1 Introduction

With the advancement of science and technology, the development of electronic products and wireless communication technology has also brought about electromagnetic wave (EMW) pollution problems, which seriously endanger human health [1,2,3,4,5,6,7]. Therefore, exploring advanced and multifunctional EMW absorbing materials is the most effective way to solve this problem. The characteristics of excellent EMW absorption materials are light weight, thin thickness, strong absorption capacity and wide effective absorption bandwidth [8, 9]. Expanded graphite (EG), graphene (GR), carbon nanofibers (CNF), conductive polymers, carbon-based biomaterials, metal–organic framework, and their derivatives have attracted much attention as they have the advantages of low density, rich functional groups, and tunable properties [10,11,12,13,14,15]. However, single-component materials have disadvantages such as poor dielectric properties and impedance mismatch. Therefore, the multi-scale structural design of EMW absorbent materials and construction of multi-component composite materials are effective ways to solve this problem [16,17,18,19,20].

Constructing multi-material systems and forming several heterogeneous interfaces to increase the interface polarization effect is a typical tactic. Multi-component composites could fully utilize the advantages of each component to adjust the impedance matching and enrich the electromagnetic loss mechanism. The electronic and crystal structure asymmetries at non-uniform interfaces lead to lattice distortions, charge mismatches and energy band migration, which dissipate EMW energy [21,22,23,24,25]. As a traditional carbon material, carbon fiber (CF) has a high aspect ratio and electrical conductivity, making it a widely employed component in the EMW absorption field [26,27,28,29,30]. Unfortunately, the impedance matching capability of single-component CF is poor, resulting in a large amount of electromagnetic waves being reflected rather than absorbed during transmission. This limits the prospects for its application in EMW absorption. The introduction of nanocompounds into the composite system can endow the composites with multilayered structure, excellent impedance matching, and enriched EMW loss mechanisms. Semiconductor materials have adjustable conductivity and excellent dielectric properties and can be combined with other materials to form heterogeneous structures [31, 32]. These heterogeneous interfaces can enhance the interface polarization effect and further enhance the EMW absorption performance. Therefore, carbon fibers were composited with semiconductor materials in order to tune the composition, interfaces, and defects of the materials to optimize the EMW absorption properties of the materials [33,34,35,36]. Tin dioxide (SnO2) semiconductor materials have good chemical stability and unique structural features. It has good electrical conductivity and dielectric properties, and can effectively regulate the dielectric loss of the material. Tin disulfide (SnS2) and stannous sulfide (SnS) have a unique layered structure, large layer spacing, high reversible capacity, and excellent electron transport properties [37,38,39,40]. These characteristics contribute to the multiple dissipation of EMWs, thus having great application potential in the field of EMW absorption.

Hence, in this work, a variety of stannide-modified CF composites (SnS/SnS2/SnO2/CF) were constructed. The CF skeleton was first prepared by electrostatic spinning and high-temperature calcination processes. Then, SnS/SnS2/SnO2/CF with a large number of non-uniform interfaces were successfully prepared by hydrothermal and thermal decomposition reduction processes. The degree of sulfidation and thermal reduction were adjusted by controlling the content of the sulfur source and thermal decomposition temperature, to realize the coexistence of multi-component tin compounds. Multi-component composites have a large number of non-uniform interfaces that can effectively regulate impedance matching. Strong interactions or synergistic effects are generated at the interfaces to realize the interface polarization loss effect and multiple attenuation of EMW, thus enhancing the EMW loss capability. The SnS/SnS2/SnO2/CF composite achieved minimum reflection loss (RLmin) = − 46.74 dB at a matched thickness of 2.0 mm and the maximum effective absorption bandwidth (EAB)max = 5.28 GHz at a matched thickness of 1.7 mm. The SnS/SnS2/SnO2/CF epoxy composite coatings were tested using a three-electrode system, and the results showed that the composite coatings have long-term corrosion resistance. Therefore, this work provides a valuable reference for the preparation of EMW absorbers with thin thickness, lightweight, and excellent absorption performance that can be applied under harsh and complex conditions.

2 Experimental Section

2.1 Materials

Tin (II) chloride dihydrate (SnCl2·2H2O), Trisodium citrate dihydrate (Na3C6H5O7·2H2O), thioacetamide (TAA, C2H5NS), acetic acid (CH3COOH), N, N-dimethylformamide (DMF, 99%), polyacrylonitrile (PAN, Mw = 150,000), anhydrous ethanol (CH3CH2OH, AR). Deionized water. All chemicals are analytical grade and are used directly without further purification.

2.2 Preparation of Carbon Fiber

First weigh 1 g PAN and add it to a 15 mL beaker. Then add 10 mL DMF and stir for 12 h to obtain PAN spinning solution. Transfer the spinning solution into a 10 mL syringe and attach a 19 G needle. Subsequently, the syringe was placed in the electrostatic spinning device and the parameters were adjusted (Voltage: 19 kV; Injection rate: 0.9 mL h−1; Distance between syringe and collection device: 17 cm). The obtained fiber membrane was placed in a vacuum oven at 70 °C to dry overnight. Subsequently, the fiber membrane was placed in a muffle furnace and heated up to 260 °C at 2 °C min−1 and held for 2 h. Finally, the pre-oxidized fibers were carbonized in a tube furnace at 900 °C for 3 h, and the heating rate was 5 °C min−1 to obtain CF.

2.3 Preparation of SnO2/CF

First, 0.3 g of SnCl2·2H2O and 0.818 g of Na3C6H5O7·2H2O were weighed and dissolved in a mixed solution of 15 mL of CH3CH2OH and 15 mL of deionized water and stirred for 1 h. The obtained solution is then transferred to a 50 mL teflon lining and the appropriate amount of CF is added. Subsequently, SnO2/CF was obtained by keeping it at 180 °C for 12 h. Finally, wash with water and ethanol and dry overnight in a vacuum oven at 70 °C.

2.4 Preparation of SnS2/SnO2/CF

First, 0.09 g of TAA was weighed, then 2 mL of CH3COOH and 15 mL of deionized water were added and stirred for 40 min to form a homogeneous solution. The solution was then transferred to a 50 mL teflon lining, with the addition of an appropriate amount of SnO2/CF, and held at 180 °C for 12 h to obtain SnS2/SnO2/CF. Finally, it was dried overnight in a vacuum oven at 70 °C. SnS2/CF was prepared by adding 0.25 g TAA, and the other steps were consistent with the above method.

2.5 Preparation of SnS/SnS2/SnO2/CF

SnS2/SnO2/CF was placed under H2/Ar atmosphere at a heating rate of 2.5 °C min−1 to 410 °C and held for 1 h to obtain SnS/SnS2/SnO2/CF. SnS/SnS2/CF − 0.25–410 was prepared by placing SnS2/CF in a tube furnace by the same procedure. SnS/SnS2/CF, SnS/CF and SnS/SnS2/CF − 0.25–450, SnS/CF − 0.25–500 were prepared from SnS2/SnO2/CF and SnS2/CF at 450 and 500 °C, respectively, and the other conditions were the same as above.

2.6 Preparation of Composite Coatings

In this work, Q235 mild steel was used as the substrate for the epoxy coating. Before use, the surface of Q235 steel is sandpapered smooth, then ultrasonically washed with ethanol, and finally dried. The SnS/SnS2/SnO2/CF composites were mixed with epoxy resin at 2 wt% loading, and n-butanol and xylene solvents were added and stirred evenly. Then, the polyimide curing agent was added and stirred for 0.5 h to make the sample and epoxy resin evenly mixed. Subsequently, the prepared coating solution was placed in a vacuum drying oven to remove bubbles. Finally, the composite coating solution was uniformly applied to the surface of the Q235 steel using a coating tool and placed in an oven to allow it to dry. Pure epoxy coatings were prepared without sample addition.

2.7 Characterization

The crystal structure of the samples was described by X-ray diffraction (XRD, Rigaku Ultima IV, Cu Ka radiation, (λ = 0.51418)). The morphology, elemental distribution, and microstructure of the materials were observed by field emission scanning electron microscopy (F-SEM, JEOL JSM-7800F) transmission electron microscopy (TEM, JEOL JEM-2100) and high magnification transmission electron microscopy (HRTEM). Thermogravimetric analysis (TGA) was performed on an SDT Q600 analyzer with an air ramp rate of 10 °C min−1 from room temperature to 900 °C. Raman spectra are collected at 532 nm using a Renishaw InviaPlus micro-Raman spectroscopy system with a 50 mW DPSS laser. The surface elemental distribution of the material was characterized by X-ray photoelectron spectroscopy (XPS) on a Thermo Fisher ESCALAB 250Xi spectrometer using a 1486.6 eV Al Ka X-ray source. The EM parameters of each sample were determined using a vector network analyzer (Agilent N5234A, USA) using the coaxial method at test frequencies from 2 to 18 GHz, and the EMW absorption characteristics of the samples were further calculated. The sample was loaded 50 wt% in the oxygen resin, and n-butanol and xylene were added to the epoxy resin in the ratio of 3:7 and stirred well. Then an amount of polyimide curing agent was added and stirred well. Put the mixture into the preparation mold, and after the subsequent curing process, make a ring with an inner diameter of 3.04 mm and an outer diameter of 7 mm. The reflection loss of the material can be calculated by the following equations [41,42,43,44]:

$$ \begin{array}{*{20}c} {{\text{Z}}_{{{\text{in}}}} = Z_{0} \sqrt {\frac{{\mu_{r} }}{{\varepsilon_{r} }}} {\text{tan}}h\left| {j\left( {\frac{2\pi fd}{c}} \right)\sqrt {\varepsilon_{r} \mu_{r} } } \right|} \\ \end{array} $$
(1)
$$ \begin{array}{*{20}c} {{\text{RL}}\left( {{\text{d}}B} \right) = 20{\text{log}}\left| {\frac{{Z_{{{\text{in}}}} - Z_{0} }}{{Z_{{{\text{in}}}} + Z_{0} }}} \right|} \\ \end{array} $$
(2)

where Zin and Z0 are the input impedance and free-space characteristic impedance of the standard wave-absorbing material, respectively, f is the frequency of the EMW, d is the thickness of the sample, and c is the velocity of the EMW in free space.

Through a typical three-electrode system, electrochemical corrosion experiments of samples and epoxy composite coatings in seawater solutions were measured on an electrochemical workstation CHI 760E. The Pt sheet was used as a counter electrode, Ag/AgCl as the reference electrode, and the coated electrode as the working electrode. The open circuit potential (OCP) behavior was recorded in the frequency range of 10–2 ~ 10–5 Hz and electrochemical impedance spectroscopy (EIS) measurements were performed. The seawater solution used was taken from the coast of Qingdao.

3 Results and Discussion

The preparation process of SnS/SnS2/SnO2/CF composites is illustrated in Fig. 1. Firstly, CF was obtained by electrostatic spinning and high-temperature calcination processes. Then, SnO2 nanosheets were grown on the surface of CF by hydrothermal method. In detail, SnCl2 dissolves to form Sn2+ and Cl ions, and trisodium citrate ionizes in solution to release citrate ions (C6H5O73−) and sodium ions (Na+). Then, C6H5O73− complexes with Sn2+ ions to form a stable tin-citrate complex. Under high temperature and pressure, Sn2+ is oxidized to Sn4+. Sn4+ is further hydrolyzed and condensed to ultimately form SnO2, obtaining SnO2/CF. Subsequently, SnS2/SnO2/CF was prepared by a hydrothermal method by controlling the amount of TAA added. Specifically, SnO2/CF reacts in a mixed solution of acetic acid and water, and TAA releases H2S at high temperatures. CH3COOH reacts with SnO2 to dissociate Sn4+ with Sn4+ reacting with H2S to form SnS2 to obtain SnS2/SnO2/CF. Finally, under hydrogen/argon conditions, H2 acts as a reducing agent to effectively reduce SnS2 to SnS. Since the reduction temperature was low, SnS2 was partially reduced to SnS, and the coexistence of SnS2 and SnS was realized, thus obtaining the target product SnS/SnS2/SnO2/CF with a lamellar structure.

Fig. 1
figure 1

Schematic diagram of the synthesis process of SnS/SnS2/SnO2/CF

In Fig. 2a, the XRD images of SnO2/CF, SnS2/SnO2/CF, and SnS/SnS2/SnO2/CF are shown. The wide diffraction peaks at around 20°-30° correspond to amorphous carbon peaks. Diffraction peaks of SnO2 at 26.61°, 33.89°, 37.95°, 42.63°, 51.78°, 61.87°, 64.71°, and 78.71° correspond to its (110), (101), (200), (201), (211), (310), (112), and (310) crystal planes (PDF#41–1445), respectively. The diffraction peaks located at 15.05°, 28.19°, 32.12°, 41.88°, 49.96°, and 52.45° correspond to the (001), (100), (101), (102), (110), and (111) crystal planes of SnS2 (PDF#23–0677), respectively. The successful synthesis of SnS2/SnO2/CF composites was confirmed. The diffraction peaks of the (110), (120), (021), (101), (111), (040), (131), (002), and (211) crystal planes of SnS were located at 22.01°, 26.01°, 27.47°, 30.47°, 31.53°, 31.97°, 39.04°, 45.49°, and 48.51° (PDF#39–0354), respectively. Consequently, the above information confirms the successful synthesis of the multi-component tin compound SnS/SnS2/SnO2/CF composite. Figure 2b shows the XRD patterns of SnS/SnS2/CF and SnS/CF composites obtained at different pyrolysis temperatures of SnS2/SnO2/CF. The generation of SnS/SnS2/CF is due to the thermal decomposition of SnS2 at elevated temperatures to produce H2S and SnS. Subsequently, H2S further reacted with SnO2 to form SnS2 to obtain SnS/SnS2/CF composites. Further increasing the temperature, the H2 reducing ability is enhanced, and the conversion rate of SnS2 to SnS is significantly increased. At this point SnS2 can be completely reduced to SnS, giving SnS/CF. In this work, we also investigated the complete sulfidation of SnO2 to obtain SnS2/CF composites. The high-temperature thermal decomposition and reduction processes were then carried out at 410, 450, and 500 °C to obtain SnS/SnS2/CF − 0.25–410, SnS/SnS2/CF − 0.25–450, and SnS/CF − 0.25–500 (Fig. S1a). SnO2 is fully sulfidized and then reduced by thermal decomposition. At 410 and 450 °C this leads to partial reduction of SnS2 and at 500 °C SnS2 is completely reduced to SnS. The thermogravimetric analysis images of SnO2/CF, SnS2/SnO2/CF, and SnS/SnS2/SnO2/CF are shown in Fig. 2c. The mass loss of SnO2/CF is mainly due to the loss of carbon under air atmosphere. The loss of mass of SnS2/SnO2/CF and SnS/SnS2/SnO2/CF is due to the oxidation of SnS and SnS2 and the carbon combustion [45,46,47,48]. In the Raman spectrum (Fig. 2d), the D band (1350 cm−1) and G band (1590 cm−1) reflect the ordered/disordered carbon arrangement of the composite material. As can be seen in Figs. 2d and S1b, the graphitization degree of the samples is relatively high, which is mainly due to the influence of the CF skeleton. The crisscrossed CF network can enhance the effective transfer of electrons and has good conductivity loss, thereby promoting the absorption of EMW. The molecular structure and chemical state of the surface of the samples were analyzed by XPS. As shown in Fig. 2e, the characteristic peaks of Sn, O, C, and S can be observed in the range of 0–1200 eV. Sn (MNN) and O (KLL) represent the existence of Sn and O spiral peaks [49, 50]. C 1s at 284.62, 286.2, and 289.2 eV with three peaks representing C–C/C = C, C–O, and C = O, respectively (Fig. 2f) [51]. The diffraction peaks at positions 525.29, 530.96, and 531.99 eV in Fig. 2g are absorbed oxygen, Sn–O, and lattice oxygen, respectively [52]. S 2p has two diffraction peaks at 162.7 eV (S 2p1/2) and 161.5 eV (S 2p3/2) (Fig. 2h) [53]. Where based on previous reports [54, 55], the peak of S 2p1/2 corresponding to sulfur at the low coordination site is associated with S vacancies. The diffraction peaks of Sn 3d are shown in Fig. 2i, and the diffraction peaks at 486.5 and 494.9 eV correspond to Sn 3d5/2 and Sn 3d3/2, respectively [56]. The binding energy of S 2p in SnS/SnS2/SnO2/CF composites was decreased relative to that of pure SnS2 (161.69 and 162.85 eV). In addition, the Sn 3d binding energy in the SnS/SnS2/SnO2/CF composites was red-shifted compared to the pure SnS2 nanosheets (486.88 and 495.37 eV). All these evidences suggest the presence of sulfur vacancies [57,58,59,60]. The XPS spectra of SnO2/CF (Fig. S2a–d) displayed the elemental compositions (Sn, O, and C) of SnO2/CF. Among them, the content of O is higher, which confirms the successful synthesis of SnO2/CF. Figure S2e-i shows the XPS image of SnS2/SnO2/CF, where the presence of the elements Sn, O, C, and S can be seen in the total spectrum. The XPS spectra of SnS/SnS2/CF and SnS/CF are shown in Fig. S3, where the presence of elemental O is a result of unavoidable oxidation in air. The three diffraction peaks in the C 1s energy spectrum are attributed to C–C/C = C, C–O, and C = O. The diffraction peaks in the S 2p energy spectrum correspond to S 2p1/2 and S 2p3/2, and the diffraction peaks in Sn 3d correspond to Sn 3d5/2 and Sn 3d3/2.

Fig. 2
figure 2

a, b XRD image of SnO2/CF, SnS2/SnO2/CF, SnS/SnS2/SnO2/CF, SnS/SnS2/CF, and SnS/CF. c Thermogravimetric images of SnO2/CF, SnS2/SnO2/CF, SnS/SnS2/SnO2/CF. d Raman image of SnO2/CF, SnS2/SnO2/CF, SnS/SnS2/SnO2/CF, SnS/SnS2/CF, and SnS/CF. e-i XPS spectra of SnS/SnS2/SnO2/CF sample

Microscopic morphology and structure of the samples were observed by SEM and TEM. Figure S4a shows PANF with smooth surface and uniform size. Following pre-oxidation and high-temperature carbonization the obtained CF surface had some folds (Fig. S4b). SnO2 was closely wrapped around CF, forming SnO2/CF with a diameter of about 1 μm (Fig. 3a, b). The EDS image of SnO2/CF (Fig. S4c) shows the uniform distribution of Sn, O, and C elements along the fiber axis. As shown in Fig. 3c, SnO2 with a thickness of 260 nm is wrapped on the CF surface, and the morphology of the nanosheets can be seen more clearly at the edge. In the HRTEM image of SnO2/CF (Fig. 3d), the measured lattice spacing is 0.334 nm, corresponding to the (110) crystal plane of SnO2. After the sulfidation of SnO2/CF, part of SnO2 was converted into larger SnS2 nanosheets to obtain SnS2/SnO2/CF (Fig. 3e, f). In the TEM image of SnS2/SnO2/CF (Fig. 3g), SnS2/SnO2 is located in the outer layer of the fiber with a thickness of about 240 nm. In the HRTEM image of SnS2/SnO2/CF (Fig. 3h), the measured lattice spacings are 0.264 and 0.295 nm, indexed to the (101) crystal plane of SnO2 and the (002) crystal plane of SnS2, respectively. The existence of heterogeneous interfaces, lattice distortions, and discontinuous lattice fringes in SnS2/SnO2/CF can also be observed. The thickness of SnS/SnS2/SnO2/CF nanosheets obtained by thermal decomposition and reduction at 410 °C decreased slightly, but the overall change was not significant (Fig. 3i, j). The nanosheet structure can promote the scattering of EMW and extend the propagation path, thereby promoting the loss of EMW. Compared with the product of the previous step, the thickness of the fiber outer layer and the morphology of the nanosheets in SnS/SnS2/SnO2/CF remained basically unchanged (Fig. 3k). In the HRTEM image (Fig. 3l), SnS can be seen precipitating at the edge of the material. The measured lattice stripe spacings of 0.334, 0.316, and 0.293 nm point to the (110) crystal plane of SnO2, the (100) crystal plane of SnS2, and the (101) crystal plane of SnS, respectively. All information confirms the successful synthesis of SnS/SnS2/SnO2/CF. In SnS/SnS2/SnO2/CF materials, multiple materials contact to form heterogeneous interfaces with unique electronic properties and bonding states [61]. At the heterogeneous interfaces, non-homogeneous interfaces greatly impede the flow of charges [62]. The difference in dielectric properties leads to uneven charge distribution, resulting in an electric dipole moment, which enhances interfacial polarization and losses EMW. The amorphous structure results in the generation of discontinuous lattice streaks. Lattice distortion and discontinuous lattice fringes lead to the destruction of lattice symmetry in SnS/SnS2/SnO2/CF, resulting in lattice mismatches, and changes in electron distribution and charge transfer pathways. Furthermore, the presence of sulfur vacancies can also be inferred from the broken lattice at the marked circles. Sulfur vacancies can introduce defect energy levels that act as capture centers for electrons and holes, thereby changing the electronic and energy band structure of the material [63,64,65]. Thereby promoting interfacial polarization and EMW loss. The selected area electron diffraction (SAED) of SnO2/CF, SnS2/SnO2/CF, and SnS/SnS2/SnO2/CF are shown in Fig. S5. Figure 3m-p shows the SnS/SnS2/CF and SnS/CF composites obtained by thermal decomposition of SnS2/SnO2/CF at 450 and 500 °C. As the temperature increases, the nanosheets on the fiber surface begin to be destroyed. At 450 °C, the nanosheets on the surface of SnS/SnS2/CF material gradually fall off, but the basic nanosheet structure can be maintained. After thermal decomposition at 500 °C, the nanosheet structure on the fiber surface of SnS/CF material is destroyed. The nanosheets melt at high temperatures and form small stone-like structures. After SnO2 was thoroughly sulfurized, large SnS2 nanosheets tightly wrapped CF to form SnS2/CF composites (Fig. S6a). SnS2/CF contains Sn, S, and C elements distributed along the fiber (Fig. S6e, f). Subsequently, high-temperature thermal decomposition reduction was carried out at 410, 450, and 500 °C, to obtain SnS/SnS2/CF − 0.25–410 (Fig. S6b), SnS/SnS2/CF − 0.25–450 (Fig. S6c), and SnS/CF − 0.25–500 (Fig. S6d), respectively. The edges of the nanosheets begin to melt by heat at 410 °C. When the temperature rises to 450 °C the nanosheets wrapped around the fibers begin to fall off, and as the temperature continues to rise to 500 °C, the nanosheet structure of the fibers completely collapses. This demonstrated that the SnO2 component also works as an important role in structural stabilization.

Fig. 3
figure 3

a, b SEM, c TEM, d HRTEM image of SnO2/CF. e, f SEM, g TEM, h HRTEM image of SnS2/SnO2/CF. i, j SEM, k TEM, l HRTEM image of SnS/SnS2/SnO2/CF. m, n SEM images of SnS/SnS2/CF. o, p SEM image of SnS/CF. q EDS mapping image of SnS/SnS2/SnO2/CF

Based on the EMW theory, the absorption performance of a wave-absorbing material is mainly intimately associated with its relative complex permittivity (εr = ε'-jε") and magnetic permeability (μr = μ′-jμ") [66,67,68]. The real part of the dielectric constant (ε′) reflects the energy storage capacity of the material, while the imaginary part (ε″) reveals its loss capacity [69, 70]. In this study, the magnetic storage and loss capacity of all samples can be neglected due to the lack of magnetic components. Therefore, the real (µ') and imaginary (µ″) parts of permeability are taken as 1 and 0, respectively [71, 72]. For the fiber composites prepared in this work, the complex permittivity exhibits a frequency dependence [73]. As shown in Fig. 4a, the ε′ values of all samples are in the range of 8.81 ~ 14.81 and show a downward trend in the test frequency range. In Fig. 4b, it is shown that the ε" values of all the samples show an increasing trend in the range of 2–18 GHz. Among them, the ε′ and ε″ values of SnS/CF are the largest. This is attributed to the unique layered structure, large interlayer spacing, high reversible capacity, and excellent conductivity of SnS. Especially note that the sample's dielectric curves with multiple polarization relaxation peaks can be observed in the ε' and ε'' plots This indicates the presence of a polarization relaxation process loss EMW. The dielectric loss tangent (tanδε) values of SnS/SnS2/CF and SnS/CF are larger, showing that they have stronger dielectric loss capability (Fig. 4c). The relevant electromagnetic parameters of SnS2/CF, SnS/SnS2/CF − 0.25–410, SnS/SnS2/CF − 0.25–450, and SnS/CF − 0.25–500 are shown in Fig. S7. Where SnS2/CF has the lowest dielectric parameter, indicating the worst dielectric loss capability. The dielectric parameters of the other samples do not differ much. The contribution of dielectric loss is mainly composed of conduction loss and polarization loss, while the polarization loss is mainly contributed by interface polarization and dipole polarization. The dielectric property related to the conduction loss can be calculated by Eq. (3) [74, 75]:

$$ \begin{array}{*{20}c} {\sigma = \omega \varepsilon_{0} \varepsilon^{\prime\prime}} \\ \end{array} $$
(3)

where ω (ω = 2πf) is the angular frequency and ε0 (ε0 = 8.854 × 10–12 F m−1) is the free-space dielectric constant. The conductivity of SnS2/SnO2/CF is relatively small, and other samples have multiple peaks at high frequencies (Fig. 4d). This is due to the influence of polarization relaxation. The polarization relaxation process is further investigated by Deby’s theory and combined with Eqs. (4) and (5) to study the Cole–Cole semicircle [76, 77]:

$$ \begin{array}{*{20}c} {\varepsilon^{\prime} = \varepsilon_{\infty } + \frac{{\varepsilon_{s} - \varepsilon_{\infty } }}{{1 + \left( {2\pi f} \right)^{2} r^{2} }}} \\ \end{array} $$
(4)
$$ \begin{array}{*{20}c} {\varepsilon^{\prime\prime} = \frac{{\omega \tau \left( {\varepsilon_{s} - \varepsilon_{\infty } } \right)}}{{1 + \left( {2\pi f} \right)^{2} r^{2} }}} \\ \end{array} $$
(5)

where εs is the static dielectric constant and ε is the relative dielectric constant at infinite frequency. The long smooth straight line on the ε′-ε″ image curve is related to the conduction loss (tail solid line), while the semicircle represents the polarization relaxation process (dotted circle) [78]. SnS/SnS2/SnO2/CF has more semicircles compared to the other materials (Fig. 4a1-a5), which indicates that it has more polarization relaxation processes. This is attributed to the fact that it has more component materials and more heterogeneous interfaces. Due to the differences in work functions of different components in the composite system, carriers tend to migrate from the phase with higher work function to the phase with lower surface work function. The surface electron affinity of SnO2 oxide is high, and the electron escapes from the surface requires high energy, and its work function is high. SnS2 increases the negative charge density because of the presence of sulfur atoms on the surface, resulting in a higher work function. The electronic structure of SnS is closer to the metallic state, so the work function is lower [79]. Carrier migration between different components is accompanied by EM energy consumption. The electron transfer at the interface between the two materials leads to the formation of a built-in electric field and a space charge region, which affects charge transport and interface polarization [80]. Opposite charges accumulate at the heterogeneous interface, forming a built-in electric field that promotes electron transport and dissipates EMW through interfacial polarization effects (Fig. 4e). Meanwhile, the presence of S vacancies in SnS/SnS2/SnO2/CF can act as the active site of the dipole. The center of symmetry is deflected in the presence of an applied electromagnetic field, producing a strong dipole polarization. We further study the polarization relaxation process by Eq. (6) [81,82,83]:

$$ \begin{array}{*{20}c} {\varepsilon^{\prime} = \frac{1}{2\pi \tau }\frac{{\varepsilon^{\prime\prime}}}{f} + \varepsilon_{\infty } } \\ \end{array} $$
(6)
Fig. 4
figure 4

a ɛʹ, b ɛ″, c tanδɛ, d dielectric conductivity, a1-a5 Cole–Cole, b1-b5 ɛʹ versus ɛ″/f of SnO2/CF, SnS2/SnO2/CF, SnS/SnS2/SnO2/CF, SnS/SnS2/CF, and SnS/CF. e Interface polarization diagram of SnS/SnS2/SnO2/CF

Followed by the relationship between ε′ and ε″/f can be concluded the time τ of polarization relaxation. As shown in Fig. 4b1-b5, the relaxation times of SnO2/CF, SnS2/SnO2/CF, SnS/SnS2/SnO2/CF, SnS/SnS2/CF, SnS/CF were 0.10, 0.12, (0.014 and 0.082), 0.10, (0.02 and 0.09), respectively. It is worth noting that different materials have different polarization relaxation times. Among them, SnS/SnS2/SnO2/CF and SnS/CF composites have multiple polarization relaxation times, which indicates the existence of multiple polarization relaxation processes [84]. This is attributed to the fact that each polarization relaxation process possesses a unique relaxation time and exhibits multiple EMW loss characteristics. Defects and other heteroatoms in the material act as polarization sites capable of inducing dipole polarization phenomena in the presence of an applied magnetic field. Simultaneously, the interfacial polarization is induced due to the interaction of SnS, SnS2, SnO2, and CF with each other at the interface, which further enhances the EMW absorption ability of the material. The attenuation constant α is used to characterize the attenuation properties of the material to EMW and is calculated using Eq. (7) [85,86,87]:

$$ \begin{array}{*{20}c} {\alpha = \frac{\sqrt 2 \pi f}{c} \times \sqrt {\left( {\mu^{\prime\prime}\varepsilon^{\prime\prime} - \mu^{\prime}\varepsilon{\prime} } \right) + \sqrt {\left( {\mu^{\prime}\varepsilon^{\prime\prime} + \mu^{\prime\prime}\varepsilon^{\prime}} \right)^{2} + \left( {\mu^{\prime\prime}\varepsilon^{\prime\prime} - \mu^{\prime}\varepsilon^{\prime}} \right)^{2} } } } \\ \end{array} $$
(7)

As shown in Fig. S8a, the change in the α value of the material is consistent with the changing trend of the dielectric constant, which proves that dielectric loss is the main mechanism of EMW attenuation. SnS/SnS2/CF and SnS/CF have large α values, indicating that they have strong EMW attenuation ability. However, EMW absorption performance is determined by several factors, especially impedance matching. The |Zin/Z0| values of SnS/SnS2/SnO2/CF and SnS/SnS2/CF are closest to 1, which may have excellent impedance matching performance (Fig. 8b).

As shown in Figs. 5a1-a5 and S9, S10, SnS/CF has the best EMW absorption performance (RLmin = 37.44 dB, EABmax = 4.80 GHz) in the binary system material. Among the ternary system materials, SnS/SnS2/CF has the most excellent EMW absorption performance, with an EABmax of 5.20 GHz, while SnS2/SnO2/CF has an EABmax of only 4.32 GHz. This is because SnS2 generates S vacancies during its reduction to SnS, which promotes the generation of polarization relaxation. Nevertheless, the EMW absorption performance of SnS2/SnO2/CF is still better than that of SnS/CF. This suggests that the EMW absorption performance can be enhanced by constructing multiple heterointerfaces by increasing the components. The SnS/SnS2/SnO2/CF composites are obtained by further increasing the components. The EMW absorption performance is further enhanced due to the increased heterogeneous interface and S vacancies effect, with an RLmin of − 46.74 dB at a matched thickness of 2.0 mm and an EABmax of 5.28 GHz at 1.7 mm. The results show that the EMW attenuation performance can be effectively improved by adjusting the multiple loss mechanism in the composite system to construct multiple heterogeneous interfaces. The EMW absorption performance of a material has an important relationship with impedance matching, which indicates the ability of EMW to enter the material without being reflected [88, 89]. Zin/Z0 in the region of 0.8–1.2 is the ideal impedance matching performance. Figure 5b1-b5 shows the area occupied by SnO2/CF, SnS2/SnO2/CF, SnS/SnS2/SnO2/CF, SnS/SnS2/CF, and SnS/CF impedance matches in the range of 0.8–1.2, are 5.87%, 5.71%, 8.15%, 7.22%, and 6.36%, respectively. Combined with Fig. 5d3 it can be seen that SnS/SnS2/SnO2/CF has the largest area, which indicates the optimum impedance matching. Its superior impedance matching performance comes from its own characteristics and non-uniform interface structure. The EAB at various thicknesses are shown in Fig. 5c1-c5, and it can be seen that at thin matching thicknesses, the EAB of the material is mainly concentrated in the X and Ku bands. The relationship between the matching thickness of the material and the reflection loss and frequency is studied by the quarter-wavelength theory, the equation is as follows [90,91,92,93]:

$$ \begin{array}{*{20}c} {t_{m} = \frac{n\lambda }{4} = \frac{nc}{{4f_{m} \sqrt {\left| {\mu_{r} } \right|\left| {\varepsilon_{r} } \right|} }}\left( {n = 1,3,5 \ldots } \right)} \\ \end{array} $$
(8)

where λ is the wavelength of EMW, c is the propagation speed of EMW in vacuum, and |μr| and |εr| are the moduli of μr and εr, respectively. The matched thickness of RL moves in a decreasing direction with increasing frequency and falls exactly on the λ/4 curve (Fig. S11). This suggests that RL corresponds to a matching thickness that corresponds perfectly to the quarter-wavelength theory. The SnS/SnS2/SnO2/CF composite exhibits the smallest RL and the largest EAB at the thin thickness (Fig. 5d1, d2), which further confirms its excellent EMW absorption properties. The SnS/SnS2/SnO2/CF sample has the strongest RL, covering part of the X-band and Ku-band (Fig. 5d4). By comparing the relationships among RL, EAB, and components of all samples in this work, it was found that SnS/SnS2/SnO2/CF samples with the most components had the best EMW absorption performance (Fig. 5d5). Comparing the SnS/SnS2/SnO2/CF material prepared in this study with other CF-based and Sn-based materials, it can be seen that it has a small RL and a large EAB. It has significant advantages in EMW absorption (Fig. S12, Table S1). Therefore, it is an effective strategy to improve the absorption performance of EMW by rationally adjusting the microstructure and chemical components.

Fig. 5
figure 5

EMW absorption performance of SnO2/CF, SnS2/SnO2/CF, SnS/SnS2/SnO2/CF, SnS/SnS2/CF, SnS/CF. a1-a5 3D RL, b1-b5 2D impedance matching, c1-c5 impedance matching at different thicknesses, d1 3D RL and d2 EAB at 1.6 to 2.5 mm, d3 3D impedance matching and d4 2D RL at matching thickness. d5 Radar map of RL, EAB, and component relationships for all samples

The radar cross section (RCS) of the composites was simulated by CST Studio Suite software at 12.16 GHz to explore the application prospect of the materials in the real environment [94,95,96,97]. The 3D radar wave scattering signals of the perfect electric conductor (PEC) and all samples are shown in Figs. 6a and S13a. It can be clearly seen that the radar scattering signal intensity of the composite material is reduced compared with that of PEC. It is particularly noteworthy that the radar scattering signal intensity of the SnS/SnS2/SnO2/CF sample is the lowest, indicating that it has a strong EMW attenuation ability. Figures 6b, c and S13b, c show the RCS values of pure PEC and all samples in different angle ranges. In the range of − 90° ~ 90°, the RCS values of all the samples were reduced to different degrees relative to the pure PEC. The RCS results show that the SnS/SnS2/SnO2/CF sample have good attenuation ability for EMW and can effectively suppress the radar scattering signals. The EMW attenuation capability of the prepared samples was further investigated by calculating the RCS reduction values at different detection angles (0°, 30°, 60°, 90°) (Fig. 6d). The maximum value of reduction in RCS value for SnS/SnS2/SnO2/CF sample is 26.5 dB m2 at 0°. This further confirms that the SnS/SnS2/SnO2/CF sample has strong EMW attenuation capability and has great potential in real-world situations such as radar stealth [98,99,100].

Fig. 6
figure 6

a1–a6 3D radar wave scattering signals, b1–b5 RCS simulation curves, and c1–c5 RCS simulation results of polar coordinates of PEC, SnO2/CF, SnS2/SnO2/CF, SnS/SnS2/SnO2/CF, SnS/SnS2/CF, SnS/CF. d RCS reduction value of the all samples

Figure 7 shows the mechanism of the EMW absorption of SnS/SnS2/SnO2/CF composites. Firstly, the nanosheet structure can effectively scatter EMW, which effectively prolongs the propagation path of EMW and promotes the absorption of EMW. Secondly, the CF skeletons are composite with tin compounds with electrical conductivity to form an efficient conduction network. As EMW propagates through the conduction network, electrons in the material are stimulated, prompting them to migrate and jump, which in turn creates an electric current. This process dissipates EMW energy in the form of heat, which significantly improves the conduction loss capability of the material [101]. The introduction of tin compounds changes the dielectric parameters of the sample and optimizes the impedance matching. Furthermore, the presence of multiple non-uniform interfaces promotes the accumulation of positive and negative charges at the interfaces, creating a model that resembles an equivalent capacitor structure. The charges at the interface cannot be fully synchronized to respond to the alternating frequency of the EMW under the action of the electromagnetic field. This results in a polarization relaxation process, which in turn enhances the dielectric loss capability of the material to EMW. The material also has vacancies and lattice defects capable of producing dipoles. Under the induction of an electromagnetic field, the dipole moments are rearranged with the direction of the electric field, and the sum of the dipole moment vectors is not zero, giving rise to the dipole polarization effect [102,103,104,105]. In summary, SnS/SnS2/SnO2/CF composites absorb EMW through the synergistic action of multiple loss mechanisms.

Fig. 7
figure 7

Electromagnetic wave absorption mechanism

For applications in complex environments, especially in the marine environment, this places higher demands on EMW absorbers. The coating’s corrosion protection is assessed by a chemical measurement technique using a three-electrode system. Firstly, the variation of open circuit potential (OCP) of the material with immersion time was investigated. Generally, higher OCP values have a lower tendency to corrode [106,107,108,109,110]. The composite coating with SnS/SnS2/SnO2/CF samples has the highest OCP value, and the bare steel coating has the lowest OCP value (Fig. 8a1-c1). The results show that the corrosion resistance of SnS/SnS2/SnO2/CF composite coating is the best, followed by pure epoxy coating, and bare steel coating is the worst. With the extension of immersion time, after two weeks of immersion, the corrosive medium gradually penetrated the coating through the coating defects. This results in a reduction in the corrosion resistance of the coating, along with a corresponding reduction in the OCP value. Figure 8a2-c2 shows the electrochemical impedance spectroscopy (EIS) of SnS/SnS2/SnO2/CF and epoxy resin composite coating, pure epoxy resin, and bare steel, respectively. From the plots, it can be shown that the capacitive arc radii of all three coatings exhibit larger values on the first day. Noteworthy, the SnS/SnS2/SnO2/CF composite coating reached a capacitive arc maximum of 1.40 × 105, which proves their strong corrosion resistance. The bare steel has the smallest capacitive arc radius, indicating the worst corrosion resistance. With time the radius of the capacitive arc decreases, which shows a gradual decrease in the corrosion protection of the coating. However, the coating containing SnS/SnS2/SnO2/CF always maintained the largest capacitive arc radius throughout the process, which fully demonstrated its excellent corrosion resistance. The radius of the capacitor arc decreases sharply from day 5 and changes little in the following days. This indicates that the coating is beginning to be damaged and that the corrosion of the coating becomes more severe with increasing immersion time. This is mainly due to the penetration of oxygen, water, and corrosive ions (Cl) into the coating, which ultimately results in the corrosion process occurring at the junction of the coating and the Q235 steel substrate. Normally, the impedance modulus at the lowest frequency of the coating (|Z|0.01 Hz) is used as a semi-quantitative indicator to assess the corrosion resistance of the coating. The impedance modulus in the Bode plot reflected the ability of the coating to suppress the current between the cathode and anode. The larger the impedance modulus of the coating, the more effective the corrosion protection [111, 112]. The |Z|0.01Hz values of SnS/SnS2/SnO2/CF composite coating, pure epoxy, and bare steel are presented in Fig. 8a3-c3. On the first day, their impedance modulus at low frequencies was 1 × 105.65, 1 × 105.08, and 1 × 102.74 Ω cm2, respectively. Among them, the impedance modulus of the SnS/SnS2/SnO2/CF composite coating was the largest, which indicated that they had the strongest ability to block the penetration of corrosive substances. The impedance modulus decreases with increasing immersion time. After soaking for 14 days, the impedance modulus of the three coatings decreased to 1 × 103.19, 1 × 103.61, and 1 × 102.29 Ω cm2, respectively. This phenomenon indicates that the corrosion resistance of the coating decreases. The change of coating properties was studied by studying the phase Angle of each coating, and the phase Angle of the coating decreased with the increase of soaking time (Fig. 8a4-c4). There is also a gradual shift of the time constant toward the low-frequency region, both of which indicate that corrosion of the coating has occurred [113,114,115,116].

Fig. 8
figure 8

a1–c1 OCP values, a2–c2 Nyquist plots, a3–c3 Bode plots, a4–c4 Bode plots of SnS/SnS2/SnO2/CF composite coating, pure epoxy coating, and Q235 steel

The corresponding equivalent circuit model when the corrosion medium before reaching the metal substrate is shown in Fig. 9a. On the contrary, when the corrosive medium reaches the metal matrix, it reacts with the metal and charge transfer occurs, resulting in corrosion of the Q235 steel substrate (Fig. 9b). Figure S14 shows SEM images of SnS/SnS2/SnO2/CF composite coating and pure epoxy coating after 14 days of immersion. Following prolonged immersion, the SnS/SnS2/SnO2/CF composite coating was able to maintain intact coating morphology with and without significant damage. The pure epoxy resin coating has been damaged, and a large number of bubbles are produced on the surface, indicating that the coating has been severely corroded. The above information indicates that the addition of SnS/SnS2/SnO2/CF composites to epoxy resin is beneficial to improve the corrosion resistance of the coating. Figure 9c shows the anti-corrosion mechanism of SnS/SnS2/SnO2/CF composite coating. Adding SnS/SnS2/SnO2/CF to epoxy resin coating can improve the density of the coating. It can effectively prevent water molecules, oxygen, Cl and other corrosive media from passing through the coating. The addition of fiber materials can improve the mechanical strength of the coating, thus increasing the coating’s service life and protective properties. Carbon fiber composites also have a certain degree of electrical conductivity, achieving cathodic protection and reducing the corrosion rate of the substrate. Thereby achieving anti-corrosion properties [117,118,119,120,121].

Fig. 9
figure 9

a, b Equivalent circuit models of coatings at different stages. c Corrosion prevention mechanism diagram of SnS/SnS2/SnO2/CF composite coating

4 Conclusions

In summary, SnS/SnS2/SnO2/CF one-dimensional (1D) composites with multi-heterogeneous interfaces were ingeniously constructed by component modulation engineering. The 1D structure establishes the conduction network and promotes the conduction loss. The multi-component materials form multiple heterogeneous interfaces to enhance polarization relaxation and effectively regulate impedance matching, thereby achieving excellent EMW absorption performance. The SnS/SnS2/SnO2/CF composite material achieves strong absorption with an RLmin of − 46.74 dB at a thickness of 2.0 mm. The EABmax at 1.7 mm is 5.28 GHz, realizing a wide bandwidth. At the same time, the coatings with added SnS/SnS2/SnO2/CF composites were shown by electrochemical tests to have excellent corrosion resistance, which hindered the corrosion reaction by delaying the penetration of corrosive media. In this work, high-efficiency EMW absorption and good corrosion protection were achieved by modulating the material components. This provides a new direction for the development of EMW absorbers with excellent EMW absorption properties and corrosion resistance.