3 are about 34.7%, 45.0%, and 65.7%, respectively. In this experiment, PVDF is used as a carbon source, and different amount of PVDF means the different content of carbon source. The carbon content of three samples is reduced in a high temperature oxidation process, resulting in different thermal weight loss. The residual weight values of S1, S2, and S3 show an increasing trend and this can be attributed to a decrease in the amount of added carbon source.
Figure 4d shows hysteresis loops for all samples measured using VSM at room temperature. Because the FeNi3 nanoalloy is a kind of soft magnetic material, S1, S2, and S3 all exhibit typical soft magnetic behavior [26]. The values of the saturation magnetization (Ms) of S1, S2, and S3 are 10.6, 12.6, and 14.4 emu/g, respectively. Relative coercivity (Hc) values are 49.6, 50.1, and 50.3 Oe, respectively. Ms of S1 is the lowest among that of the three samples due to its high content of the carbon source. We found that the Ms value of FeNi3@C is significantly lower than the Ms value of pure FeNi3 (Ms = 85.2 emu/g) [28]. The decrease Ms can be ascribed to the presence of nonmagnetic carbon materials. Furthermore, the 3D honeycomb-like structure that has more defects in the crystalline structure may cause surface imperfections in the FeNi3@C composites. Other affects for losing Ms comes from the oxidation vulnerability and nanoscale lattice mismatch in ultrafine FiNi3 nanocrystals. This thus leads to a much smaller Ms value [29,30,31]. Additionally, the grain size and shape anisotropy influence the Hc values of magnetic materials [32]. The higher coercivity of FeNi3@C composites may be due to the small grain size of FeNi3 nanocrystals.
The XPS measurements were used for further analyzing the surface chemical compositions and elemental valence states of the samples. The presence of Fe, Ni, and C in the synthesized FeNi3@C composites was confirmed by the spectrum of the survey scan (Figure 5a). The survey scan shows five peaks with binding energies of approximately 285, 531, 641, 712, and 856 eV, and these are assigned to C 1s, O 1s, Ni Auger, Fe 2p, and Ni 2p, respectively. These indicate the presence of C, O, Fe, and Ni in the sample. The existence of oxygen may be due to the fact that the original material PVDF contains some oxygen-containing functional groups [33]. However, we did not find oxygen element in the EDS mapping, indicating that the oxygen content in the sample is extremely low. The C1s spectrum (Figure 5b) can be split into three peaks that are located at 284.6, 286.3, and 288.8, and these correspond to C − C/C = C, C − O, and O − C = O, respectively. Figure 5c displays the Fe 2p spectrum of S2, and it can be fitted by two peaks. One peak at 711.6 eV corresponds to Fe 2p3/2 and one peak at 724.5 eV corresponds to Fe 2p1/2. As seen in Figure 5d, the fitting peak with binding energies of 856.5 and 874.3 eV should be assigned to Ni 2p3/2 and Ni 2p1/2, respectively. Also, the peaks at 853.1 and 870.3 eV can be attributed to the Ni 2p3/2 and Ni 2p1/2 of Ni2+ states, respectively [4].
In order to explore the influence of the filler loading ratios on the MA performance, we took the S2 sample as an example to measure the electromagnetic parameters. The content of S2 in the paraffin matrix varied from 10 wt% to 40 wt%, and was labeled as S2–10 wt%, S2–20 wt%, S2–30 wt% and S2–40 wt%, respectively. In Figure 6c, when the thickness is 3.0 mm, the maximum RL value of S2–30 wt% can reach −40.6 dB. However, for layers of the same thickness, the maximum RL values of S2–10 wt%, S2–20 wt%, and S2–40wt% are −9.13 dB, −12.62 dB, and −20.68 dB, respectively. Obviously, S2–30 wt% has the largest RL value at the same thickness. Therefore, the 30% filler loading ration was chosen for comparing the MA properties of other samples in the following discussion.
RL values are usually used to assess the performance of MA materials. These values can be calculated according to transmission line theory, and the equations are as follows [2,34]:
RL ( dB ) = 20   log 10 | Z i n Z 0 Z i n + Z 0 |
Z i n = Z 0 μ r / ε r tanh [ j ( 2 π f d / c ) × μ r × ε r   ]
where Z i n represents the input impedance at the surface of the absorber, Z 0 is the impedance of free space, d is thickness of the sample, f is the frequency of the electromagnetic wave, and c is the speed of light. When the RL values are −10 and −20 dB, it means that 90% and 99% of the incident EM waves are attenuated [35,36]. Also, the frequency ranges of RL values below −10 dB are valuable for engineering applications [10].
Figure 7 presents RL curves for S1, S2, and S3 at various thicknesses, each with filler loadings of 30 wt%. It is clear that with an increase in thickness, the RL peaks move toward the lower frequency region, and this is consistent with other reports [2,37]. All samples are consistent with quarter-wavelength matching model, where the relationship between sample thickness (d) and absorption peak frequency (f) can be expressed as the following equation [38]:
d   = n c / 4 f | ε r | × | μ r | ( n = 1 , 2 , )
When d and f satisfy this formula, the value of RL can be calculated. As seen in Figure 7a, the maximum RL value of S1 can reach −36.43 dB at 7.72 GHz when the thickness is 4.0 mm. In Figure 7c, the RL value of S3 is only −16.35 dB at 13.56 GHz when the thickness is 3.5 mm. In addition, the S2 possess the maximum RL value of −40.6 dB at 10.04 GHz and an ultra-wide bandwidth (RL ≤ −10 dB) of 13.0 GHz in the range of 5.0–18.0 GHz, which we can see from Figure 7b. This suggests that the S2 exhibits more excellent MA performance compare to S1 and S3. To better assess the MA properties, Figure 7d–f presents the 3D RL diagram of all the FeNi3@C samples with different thicknesses. Furthermore, the RL values calculated at a thickness of 3.0 mm are shown in Figure S3. The maximum RL value first decreases and then increases as the carbon source decreases. As a result, the S2 sample with 0.15 g of added PVDF has the optimal MA performance. This can be attributed to the gradual formation of a conductive network that occurs when the content of the carbon source increases, and this results in conductive losses that dissipate incident EM waves [22]. However, high conductivity may lead to poor MA properties when the filler loading ratio of the carbon source is increasing. These results suggest that the amount of added PVDF had a remarkable effect on the MA properties of the absorber, and thus, the amount of added PVDF should be well optimized.
It is known that, the MA properties of an absorber are highly dependent on the relative complex permittivity ( ε r = ε j ε ) and relative complex permeability ( μ r = μ j μ ) . Generally, the real parts ( ε and μ ) and imaginary parts ( ε and μ ) represent the storage and dissipation capabilities of electric and magnetic energy, respectively [39]. The dielectric loss tangents ( tan δ ε = ε / ε ) and magnetic loss tangents ( tan δ μ = μ / μ ) are widely used to evaluate the dielectric loss capability and magnetic loss capability of the absorber.
Figure 8 displays vital parameters ( ε , ε ,   μ ,   and μ ) for all of the samples. In Figure 8a, it is obvious that with the increase of PVDF loading, the values of the real part of the relative complex permittivity first increases and then decreases. Also, typical frequency dispersion behavior for all of the samples is shown in the Figure 8a,b) [32]. Over the entire frequency range, the values of ε and ε for S2 dropped from 10.02 to 4.61 and 6.70 to 0.27, respectively, with slight fluctuates, and a similar decrease is observed in S1 and S3. Liu et al. [35] found that moderate complex permittivity is more favorable for the MA performances of an absorber than a lower or higher complex permittivity. This is consistent with the high microwave absorption properties of S2, as reported in our study. The dielectric loss factor is plotted versus frequency (Figure 8c), and a higher value of tan δ ε indicates a higher ability to dissipate the microwave energy. In alternating electromagnetic fields, there are different variations in tan δ ε and these indicate that multiple polarization relaxation processes occurring in the FeNi3@C composites.
Generally, the dielectric loss capability mainly derives from conductivity loss and polarization relaxation behaviors according to free electron theory and Debye theory [40,41]. Conductivity loss may result from the conductive network, which is caused by the 3D honeycomb-like structure. Also polarization relaxation can be divided into four polarization modes: electronic, ionic, dipole orientation, and interfacial polarization [32,42]. Ionic polarization and electronic polarization are not considered in our study because they usually occur in the high frequency range of 103–106 GHz [42]. Thus, the dielectric loss mainly originates from interfacial polarization and dipolar polarization, and these are caused by defects during carbonization. Rich interfaces between FeNi3 cores and the graphite layer as well as between the graphite layer and paraffin lead to sufficient interfacial polarization for dielectric loss [32]. To further prove the diversified dielectric loss mechanism, curves of ε versus ε were plotted, as shown in Figure 9. According to Debye theory, ε and ε meet the formula [12,43]:
( ε ε s + ε 2 ) 2 + ( ε ) 2 = ( ε s ε 2 ) 2
where ε s is static permittivity and ε is relative permittivity at the high-frequency limits. Each semicircle in the curve of ε ε is generally denoted as the Debye or Cole–Cole semicircle. There are different numbers of distorted Cole–Cole semicircles belonging to S1–S3, and each semicircle corresponds to one Debye dipolar relaxation process. This reveals that carbon source content will influence the Debye relaxation process due to different matching between the ε and ε values. These results well confirm that the dielectric loss in the 3D honeycomb-like FeNi3@C composites mainly comes from multiple-relaxations. It can be found more Cole–Cole semicircles of S2 and S3, compared to ones of S1. It is widely accepted that dielectric losses are not the only determinant in the evaluation of MA performances in ferrite-carbon composite absorbers [19]. Permeability and magnetic loss also need to be analyzed, which are the other crucial factors responsible for MA performances. The excellent performance of S2 benefits from the combined effect of dielectric loss and magnetic loss.
Figure 8d,e show the μ and μ values for all the samples in the frequency range of 2.0–18.0 GHz. Figure 8d shows the μ values for all the samples are in the range of 1.0–1.3 with some fluctuations. In particular, the S2 and S3 samples show obvious resonance peaks at different frequencies. The μ″ value of S2 is the largest among the three samples, and this indicates higher dissipation of magnetic energy. At the same time, the μ values for S1 and S3 exhibit pretty similar trends. However, the μ value for S3 becomes negative in the frequency range of 9.04–18.0 GHz. Generally speaking, the electromagnetic wave absorption materials consume magnetic energy of the incident EM wave, and thus, there are rarely negative μ values [32]. However, negative values may result from the implicit Fabry–Perot resonance, porous structure of FeNi3@C, and Snoek limit at high frequencies [15,44,45,46]. According to the Maxwell equation, movement of electric charge first generates an alternating electric field and then forms a magnetic field. Moreover, the negative value of μ indicates that the redundant magnetic energy that radiated out by the energy of the internal and external magnetic fields overloads the loss capability of FeNi3 nanocrystals [32,47]. The magnetic loss factors ( tan δ μ ) for all samples are presented in the Figure 8f, and the shape and tendency of tan δ μ are similar to μ . In addition, tan δ μ for S2 is higher than that for other samples, and this indicates that S2 has better magnetic loss capability. Interestingly, it is found that the enhanced MA performance of S2 is mainly caused by dielectric loss at 2–8.0 GHz, and magnetic loss plays a major role at high frequency, as seen in Figure 8c,f. It is known that, magnetic loss always results from natural ferromagnetic resonance, domain wall resonance, eddy current effects, and hysteresis [16]. Hysteresis loss can be ignored because of the weak field, and domain wall resonance loss can also be neglected because this resonance only occurs at a much lower frequency (MHz) [48,49]. Thus, the main loss mechanisms of absorbers are the eddy current effect and natural ferromagnetic resonance. The eddy current loss can be expressed as follows: C 0 = μ ( μ ) 2 f 1 . If the eddy current effect is the only factor that results in magnetic loss, the value of C 0 is nearly constant regardless of changes in frequency [16]. As seen in Figure 9d, the values of   C 0 for all of the samples obviously fluctuate with an increase in frequency, and this reveals that the eddy current effect and natural ferromagnetic resonance both contribute to magnetic loss. As reported in many studies, there is a further increase in the resonance frequency because of the shape anisotropy that result from the small size effect of nanoparticles [49,50].
Nowadays, the preparation of FeNi3@C composites has been reported, and their MA properties have been investigated. Ding et al. [8] have reported FeNi3@C–paraffin composites containing 10 wt% for MA absorber with an RL of −47.6 dB. The absorption bandwidth was about 3.12 GHz in the range of 7.28–10.40 GHz for a layer of 3.0 mm thickness. Sun et al. [9] found FeNi3@C nanowires–paraffin composites containing 40 wt% FeNi3@C nanowires had an excellent MA performance. For a 2.0 mm thickness layer, an optimal RL of −43.3 dB was obtained and RL exceeding −10 dB was observed at 8.7–13.78 GHz. In our work, when the filler loading ratio of FeNi3@C/paraffin composites is only 30 wt%, the RL value can reach −40.6 dB at 10.04 GHz. Meanwhile, an ultra-wide absorption frequency bandwidth of 13.0 GHz (5.0–18.0 GHz over −10 dB) is obtained. Table 1 shows the MA properties of some reported FeNi3-based composites. Compared to recent reports, the 3D honeycomb-like FeNi3@C exhibits outstanding MA properties.
Based on the above, the S2 sample has excellent MA properties and a broad bandwidth as a result of the synergistic effect between the dual loss (dielectric and magnetic loss) of the carbon layer and FeNi3 nanocrystals. However, beyond that, the unique structure also contributes to the improved MA properties. In Figure 10, when incident microwave penetrates into the absorbent, multiple reflections and scatterings occur in the 3D honeycomb-like structure, and this extends the propagation of the microwave pathway to largely attenuate the incident EM wave [4,10,22]. Dielectric loss is induced in dipole polarization and interfacial polarization, and these are respectively caused by a large number of defects and interfaces. Also, the 3D honeycomb-like structure forms a conductive network which is good for improving electrical loss. As expected, the 3D honeycomb-like FeNi3@C composites exhibit optimum MA performance.

4. Conclusions

In summary, 3D honeycomb-like FeNi3@C composites were successfully synthesized via a facile method that included drying and carbonization. During drying, metal ions evenly fill in the gaps of the PS spheres, and the template PS spheres are removed as a result of the formation of plentiful pores in the carbonization process. The 3D honeycomb-like FeNi3@C composites (S2) possess superb MA performances as expected. The optimal RL value is −40.6 dB at 3 mm coating thickness, and response absorption bandwidth (RL ≤ −10 dB) is 13.0 GHz when the thicknesses are only 2.0–4.5 mm. This means that 90% of the incident waves can be consumed by the absorber when the frequency is in the range of 5.0–18.0 GHz. The excellent MA performance can be ascribed to the combined effect of dual loss components, multiple interfacial polarizations, dipolar polarization, and the distinctive 3D honeycomb-like structure. Based on the above analyses and results, the 3D honeycomb-like FeNi3@C composites are promising for exploring new high-performance EM wave absorbing materials.

Supplementary Materials

The following are available online at https://www.mdpi.com/2079-4991/10/4/598/s1, Figure S1: SEM images of (a,b) S1, (c,d) S3. Figure S2: (a) Real parts of complex permittivity and (d) permeability, (b) imaginary parts of complex permittivity and (e) permeability, (c) dielectric and (f) magnetic loss tangents of S2 with the filler loadings of 10, 20, 30, and 40 wt%, respectively. Figure S3: The RL values of the S1–S3 with a thickness of 3.0 mm.

Author Contributions

Data curation, C.H.; formal analysis, J.S.; methodology, D.Z. and Y.D.; supervision, H.Z.; visualization, G.Z.; writing—Original draft preparation, C.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Program of Guangdong Province of China, grant number (2019A050510012); the Science and Technology Program of Guangzhou City of China, grant number (201802020005, 201803010071); the Guangzhou emerging industry development fund project of Guangzhou development and reform commission.

Acknowledgments

This work is supported by the Science and Technology Program of Guangdong Province of China (Grant No.2019A050510012), the Science and Technology Program of Guangzhou City of China (Grant No.201802020005, 201803010071), and by the Guangzhou emerging industry development fund project of Guangzhou development and reform commission.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, X.; Cui, X.; Chen, Y.; Zhang, X.-J.; Yu, R.; Wang, G.-S.; Ma, H. Modulation of electromagnetic wave absorption by carbon shell thickness in carbon encapsulated magnetite nanospindles–poly(vinylidene fluoride) composites. Carbon 2015, 95, 870–878. [Google Scholar] [CrossRef]
  2. Xiong, L.; Yu, M.; Liu, J.; Li, S.; Xue, B. Preparation and evaluation of the microwave absorption properties of template-free graphene foam-supported Ni nanoparticles. RSC Adv. 2017, 7, 14733–14741. [Google Scholar] [CrossRef] [Green Version]
  3. Yan, L.; Hong, C.; Sun, B.; Zhao, G.; Cheng, Y.; Dong, S.; Zhang, D.; Zhang, X. In Situ Growth of Core–Sheath Heterostructural SiC Nanowire Arrays on Carbon Fibers and Enhanced Electromagnetic Wave Absorption Performance. ACS Appl. Mater. Interfaces 2017, 9, 6320–6331. [Google Scholar] [CrossRef] [PubMed]
  4. Feng, J.; Zong, Y.; Sun, Y.; Zhang, Y.; Yang, X.; Long, G.; Wang, Y.; Li, X.; Zheng, X. Optimization of porous FeNi3/N-GN composites with superior microwave absorption performance. Chem. Eng. J. 2018, 345, 441–451. [Google Scholar] [CrossRef]
  5. Wang, C.; Ding, Y.; Yuan, Y.; He, X.; Wu, S.; Hu, S.; Zou, M.; Zhao, W.; Yang, L.; Cao, A.; et al. Graphene aerogel composites derived from recycled cigarette filters for electromagnetic wave absorption. J. Mater. Chem. C 2015, 3, 11893–11901. [Google Scholar] [CrossRef]
  6. Lv, H.; Zhang, H.; Zhang, B.; Ji, G.; He, Y.; Lin, Q. A proposed electron transmission mechanism between Fe3+/Co2+ and Fe3+/Fe3+ in the spinel structure and its practical evidence in quaternary Fe0.5Ni0.5Co2S4. J. Mater. Chem. C 2016, 4, 5476–5482. [Google Scholar] [CrossRef]
  7. Golchinvafa, S.; Masoudpanah, S.M. Magnetic and microwave absorption properties of FeNi3/NiFe2O4 composites synthesized by solution combustion method. J. Alloy. Compd. 2019, 787, 390–396. [Google Scholar] [CrossRef]
  8. Ding, X.; Huang, Y.; Wang, J. Synthesis of FeNi3 nanocrystals encapsulated in carbon nanospheres/reduced graphene oxide as a light weight electromagnetic wave absorbent. RSC Adv. 2015, 5, 64878–64885. [Google Scholar] [CrossRef]
  9. Sun, Y.; Liu, X.; Feng, C.; Fan, J.; Lv, Y.; Wang, Y.; Li, C. A facile synthesis of FeNi3@C nanowires for electromagnetic wave absorber. J. Alloy. Compd. 2014, 586, 688–692. [Google Scholar] [CrossRef]
  10. Huang, T.; Wu, Z.; Yu, Q.; Tan, D.; Li, L. Preparation of hierarchically porous carbon/magnetic particle composites with broad microwave absorption bandwidth. Chem. Eng. J. 2019, 359, 69–78. [Google Scholar] [CrossRef]
  11. Luo, H.; Gong, R.; Wang, X.; Song, K.; Zhu, C.; Wang, L. Synthesis and excellent microwave absorption properties of reduced graphene oxide/FeNi3/Fe3O4 composite. New J. Chem. 2016, 40, 6238–6243. [Google Scholar] [CrossRef]
  12. Xiang, Z.; Song, Y.; Xiong, J.; Pan, Z.; Wang, X.; Liu, L.; Liu, R.; Yang, H.; Lu, W. Enhanced electromagnetic wave absorption of nanoporous Fe3O4@carbon composites derived from metal-organic frameworks. Carbon 2019, 142, 20–31. [Google Scholar] [CrossRef]
  13. Che, R.C.; Peng, L.-M.; Duan, X.F.; Chen, Q.; Liang, X.L. Microwave Absorption Enhancement and Complex Permittivity and Permeability of Fe Encapsulated within Carbon Nanotubes. Adv. Mater. 2004, 16, 401–405. [Google Scholar] [CrossRef]
  14. Chen, Y.-H.; Huang, Z.-H.; Lu, M.-M.; Cao, W.-Q.; Yuan, J.; Zhang, D.-Q.; Cao, M.-S. 3D Fe3O4 nanocrystals decorating carbon nanotubes to tune electromagnetic properties and enhance microwave absorption capacity. J. Mater. Chem. A 2015, 3, 12621–12625. [Google Scholar] [CrossRef]
  15. Li, D.; Guo, K.; Wang, F.; Wu, Z.; Zhong, B.; Zuo, S.; Tang, J.; Feng, J.; Zhuo, R.; Yan, D.; et al. Enhanced microwave absorption properties in C band of Ni/C porous nanofibers prepared by electrospinning. J. Alloy. Compd. 2019, 800, 294–304. [Google Scholar] [CrossRef]
  16. Ding, D.; Wang, Y.; Li, X.; Qiang, R.; Xu, P.; Chu, W.; Han, X.; Du, Y. Rational design of core-shell Co@C microspheres for high-performance microwave absorption. Carbon 2017, 111, 722–732. [Google Scholar] [CrossRef]
  17. Kuang, D.; Hou, L.; Wang, S.; Luo, H.; Deng, L.; Mead, J.L.; Huang, H.; Song, M. Large-scale synthesis and outstanding microwave absorption properties of carbon nanotubes coated by extremely small FeCo-C core-shell nanoparticles. Carbon 2019, 153, 52–61. [Google Scholar] [CrossRef]
  18. Zhou, N.; An, Q.; Xiao, Z.; Zhai, S.; Shi, Z. Rational Design of Superior Microwave Shielding Composites Employing Synergy of Encapsulating Character of Alginate Hydrogels and Task-Specific Components (Ni NPs, Fe3O4/CNTs). ACS Sustain. Chem. Eng. 2017, 5, 5394–5407. [Google Scholar] [CrossRef]
  19. Wei, S.; Wang, X.; Zhang, B.; Yu, M.; Zheng, Y.; Wang, Y.; Liu, J. Preparation of hierarchical core-shell C@NiCo2O4@Fe3O4 composites for enhanced microwave absorption performance. Chem. Eng. J. 2017, 314, 477–487. [Google Scholar] [CrossRef]
  20. Liu, Q.; Xu, X.; Xia, W.; Che, R.; Chen, C.; Cao, Q.; He, J. Dependency of magnetic microwave absorption on surface architecture of Co20Ni80 hierarchical structures studied by electron holography. Nanoscale 2015, 7, 1736–1743. [Google Scholar] [CrossRef]
  21. Shen, Y. Preparation of microwave absorbing Co-C nanofibers with robust superhydrophobic properties by electrospinning. J. Mater. Sci. 2019, 13, 3365–3377. [Google Scholar] [CrossRef]
  22. Zhao, B.; Zhang, X.; Deng, J.; Bai, Z.; Liang, L.; Li, Y.; Zhang, R. A novel sponge-like 2D Ni/derivative heterostructure to strengthen microwave absorption performance. Phys. Chem. Chem. Phys. 2018, 20, 28623–28633. [Google Scholar] [CrossRef] [PubMed]
  23. Fathi, H.; Masoudpanah, S.M.; Alamolhoda, S.; Parnianfar, H. Effect of fuel type on the microstructure and magnetic properties of solution combusted Fe3O4 powders. Ceram. Int. 2017, 43, 7448–7453. [Google Scholar] [CrossRef]
  24. Biglari, Z.; Masoudpanah, S.M.; Alamolhoda, S. Solution Combustion Synthesis of Ni/NiO/ZnO Nanocomposites for Photodegradation of Methylene Blue Under Ultraviolet Irradiation. J. Electron. Mater. 2018, 47, 2703–2709. [Google Scholar] [CrossRef]
  25. Zhou, X.-M.; Wei, X.-W. Single Crystalline FeNi3 Dendrites: Large Scale Synthesis, Formation Mechanism, and Magnetic Properties. Cryst. Growth Des. 2009, 9, 7–12. [Google Scholar] [CrossRef]
  26. Ding, X.; Huang, Y.; Li, S.; Zhang, N.; Wang, J. FeNi3 nanoalloy decorated on 3D architecture composite of reduced graphene oxide/molybdenum disulfide giving excellent electromagnetic wave absorption properties. J. Alloy. Compd. 2016, 689, 208–217. [Google Scholar] [CrossRef]
  27. Bose, S.; Kuila, T.; Mishra, A.K.; Kim, N.H.; Lee, J.H. Dual role of glycine as a chemical functionalizer and a reducing agent in the preparation of graphene: An environmentally friendly method. J. Mater. Chem. 2012, 22, 9696–9703. [Google Scholar] [CrossRef]
  28. Yan, S.J.; Zhen, L.; Xu, C.Y.; Jiang, J.T.; Shao, W.Z. Microwave absorption properties of FeNi3 submicrometre spheres and SiO2@FeNi3 core–shell structures. J. Phys. D Appl. Phys. 2010, 43, 245003. [Google Scholar] [CrossRef]
  29. He, F.; Fan, J.; Ma, D.; Zhang, L.; Leung, C.; Chan, H.L. The attachment of Fe3O4 nanoparticles to graphene oxide by covalent bonding. Carbon 2010, 48, 3139–3144. [Google Scholar] [CrossRef]
  30. Liu, Q.; Ren, W.; Chen, Z.-G.; Liu, B.; Yu, B.; Li, F.; Cong, H.; Cheng, H.-M. Direct synthesis of carbon nanotubes decorated with size-controllable Fe nanoparticles encapsulated by graphitic layers. Carbon 2008, 46, 1417–1423. [Google Scholar] [CrossRef]
  31. Liu, Y.; Jiang, W.; Li, S.; Li, F. Electrostatic self-assembly of Fe3O4 nanoparticles on carbon nanotubes. Appl. Surf. Sci. 2009, 255, 7999–8002. [Google Scholar] [CrossRef]
  32. Liu, D.; Du, Y.; Xu, P.; Liu, N.; Wang, Y.; Zhao, H.; Cui, L.; Han, X. Waxberry-like hierarchical Ni@C microspheres with high-performance microwave absorption. J. Mater. Chem. C 2019, 7, 5037–5046. [Google Scholar] [CrossRef]
  33. Yang, Y.; Centrone, A.; Chen, L.; Simeon, F.; Alan Hatton, T.; Rutledge, G.C. Highly porous electrospun polyvinylidene fluoride (PVDF)-based carbon fiber. Carbon 2011, 49, 3395–3403. [Google Scholar] [CrossRef]
  34. Saeedi Afshar, S.R.; Hasheminiasari, M.; Masoudpanah, S.M. Structural, magnetic and microwave absorption properties of SrFe12O19/Ni0.6Zn0.4Fe2O4 composites prepared by one-pot solution combustion method. J. Magn. Magn. Mater. 2018, 466, 1–6. [Google Scholar] [CrossRef]
  35. Liu, T.; Xie, X.; Pang, Y.; Kobayashi, S. Co/C nanoparticles with low graphitization degree: A high performance microwave-absorbing material. J. Mater. Chem. C 2016, 4, 1727–1735. [Google Scholar] [CrossRef]
  36. Tian, X.; Meng, F.; Meng, F.; Chen, X.; Guo, Y.; Wang, Y.; Zhu, W.; Zhou, Z. Synergistic Enhancement of Microwave Absorption Using Hybridized Polyaniline@helical CNTs with Dual Chirality. ACS Appl. Mater. Interfaces 2017, 9, 15711–15718. [Google Scholar] [CrossRef]
  37. Zhang, X.; Ji, G.; Liu, W.; Zhang, X.; Gao, Q.; Li, Y.; Du, Y. A novel Co/TiO2 nanocomposite derived from a metal–organic framework: Synthesis and efficient microwave absorption. J. Mater. Chem. C 2016, 4, 1860–1870. [Google Scholar] [CrossRef]
  38. Hu, Q.; Yang, R.; Mo, Z.; Lu, D.; Yang, L.; He, Z.; Zhu, H.; Tang, Z.; Gui, X. Nitrogen-doped and Fe-filled CNTs/NiCo2O4 porous sponge with tunable microwave absorption performance. Carbon 2019, 153, 737–744. [Google Scholar] [CrossRef]
  39. 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]
  40. Wen, B.; Cao, M.-S.; Hou, Z.-L.; Song, W.-L.; Zhang, L.; Lu, M.-M.; Jin, H.-B.; Fang, X.-Y.; Wang, W.-Z.; Yuan, J. Temperature dependent microwave attenuation behavior for carbon-nanotube/silica composites. Carbon 2013, 65, 124–139. [Google Scholar] [CrossRef]
  41. Lu, M.-M.; Cao, W.-Q.; Shi, H.-L.; Fang, X.-Y.; Yang, J.; Hou, Z.-L.; Jin, H.-B.; Wang, W.-Z.; Yuan, J.; Cao, M.-S. Multi-wall carbon nanotubes decorated with ZnO nanocrystals: Mild solution-process synthesis and highly efficient microwave absorption properties at elevated temperature. J. Mater. Chem. A 2014, 2, 10540. [Google Scholar] [CrossRef]
  42. Tian, C.; Du, Y.; Xu, P.; Qiang, R.; Wang, Y.; Ding, D.; Xue, J.; Ma, J.; Zhao, H.; Han, X. Constructing Uniform Core–Shell PPy@PANI Composites with Tunable Shell Thickness toward Enhancement in Microwave Absorption. ACS Appl. Mater. Interfaces 2015, 7, 20090–20099. [Google Scholar] [CrossRef]
  43. Zhang, Z.; Lv, Y.; Chen, X.; Wu, Z.; He, Y.; Zhang, L.; Zou, Y. Porous flower-like Ni/C composites derived from MOFs toward high-performance electromagnetic wave absorption. J. Magn. Magn. Mater. 2019, 487, 165334. [Google Scholar] [CrossRef]
  44. Xiang, J.; Li, J.; Zhang, X.; Ye, Q.; Xu, J.; Shen, X. Magnetic carbon nanofibers containing uniformly dispersed Fe/Co/Ni nanoparticles as stable and high-performance electromagnetic wave absorbers. J. Mater. Chem. A 2014, 2, 16905–16914. [Google Scholar] [CrossRef]
  45. Hou, Z.-L.; Zhang, M.; Kong, L.-B.; Fang, H.-M.; Li, Z.-J.; Zhou, H.-F.; Jin, H.-B.; Cao, M.-S. Microwave permittivity and permeability experiments in high-loss dielectrics: Caution with implicit Fabry-Pérot resonance for negative imaginary permeability. Appl. Phys. Lett. 2013, 103, 162905. [Google Scholar] [CrossRef]
  46. Maglione, M.; Subramanian, M.A. Dielectric and polarization experiments in high loss dielectrics: A word of caution. Appl. Phys. Lett. 2008, 93, 032902. [Google Scholar] [CrossRef] [Green Version]
  47. Sun, X.; He, J.; Li, G.; Tang, J.; Wang, T.; Guo, Y.; Xue, H. Laminated magnetic graphene with enhanced electromagnetic wave absorption properties. J. Mater. Chem. C 2013, 1, 765–777. [Google Scholar] [CrossRef]
  48. Wu, M.; Zhang, Y.D.; Hui, S.; Xiao, T.D.; Ge, S.; Hines, W.A.; Budnick, J.I.; Taylor, G.W. Microwave magnetic properties of Co50/(SiO2)50 nanoparticles. Appl. Phys. Lett. 2002, 80, 4404–4406. [Google Scholar] [CrossRef]
  49. Zhang, J.; Wang, P.; Chen, Y.; Wang, G.; Wang, D.; Qiao, L.; Wang, T.; Li, F. Microwave Absorption Properties of Co@C Nanofiber Composite for Normal and Oblique Incidence. J. Electron. Mater. 2018, 47, 4703–4709. [Google Scholar] [CrossRef]
  50. Qiang, R.; Du, Y.; Zhao, H.; Wang, Y.; Tian, C.; Li, Z.; Han, X.; Xu, P. Metal organic framework-derived Fe/C nanocubes toward efficient microwave absorption. J. Mater. Chem. A 2015, 3, 13426–13434. [Google Scholar] [CrossRef]
  51. Yu, K.-C.; Ma, C.-C.M.; Teng, C.-C.; Huang, Y.-L.; Yuen, S.-M.; Liao, S.-H.; Lee, S.-H.; Chang, C.-C. Preparation and microwave absorbency of Fe/epoxy and FeNi3/epoxy composites. J. Alloy. Compd. 2011, 509, 8427–8432. [Google Scholar] [CrossRef]
Figure 1. (a) The photo of FeNi3@C pristine powder. (b) The photo of FeNi3@C/paraffin ring-shaped sample.
Figure 1. (a) The photo of FeNi3@C pristine powder. (b) The photo of FeNi3@C/paraffin ring-shaped sample.
Nanomaterials 10 00598 g001
Figure 2. Schematic illustration of the synthesis of 3D honeycomb-like FeNi3@C composites.
Figure 2. Schematic illustration of the synthesis of 3D honeycomb-like FeNi3@C composites.
Nanomaterials 10 00598 g002
Figure 3. SEM images of S2 at (a) low, (b) medium, and (c) high magnification. (df) TEM images of S2 at different resolutions. EDS mapping images of (g) C, (h) Ni, and (i) Fe elements.
Figure 3. SEM images of S2 at (a) low, (b) medium, and (c) high magnification. (df) TEM images of S2 at different resolutions. EDS mapping images of (g) C, (h) Ni, and (i) Fe elements.
Nanomaterials 10 00598 g003
Figure 4. (a) XRD patterns of the S1, S2, and S3 samples. (b) Raman spectra of the S1, S2, and S3 samples. (c) Thermogravimetric (TG) curves of all of the samples under an atmosphere of air. (d) Hysteresis loops of the S1, S2, and S3 samples.
Figure 4. (a) XRD patterns of the S1, S2, and S3 samples. (b) Raman spectra of the S1, S2, and S3 samples. (c) Thermogravimetric (TG) curves of all of the samples under an atmosphere of air. (d) Hysteresis loops of the S1, S2, and S3 samples.
Nanomaterials 10 00598 g004
Figure 5. X-ray photoelectron spectroscopy (XPS) spectra of the sample of S2: (a) survey scan of FeNi3@C composites; (b) C 1s spectrum; (c) Fe 2p spectrum; (d) Ni 2p spectrum.
Figure 5. X-ray photoelectron spectroscopy (XPS) spectra of the sample of S2: (a) survey scan of FeNi3@C composites; (b) C 1s spectrum; (c) Fe 2p spectrum; (d) Ni 2p spectrum.
Nanomaterials 10 00598 g005
Figure 6. Calculated reflection loss (RL) curves of the S2–paraffin composites with different absorber thicknesses: (a) S2–10 wt%, (b) S2–20 wt%, (c) S2–30 wt%, and (d) S2–40 wt%.
Figure 6. Calculated reflection loss (RL) curves of the S2–paraffin composites with different absorber thicknesses: (a) S2–10 wt%, (b) S2–20 wt%, (c) S2–30 wt%, and (d) S2–40 wt%.
Nanomaterials 10 00598 g006
Figure 7. Frequency dependence of RL curves at different thicknesses of (a) S1, (b) S2, and (c) S3 with a filler loading ratio of 30 wt%. Three-dimensional diagrams of RL values for (d) S1, (e) S2, and (f) S3.
Figure 7. Frequency dependence of RL curves at different thicknesses of (a) S1, (b) S2, and (c) S3 with a filler loading ratio of 30 wt%. Three-dimensional diagrams of RL values for (d) S1, (e) S2, and (f) S3.
Nanomaterials 10 00598 g007
Figure 8. (a) Real and (b) imaginary parts of complex permittivity for S1, S2, and S3. (d) Real and (e) imaginary parts of complex permeability for S1, S2, and S3. (c) Dielectric loss tangents and (f) magnetic loss tangents of S1, S2, and S3.
Figure 8. (a) Real and (b) imaginary parts of complex permittivity for S1, S2, and S3. (d) Real and (e) imaginary parts of complex permeability for S1, S2, and S3. (c) Dielectric loss tangents and (f) magnetic loss tangents of S1, S2, and S3.
Nanomaterials 10 00598 g008
Figure 9. Curves of ε versus ε (Cole–Cole semicircles) for (a) S1, (b) S2, and (c) S3. (d) Frequency dependences of μ ( μ ) 2 f 1 values for S1–S3.
Figure 9. Curves of ε versus ε (Cole–Cole semicircles) for (a) S1, (b) S2, and (c) S3. (d) Frequency dependences of μ ( μ ) 2 f 1 values for S1–S3.
Nanomaterials 10 00598 g009
Figure 10. Schematic diagram of microwave absorption for the 3D honeycomb-like FeNi3@C.
Figure 10. Schematic diagram of microwave absorption for the 3D honeycomb-like FeNi3@C.
Nanomaterials 10 00598 g010
Table 1. Some FeNi3-based composites for microwave absorption (MA) materials reported in publications.
Table 1. Some FeNi3-based composites for microwave absorption (MA) materials reported in publications.
SamplesMinimum RL Value (dB)Frequency Range (RL ≤ −10 dB) GHzThickness (mm)Content (wt%)Reference
FeNi3/N-GN−34.00 3.60–18.00 3.2350[4]
FeNi3/NiFe2O4−15.00 unknown1.4080[7]
FeNi3@C−47.607.28–10.403.0010[8]
FeNi3@C−43.308.70–13.782.0040[9]
RGO/FeNi3/Fe3O4−46.60 6.10–15.361.9060[11]
FeNi3@RGO/MoS2−30.39 unknown2.0040[26]
FeNi3/epoxy−20.0013.10–16.201.60unknown[51]
S1−36.43 5.40–18.00 4.0030This work
S2−40.60 5.00–18.00 3.0030This work

Share and Cite

MDPI and ACS Style

Han, C.; Zhang, H.; Zhang, D.; Deng, Y.; Shen, J.; Zeng, G. Ultrafine FeNi3 Nanocrystals Embedded in 3D Honeycomb-Like Carbon Matrix for High-Performance Microwave Absorption. Nanomaterials 2020, 10, 598. https://doi.org/10.3390/nano10040598

AMA Style

Han C, Zhang H, Zhang D, Deng Y, Shen J, Zeng G. Ultrafine FeNi3 Nanocrystals Embedded in 3D Honeycomb-Like Carbon Matrix for High-Performance Microwave Absorption. Nanomaterials. 2020; 10(4):598. https://doi.org/10.3390/nano10040598

Chicago/Turabian Style

Han, Congai, Haiyan Zhang, Danfeng Zhang, Yunfei Deng, Junyao Shen, and Guoxun Zeng. 2020. "Ultrafine FeNi3 Nanocrystals Embedded in 3D Honeycomb-Like Carbon Matrix for High-Performance Microwave Absorption" Nanomaterials 10, no. 4: 598. https://doi.org/10.3390/nano10040598

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