Jian
Xu
a,
Wenjun
Ma
a,
Peng
He
a,
Yukang
Zhou
a,
Xiaoyun
Liu
a,
Yi
Chen
*b,
Peiyuan
Zuo
a and
Qixin
Zhuang
*a
aThe Key Laboratory of Specially Functional Polymeric Materials and Related Technology, Ministry of Education, School of Materials Science and Engineering, East China University of Science and Technology, Shanghai 200237, P. R. China. E-mail: qxzhuang@ecust.edu.cn
bShanghai Spaceflight Precision Machinery Institute, Shanghai, 201108, P. R. China. E-mail: 18801972192@139.com
First published on 24th November 2022
Three-dimensional (3D) network structures assembled from two-dimensional (2D) MXene nanomaterials hold enormous potential for microwave absorption (MA) due to strong dielectric loss ability. However, the uncontrollable oversized conductive paths caused by self-stacking issues and the natural lack of magnetic loss hinder the further MA application of the assembled 3D MXene. Herein, we report a facile water-induced self-assembly method to develop a series of 3D hierarchical magnetic nanocrystal@C@MXene hybrids with tunable network structures. The design of the hierarchical structure can isolate each electromagnetic component, thereby ensuring the efficient synergy of the high-density ultrafine magnetic components and high-loss dielectric components. The resulting massive heterointerfaces greatly enhance the interface polarization and reinforce the dielectric loss capability. Moreover, through control over the MXene load, the size of 3D networks can be precisely regulated to adjust the impedance matching. According to experimental and finite element simulation, all the 3D hierarchical magnetic hybrids were found to possess impressive MA properties. Specifically, ZnFe2O4@C@MXene exhibits a minimum reflection loss value of −62.59 dB over an effective absorption bandwidth of 4.42 GHz at a thickness of only 1.33 mm. This work provides a flexible route for constructing 3D hierarchical network materials to realize highly efficient MA performance with thin thickness.
An effective solution strategy to enhance electromagnetic synergy is to incorporate magnetic components to build 3D network structures through interfacial engineering.18,19 Apart from introducing magnetic losses, the interfacial engineering strategy can anchor MXenes on the heterogeneous interface and construct a controllable conductive network.20 The typical method for forming MXene-based network structures is to assemble MXenes with other components via hydrogen bonds or electrostatic force.21 However, normal magnetic materials cannot provide an active surface for assembly.22 Accordingly, some strategies based on surface modification have been exploited. These strategies rely on organic ligands, like cetrimonium bromide (CTAB), oleylamine (OA), and poly(diallyldimethylammonium chloride) (PDDA), to assist the self-assembly of magnetic structures and MXene.23–26 Despite the fact that various magnetic components were successfully assembled with MXene through these time-consuming strategies, the resulting irregular network structure is unfavorable for MA performance.27 Therefore, it remains a significant challenge to fabricate a controllable 3D magnetic network through a more feasible and simpler approach.
Proper selection of magnetic components will aid in the construction of 3D hierarchical structures and the enhancement of MA performance.19 The magnetic nanoparticles are the preferred choice due to their simple synthesis and high magnetic loss.28 Meanwhile, reducing the size of magnetic nanoparticles, which increases the active surface area for assembling with MXenes, can also enhance interface loss and reduce the local accumulation of MXenes.9 Nevertheless, ultrafine magnetic nanoparticles of small size tend to assemble on relatively larger-sized nanosheets instead of isolating MXenes and forming networks.29,30 Besides, the self-aggregation of ultrafine nanoparticles further exacerbates the self-restacking of MXenes, weakening the advantages of magnetic nanostructures and 3D networks.31 Accordingly, the construction of controllable MXene networks with high-density and uniformly dispersed magnetic nanocomponents still needs more investigations in depth.
Herein, we provide a general synthesis method of 3D hierarchical magnetic nanocrystal@C@MXene hybrids, including isolation structures composed of ultrafine magnetic nanocrystals densely isolated in spherical carbon matrixes and the controllable MXene networks anchored on matrix interfaces. The orderly isolation can efficiently coordinate the superiority of each electromagnetic component, while massive heterointerfaces created by the continuous isolation structure can greatly enhance the interface polarization and reinforce the dielectric loss capability. More importantly, the involved water-induced self-assembly method, which not only utilizes water as a solvent but also activates and participates in the assembly, can assemble 2D nanosheets into 3D MXene networks. Therefore, the size of networks can be orderly regulated, thus enabling effective modulation of impedance matching and attenuation paths. As expected, the hierarchical hybrids with tunable MXene networks were found to possess impressive MA properties. Specifically, all obtained 3D hierarchical ZnFe2O4@C@MXene materials with different sized networks show excellent MA performance. Moreover, regulating the size of network structures can precisely adjust the impedance matching and prolong the attenuation paths, thus greatly promoting the MA performance at ultrathin thicknesses. ZnFe2O4@C@MXene with 3D conductive networks delivers a minimum reflection loss (RLmin) value of −62.59 dB at 14.16 GHz over an effective absorption bandwidth (EAB, RL < −10 dB) of 4.42 GHz at an ultrathin thickness of only 1.33 mm. In short, the present work provides a flexible route for the structural design of 3D hierarchical network materials and promotes the further development of ultrathin MA materials.
The synthesis of hierarchical ZnFe2O4@C@MXene (denoted as ZFCM) is taken as a model system. Monodisperse ZnFe-G spheres (diameter size: ca. 400 nm) were obtained by a solvothermal method.32 After a facile water induction, the colour of the corresponding dispersion changed from yellow to orange (Fig. 2a), implying the conversion of ZnFe-G to activated ZnFe-GH. And the obtained ZnFe-GH exhibits a similar morphology to ZnFe-G (Fig. 2b and S4†). Ti3C2Tx MXene (lateral size: ca. 0.5–1.5 μm, Fig. 2c) is prepared by a wet chemistry method.33 Subsequently, the slowly added MXene nanosheets were rapidly assembled with the ZnFe-GH spheres driven by electrostatic force (Fig. S3†). After sitting for a few minutes, the obtained hybrids quickly settled at the bottom of the container, resulting in a colourless supernatant (Fig. 2a), which confirmed the successful assembly of ZnFe-GH spheres and MXene.
The morphology of ZnFe-GH@MXene-3 was observed with a scanning electron microscope (SEM) and transmission electron microscope (TEM). From Fig. 2d and e, the ZnFe-GH spheres are perfectly assembled with few MXene nanosheets due to the excellent flexibility of MXene. Simultaneously, the edges of the assembled nanosheets are self-folded through van der Waals forces, hydrogen bonds, and the centrifugal force generated by stirring (Fig. 2f), resulting in petal-like MXene shells that can provide local conductive networks anchored on the spherical surface.9,34 In addition, high-resolution TEM (HRTEM, Fig. 2g) verifies that the MXene structure remains stable after assembly, showing periodic lattice fringes with an interlayer distance of 0.26 nm corresponding to the (100) plane of Ti3C2Tx MXene.35Fig. 2h shows the Fourier transform infrared spectroscopy (FTIR) spectra of water-modified ZnFe-GH and self-assembled ZnFe-GH@MXene. The weakening of CH2/CH3 and C–O vibrations is observed in the FTIR spectrum of ZnFe-GH, indicating the partial removal of glycerate ligands.36 And the pronounced peaks ascribed to Ti–O, C–H, C–O, and CO vibrations are observed in the FTIR spectrum of ZnFe-GH@MXene, further confirming the synthesis of ZnFe-GH@MXene.8
X-ray photoelectron spectroscopy (XPS) and X-ray diffraction (XRD) were conducted to investigate the mechanism of the water-induced self-assembly method. The C 1s core-level XPS spectra of ZnFe-GH show a lower proportion of C–O and CO relative to ZnFe-G (Fig. 2j), due to the surface deconstruction caused by the attack of OH− ions in water. In the XRD pattern of ZnFe-G, there is only an obvious diffraction peak of about 12°, which is the characteristic of metal glycerate (Fig. 2i).37 After water dispersion, the characteristic peak transformed into a board peak. And the corresponding partial enlargement of the XRD pattern of ZnFe-GH displays two new peaks at 34° and 61°, suggesting that a trace amount of hydroxide was synthesized in situ.38 Since such self-reconstruction is a kind of time-dependent reaction, only a small amount of M–OH is generated on the surface of M–G in a short time.39 Notably, the characteristic peak (002) of MXene is nearly invisible in the XRD pattern of ZnFe-GH@MXene, illustrating that MXene nanosheets are effectively isolated by ZnFe-GH spheres, resulting in the expanding of layer spacing.26
Through the fine-tuning of the MXene loading, the size of MXene networks can be precisely regulated from the nanoscale to the microscale. Different from the 2D self-stacking structure of the pristine MXene, the petal-like MXene shells of preliminary assembled hybrids provide 3D connection spaces for the interaction with neighboring nanosheets (Fig. S6†). After subsequent thermal annealing of ZnFe-GH@MXene, the hierarchical magnetic ZFCM is generated. As shown in Fig. 3, all the samples perfectly inherit the 3D network morphologies of the assembled precursors. The ZFCM-3 hybrids with petal-like shells show good dispersion (Fig. 3d and i), and the morphology of the exposed core is exactly the same as the morphology of ZnFe2O4@C (denoted as ZFC, Fig. 3h), which is synthesized by annealing ZnFe-GH in a similar method (Fig. S8†). Likewise, the XRD patterns of the ZFCM series are similar to those of ZFC (Fig. S7†), indicating that the ZnFe-GH cores of the assembled precursors were successfully transformed into ZFC. As the MXene content in the assembled precursors is increased to 5%, the obtained ZFC cores are fully encapsulated and interconnected by MXene nanosheets, resulting in many local interconnected networks (Fig. 3e and j). By continuously increasing the load of MXene, a microscale 3D network structure was obtained for ZFCM-7, which is due to the bridge connection between MXene nanosheets and nanoscale 3D networks (Fig. 3f and k). Accordingly, it can be concluded that the morphologies of the resulting conductive networks can be directly controlled by adjusting the MXene load of assembled precursors, thereby modifying the dielectric properties of ZFCM for meeting the requirements of attenuating microwaves effectively.
In the TEM image of ZFCM-3 (Fig. 3n and o), the core of the hybrid exhibits a similar morphology to ZFC (Fig. 3m), which is consistent with the SEM results. In the HRTEM of the exposed core (Fig. 3p), the ultrafine nanocrystals with periodic lattice fringes are orderly separated by the carbon matrix, and the lattice space of 0.48 nm is attributed to the (111) plane of ZnFe2O4 (JCPDS no. 22-1012, d = 0.487 nm).40 The uniform carbon matrix derived from organic residues and massive ZnFe2O4 nanocrystals together form continuous heterogeneous interfaces. It is further verified by the high-density distribution of nanocrystals in the core of ZFCM, as evidenced in the EDS mapping of ZFCM-3 and ZFC (Fig. 3r and S9†).
For comparison, 3D networks composed only of MXene (denoted as 3D MXene) were synthesized via a hard template method. Compared to ZFCM-7 with core-supported 3D networks, the obtained 3D MXene exhibits obvious collapse and local aggregation, which would lead to undesired local impedance mismatch (Fig. 3l, q and S11†).19
As shown in Fig. 4a, the ZFCM series show good magnetism due to the isolated magnetic nanocrystals. From the hysteresis loops of the obtained samples, ZFC exhibits the strongest static magnetic properties (Fig. 4a). With the increase of MXene load, the saturation magnetization value and coercivity of the ZFCM series decreased, indicating that the structural adjustment of ZFCM affects the static magnetic properties of the resulting hybrids.
Fig. 4b shows the Raman spectrum of ZFC, MXene, and ZFCM-7. ZFCM-7 presents peaks centered at 198, 340, 398, 511, 575, 670, 1098, 1347, and 1580 cm−1. The peaks at 198, 398, and 575 cm−1 come from the contribution of MXene.41 The peaks at 340, 511, 670, and 1098 cm−1 are assigned to the contribution of ZnFe2O4. Besides, the two broad bands at 1347 and 1580 cm−1 are the characteristic D and G bands of carbon, which come from the carbon matrix of ZFC and the carbon exposed by mild oxidation of MXene42,43.
X-ray photoelectron spectroscopy (XPS, Fig. 4c) reveals the coexistence of C, Ti, F, Fe, Zn, and O elements in ZFCM-7. Fig. 4d shows the C 1s core-level XPS spectrum of ZFCM-7, which could be fitted with five components ascribed to C–Ti (282.1 eV), C–Ti–Tx (283.5 eV), C–C (284.8 eV), C–O (286.1 eV), and CO (288.9 eV), respectively.44 The emergence of C–Ti and C–Ti–Tx relative to ZFC further validates the successful synthesis of ZFCM-7. The Fe 2p core-level XPS spectrum (Fig. 4e) of ZFCM-7 could be fitted with two components ascribed to Fe2+ (711.1 and 724.5 eV) and Fe3+ (714.0 and 727.8 eV). The Zn 2p core-level XPS spectrum could be fitted with one component ascribed to Zn2+ (1022.1 and 1045.2 eV).45 The Ti 2p core-level XPS spectrum (Fig. 4f) of ZFCM-7 could be fitted with four components ascribed to Ti–C (455.1 eV), Ti2+ (456.0 eV), Ti3+ (457.0 eV), and Ti–O (458.9 eV), respectively.8 The oxidation of Ti atoms on the surface accounts for a higher proportion of Ti–O relative to the pristine Ti3C2Tx MXene, which could be ascribed to the mild oxidation of MXene by its oxygen-containing functional groups and the decomposition products of the assembled organometallic precursor.34
To further validate the generality of the synthetic strategy, a series of hierarchical magnetic nanocrystal@C@MXene hybrids, including Fe3O4@C@MXene, CoO@C@MXene, and CuCo@C@MXene, were synthesized simply by using the respective precursors. All samples show strong magnetism under an external magnetic field (Fig. S2†), which is consistent with the VSM results (Fig. S12†). The formation of hybrids is confirmed by the corresponding XPS spectra and XRD patterns (Fig. S13 and S15†). The SEM images of the obtained magnetic hybrids exhibit a similar spherical morphology with an obvious petal-like MXene shell. TEM images provide direct evidence that the hierarchical structures are formed, while there is a slight difference in the morphological details of magnetic nanocrystals, which is due to the variations in the doping atoms of precursors (Fig. S14 and S15†). Moreover, the corresponding hybrids with microscale 3D networks were easily synthesized by increasing the MXene load in the assembled precursors (Fig. S16†), realizing the transformation from a nanoscale 3D network to a microscale 3D network. Our preliminary results suggest that the present approach is simple and general, and can be applied to synthesize hierarchical magnetic hybrids with tunable conductive networks and adjustable element doping.
To evaluate the MA performances of hierarchical structures with tunable networks, the electromagnetic parameters of samples were analyzed in the range of frequencies 2–18 GHz. In general, the real part of complex permittivity (ε′) and complex permeability (μ′) reflects the ability of absorbers to store electromagnetic energy, while the imaginary part of complex permittivity (ε′′) and complex permeability (μ′′) represents the loss capability of energy.46 As shown in Fig. 5a and b, due to the introduction of MXene networks, both the ε′ and ε′′′ of ZFCM series are significantly elevated. The ε′ plots of all samples showed a decreasing trend with increasing frequency in the range of 2–18 GHz, which is consistent with Debye theory.47 According to Debye theory, the conduction part and the polarization part
can be separated from ε′′ and investigated as follows:48,49
![]() | (1) |
![]() | (2) |
![]() | (3) |
All ZFCM series exhibit undulant curves containing many semicircles (Fig. 5g and S19†), which may arise from interfacial polarization and dipolar polarization.23,46 The continuous interfaces between the three phases of MXene, carbon matrixes, and magnetic nanocrystals can significantly trigger the interfacial polarization.50 The residual end groups, the intrinsic defects, and the oxidation-generated defects of MXene as well as the defects in carbon matrixes can lead to the asymmetry of electron spatial distribution and the formation of dipole moment, resulting in dipolar polarization under microwave irradiation.52–54
Due to the introduction of magnetic ZnFe2O4 nanocrystals isolated by the carbon matrix, multiple magnetic losses are introduced into all ZFCM series, showing magnetic spectrum curves (Fig. 5d and e).55 The μ′ plots of ZFCM series exhibit higher μ′ in the high frequency range, representing its better capability of storing magnetic energy in this range.56 Interestingly, the μ′ and μ′′ plots of ZFCM-7 have the most obvious fluctuation, probably attributed to the magnetic coupling effect of the assembled microscale 3D networks.57 In the tested frequency range, the magnetic loss is mainly derived from magnetic resonance and eddy current loss. Normally, magnetic resonance can be divided into natural resonance and exchange resonance, which always takes place in the range of 0.1–10 GHz and over the 10 GHz range, respectively.58 The magnetic loss tangent value (tanδμ = μ′′/μ′) plots of ZFCM series exhibit strong fluctuations in the range of 2–6 GHz, which can be recognized as a symbol of natural resonance.59 And the fluctuations of tan
δμ plots correspond to the magnetic exchange resonance in the range of 12–18 GHz. The magnetic eddy current loss can be evaluated by using μ′′(μ′)−2f−1, where μ′ and μ′′ correspond to the real and imaginary parts of complex permeability, respectively, and f represents the electromagnetic field frequency. When eddy current loss exists, μ′′(μ′)−2f−1 does not change with frequency. In Fig. 6a, all μ′′(μ′)−2f−1 plots of ZFCM series remain constant at 9–18 GHz, revealing the existence of eddy current loss in the high frequency range (Fig. 5h).60 Meanwhile, an attenuation constant (α) is introduced to describe the loss capacity of an absorber as described in eqn (4):51
![]() | (4) |
![]() | ||
Fig. 6 3D RL values of (a) ZnFe2O4, (b) ZFC, (c) ZFCM-3, (d) ZFCM-5, (e) ZFCM-7, and (f) 3D MXene. (g) 2D contour maps of the |Zin/Z0| values of ZFC, ZFCM series, and 3D MXene. |
It can be concluded from the formula that the attenuation constant (α) increases with the dielectric loss and magnetic loss capacity. As shown in Fig. 5i, the α values of ZFCM series are significantly improved with the increase of MXene load, demonstrating that the loss capacity can be enhanced by regulating the size of MXene networks.11 In addition, the 3D MXene composed entirely of the MXene exhibits the highest α value, which is due to its excellent dielectric properties.
To intuitively evaluate the electromagnetic absorbing capacity of materials, the reflection loss (RL) was calculated based on the transmission line equation as follows:
![]() | (5) |
![]() | (6) |
The plots of RL versus frequency and thickness are shown in Fig. 6 and S21.† For comparison with multiphase hybrids, pristine ZnFe2O4 was prepared (Fig. S10†). The pristine ZnFe2O4 exhibits negligible absorption of electromagnetic waves (Fig. 6a) due to low dielectric properties and impedance mismatch (Fig. S20†). The ZFC with carbon matrixes displays an RLmin value of −16.38 dB with a thickness of 2.50 mm (Fig. 6b). On account of the highly efficient synergistic effect among the MXene shell, carbon matrixes, and magnetic nanocrystals, ZFCM-3 and ZFCM-5 display RLmin values of −57.66 dB at 7.24 GHz and −61.19 dB at 8.81 GHz, respectively (Fig. 6c and d). Notability, ZFCM-7 with a microscale 3D network structure represents the most prominent RLmin value of −62.59 dB at 14.16 GHz and an excellent EAB of 4.42 GHz, and the corresponding fitting thickness reaches an astonishing 1.33 mm, realizing high MA under ultrathin thickness (Fig. 6e). Conversely, the 3D MXene with an excellent attenuation constant only offers an unsatisfactory RLmin value of −7.94 GHz, which originates from the incompatibility between the high dielectric constant and impedance matching (Fig. 6f and S22†).11
To further explore the effect of the tunable 3D structure on the impedance matching of ZFCM, the normalized characteristic impedance (|Zin/Z0|) is introduced, which can intuitively evaluate the degree of impedance matching. According to transmission line theory, when the value of |Zin/Z0| is close to or equal to 1, the impedance matching is favorable, which demonstrates that the electromagnetic wave can successfully enter the absorber instead of turning into an undesired reflection.62 In the 2D contour maps of |Zin/Z0| values (Fig. 6g), the deep blue region belongs to the region with good impedance matching. Different from the complete impedance mismatch of 3D MXene and ZnFe2O4 (Fig. S20b†), ZFC achieves impedance matching in a large area. However, the |Zin/Z0| value of ZFC changes rapidly with frequency, resulting in a relatively high RLmin and an imperfect EAB of 3.48 GHz. After being assembled with the MXene, the matching region of the obtained ZFCM moves towards higher frequency and lower thickness with the increase of MXene load. Notably, ZFCM-7 with microscale 3D networks exhibits a mostly continuous impedance matching region, concentrating in the ultrathin thickness and the high frequency region, which is consistent with the RL result. Moreover, the MA performance of CuCo@C@MXene with microscale 3D networks (denoted as CCCM-7) was also tested. Similar to the impedance performance of ZFCM-7, the impedance matching region of CCCM-7 is concentrated in the high frequency and low thickness region, resulting in a highly efficient MA under an ultrathin thickness of 1.40 mm, and the corresponding EAB reaches 5.60 GHz, further elucidating the superiority of the 3D hierarchical network structure (Fig. S23†).
Overall, based on the details (the RLmin values and the broadest EAB) of the MA properties exhibited by ZFCM with different morphologies, it is obvious to conclude that the facile synthesis strategy of 3D hierarchical structures can synergistically integrate the advantages of MXenes, carbon matrixes and magnetic nanocrystals to achieve a superior MA performance (Fig. 7a). Besides, the step-by-step regulation of the MXene networks realizes the thinning of absorbers (Fig. S24†). The special RL values (SRL = RLmin/thickness) and the special EAB (SEAB = EAB/thickness) are introduced to reasonably evaluate MA ability considering the thickness.63,64 As shown in Fig. 7b and Tables S1 and S2,† compared to other MXene-based absorbers and 3D hierarchical absorbers, ZFCM-7 prepared in this work exhibits better wave absorption properties at a lower thickness, achieving a balance of both strong absorption and a wide EAB.
To further investigate the influence of the hierarchical magnetic structure on the electromagnetic wave loss paths and loss methods, the electric field intensity distribution and magnetic loss distribution of 3D MXenes and ZFCM-7 were simulated by the finite element method based on COMSOL Multiphysics. As shown in Fig. 7c, an inhomogeneous electric field is concentrated inside the cavity structure of 3D MXene. The relatively high local field strength performance is attributed to the reflection and scattering of electromagnetic waves inside the cavity, which reveals that some of the electromagnetic waves can penetrate the MXene layer and propagate to the interior of the hollow sphere. As for ZFCM-7 with multiple heterogeneous interfaces, electric fields with strong energy density are present at the massive interfaces and magnetic nanocrystals. This is attributed to the local electric field distortion caused by the dielectric constant difference among the multiple heterostructures, which can provide evidence for the existence of strong interfacial polarization.65 In the overall view and cross-section view of simulated magnetic loss, compared with the negligible magnetic loss performance of 3D MXenes, ZFCM-7 exhibits a significant magnetic loss distribution. Notably, the magnetic loss exhibits an array-like dispersion with high energy density, which is derived from the high-density carbon-encapsulated magnetic nanocrystals. At the same time, it can be found that strong magnetic loss distribution is also intense in the MXene shell and the carbon matrix, which might be attributed to the magnetic coupling between neighboring encapsulated magnetic nanocrystals.19 The heterojunction composed of highly separated magnetic nanocrystals and the carbon matrix effectively suppresses the skin effect, enhances their magnetic interaction, and induces the formation of high-density magnetic coupling networks, which can further promote the magnetic loss.66 Hence, the construction of the 3D hierarchical magnetic structure successfully introduced various loss modes such as interface polarization and magnetic coupling, which effectively extended the microwave absorption path.
Accordingly, with the results of the above experiments and simulation analysis, the absorbing mechanism of the 3D hierarchical magnetic structure is initially proposed as shown in Fig. 7d. First, benefiting from the water-induced self-assembly strategy, the original self-stacked structure of MXene is greatly destroyed. The surfaces of the obtained samples are composed of randomly oriented and loosely connected MXenes, which facilitates the efficient entry of electromagnetic waves rather than reflection. Simultaneously, the external MXene nanosheets formed a microscale 3D conductive network, which can enable the free migration of excited electrons generated by electromagnetic waves. Second, the hierarchical structure endows samples with continuous heterointerfaces, resulting in a rich interface polarization loss between the three phases of MXene, carbon matrixes, and magnetic nanocrystals. The residual end groups on the MXene surfaces, as well as the defects of MXenes and carbon matrixes, can lead to the asymmetry of electron spatial distribution and the formation of dipolar polarization sites, providing dipole polarization losses. Third, the isolated magnetic nanocrystals with a high density provide multiple magnetic losses including magnetic resonance and eddy current loss, further eroding the energy of electromagnetic waves. Moreover, the folded MXene on the 3D shell promotes multilevel reflection and scattering, prolonging the path of attenuation. As a result, the unique hierarchical structure delivers excellent MA performance at ultrathin thickness through the synergistic high efficiency of high-density magnetic components and high-loss dielectric networks.
Footnote |
† Electronic supplementary information (ESI) available: Experimental sections; zeta potentials, digital images, SEM images, TEM images, elemental mapping, XPS spectra, HRTEM images, XRD patterns, electrical conductivity values, and microwave absorption performance data of related samples. Table of microwave performance comparison. See DOI: https://doi.org/10.1039/d2ta07718c |
This journal is © The Royal Society of Chemistry 2023 |