Next Article in Journal
Design, Synthesis and Evaluation of Praziquantel Analogues and New Molecular Hybrids as Potential Antimalarial and Anti-Schistosomal Agents
Previous Article in Journal
The Oxygen Evolution Reaction at MoS2 Edge Sites: The Role of a Solvent Environment in DFT-Based Molecular Simulations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multifunctional Hollow Porous Fe3O4@N-C Nanocomposites as Anodes of Lithium-Ion Battery, Adsorbents and Surface-Enhanced Raman Scattering Substrates

1
College of Chemistry and Chemical Engineering, Anhui University, Hefei 230601, China
2
Department of Chemical Engineering, Hefei Normal University, Hefei 230601, China
3
School of Biomedical Engineering, Anhui Medical University, Hefei 230032, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(13), 5183; https://doi.org/10.3390/molecules28135183
Submission received: 12 May 2023 / Revised: 19 June 2023 / Accepted: 23 June 2023 / Published: 3 July 2023

Abstract

:
At present, it is still a challenge to prepare multifunctional composite nanomaterials with simple composition and favorable structure. Here, multifunctional Fe3O4@nitrogen-doped carbon (N-C) nanocomposites with hollow porous core-shell structure and significant electrochemical, adsorption and sensing performances were successfully synthesized through the hydrothermal method, polymer coating, then thermal annealing process in nitrogen (N2) and lastly etching in hydrochloric acid (HCl). The morphologies and properties of the as-obtained Fe3O4@N-C nanocomposites were markedly affected by the etching time of HCl. When the Fe3O4@N-C nanocomposites after etching for 30 min (Fe3O4@N-C-3) were applied as the anodes for lithium-ion batteries (LIBs), the invertible capacity could reach 1772 mA h g−1 after 100 cycles at the current density of 0.2 A g−1, which is much better than that of Fe3O4@N-C nanocomposites etched, respectively, for 15 min and 45 min (948 mA h g−1 and 1127 mA h g−1). Additionally, the hollow porous Fe3O4@N-C-3 nanocomposites also exhibited superior rate capacity (950 mA h g−1 at 0.6 A g−1). The excellent electrochemical properties of Fe3O4@N-C nanocomposites are attributed to their distinctive hollow porous core-shell structure and appropriate N-doped carbon coating, which could provide high-efficiency transmission channels for ions/electrons, improve the structural stability and accommodate the volume variation in the repeated Li insertion/extraction procedure. In addition, the Fe3O4@N-C nanocomposites etched by HCl for different lengths of time, especially Fe3O4@N-C-3 nanocomposites, also show good performance as adsorbents for the removal of the organic dye (methyl orange, MO) and surface-enhanced Raman scattering (SERS) substrates for the determination of a pesticide (thiram). This work provides reference for the design and preparation of multifunctional materials with peculiar pore structure and uncomplicated composition.

1. Introduction

Transition metal oxides (TMOs) and their related nanocomposites are widely used in the anodes of lithium-ion batteries because of their high theoretical specific capacity and excellent conductivity [1]. In recent years, their applications in other aspects have gradually attracted people’s attention, such as surface-enhanced Raman scattering [2] (SERS) and removal of pollutants [3].
Lithium-ion batteries (LIBs) are widely used in our society [4,5]. However, the commercial anode material is graphite, which cannot satisfy the ever-growing requirements of social development due to its low theoretical capacity (372 mA h g−1). Therefore, significant research attention has been focused on exploring the TMOs that have a higher capacity and long cycle life, such as CuO [6], Fe2O3 [7], Co3O4 [8], TiO2 [9] and MnO2 [10], etc. Among numerous alternatives, Fe3O4 has been identified as the most promising anode material for LIBs owing to its high theoretical capacity (926 mA h g−1), low cost, non-toxicity and abundant natural reserves [11,12,13,14]. However, its enormous volume expansion, as well as dramatic electrode pulverization, could cause serious capacity loss along with poor cycling behavior, and severely limit the practical use of Fe3O4 in LIBs; many strategies have been proposed to address these issues. Compositing with conductive matrices, such as carbon, has been proven to be a useful method to enhance electrochemical properties [15,16]. Such carbon coatings can improve the electron conductivity and relieve the volume variation during the lithiation/delithiation process. For instance, Lin et al. prepared a Fe3O4/carbon nanotube composite and confirmed a discharge capacity of 930 mA h g−1 after 100 cycles at 0.1 A g−1 [17]. In addition, some studies have demonstrated that N-doping could also increase the electronic conductivity and lithium storage properties of carbon-based materials and thus enhance their electrochemical properties [18,19]. Preparing hollow/porous and core-shell nanostructures is another efficient strategy, which could provide rapid electron channels and void spaces to accommodate the huge volume expansion of the active materials, thus improving the cycling performance of the composites. For example, He et al. synthesized porous graphene-doped carbon/Fe3O4 nanofibers, displaying a discharge capacity of 872 mA h g−1 at 100 mA g−1 after 100 cycles [20].
Protecting the human environment and reducing water pollution have always been a research focus of scientists. The direct discharge of dye pollutants such as methyl orange, methyl red and rhodamine, etc., in industry will pollute water sources and threaten human health. Therefore, the effective removal of dye pollutants in water is very important. Adsorption is commonly used in industrial wastewater purification methods. Additionally, porous Fe3O4 nanocomposite materials used as adsorbents could supply a larger surface area for adsorption, and be easily recycled using the magnetic properties of Fe3O4 [21].
Recently, SERS has attracted more attention because it is an effective detection method with fast analysis speed and favorable specificity, and therefore could be used in chemical, biological and environmental analysis. Compared with precious metal substrates with worse stability, exorbitant price and poor reproducibility, TMO substrates with low price and better chemical stability have gradually become study hotspots, such as Fe2O3 [22], Cu2O [23] and TiO2 [24]. In addition, TMO substrates with rough surfaces and porous structures can also provide more active sites for the detected molecules, thus improving the Raman response substrates [23].
Herein, combining the above analyses, we developed a self-template mechanism for the fabrication of hollow porous Fe3O4@N-C core-shell nanocomposites (Scheme 1). Firstly, Fe3O4 nanospheres were prepared with the hydrothermal method, using FeCl3·6H2O as the raw material, Na3C6H5O7·2H2O as the stabilizer, CO(NH2)2 as the alkali source and (C3H5NO)n as the polymer surfactant. Then, Fe3O4 nanospheres were coated with dopamine hydrochloride to form a core–shell structure, followed by calcination in nitrogen and subsequent HCl etching. Thus, the hollow porous Fe3O4@N-C nanocomposites were obtained. The influences of different etching times (15 min, 30 min and 45 min) on the morphologies and performances of the Fe3O4@N-C nanocomposites (expressed as Fe3O4@N-C-2, Fe3O4@N-C-3 and Fe3O4@N-C-4, respectively) were studied. The Fe3O4@N-C-3 nanocomposites exhibited excellent specific capacity, high-rate capacity and cycling performance as anode materials for LIBs. Meanwhile, we investigated the performances of the prepared nanocomposites in SERS detection and organic dye pollutant adsorption, obtaining the desired results. Therefore, the prepared Fe3O4@N-C nanocomposites have potential application value in energy storage, rapid SERS detection and pollutant treatment.

2. Results and Discussion

Figure 1a exhibits the XRD patterns of the Fe3O4@N-C nanocomposites with different etching times by HCl solution. All the diffraction peaks for the four samples at 2θ of 18.22°, 30.36°, 35.77°, 37.05°, 43.35°, 53.76°, 57.16° and 62.55° could be indexed to the (111), (220), (311), (222), (400), (422), (511) and (440) crystal planes of Fe3O4 (JCPDS NO. 19-0629), respectively. No additional diffraction peaks from possible impurities, such as Fe2O3, could be detected, indicating the successful synthesis and the high purity of the Fe3O4 crystals. The sample contained less carbon; hence, the carbon peak was difficult to observe in the XRD patterns. In order to further investigate the formation of carbon on the surface of Fe3O4 and explore the change of the relative content of carbon, Raman spectra (Figure 1b) of different samples were measured. The four samples display two distinct peaks at 1352 and 1588 cm−1, which individually match to the D and G bands. The low-frequency D band could be attributed to structural flaws, while the G band was related to the vibration of sp2 carbon domains in both rings and chains [25]. The intensity ratios of the D band to the G band (ID/IG) of Fe3O4@N-C-1, Fe3O4@N-C-2, Fe3O4@N-C-3 and Fe3O4@N-C-4 were 0.85, 0.79, 0.78 and 0.76, respectively. With the increase of acid etching time, samples showed relatively high orderliness and graphitizing grade.
To elucidate the morphologies and microstructures of the synthesized samples, SEM and TEM measurements were made. The SEM images (Figure 2a,c,e) show that the three samples (Fe3O4@N-C-2, Fe3O4@N-C-3 and Fe3O4@N-C-4) possess a similar spherical structure with a rough surface, and their average diameters (AD) of about 313 nm, 300 nm and 276 nm can be observed from the particle size distributions shown in the insets of Figure 2a,c,e, respectively. The sizes and morphologies of the three samples in the TEM images (Figure 2b,d,f) are consistent with those in the SEM images. The results demonstrate that with the increase of HCl etching time, the average diameter of samples gradually decreases (Figure 2b,d), and that some spherical structures of the Fe3O4@N-C-4 was even destroyed, resulting in the formation of smaller particles (Figure 2e). From the TEM images, it can be found that Fe3O4@N-C-3 nanocomposites have spherical hollow porous core-shell structures. Furthermore, the enlarged TEM image of Fe3O4@N-C-3 (Figure 2g) reveals that a N-doped carbon layer slightly increases the diameter of the sphere with a thickness of around 5 nm. Additionally, a lattice fringe with an interplanar distance of 0.25 nm is visible in the corresponding high-resolution TEM image of Fe3O4@N-C-3 (Figure 2h), which could be ascribed to the (311) plane of Fe3O4. The carbon is marked by an arrow, implying that the product consists of Fe3O4 and carbon. The hollow porous nanospheres structure can offer plentiful mesoporous channels for solid–liquid contact.
To further prove the existence of different functional groups and the change of the peak intensity, the FT-IR spectra of the four samples were taken at a range of 400–4000 cm−1. As demonstrated in Figure 3a, the peaks at 3431 cm−1 belong to the stretching vibration of –OH while the bands appearing at 2970 and 1447 cm−1 belong to the C-H groups and C-N groups, respectively. The absorptions at 1049 and 877 cm−1 are ascribed to the C-H in-plane and out-of-plane bending vibrations of 1, 4-substituted benzene. The peak at 586 cm−1 is assigned to the Fe-O group in Fe3O4 [20,26]. The FT-IR measurements demonstrate the prepared Fe3O4@N-C nanocomposites contain Fe3O4 and nitrogen-doped carbon. In addition, TG analysis was implemented to estimate the carbon content in the four samples, and the results are exhibited in Figure 3b. Obviously, the four samples started to decompose at 300 °C because of the oxidation of Fe3O4 to Fe2O3. Weight loss was evident in the temperature range of 300–450 °C, which may be ascribed to carbon oxidation. Based on the TG curves, the carbon contents calculated in Fe3O4@N-C-1, Fe3O4@N-C-2, Fe3O4@N-C-3 and Fe3O4@N-C-4 were about 11.25%, 11.67%, 16.25% and 20.84%, respectively, which shows a great influence on the electrochemical performance of Fe3O4@N-C nanocomposites.
In order to determine the surface elemental composition and valence states of the Fe3O4@N-C-3 nanocomposites, XPS characterizations were further implemented. In Figure 4a, XPS analysis of Fe3O4@N-C-3 indicates that the Fe, O, C and N atoms are presented in the nanocomposites with atomic ratios of 6.05%, 17.83%, 68.75% and 7.37%, respectively. Additionally, three peaks at 285.5 eV, 400 eV, 533.5 eV and 700 eV belong to C 1s, N 1s, O 1s and Fe 2p, respectively. The high-resolution core level XPS spectra of Fe 2p [27] are exhibited in Figure 4b. There are double peaks corresponding to Fe 2p 3/2 and Fe 2p 1/2 energy level located at 710.98 eV and 724.03 eV, respectively, demonstrating the existence of Fe3O4 rather than Fe2O3 [26]. The high-resolution C 1s XPS spectrum of the Fe3O4@N-C-3 nanocomposites (Figure 4c) displays an asymmetric broad peak that corresponds to the C-C (284.68 eV), C-N (285.98 eV) and C-O (286.63 eV) functional groups on the surface, respectively, which indicates the presence of multiple chemical states of carbon and the different types of functional groups on the surface. These results indicate the formation of an N-C covalent bond and the presence of N-doped C in the nanocomposites. In the N 1s spectra (Figure 4d), the two peaks located at 400.18 eV and 398.63 eV correspond to pyrrolic nitrogen and pyridinic nitrogen groups, respectively. According to [27], the pyrrole nitrogen and pyridine structures can provide suitable channels for Li+ insertion.
The N2 adsorption-desorption isotherms and the related pore size distributions calculated by the Barrett-Joyner-Halenda (BJH) method for the four samples are shown in Figure 5. The adsorption/desorption curves have a typical type IV shape (Figure 5a), corresponding to mesoporous structures with the distributions of pore size primarily around 10–20 nm (Figure 5b). The Brunauer-Emmett-Teller (BET) specific surface areas of Fe3O4@N-C-1, Fe3O4@N-C-2, Fe3O4@N-C-3 and Fe3O4@N-C-4 were calculated to be around 38, 37, 43 and 41 m2 g−1, respectively. By comparison, the Fe3O4@N-C-3 nanocomposites possessed the highest specific surface area, which can shorten ion and electron transport channels and provide sufficient contact of the nanocomposites with the electrolyte. Additionally, this special hollow porous structure can also help to accommodate the volume change during the process of charging and discharging.
The electrochemical properties of different Fe3O4@N-C samples were tested as the anode materials for LIBs. Figure 6a exhibits the first four cyclic voltammetry (CV) curves of Fe3O4@N-C-3 at a scan rate of 0.2 mV s−1. In the first cathodic curve, the reduction peak located around 0.54 V can be attributed to the reduction reaction of Fe3+ or Fe2+ to Fe0 (Fe3O4+2Li++2e→Li2Fe3O4, Li2Fe3O4+6Li++6e→4Li2O+3Fe), as well as the formation of a solid electrolyte interface (SEI) film [28]. In the first anodic scan, the peak at 1.75 V can be explained as the reversible oxidation of Fe0 to Fe2+ and Fe3+ (3Fe+4Li2O+8e→Fe3O4+8Li). After the first cycle, the CV curves almost overlap, showing that the stable SEI film is well generated on the surfaces of the electrode material, resulting in the enhanced stability and high coulombic efficiency of the Fe3O4@N-C-3 nanocomposites.
Figure 6b presents the 1st, 2nd, 50th and 100th discharge and charge curves for the Fe3O4@N-C-3 nanocomposites in the potential range of 0.01–3.0 V at a current density of 0.2 A g−1. The first discharge and charge capacities of the Fe3O4@N-C-3 electrode were, respectively, 1591 and 1452 mA h g−1, demonstrating an initial coulombic efficiency of 91%. The loss of reversible capacity might be attributed to the disintegration of the electrolyte, development of SEI film, trapping of lithium inside the active material and some other factors [29].
The cycling performances of the hollow porous Fe3O4@N-C nanocomposites are revealed in Figure 6c. It can be seen that the discharge capacity decreases in the first several dozens of cycles and then starts to increase; an analogous circumstance has also appeared in previous reports [25,30,31]. However, the mechanism is not clear. The reasons may be as follows: (1) the improved Li-diffusion kinetics of the reactivated electrode; (2) the activation of active materials; and (3) the structure refinement and formation of a thin and stable SEI without fracture [32,33]. The Fe3O4@N-C-3 nanocomposites exhibit excellent cycling stability; a high reversible capacity of 1772 mA h g−1 could be acquired after 100 cycles at a current density of 0.2 A g−1. It should be noted that the capacity is higher than the other samples, such as Fe3O4 (79 mA h g−1), Fe3O4@N-C-1 (540 mA h g−1), Fe3O4@N-C-2 (948 mA h g−1) and Fe3O4@N-C-4 (1127 mA h g−1). Furthermore, the cycling performance of Fe3O4@N-C-3 is also higher than those of the similar products reported previously (Table 1).
Figure 6c also demonstrates the coulombic efficiency of the composites during the cycling process. The coulombic efficiency of Fe3O4@N-C-3 nanocomposite after 100 cycles could reach 99.5%, demonstrating that the nanocomposites have excellent electrochemical performance. The rate capability is a crucial feature for high-performance LIBs to measure the reversible capacity and cycling stability. In Figure 6d, the Fe3O4@N-C-3 anode has reversible capacities of 1200, 1050 and 950 mA h g−1 at current rates of 0.2, 0.4 and 0.6 A g−1, respectively. The discharge capacity of Fe3O4@N-C-3 recovered to the greater value of 1700 mA h g−1 when the current density returned to 0.2 A g−1. These findings reveal that the hollow porous nanosphere structure can preserve the integrity of electrode as the current density increases, while the outer N-doped carbon layer also prevents large volume expansion during cycling, thus exhibiting an excellent rate performance.
Nyquist plots were fitted in order to better understand the character of the enhanced electrochemical performance. Figure 7 exhibits EIS spectroscopy carried out in a frequency range from 100 KHz to 0.01 Hz. The curves of five samples are all composed of a broad semicircle in the high-medium-frequency region and a long low-frequency line, which expresses the charge-transfer impedance (Rct) at the electrode/electrolyte interface and the Warburg impedance (Zw) related to the Li+ diffusion process in the electrode materials. The constant phase element (CPEct) represents the capacitance at the electrode-electrolyte interface. Compared to the pure Fe3O4, the electrical conductivity of the Fe3O4@N-C nanocomposites has been optimized. At the same time, the Fe3O4@N-C-3 electrode exhibits lower charge transfer resistance than other samples, which is related to the appropriate thickness of the N-C shell, high specific surface area and mesoporous structure. With these benefits, the conductivity of nanocomposites can be effectively increased, and the contact area between the electrode and the electrolyte can be expanded.
The adsorption capacities of Fe3O4@N-C nanocomposites to MO are shown in Figure 8a–d. It can be observed that the absorption peak intensity of MO at 465 nm rapidly declines with the extension of the adsorption duration in the presence of different Fe3O4@N-C nanocomposites. After about 100 min, the absorption peak of MO at 460 nm almost disappears. The results indicate that different Fe3O4@N-C nanocomposites all have good adsorption performances towards MO, which is due to the cooperation of the special hollow porous core-shell structure with the N-doped carbon layer [38] of the samples. From Figure 8e, we can more intuitively understand the relationship of MO adsorption behavior with time. It can be seen that the removal efficiency of MO over Fe3O4@N-C-3 at the adsorption of 100 min reaches 98.65%, while those over Fe3O4@N-C-1, Fe3O4@N-C-2 and Fe3O4@N-C-4 are 96.57%, 95.54% and 93.40%, respectively, which is due to the larger surface area increasing the contact between Fe3O4@N-C-3 and MO. The inset of Figure 8f demonstrates that the Fe3O4@N-C-3 has strong magnetism and is easy to recover and recycle. In Figure 8f, the removal efficiency of MO in the presence of Fe3O4@N-C-3 decreases to 86.92% after five cycles, indicating that the Fe3O4@N-C-3 nanocomposites as adsorbents have strong recyclability and adsorption stability, and are beneficial to the treatment of industrial dye wastewater.
In view of the fact that Fe3O4 is a semiconductor material with good SERS property [39], therefore, we selected the pesticide thiram as the probe to detect the SERS-enhanced effects of different Fe3O4@N-C substrates. Raman spectra (Figure 9a) show that thiram on silicon substrate had a strong feature peak at 558 cm−1, which belongs to the characteristic band of S-S stretching vibration [40]. There are some other peaks at 849 cm−1, 972 cm−1, 1145cm−1, 1375 cm−1 and 1466 cm−1, which are assigned to υ (C-N-C) and υ (C-S), υ (S-C-S) and ρ (CH3), ρ (CH3) and υ (C-N), υ (C-N) and ρ (CH3) [41,42], respectively. Since the characteristic peak intensities of thiram after 1100 cm−1 could be influenced by the distinct D and G bands of carbon at 1352 cm−1 and 1588 cm−1, we chose the intensity changes of 558 cm−1, 849 cm−1 and 972 cm−1 peaks (Figure 9b) as the references to intuitively reflect the SERS enhancement effect of Fe3O4@N-C substrates on thiram. We found that the peak intensities of the thiram are almost universally enhanced on Fe3O4@N-C substrates, which may be on account of the hollow porosity of the Fe3O4@N-C nanocomposites enlarging the contact between the probe molecules and the substrates. It can be seen by combining Figure 9a,b, the peaks of the pesticide thiram are the strongest on the Fe3O4@N-C-3 substrate. According to BET analysis results, Fe3O4@N-C-3 has the largest specific surface area, which can offer more surface-active sites and facilitate the adsorption of probe molecules, resulting in better SERS detection performance of Fe3O4@N-C-3 for the thiram. This result indicates that the prepared Fe3O4@N-C-3 nanocomposites could be used to detect thiram with the SERS technique. Related studies will be reported separately.

3. Materials and Methods

3.1. Materials and Chemicals

Sodium citrate (Na3C6H5O7∙2H2O), iron (III) chloride hexahydrate (FeCl3∙6H2O, 99%), urea (CO(NH2)2) and polyacrylamide ((C3H5NO)n) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). All reagents were of analytical grade and used without further purification.

3.2. Synthesis of Fe3O4 Nanospheres

Firstly, FeCl3·6H2O (6.6 mmol) was dissolved into the deionized water (80 mL) under ultrasound irradiation for 20 min at room temperature to form yellow transparent solution, and then 8.0 mmol of Na3C6H5O7·2H2O and 12.0 mmol of urea were added to the solution under stirring for 30 min. Subsequently, 8.6 mmol of polyacrylamide was added into above solution under continuous stirring until it was dissolved totally, and a light green clear solution formed. The above solution was transferred into a 100 mL Teflon-lined stainless steel autoclave and heated at 200 °C for 12 h. The sediment was collected after being cooled down to room temperature, washed with deionized water and alcohol three times each and dried at 80 °C for 12 h to obtain Fe3O4 black powders.

3.3. Synthesis of Hollow Porous Fe3O4@N-C Nanosphere Composites

Overall, 0.5 g of Fe3O4 power and 0.1 g of dopamine hydrochloride were decentralized successively in 20 mL of ethanol under stirring for 30 min, and the mixture was dried under vacuum at 80 °C. The carbon-precursor-coated Fe3O4 hollow nanospheres were further heated in a tube furnace at 500 °C in nitrogen atmosphere for 4 h to acquire Fe3O4@N-C nanospheres (Fe3O4@N-C-1). In the following step, 1.0 g Fe3O4@N-C nanospheres were etched in 75 mL hydrochloric acid (HCl) solution for 15 min (Fe3O4@N-C-2), 30 min (Fe3O4@N-C-3) and 45 min (Fe3O4@N-C-4), respectively.

3.4. Characterization

Field-emission scanning electron microscopy (FE-SEM, S4800, Hitachi, Tokyo, Japan) and high-resolution transmission electron microscopy (HR-TEM, JEM2100, JEOL Ltd., Beijing, China) were used to investigate the morphologies of the samples. The crystal structures of the samples were determined by X-ray diffraction (XRD) (DX-2700) using Cu-Kα (30 kV, 25 mA, λ = 1.5406 Å) radiation over a 2θ range of 5–80°. A Dilor LABRAM-1B multi-channel confocal microspectrometer with an excitation wavelength of 532 nm was applied to record the Raman spectra. The surface chemical compositions of samples were investigated using X-ray photoelectron spectroscopy (XPS, ESCALAB 250, Thermo Scientific, Waltham, MA, USA). The thermogravimetry (TG) of the nanocomposites was performed in air atmosphere from room temperature to 800 °C at a heating rate of 10 °C min−1. Fourier transform infrared spectrometry (FT-IR) (Thermo Nicolet NEXUS 670, Thermo Scientific, Waltham, MA, USA). with a wavelength range from 400 to 4000 cm−1 was applied to identify the functional groups of samples. On the 3H-2000BET-A system, N2 adsorption-desorption isotherms were determined at liquid nitrogen temperature.

3.5. Electrochemical Measurements

The active material powders, acetylene black (Super P) and polyvinylidene fluoride (PVDF), were mixed in a 8:1:1 mass ratio, then added into N-methyl-2-pyrrolidinone (NMP) solvent to form a homogeneous slurry with stirring, subsequently spread on a copper foil substrate and finally dried in vacuum at 80 °C for 10 h to remove the solvent. Thus, a working electrode was obtained. The 2016 coin-type cells were assembled in an argon-filled glove box using metallic lithium as the counter electrode and polyethylene film (Celgard, 2400) as the separator. The nonaqueous electrolyte was used in the 1 M solution of LiPF6 dissolving in the 1:1 wt% solution of ethylene carbonate (EC) to dimethyl carbonate (DMC). Electrochemical measurements were performed at room temperature on CT-3008W-5V 10 mA battery testing equipment (Neware Technology Co. Ltd., Shenzhen, China) at various current densities ranging from 0.01 to 3.0 V. Using battery testing equipment (NEWARE BTS-3000), the galvanostatic charge-discharge profiles were determined at various current densities within the voltage window of 0.01 to 3.0 V (vs. Li/Li+). Cyclic voltammetry (CV) measurements were performed on an electrochemical workstation (CHI660D, Shanghai CH Instruments Co., Ltd., Shanghai, China) over the potential between 0.01 V and 3.0 V at a scanning rate of 0.2 mV s−1. Electrochemical impedance spectroscopy (EIS) was measured in the 100 KHz to 0.01 Hz frequency range.

3.6. Adsorption Measurements

As a typical pollutant, methyl orange (MO) was employed to study sample adsorption behaviors. The details are as follows: 10 mg of the sample was added to 50 mL of 10−5 M MO aqueous solution, and stirred at a constant speed at room temperature. At 0, 1, 5, 10, 15, 20, 30, 40, 60, 80 and 100 min, the upper clarified solution after reaction was taken out and the ultraviolet visible spectrophotometer (SHIMADZU, UV-1800, Shanghai, China) was applied to measure the absorbance of MO at 465 nm in the solution.
Through the above adsorption experiments, the sample with the best adsorption performance was selected for investigating the recycle adsorption stability. The first adsorption was carried out according to the above steps. After 100 min, the sample was separated from the solution with an external magnet, rinsed with water and ethanol three times, respectively, and then re-decentralized into 50 mL of 10−5 M MO aqueous solution for the adsorption experiment. The adsorption was carried out five times.
The MO removal efficiency of the sample can be computed by Formula (1):
Removal   efficiency   ( % ) = C 0 C t C 0 × 100 %
where C0 and Ct represent the MO concentration in the original and after being adsorbed by the sample for a period of time, respectively.

3.7. SERS Measurements

For the preparation of the SERS substrates, 10 mg of the samples was dispersed into 5 mL of anhydrous alcohol, followed by dripping a drop of the dispersions onto the silicon substrate and then natural room-temperature drying. Next, the above substrate was covered with a drop of a 10−3 M thiram aqueous solution, which was dried at room temperature. The blank group was obtained by dropping the 10−3 M thiram aqueous solution on the silicon substrate without samples. A Dilor LABRAM-1B multi-channel confocal micro spectrometer (Dilor, Lille, France) with 532 nm wavelength and 0.5 mW of incident excitation power was used to record the Raman spectra.

4. Conclusions

In summary, multifunctional hollow porous Fe3O4@N-C core-shell nanocomposites have been fabricated successfully by a simple self-template method, which involved the preparation of Fe3O4 nanospheres via the solvothermal method, in situ formation of dopamine coating and subsequent calcination in N2. Then, the Fe3O4@N-C nanocomposites were etched in hydrochloric acid solution by controlling the reaction time, and finally a series of hollow porous Fe3O4@N-C nanocomposites were acquired. The unique hollow porous core-shell structure with appropriate N-doped carbon coating can shorten the ion/electron transport channel, reduce the volume change during the repeated insertion/extraction of lithium ions and improve the structural stability. The typical Fe3O4@N-C-3 nanocomposites as an anode for LIBs exhibit excellent specific capacity, rate capacity and cycling performance. Meanwhile, Fe3O4@N-C nanocomposites also show expansive application prospects in SERS determination and the adsorption of pollutants in water. The removal rate of MO reached 98.65% after 100 min when Fe3O4@N-C-3 was used as the adsorbent and the Raman spectral response intensity of the thiram increased obviously on Fe3O4@N-C substrates. In the future, the electrochemical performance of a full battery containing Fe3O4@N-C anode materials, the faster and more selective adsorption of the samples towards organic pollutants in wastewater and the sensitivity of Fe3O4@N-C SERS substrates for detecting pesticide residues could be further investigated to explore the practical application possibility of products.

Author Contributions

Conceptualization, C.Q., M.Z. and P.W.; writing-original draft, C.Q.; methodology and data analysis, M.Z.; figure analysis, T.F.; formal conceptualization and visualization, Y.Z.; resources and funding acquisition, P.W.; supervision, A.X.; writing-review and editing, Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Nature Science Foundation of China (21173001, 21371003 and 21571002), horizontal project of Hefei Normal University (50122149), Inflammation and Immune Mediated Diseases Laboratory of Anhui Province, Anhui Institute of Innovative Drugs (IMMDL202101), and Major Project of Anhui Provincial Department of Education (2022AH040097).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

There are no conflict of interest to declare.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Ye, H.; Zheng, G.; Yang, X.; Zhang, D.; Zhang, Y.; Yan, S.; You, L.; Hou, S.; Huang, Z. Application of different carbon-based transition metal oxide composite materials in lithium-ion batteries. J. Electroanal. Chem. 2021, 898, 115652. [Google Scholar] [CrossRef]
  2. Wang, X.; Yang, J.; Zhou, L.; Song, G.; Lu, F.; You, L.; Li, J. Rapid and ultrasensitive surface enhanced Raman scattering detection of hexavalent chromium using magnetic Fe3O4/ZrO2/Ag composite microsphere substrates. Colloids Surf. A Physicochem. Eng. Asp. 2021, 610, 125414. [Google Scholar] [CrossRef]
  3. Dehmani, Y.; Dridi, D.; Lamhasni, T.; Abouarnadasse, S.; Chtourou, R.; Lima, E.C. Review of phenol adsorption on transition metal oxides and other adsorbents. J. Water Process Eng. 2022, 49, 102965. [Google Scholar] [CrossRef]
  4. Molaiyan, P.; Dos Reis, G.S.; Karuppiah, D.; Subramaniyam, C.M.; García-Alvarado, F.; Lassi, U. Recent Progress in Biomass-Derived Carbon Materials for Li-Ion and Na-Ion Batteries—A Review. Batteries 2023, 9, 116. [Google Scholar] [CrossRef]
  5. Abdollahifar, M.; Molaiyan, P.; Lassi, U.; Wu, N.L.; Kwade, A. Multifunctional Behaviour of Graphite in Lithium–sulfur Batteries. Renew. Sustain. Energy Rev. 2022, 169, 112948. [Google Scholar] [CrossRef]
  6. Yong, L.; Duan, C.; Jiang, Z.; Zhou, X.; Wang, Y. CuO/rGO nanocomposite as an anode material for high-performance lithium-ion batteries. Mater. Res. Express 2021, 8, 055505. [Google Scholar]
  7. Dong, H.; Deng, M.; Sun, D.; Zhao, Y.; Liu, H.; Xie, M.; Dong, W.; Huang, F. Amorphous Lithium-Phosphate-Encapsulated Fe2O3 as a High-Rate and Long-Life Anode for Lithium-Ion Batteries. ACS Appl. Energy Mater. 2022, 5, 3463–3470. [Google Scholar] [CrossRef]
  8. Sun, Y.; Huang, F.; Li, S.; Shen, Y.; Xie, A. Novel porous starfish-like Co3O4@nitrogen-doped carbon as an advanced anode for lithium-ion batteries. Nano Res. 2017, 10, 3457–3467. [Google Scholar] [CrossRef]
  9. Yi, Z.; Lin, N.; Xu, T.; Qian, Y. TiO2 coated Si/C interconnected microsphere with stable framework and interface for high-rate lithium storage. Chem. Eng. J. 2018, 347, 214–222. [Google Scholar] [CrossRef]
  10. Gao, Z.; Yang, Y.; Qin, J.; Su, Z. A core-shell porous MnO2/Carbon nanosphere composite as the anode of lithium-ion batteries. J. Power Sources 2021, 491, 229577. [Google Scholar]
  11. Jang, J.; Song, S.; Kim, H.; Moon, J.; Ahn, H.; Jo, K.; Bang, J.; Kim, H.; Koo, J. Janus Graphene Oxide Sheets with Fe3O4 Nanoparticles and Polydopamine as Anodes for Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2021, 13, 14786–14795. [Google Scholar] [CrossRef]
  12. Kopuklu, B.; Kopuklu, A.; Yurum, A.; Kopuklu, A. High stability graphene oxide aerogel supported ultrafine Fe3O4 particles with superior performance as a Li-ion battery anode. Carbon 2021, 174, 158–172. [Google Scholar] [CrossRef]
  13. Hao, S.; Li, Q.; Qu, J.; An, F.; Zhang, Y.; Yu, Z. Neuron-Inspired Fe3O4/Conductive Carbon Filament Network for High-Speed and Stable Lithium Storage. ACS Appl. Mater. Interfaces 2018, 10, 17923–17932. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, K.; Zhang, Q.; Gao, X.; Chen, X.; Wang, Y.; Li, W.; Wu, J. Effect of absorbers’ composition on the microwave absorbing performance of hollow Fe3O4 nanoparticles decorated CNTs/graphene/C composites. J. Alloys Compd. 2018, 748, 706–716. [Google Scholar] [CrossRef]
  15. Zhang, K.; Zhang, Q.; Gao, X.; Chen, X.; Shi, J.; Wu, J. Ellipsoidal Fe3O4@C nanoparticles decorated fluffy structured graphene nanocomposites and their enhanced microwave absorption properties. J. Mater. Sci. Mater. Electron. 2018, 29, 6785–6796. [Google Scholar] [CrossRef]
  16. Wang, J.; Zhang, M.; Xu, J.; Zheng, J.; Hayat, T.; Alharbi, N.S. Formation of Fe3O4@C/Ni microtubes for efficient catalysis and protein adsorption. Dalton Trans. 2018, 47, 2791–2798. [Google Scholar] [CrossRef] [PubMed]
  17. Lin, Q.; Wang, J.; Zhong, Y.; Sunarso, J.; Tadé, M.O.; Li, L.; Shao, Z. High performance porous iron oxide-carbon nanotube nanocomposite as an anode material for lithium-ion batteries. Electrochim. Acta 2016, 212, 179–186. [Google Scholar] [CrossRef]
  18. Bhattacharjya, D.; Park, H.; Kim, M.; Choi, H.; Inamdar, S.N.; Yu, J. Nitrogen-Doped Carbon nanoparticles by flame synthesis as anode material for rechargeable lithium-ion batteries. Langmuir 2014, 30, 318–324. [Google Scholar] [CrossRef]
  19. Peng, C.; Zhou, S.; Zhang, X.; Zeng, T.; Zhang, W.; Li, H.; Liu, X.; Zhao, P. One pot synthesis of nitrogen-doped hollow carbon spheres with improved electrocatalytic properties for sensitive H2O2 sensing in human serum. Sens. Actuators B Chem. 2018, 270, 530–537. [Google Scholar] [CrossRef]
  20. He, J.; Zhao, S.; Lian, Y.; Zhou, M.; Wang, L.; Ding, B.; Cui, S. Graphene-doped carbon/Fe3O4 porous nanofibers with hierarchical band construction as high-performance anodes for lithium-ion batteries. Electrochim. Acta 2017, 229, 306–315. [Google Scholar] [CrossRef]
  21. Tran, T.; Phan, T.; Nguyen, D.; Nguyen, T.; Nguyen, D.; Vo, D.; Bach, L.; Nguyen, D. Recyclable Fe3O4@C nanocomposite as potential adsorbent for a wide range of organic dyes and simulated hospital effluents. Environ. Technol. Innov. 2021, 20, 101122. [Google Scholar] [CrossRef]
  22. Xu, J.; Li, X.; Wang, Y.; Guo, R.; Shang, S.; Jiang, S. Flexible and reusable cap-like thin Fe2O3 film for SERS applications. Nano Res. 2019, 12, 381–388. [Google Scholar] [CrossRef]
  23. Yang, Y.; Gao, X.; Yang, S.; Shen, Y.; Xie, A. Synthesis and superior SERS performance of porous octahedron Cu2O with oxygen vacancy derived from MOFs. J. Mater. Sci. 2021, 56, 9702–9711. [Google Scholar] [CrossRef]
  24. Liu, L.; Pan, F.; Liu, C.; Huang, L.; Li, W.; Lu, X. TiO2 nanofoam–nanotube array for surface-enhanced Raman scattering. ACS Appl. Nano Mater. 2018, 1, 6563–6566. [Google Scholar] [CrossRef] [Green Version]
  25. Xu, J.; Zhang, J.; Wang, C. Carbon encapsulated Fe3O4 nanospheres with high electrochemical performance as anode materials for Li-ion battery. Int. J. Appl. Ceram. Technol. 2017, 14, 938–947. [Google Scholar] [CrossRef]
  26. Han, F.; Ma, L.; Sun, Q.; Lei, C.; Lu, A. Rationally designed carbon-coated Fe3O4 coaxial nanotubes with hierarchical porosity as high-rate anodes for lithium ion batteries. Nano Res. 2014, 7, 1706–1717. [Google Scholar] [CrossRef]
  27. Jiang, Z.; Jiang, Z. Fabrication of Nitrogen-Doped Holey Graphene Hollow Microspheres and Their Use as an Active Electrode Material for Lithium Ion Batteries. ACS Appl. Mater. Interfaces 2014, 6, 19082–19091. [Google Scholar] [CrossRef]
  28. Lei, C.; Han, F.; Sun, Q.; Li, W.; Lu, A. Confined nanospace pyrolysis for the fabrication of coaxial Fe3O4@C hollow particles with a penetrated mesochannel as a superior anode for Li-ion batteries. Chemistry 2014, 20, 139–145. [Google Scholar] [CrossRef]
  29. Dang, R.; Jia, X.; Liu, X.; Ma, H.; Gao, H.; Wang, G. Controlled synthesis of hierarchical Cu nanosheets@CuO nanorods as high-performance anode material for lithium-ion batteries. Nano Energy 2017, 33, 427–435. [Google Scholar] [CrossRef]
  30. Wang, L.; Yu, Y.; Chen, P.; Zhang, D.; Chen, C. Electrospinning synthesis of C/Fe3O4 composite nanofibers and their application for high performance lithium-ion batteries. J. Power Sources 2008, 183, 717–723. [Google Scholar] [CrossRef]
  31. Wu, H.; Du, N.; Wang, J.; Zhang, H.; Yang, D. Three-dimensionally porous Fe3O4 as high-performance anode materials for lithium–ion batteries. J. Power Sources 2014, 246, 198–203. [Google Scholar] [CrossRef]
  32. Zhou, G.; Wang, D.; Li, F.; Zhang, L.; Li, N.; Wu, Z.; Wen, L.; Lu, G.; Cheng, H. Graphene-Wrapped Fe3O4 Anode Material with Improved Reversible Capacity and Cyclic Stability for Lithium Ion Batteries. Chem. Mater. 2010, 22, 5306–5313. [Google Scholar] [CrossRef]
  33. Cherian, C.; Sundaramurthy, J.; Kalaivani, M.; Ragupathy, P.; Kumar, P.; Thavasi, V.; Reddy, M.; Sow, C.; Mhaisalkar, S.; Ramakrishna, S.; et al. Electrospun α-Fe2O3 nanorods as a stable, high capacity anode material for Li-ion batteries. J. Mater. Chem. 2012, 22, 12198–12204. [Google Scholar] [CrossRef]
  34. Zhao, Y.; Li, J.; Wu, C.; Ding, Y.; Guan, L. A Yolk-Shell Fe3O4@C Composite as an Anode Material for High-Rate Lithium Batteries. ChemPlusChem 2012, 77, 748–751. [Google Scholar] [CrossRef]
  35. Su, L.; Zhong, Y.; Zhou, Z. Role of transition metal nanoparticles in the extra lithium storage capacity of transition metal oxides: A case study of hierarchical core–shell Fe3O4@C and Fe@C microspheres. J. Mater. Chem. A 2013, 1, 15158–15166. [Google Scholar] [CrossRef]
  36. Deng, W.; Ci, S.; Li, H.; Wen, Z. One-step ultrasonic spray route for rapid preparation of hollow Fe3O4/C microspheres anode for lithium-ion batteries. Chem. Eng. J. 2017, 330, 995–1001. [Google Scholar] [CrossRef]
  37. Jiang, F.; Liu, Y.; Wang, Q.; Zhou, Y. Hierarchical Fe3O4@NC composites: Ultra-long cycle life anode materials for lithium ion batteries. J. Mater. Sci. 2017, 53, 2127–2136. [Google Scholar] [CrossRef]
  38. Nsabimana, A.; Kitte, S.; Wu, F.; Qi, L.; Liu, Z.; Zafar, M.; Luque, R.; Xu, G. Multifunctional magnetic Fe3O4/nitrogen-doped porous carbon nanocomposites for removal of dyes and sensing applications. Appl. Surf. Sci. 2018, 467, 89–95. [Google Scholar] [CrossRef]
  39. Zhao, D.; Lin, K.; Wang, L.; Qin, Z.; Zhao, X.; Du, K.; Han, L.; Tian, F.; Chang, Y. A physical approach for the estimation of the SERS enhancement factor through the enrichment and separation of target molecules using magnetic adsorbents. RSC Adv. 2020, 10, 20028–20037. [Google Scholar] [CrossRef]
  40. Li, L.; Zhao, A.; Wang, D.; Guo, H.; Sun, H.; He, Q. Fabrication of cube-like Fe3O4@SiO2@Ag nanocomposites with high SERS activity and their application in pesticide detection. J. Nanoparticle Res. 2016, 18, 178. [Google Scholar] [CrossRef]
  41. Yao, C.; Gao, X.; Liu, X.; Shen, Y.; Xie, A. In-situ preparation of Ferrero (R) chocolate-like Cu2O@Ag microsphere as SERS substrate for detection of thiram. J. Mater. Res. Technol. 2021, 11, 857–865. [Google Scholar] [CrossRef]
  42. Liu, Z.; Wang, Y.; Deng, R.; Yang, L.; Yu, S.; Xu, S.; Xu, W. Fe3O4@Graphene Oxide@Ag Particles for Surface Magnet Solid-Phase Extraction Surface-Enhanced Raman Scattering (SMSPE-SERS): From Sample Pretreatment to Detection All-in-One. ACS Appl. Mater. Interfaces 2016, 8, 14160–14168. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Preparation process and multi-applications of the hollow porous Fe3O4@N-C nanocomposites.
Scheme 1. Preparation process and multi-applications of the hollow porous Fe3O4@N-C nanocomposites.
Molecules 28 05183 sch001
Figure 1. (a) XRD patterns and (b) Raman spectra of different samples.
Figure 1. (a) XRD patterns and (b) Raman spectra of different samples.
Molecules 28 05183 g001
Figure 2. (a,c,e) SEM and (b,d,fh) TEM images of (a,b) Fe3O4@N-C-2, (c,d,g) Fe3O4@N-C-3 and (e,f) Fe3O4@N-C-4, respectively; (g) high-magnification and (h) high-resolution TEM images of Fe3O4@N-C-3. The insets in figures (a,c,e) present the particle size distributions of Fe3O4@N-C-2, Fe3O4@N-C-3 and Fe3O4@N-C-4 nanocomposites.
Figure 2. (a,c,e) SEM and (b,d,fh) TEM images of (a,b) Fe3O4@N-C-2, (c,d,g) Fe3O4@N-C-3 and (e,f) Fe3O4@N-C-4, respectively; (g) high-magnification and (h) high-resolution TEM images of Fe3O4@N-C-3. The insets in figures (a,c,e) present the particle size distributions of Fe3O4@N-C-2, Fe3O4@N-C-3 and Fe3O4@N-C-4 nanocomposites.
Molecules 28 05183 g002
Figure 3. (a) FT-IR spectra and (b) TG curves of different samples.
Figure 3. (a) FT-IR spectra and (b) TG curves of different samples.
Molecules 28 05183 g003
Figure 4. (a) XPS survey spectrum of the Fe3O4@N-C-3 nanocomposites, (b) Fe 2p, (c) C 1s and (d) N 1s core-level spectra of the Fe3O4@N-C-3 nanocomposites.
Figure 4. (a) XPS survey spectrum of the Fe3O4@N-C-3 nanocomposites, (b) Fe 2p, (c) C 1s and (d) N 1s core-level spectra of the Fe3O4@N-C-3 nanocomposites.
Molecules 28 05183 g004
Figure 5. (a) N2 adsorption-desorption isotherms and (b) pore size distributions of different samples.
Figure 5. (a) N2 adsorption-desorption isotherms and (b) pore size distributions of different samples.
Molecules 28 05183 g005
Figure 6. (a) Cyclic voltammograms at a scan rate of 0.2 mV s−1 and (b) charge-discharge profiles at 0.2 A g−1 of the Fe3O4@N-C-3; (c) cycling performance of the five samples for 100 cycles at a current density of 0.2 A g−1; (d) rate performance of Fe3O4@N-C-3 at different current densities.
Figure 6. (a) Cyclic voltammograms at a scan rate of 0.2 mV s−1 and (b) charge-discharge profiles at 0.2 A g−1 of the Fe3O4@N-C-3; (c) cycling performance of the five samples for 100 cycles at a current density of 0.2 A g−1; (d) rate performance of Fe3O4@N-C-3 at different current densities.
Molecules 28 05183 g006
Figure 7. Nyquist plots of the Fe3O4@N-C electrodes in the frequency range between 100 kHz and 0.01 Hz. The inset presents the equivalent circuit.
Figure 7. Nyquist plots of the Fe3O4@N-C electrodes in the frequency range between 100 kHz and 0.01 Hz. The inset presents the equivalent circuit.
Molecules 28 05183 g007
Figure 8. UV-Vis absorption spectra of MO over (a) Fe3O4@N-C-1, (b) Fe3O4@N-C-2, (c) Fe3O4@N-C-3 and (d) Fe3O4@N-C-4 nanocomposites. (e) The change of removal efficiency of MO by different Fe3O4@N-C nanocomposites with time. (f) The removal rate changes of MO by the Fe3O4@N-C-3 nanocomposites for five cycles. The inset shows Fe3O4@N-C-3 in water being attracted apart with an external magnet.
Figure 8. UV-Vis absorption spectra of MO over (a) Fe3O4@N-C-1, (b) Fe3O4@N-C-2, (c) Fe3O4@N-C-3 and (d) Fe3O4@N-C-4 nanocomposites. (e) The change of removal efficiency of MO by different Fe3O4@N-C nanocomposites with time. (f) The removal rate changes of MO by the Fe3O4@N-C-3 nanocomposites for five cycles. The inset shows Fe3O4@N-C-3 in water being attracted apart with an external magnet.
Molecules 28 05183 g008
Figure 9. (a) SERS spectra of thiram before (blank) and after absorption on different Fe3O4@N-C substrates; (b) the related peak intensities at 558 cm−1, 849 cm−1 and 972 cm−1 of thiram in (a).
Figure 9. (a) SERS spectra of thiram before (blank) and after absorption on different Fe3O4@N-C substrates; (b) the related peak intensities at 558 cm−1, 849 cm−1 and 972 cm−1 of thiram in (a).
Molecules 28 05183 g009
Table 1. Contrast of the electrochemical performances of similar LIB anode materials.
Table 1. Contrast of the electrochemical performances of similar LIB anode materials.
SamplesCurrent Density mA g−1Cycle NumbersCapacity
mA h g−1
Ref.
yolk–shell Fe3O4@C150201010[34]
Fe3O4/C nanofibers500150780[20]
core–shell Fe3O4@C and Fe@C50401080[35]
3D porous Fe3O410001001382[31]
porous Fe3O4/C microspheres1001001180[36]
hierarchical Fe3O4@NC500100992[37]
hollow porous Fe3O4@N/C2001001772This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Qi, C.; Zhao, M.; Fang, T.; Zhu, Y.; Wang, P.; Xie, A.; Shen, Y. Multifunctional Hollow Porous Fe3O4@N-C Nanocomposites as Anodes of Lithium-Ion Battery, Adsorbents and Surface-Enhanced Raman Scattering Substrates. Molecules 2023, 28, 5183. https://doi.org/10.3390/molecules28135183

AMA Style

Qi C, Zhao M, Fang T, Zhu Y, Wang P, Xie A, Shen Y. Multifunctional Hollow Porous Fe3O4@N-C Nanocomposites as Anodes of Lithium-Ion Battery, Adsorbents and Surface-Enhanced Raman Scattering Substrates. Molecules. 2023; 28(13):5183. https://doi.org/10.3390/molecules28135183

Chicago/Turabian Style

Qi, Chunxia, Mengxiao Zhao, Tian Fang, Yaping Zhu, Peisan Wang, Anjian Xie, and Yuhua Shen. 2023. "Multifunctional Hollow Porous Fe3O4@N-C Nanocomposites as Anodes of Lithium-Ion Battery, Adsorbents and Surface-Enhanced Raman Scattering Substrates" Molecules 28, no. 13: 5183. https://doi.org/10.3390/molecules28135183

Article Metrics

Back to TopTop