Next Article in Journal
Effect of External Compression on the Thermal Runaway of Lithium-Ion Battery Cells during Crush Tests: Insights for Improved Safety Assessment
Next Article in Special Issue
Rational Design of Bismuth Metal Anodes for Sodium-/Potassium-Ion Batteries: Recent Advances and Perspectives
Previous Article in Journal
Conditioning Solid-State Anode-Less Cells for the Next Generation of Batteries
Previous Article in Special Issue
Low-Temperature-Aged Synthesis of CeO2-Coated Li-Rich Oxide as Cathode for Low-Cost High-Energy Density Li-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Facile Constructing Hierarchical Fe3O4@C Nanocomposites as Anode for Superior Lithium-Ion Storage

1
Fujian Provincial Key Laboratory of Functional Materials and Applications, School of Materials Science and Engineering, Xiamen University of Technology, Xiamen 361024, China
2
School of Chemistry and Material Science, Hunan Agricultural University, Changsha 410128, China
3
School of Marine Engineering, Jimei University, Xiamen 361021, China
4
College of Materials Science and Engineering, Zhejiang University of Technology, Hangzhou 310014, China
5
College of Chemical Engineering and Light Industry, Guangdong University of Technology, Guangzhou 510006, China
*
Authors to whom correspondence should be addressed.
Batteries 2023, 9(8), 403; https://doi.org/10.3390/batteries9080403
Submission received: 26 June 2023 / Revised: 12 July 2023 / Accepted: 27 July 2023 / Published: 2 August 2023
(This article belongs to the Special Issue Transition Metal Compound Materials for Secondary Batteries)

Abstract

:
Ferroferric oxide (Fe3O4) is regarded to be a promising high-capacity anode material for LIBs. However, the capacity attenuates fast and the rate performance is poor due to the dramatic pulverization and sluggish charge transfer properties. To solve these problems, a simple in situ encapsulation and composite method was successfully developed to construct carbon nanotube/nanorod/nanosheet-supported Fe3O4 nanoparticles. Owing to the hierarchical architecture design, the novel structure Fe3O4@C nanocomposites effectively enhance the charge transfer, alleviate pulverization, avoid the agglomeration of Fe3O4 nanoparticles, and also provide superior kinetics toward lithium storage, thereby showing significantly improved reversibility and rate performance. The carbon nanotube/nanorod supported core-shell structure Fe3O4@C nanocomposite displays outstanding high rate capability and stable cycling performance (reversible capability of 1006, 552 and 423 mA h g−1 at 0.2, 0.5 and 1 A g−1 after running 100, 300 and 500 cycles, respectively).

Graphical Abstract

1. Introduction

The lithium-ion battery (LIB), an effective energy storage device, is extensively applied to electric vehicles and various portable devices, e.g., mobile phones, laptops and so on [1,2,3]. However, commercial graphite as an anode material has fatal disadvantages of low specific capacity (372 mAh g−1) and deficient rate performance, so the present LIB cannot support sufficient capacity for intelligent electronic devices to satisfy people’s demand for long working time [1,4]. To solve the problem, we need to explore new suitable anode materials with high specific capacity and excellent cycling stability. In the past decades, intensive efforts have been employed to explore various kinds of new anode materials, including Si, Sn, NiSb, etc.; alloy-type [5,6,7,8,9] and transition metal oxides (e.g., Fe2O3/Fe3O4, Co3O4, SnO2, etc.) [10,11,12,13,14] and so on. Ferroferric oxide (Fe3O4) is regarded to be a promising anode material and attracts many researchers’ interest due to the merits of high reversible theoretical specific capacity (928 mAh g−1), low cost and eco-friendliness [9,10,15,16,17]. Unfortunately, the capacity of the Fe3O4 anode decays very fast, which results from the severe pulverization of Fe3O4 particles because of the huge volume variation caused by lithiation and delithiation; therefore, the Fe3O4 anode exhibits poor cycling stability [18,19,20,21]. On the other hand, the charge transfer property of the pristine Fe3O4 is poor and needs to be improved in an electrode material [22,23].
To solve these disadvantages, many strategies were explored to enhance the structural stability and improve the electrical conductivity of Fe3O4 [23,24,25,26,27]. Concerning enhancing the structural stability, the synthesis of nanostructured Fe3O4, such as nanoparticles, nanorods, hollow spheres, and so on, was intensively investigated [24,25,26,27,28,29]. The other way is designing Fe3O4@C nanocomposites, including carbon-coated Fe3O4 nanocomposites [16], graphene-wrapped nano-Fe3O4 or 3D structured Fe3O4@C nanocomposites [30,31,32]. Compared with the bulk Fe3O4, nanostructured Fe3O4 is not only beneficial in alleviating the stress but also helps to reduce the Li/Li+ diffusion distance [33,34,35]. It was reported that the single-crystalline Fe3O4 and α-Fe2O3 nanorods could alleviate stress effectively and supply a larger electrochemical reaction area [36]. With respect to Fe3O4@C nanocomposites, the carbon matrix severed as a buffer for the accommodation of strain and inhibited the agglomeration and coarsening of nano-Fe3O4 [37]. On the other hand, attributed to its good conductivity, the introduction of carbon also enhances the charge transfer of Fe3O4 [38]. As a result, the Fe3O4@C nanocomposites always display improved cycling stability. For example the Fe3O4 nanoparticles encapsulated by macroporous carbon spheres exhibited a discharge capacity of 645 mAh g−1 after 1000 cycles at 2.0 A g−1 [28].
Recently, constructions of yolk-shell or other hollow structures became a favorable methodology to realize the special structural effects [39,40,41,42]. Metal–organic frameworks (MOFs) were considered to be the ideal precursors for the preparation of functional porous materials (metal/metal oxide and carbon composites) with desired shapes [17,41,43,44]. Using ferrocene as the precursor, Liu and co-authors synthesized a pitaya-structured Fe3O4@C nanocomposite and finally obtained the same structured FeS2@C nanocomposites as a cathode material for LIBs, which displayed stable and high energy density (~1100 Wh kg−1) and ultrahigh rate capability41. The yolk-shell Fe2O3@C composite containing mesoporous derived from MIL-53(Fe), could retain 744 mAh g−1 after 500 cycles at 1 A g−1 [43] Although notable successes have been achieved, these chemical synthesis methods were severely limited by their multistep, the use of organic solvents and low yield [45] In this work, we presented a facile and suitable large-scale application method to prepare Fe3O4@C nanocomposites with controllable morphologies by carbonizing the plasma-assisted ball-milled ferrocene (C10H10Fe) precursor. The leaves-like amorphous carbon-nanosheet-encapsulated ultrafine nano-Fe3O4 and carbon nanotubes/nanorods-supported core-shell nanostructured Fe3O4 composites were prepared and exhibited outstanding electrochemical performances as an anode material for LIBs. Typically the carbon nanotubes/nanorods-supported core-shell nanostructured Fe3O4 presented impressive cycling performance (423 mAh g−1 at 1 A g−1 after 500 cycles) and outstanding high-rate capability.

2. Experiments

2.1. Material Preparation and Characterization

The synthesis of Fe3O4@C nanocomposites is illustrated in Scheme 1. First, the purchased ferrocene (C10H10Fe) powders were milled for 5 h on a plasma-assisted ball miller (PA miller, homemade) to achieve partially carbonized pretreated products. Then, these pretreated products were annealed in the tube furnace at 773 and 873 K for 30 min under the atmosphere of argon and air mixtures. The obtained black powders were the final Fe3O4@C nanocomposites. The Fe3O4@C nanocomposites achieved at 773 and 873 K were named Fe3O4@C-773 and Fe3O4@C-873, respectively, in the following.
The phase components of the Fe3O4@C composites were examined by an X-ray diffractometer (XRD, Rigaku SmartLab 2 kW, Tokyo, Japan) using Cu-Kα radiation (λ = 1.54056 Å) at 40 kV and 40 mA. The morphology and microstructure were further investigated by a field emission scanning electron microscope (SEM, ZEISS Sigma 500) and scanning transmission electron microscopy (STEM, FEI TALOS F200S attached BRUKER Super-X EDS). The carbon contents were determined using thermogravimetric analysis (TG, Setaram Labsys) at a heating rate of 2 K/min under an airflow of 50 mL/min. The pore structure and specific surface area were evaluated by the measurement of nitrogen adsorption/desorption isotherms at 77 K on a micromeritics 3Flex instrument (USA). The Raman spectra were tested on an invia apparatus (Renishaw, Wotton-under-Edge, UK). X-ray photoelectron spectroscopy (XPS, Nexsa from ThermoFisher Inc., Waltham, MA, USA) was performed to analyze the chemical components and element valance.

2.2. Electrochemical Measurements

The slurry was obtained by grinding the Fe3O4@C composites, binder-polyvinylidene fluoride (PVDF) and conductive additive (Super P) with a weight ratio of 8:1:1 in a mortar with an appropriate Nmethyl-2-pyrrolidone (NMP) on a ball mill. The prepared slurry was coated on the Cu foil, which then was dried at 85 °C for 24 h under vacuum conditions. The button batteries (Li//Fe3O4@C) were assembled in a glove box under the protection of high-purity Ar (O2 and H2O < 0.1 ppm). Lithium metal was the cathode, a celgard2400 membrane was the separator and a solution of 1 M LiPF6 dissolved in ethylene carbonate and diethyl carbonate (1:1 in volume) was the electrolyte. The galvanostatic charge/discharge measurements were performed on a LAND-CT2001A battery test system in the voltage range from 0.1 to 3.0 V (vs. Li/Li+). A CHI660D electrochemical workstation was used for the measurements of cyclic voltammetry (CV) profiles in a voltage range of 0.01 to 3.0 V by a scan rate of 1 mV s−1, and electrochemical impedance spectroscopy (EIS) was used from 100 kHz to 0.1 Hz.

3. Results and Discussion

Figure 1a shows the XRD patterns of the Fe3O4@C nanocomposites. The diffraction peaks at 18.3°, 30.0°, 35.3°, 43.1°, 57.0° and 62.5° match well with the (111), (220), (311), (400), (333) and (440) crystal plane diffractions of Fe3O4, respectively. This result indicated that Fe3O4 crystalline was formed by sintering the PA-milled C10H10Fe at 773 and 873 K for 30 min. The diffraction intensities of Fe3O4 in Fe3O4@C-873 were comparatively higher than these of Fe3O4@C-773, which could result from the better crystallinity due to the higher annealing temperature for Fe3O4@C-873. On the other hand, the (002) peak of carbon (JCPDS no. 89-7213) was confirmed in Fe3O4@C-873. However, there was no observed carbon diffraction peak in Fe3O4@C-773, which could be attributed to its amorphous structure. Figure 1b shows the TG results for two samples. From Figure 1b, it could be seen that the major weight loss, induced by the removement of carbon due to oxidation, occurred at approximately 750 K. The weight loss starting temperature of Fe3O4@C-773 was obviously lower than that of Fe3O4@C-873. This could be because the amorphous carbon was more easily oxidated in the air condition. The weight percent of carbon was approximately 29 wt.% and 8 wt.% for Fe3O4@C-773 and Fe3O4@C-873, respectively, based on the TG results. It implied that the higher annealing temperature would result in more weight loss in carbon.
Figure 2 shows the morphology and microstructure of the Fe3O4@C nanocomposites. Fe3O4@C-773 and Fe3O4@C-873 showed completely different morphologies. As observed in Figure 2a,b, Fe3O4@C-773 showed a leaves-like nanosheet structure with the Fe3O4 nanoparticles (~5 nm) dispersively embedded in the carbon leaves. The insert of Figure 2b is the STEM-selected area diffraction pattern, which contained an amorphous diffraction ring of carbon together with the (111) and (220) facet families of the polycrystalline Fe3O4. These agreed well with the above X-ray diffraction analysis results. Figure 2c is the enlarged image of the marked part in Figure 2a, and the related mapping for C, Fe, and O are illustrated in Figure 2d–f. The mapping results further confirmed that the ultrafine Fe3O4 nanoparticles were dispersively incorporated into the carbon matrix. Different from Fe3O4@C-773, Fe3O4@C-873 was characterized by a nanotube/nanorod structure and coated with larger Fe3O4 nanoparticles (~100 nm), as seen in Figure 2g (SEM) and Figure 2h (STEM). From the insert of Figure 2h, it could be clearly seen that the Fe3O4 nanoparticles were wrapped by a nanolayer of carbon (~5 nm) with a typical core-shell nanostructure. The core-shell structure could be more favorable for the release of strain caused by Li/Li+ insertion/extraction, and the carbon matrix could serve as a buffer for the accommodation of stress resulting from the sharp volume variation in Fe3O4, thereby avoiding the severe pulverization of Fe3O4 [46]. It was regarded that this novel structure greatly benefitted the enhancement of the structural stability of nano-Fe3O4.
Figure 3a illustrates the Raman spectroscopies of the Fe3O4@C nanocomposites. Fe3O4@C-773 and Fe3O4@C-873 exhibited two bands, centering at approximately 1343 cm−1 and 1589 cm−1, attributed to the D-band and G-band, respectively [47]. According to the references [17,48], the D-band was the characteristic peak of the disordered amorphous structure of carbon, and the G-band was attributed to the sp2 hybridization of carbon materials. Observably, the intensity of the D-band decreased with the increased annealing temperature of the Fe3O4@C composites. For Fe3O4@C-773, the larger value of ID/IG (>1) implied a higher density of defects and structural disorder in the amorphous carbon [49]. On the contrary, the sp2 hybridization was enhanced accompanied by the reduction of defects and structural disorder for Fe3O4@C-873 (ID/IG < 1). This confirmed that the carbon defects and structural disorder were eliminated partially by the crystallization of the amorphous carbon. The defects and structural disorder would lower the electron transfer efficiency, although they benefited from the infiltration of the electrolyte. For further investigation of the microstructure, the isothermal N2-adsorption/desorption at 77 K was carried out to evaluate the BET surface area and pore structure. For Fe3O4@C-873, the BET surface area was approximately 51.4 m2/g (Figure 3b), and there were abundant mesoporous structures with a main pore size of approximately 3 nm, together with some that were 10 nm, as illustrated in the insert of Figure 3b, which could provide active sites on the surface for the adsorption of the electrolyte [50]. However, the BET surface area of Fe3O4@C-773 (Figure 3c) was only 22.4 m2/g, far lower than the value of Fe3O4@C-873, and there was nearly no pore structure, possibly due to its leaves-like nanosheet structure. Certainly, the large surface and pore structure is more favorable to the enhancement of electrolyte entrance and infiltration and, especially, beneficial for Li/Li+ diffusion [51].
The surface element valence and chemical composition of the Fe3O4@C nanocomposites were revealed by XPS analysis. Figure 4 shows the XPS of the Fe3O4@C nanocomposites. Figure 4a shows the full XPS spectrum of Fe3O4@C-873, in which the existence of Fe, C and O species was verified by the Fe 2p, C 1s and O 1s characteristic peaks. In the fine spectra of Fe 2p (Figure 4b), the peaks centered at approximately 711 and 725 eV, attributed to the characteristic peaks of Fe 2p3/2 and Fe 2p1/2, which corresponded to Fe3+ and Fe2+, respectively [15,52]. This confirmed the existence of Fe3O4. The fine spectrum of C 1s contained four different peaks at 289.4, 287.0, 285.2 and 284.7 eV, corresponding to O-C=O, C=O, C-O, and C-C/C=C bonds, respectively [15,47,48], as seen in Figure 4c. The weak peaks derived from O-C=O, C=O and C-O bonds could be caused by the interactions of the C and O atoms located in the interface between Fe3O4 and carbon, which could be further confirmed by the fine spectrum of O 1s. In the high-resolution XPS spectra of O 1s (Figure 4d), the peak at 530.5 eV was attributed to the Fe-O bond in iron oxide, the other two peaks at 533.9 eV (O-C=O) and 532.0 eV (O-C/O=C) could be ascribed to the chemical bonds between carbon atoms and oxygen-containing groups on the surface of iron oxide [17,53,54]. The emergence of chemical interactions between C and Fe3O4 would facilitate the electron transfer between these two phases. For Fe3O4@C-773, the XPS analysis results are illustrated in Figure 4e–h. It was similar to Fe3O4@C-873, implying they had similar chemical composition and element valence states. Overall, the XPS further confirmed the preparation of Fe3O4@C nanocomposites and the existence of chemical interactions between C and Fe3O4.
Figure 5a shows the CV profiles for Fe3O4@C-873. Three cathodic peaks (1.63, 0.90 and 0.54 V) were observed in the lithiation process. The peak at 1.63 V could be attributed to the reaction of Li inserting into the lattice of Fe3O4 crystalline, which was associated with the reaction [15]:
Fe3O4 + xLi+ + xe = Lix(Fe3O4) (0 ≤ x ≤ 2)
The peak located at 0.90 V could be corresponding to the reduction of Fe3+ and Fe2+ to Fe0 and the formation of Li2O based on the reaction equation [55,56]:
Lix(Fe3O4) + (8 − x)Li+ + (8 − x)e = 3Fe + 4Li2O (0 ≤ x ≤ 2)
The peak of 0.54 V, which appeared only at the first cycle, was ascribed to the formation of the solid electrolyte interface (SEI) film that was stable in the following charge and discharge procedure [15]. For delithiation, the reduced Fe0 was gradually oxidized to Fe2+ and Fe3+, accompanied by the formation of Lix(Fe3O4) by two anodic steps at 1.70 and 1.94 V, respectively [55,56,57,58]. After the first cycle, the CV curves and the position of the peaks almost overlapped in the second and third cycles, implying good charge and discharge reversibility for Fe3O4@C-873 as a LIBs anode. For Fe3O4@C-773 (Figure 5c), its CV curves were similar to those of Fe3O4@C-873. However, the cathodic peaks slightly shifted toward a higher voltage, indicating better kinetics, which was possibly related to the ultrafine Fe3O4 nanoparticles. Figure 5b depicts the discharge and charge curves of Fe3O4@C-873 at the current densities of 0.2, 0.5, 1.0, 2.0 and 5.0 A g−1, respectively. For the profile of 0.2 A g−1, there was one plateau at approximately 0.9 V in the discharge process, while there was one plateau containing two transitions at approximately 1.5 and 2.0 V in the charging process, which was consistent with two reaction peaks in the CV profiles; the two transitions corresponded to the oxidation of Fe to Fe2+ and Fe3+, respectively. The initial discharge capacity and charge capacity were approximately 986 and 760 mAh g−1 at the current density of 0.2 A g−1. The irreversible capacity loss could mainly result from the formation of SEI, the side reaction between the electrode and the electrolyte and the decomposition of the electrolyte [15]. Certainly, the reversible capacity reduced with the increase in discharge/charge current density, but it remained at a relatively high reversible capacity of 576 mAh g−1 at 5.0 A g−1. Unfortunately the discharge capacity and charge capacity at 0.2 A g−1 were only 587 and 553 mAh g−1, and it was further reduced to 291 mAh g−1 when discharging at 5.0 A g−1 for Fe3O4@C-773 (Figure 5d). That is to say, Fe3O4@C-873 had a higher capacity and superior rate performance compared to Fe3O4@C-773.
The Li+ storage performance of the Fe3O4@C electrodes was further systematically evaluated by the extended galvanostatic discharge and charge measurements, as depicted in Figure 6. Figure 6a–d show the cycling stability of Fe3O4@C-873 at different current densities. As observed in Figure 6a, the discharge capacity reached 1006 mAh g−1 (efficiency: ~98%) from approximately 781 mAh g−1 (efficiency: ~92%) after 100 cycles at 0.2 A g−1. The maximum capacity was approximately 921 mAh g−1 (efficiency: ~98%) at the 94th cycle and then gradually attenuated to 552 mAh g−1 (efficiency: ~100%) after 300 cycles at 0.5 A g−1 (Figure 6b). When the current density further increased to 1 A g−1, it still had a reversible capacity of 423 mAh g−1 and 100% efficiency after 500 cycles (Figure 6c). For Fe3O4@C-773 (Figure 6e–h), the discharge capacity increased to 662 mAh g−1 (efficiency: ~98%) from 525 mA h g−1 (efficiency: ~92%) at the 45th cycle and then returned to 526 mAh g−1 (efficiency: ~98%) at the 100th cycle under a current density of 0.2 A g−1 (Figure 6e). The maximum capacity was approximately 548 mAh g−1 (efficiency: ~97%) at the 50th cycle and then reduced to 326 mAh g−1 (efficiency: ~100%) after 300 cycles at 0.5 A g−1 (Figure 6f). When the discharge/charge current density increased to 1 A g−1, the reversible capacity attenuated to 258 mAh g−1 (efficiency: ~100%) after continuous 500 cycles (Figure 6g). According to ref. [59], the capacity of pure Fe3O4 was only 126.8 mAh g−1 after 100 cycles at 0.1 A g−1. Significantly, the electrochemical properties and cycling stability of Fe3O4 were effectively improved by designing the unique structure Fe3O4@C nanocomposites. Figure 6d illustrates the rate performance of Fe3O4@C-873, which exhibited no capacity decay after 70 discharge and charge cycles in the current density range from 0.2 to 5 A g−1. For example, the discharge capacity was 772 mAh g−1 when undergoing 10 cycles at 0.2 A g−1. It was approximately 580 mAh g−1 after 50 cycles and the current density increased to 5 A g−1; however, it could reach 978 mAh g−1 at the 70th cycle when the current density reduced to 0.2 A g−1 again. For Fe3O4@C-773 (Figure 6h), it had a similar rate performance; the corresponding capacity was approximately 531, 300 and 655 mAh g−1, respectively. From the experiment results, we could see that two samples, especially Fe3O4@C-873, exhibited fairly good cycle stability and excellent rate performance. Certainly, it should be pointed out that Fe3O4@C-873 had a significantly higher capacity and better rate performance than Fe3O4@C-773. However, Fe3O4@C-773 more easily reached the maximum capacity with better kinetics (~50 cycles, less than ~94 cycles for Fe3O4@C-873), which could result from its large amount of defects and the superfine Fe3O4 nanoparticles, together with the leaves-like amorphous carbon structure, by enhancing the electrolyte absorption and shortening the Li/Li+ diffusion distance. For Fe3O4@C-873, the superior cycling stability and the outstanding high-rate performance were mainly attributed to the improved structural stability of nano-Fe3O4 for the robust core-shell structure supported by the nanotube/nanorod crystalline carbon.
For further understanding of the mechanism of the significantly enhanced electrochemical performances, the electrochemical impedance measurements were carried out, and the results are shown in Figure 7a. As depicted in Figure 7b, a corresponding equivalent circuit model was established to analyze the electrochemical impedance spectroscopy. In this model, Rs is the internal resistance of the analog battery (including the electrolyte impedance and the electrode impedance), Rct represents the charge-transfer resistance, CPE1 is related to the constant phase element, Zw is the Warburg resistance associated with the Li/Li+ diffusion in the active materials [55,58,59,60,61]. From Figure 7a, it could be observed that the electrochemical impedance spectroscopies for two samples contained only one observable semicircle in the test full frequency range. It was attributed to the interfacial charge transfer (Rct) combined with the double-layer capacitance CPE1. However, Fe3O4@C-873 contained a much smaller semicircle than Fe3O4@C-773. The inclined lines in the low-frequency region were derived from the Li+ diffusion into the Fe3O4@C materials. The impedance parameters were achieved by fitting the experimental Nyquist plots (Figure 7a) using the established model in Figure 7b. For Fe3O4@C-773, the impedance parameters Rs and Rct were achieved to be approximately 1.2 and 252.1 Ω, respectively. They were 4.0 and 103.0 Ω for Fe3O4@C-873. The results indicated that the Rs of Fe3O4@C-773 and Fe3O4@C-873 were very close, but the Rct of Fe3O4@C-773 was far higher than that of Fe3O4@C-873, indicating Fe3O4@C-873 had better charge transfer property. It could be attributed to the better structural integrity, such as fewer defects and good crystallinity for Fe3O4@C-873. The results suggested that the novel structure Fe3O4@C-873 nanocomposite had significantly improved electrochemical performances compared with Fe3O4.

4. Conclusions

In summary, leaves-like amorphous carbon-nanosheet-encapsulated ultrafine Fe3O4 nanoparticle composites and carbon nanotube/nanorod-supported core-shell structure Fe3O4 nanoparticle composites were fabricated via an in situ encapsulation route, which consists of PA milling the ferrocene (C10H10Fe) precursor and subsequent heating treatment. The carbon nanotube/nanorod-supported core-shell structure Fe3O4 nanocomposite has a larger specific area (51.4 m2/g), fewer defects and better crystallinity than the leaves-like amorphous carbon-nanosheet-encapsulated ultrafine Fe3O4 nanocomposite (22.4 m2/g), and the former also contains abundant mesoporous structures. The carbon nanotube/nanorod-supported core-shell structure Fe3O4 nanocomposite exhibits outstanding cycling stability and high-rate capability as anode materials for LIBs. Remarkably, it shows 1006 mAh g−1 and retains 423 mAh g−1 after 500 cycles at 1A g−1. The improved electrochemical performances could be mainly attributed to the synergetic effects of the hierarchical architecture and the nanostructure. This work supplies a feasible route for the preparation of novel structure Fe3O4@C nanocomposites as high-performance electrode materials.

Author Contributions

H.Z. and H.L. wrote the main manuscript; W.H., Y.W. and X.Y. performed equal work in preparing the samples and taking the electrochemical measurements; C.J. prepared Figure 5, Figure 6 and Figure 7; W.Z. performed the XRD analysis; C.L. performed the XPS analysis; L.D. performed the microstructure characterizations (SEM and TEM). X.X. revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Natural Science Foundation of Fujian Province, China (No. 2021J011209); the Open Fund of Fujian Provincial Key Laboratory of Functional Materials and Applications, Xiamen University of Technology, (fma2020003 and fma2022007).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, Y.; Zhu, Y.; Cui, Y. Challenges and Opportunities towards Fast-Charging Battery Materials. Nat. Energy 2019, 4, 540–550. [Google Scholar] [CrossRef]
  2. Wu, F.; Maier, J.; Yu, Y. Guidelines and Trends for Next-Generation Rechargeable Lithium and Lithium-Ion Batteries. Chem. Soc. Rev. 2020, 49, 1569–1614. [Google Scholar] [CrossRef]
  3. Shi, Y.; Zhang, J.; Bruck, A.M.; Zhang, Y.; Li, J.; Stach, E.A.; Takeuchi, K.J.; Marschilok, A.C.; Takeuchi, E.S.; Yu, G. A Tunable 3D Nanostructured Conductive Gel Framework Electrode for High-Performance Lithium Ion Batteries. Adv. Mater. 2017, 29, 1603922. [Google Scholar] [CrossRef]
  4. Goodenough, J.B. Evolution of Strategies for Modern Rechargeable Batteries. Acc. Chem. Res. 2013, 46, 1053–1061. [Google Scholar] [CrossRef]
  5. Zheng, H.; Xu, H.-S.; Hu, J.; Liu, H.; Wei, L.; Wu, S.; Huang, Y.; Li, J.; Tang, K. Construction of Hierarchical MnSe@SnSe2@N–C Nanorods for High-Performance Lithium-Ion Batteries. ACS Appl. Energy Mater. 2022, 5, 6586–6596. [Google Scholar] [CrossRef]
  6. Liu, J.; Yu, L.; Wu, C.; Wen, Y.; Yin, K.; Chiang, F.-K.; Hu, R.; Liu, J.; Sun, L.; Gu, L.; et al. New Nanoconfined Galvanic Replacement Synthesis of Hollow Sb@C Yolk–Shell Spheres Constituting a Stable Anode for High-Rate Li/Na-Ion Batteries. Nano Lett. 2017, 17, 2034–2042. [Google Scholar] [CrossRef]
  7. Li, X.; Dhanabalan, A.; Gu, L.; Wang, C. Three-Dimensional Porous Core-Shell Sn@Carbon Composite Anodes for High-Performance Lithium-Ion Battery Applications. Adv. Energy Mater. 2012, 2, 238–244. [Google Scholar] [CrossRef]
  8. Yu, L.; Liu, J.; Xu, X.; Zhang, L.; Hu, R.; Liu, J.; Yang, L.; Zhu, M. Metal–Organic Framework-Derived NiSb Alloy Embedded in Carbon Hollow Spheres as Superior Lithium-Ion Battery Anodes. ACS Appl. Mater. Interfaces 2017, 9, 2516–2525. [Google Scholar] [CrossRef]
  9. Zhang, H.; Hu, R.; Liu, H.; Sun, W.; Lu, Z.; Liu, J.; Yang, L.; Zhang, Y.; Zhu, M. A Spherical Sn–Fe3O4@graphite Composite as a Long-Life and High-Rate-Capability Anode for Lithium Ion Batteries. J. Mater. Chem. A 2016, 4, 10321–10328. [Google Scholar] [CrossRef]
  10. Hao, S.; Zhang, B.; Wang, Y.; Li, C.; Feng, J.; Ball, S.; Srinivasan, M.; Wu, J.; Huang, Y. Hierarchical Three-Dimensional Fe3O4@porous Carbon Matrix/Graphene Anodes for High Performance Lithium Ion Batteries. Electrochim. Acta 2018, 260, 965–973. [Google Scholar] [CrossRef]
  11. Marzuki, N.S.; Taib, N.U.; Hassan, M.F.; Idris, N.H. Enhanced Lithium Storage in Co3O4/Carbon Anode for Li-Ion Batteries. Electrochim. Acta 2015, 182, 452–457. [Google Scholar] [CrossRef]
  12. Kim, W.-S.; Hwa, Y.; Kim, H.-C.; Choi, J.-H.; Sohn, H.-J.; Hong, S.-H. SnO2@Co3O4 Hollow Nano-Spheres for a Li-Ion Battery Anode with Extraordinary Performance. Nano Res. 2014, 7, 1128–1136. [Google Scholar] [CrossRef]
  13. Zhao, Y.; Liu, C.; Yi, R.; Li, Z.; Chen, Y.; Li, Y.; Mitrovic, I.; Taylor, S.; Chalker, P.; Yang, L.; et al. Facile Preparation of Co3O4 Nanoparticles Incorporating with Highly Conductive MXene Nanosheets as High-Performance Anodes for Lithium-Ion Batteries. Electrochim. Acta 2020, 345, 136203. [Google Scholar] [CrossRef]
  14. Zhao, M.-Q.; Torelli, M.; Ren, C.E.; Ghidiu, M.; Ling, Z.; Anasori, B.; Barsoum, M.W.; Gogotsi, Y. 2D Titanium Carbide and Transition Metal Oxides Hybrid Electrodes for Li-Ion Storage. Nano Energy 2016, 30, 603–613. [Google Scholar] [CrossRef]
  15. Li, L.; Kovalchuk, A.; Fei, H.; Peng, Z.; Li, Y.; Kim, N.D.; Xiang, C.; Yang, Y.; Ruan, G.; Tour, J.M. Enhanced Cycling Stability of Lithium-Ion Batteries Using Graphene-Wrapped Fe3O4-Graphene Nanoribbons as Anode Materials. Adv. Energy Mater. 2015, 5, 1500171. [Google Scholar] [CrossRef]
  16. Ma, F.-X.; Hu, H.; Wu, H.B.; Xu, C.-Y.; Xu, Z.; Zhen, L.; David Lou, X.W. Formation of Uniform Fe3O4 Hollow Spheres Organized by Ultrathin Nanosheets and Their Excellent Lithium Storage Properties. Adv. Mater. 2015, 27, 4097–4101. [Google Scholar] [CrossRef]
  17. Li, Z.; Hu, X.; Li, B.; Wang, X.; Shi, Z.; Lu, J.; Wang, Z. MOF-Derived Fe3O4 Hierarchical Nanocomposites Encapsulated by Carbon Shells as High-Performance Anodes for Li-Storage Systems. J. Alloys Compd. 2021, 875, 159906. [Google Scholar] [CrossRef]
  18. Hu, X.; Ma, M.; Zeng, M.; Sun, Y.; Chen, L.; Xue, Y.; Zhang, T.; Ai, X.; Mendes, R.G.; Rümmeli, M.H.; et al. Supercritical Carbon Dioxide Anchored Fe3O4 Nanoparticles on Graphene Foam and Lithium Battery Performance. ACS Appl. Mater. Interfaces 2014, 6, 22527–22533. [Google Scholar] [CrossRef]
  19. Bulut Kopuklu, B.; Tasdemir, A.; Alkan Gursel, S.; Yurum, 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]
  20. Li, L.; Wang, H.; Xie, Z.; An, C.; Jiang, G.; Wang, Y. 3D Graphene-Encapsulated Nearly Monodisperse Fe3O4 Nanoparticles as High-Performance Lithium-Ion Battery Anodes. J. Alloys Compd. 2020, 815, 152337. [Google Scholar] [CrossRef]
  21. Mao, J.; Niu, D.; Zheng, N.; Jiang, G.; Zhao, W.; Shi, J.; Li, Y. Fe3O4 -Embedded and N-Doped Hierarchically Porous Carbon Nanospheres as High-Performance Lithium Ion Battery Anodes. ACS Sustain. Chem. Eng. 2019, 7, 3424–3433. [Google Scholar] [CrossRef]
  22. Yao, J.; Zhang, K.; Wang, W.; Zuo, X.; Yang, Q.; Tang, H.; Wu, M.; Li, G. Remarkable Enhancement in the Photoelectric Performance of Uniform Flower-like Mesoporous Fe3O4 Wrapped in Nitrogen-Doped Graphene Networks. ACS Appl. Mater. Interfaces 2018, 10, 19564–19572. [Google Scholar] [CrossRef]
  23. Guo, Y.; Zhang, D.; Yang, Y.; Wang, Y.; Bai, Z.; Chu, P.K.; Luo, Y. MXene-Encapsulated Hollow Fe3O4 Nanochains Embedded in N-Doped Carbon Nanofibers with Dual Electronic Pathways as Flexible Anodes for High-Performance Li-Ion Batteries. Nanoscale 2021, 13, 4624–4633. [Google Scholar] [CrossRef]
  24. Liu, J.; Xu, X.; Hu, R.; Yang, L.; Zhu, M. Uniform Hierarchical Fe3O4 @Polypyrrole Nanocages for Superior Lithium Ion Battery Anodes. Adv. Energy Mater. 2016, 6, 1600256. [Google Scholar] [CrossRef]
  25. Li, H.; Wang, J.; Li, Y.; Zhao, Y.; Tian, Y.; Kurmanbayeva, I.; Bakenov, Z. Hierarchical Sandwiched Fe3O4@C/Graphene Composite as Anode Material for Lithium-Ion Batteries. J. Electroanal. Chem. 2019, 847, 113240. [Google Scholar] [CrossRef]
  26. Yan, Z.; Jiang, X.; Dai, Y.; Xiao, W.; Li, X.; Du, N.; He, G. Pulverization Control by Confining Fe3O4 Nanoparticles Individually into Macropores of Hollow Carbon Spheres for High-Performance Li-Ion Batteries. ACS Appl. Mater. Interfaces 2018, 10, 2581–2590. [Google Scholar] [CrossRef]
  27. Li, M.; Du, H.; Kuai, L.; Huang, K.; Xia, Y.; Geng, B. Scalable Dry Production Process of a Superior 3D Net-Like Carbon-Based Iron Oxide Anode Material for Lithium-Ion Batteries. Angew. Chem. Int. Ed. 2017, 56, 12649–12653. [Google Scholar] [CrossRef]
  28. Park, G.D.; Hong, J.H.; Jung, D.S.; Lee, J.-H.; Kang, Y.C. Unique Structured Microspheres with Multishells Comprising Graphitic Carbon-Coated Fe3O4 Hollow Nanopowders as Anode Materials for High-Performance Li-Ion Batteries. J. Mater. Chem. A 2019, 7, 15766–15773. [Google Scholar] [CrossRef]
  29. Wang, L.; Liu, F.; Pal, A.; Ning, Y.; Wang, Z.; Zhao, B.; Bradley, R.; Wu, W. Ultra-Small Fe3O4 Nanoparticles Encapsulated in Hollow Porous Carbon Nanocapsules for High Performance Supercapacitors. Carbon 2021, 179, 327–336. [Google Scholar] [CrossRef]
  30. Ding, R.; Zhang, J.; Qi, J.; Li, Z.; Wang, C.; Chen, M. N-Doped Dual Carbon-Confined 3D Architecture RGO/Fe3O4/AC Nanocomposite for High-Performance Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2018, 10, 13470–13478. [Google Scholar] [CrossRef]
  31. Wei, W.; Yang, S.; Zhou, H.; Lieberwirth, I.; Feng, X.; Müllen, K. 3D Graphene Foams Cross-Linked with Pre-Encapsulated Fe3O4 Nanospheres for Enhanced Lithium Storage. Adv. Mater. 2013, 25, 2909–2914. [Google Scholar] [CrossRef]
  32. Liu, H.; Jia, M.; Zhu, Q.; Cao, B.; Chen, R.; Wang, Y.; Wu, F.; Xu, B. 3D-0D Graphene-Fe3O4 Quantum Dot Hybrids as High-Performance Anode Materials for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2016, 8, 26878–26885. [Google Scholar] [CrossRef]
  33. Qin, X.; Zhang, H.; Wu, J.; Chu, X.; He, Y.-B.; Han, C.; Miao, C.; Wang, S.; Li, B.; Kang, F. Fe3O4 Nanoparticles Encapsulated in Electrospun Porous Carbon Fibers with a Compact Shell as High-Performance Anode for Lithium Ion Batteries. Carbon 2015, 87, 347–356. [Google Scholar] [CrossRef]
  34. Duan, Y.; Zhang, B.; Zheng, J.; Hu, J.; Wen, J.; Miller, D.J.; Yan, P.; Liu, T.; Guo, H.; Li, W.; et al. Excess Li-Ion Storage on Reconstructed Surfaces of Nanocrystals To Boost Battery Performance. Nano Lett. 2017, 17, 6018–6026. [Google Scholar] [CrossRef]
  35. Jiao, J.; Qiu, W.; Tang, J.; Chen, L.; Jing, L. Synthesis of Well-Defined Fe3O4 Nanorods/N-Doped Graphene for Lithium-Ion Batteries. Nano Res. 2016, 9, 1256–1266. [Google Scholar] [CrossRef]
  36. Xiao, Z.; Xia, Y.; Ren, Z.; Liu, Z.; Xu, G.; Chao, C.; Li, X.; Shen, G.; Han, G. Facile Synthesis of Single-Crystalline Mesoporous α-Fe2O3 and Fe3O4 Nanorods as Anode Materials for Lithium-Ion Batteries. J. Mater. Chem. 2012, 22, 20566. [Google Scholar] [CrossRef]
  37. Kwon, Y.H.; Park, J.J.; Housel, L.M.; Minnici, K.; Zhang, G.; Lee, S.R.; Lee, S.W.; Chen, Z.; Noda, S.; Takeuchi, E.S.; et al. Carbon Nanotube Web with Carboxylated Polythiophene “Assist” for High-Performance Battery Electrodes. ACS Nano 2018, 12, 3126–3139. [Google Scholar] [CrossRef]
  38. Gao, T.; Xu, C.; Li, R.; Zhang, R.; Wang, B.; Jiang, X.; Hu, M.; Bando, Y.; Kong, D.; Dai, P.; et al. Biomass-Derived Carbon Paper to Sandwich Magnetite Anode for Long-Life Li-Ion Battery. ACS Nano 2019, 13, 11901–11911. [Google Scholar] [CrossRef]
  39. Wang, Y.; Chen, L.; Liu, H.; Xiong, Z.; Zhao, L.; Liu, S.; Huang, C.; Zhao, Y. Cornlike Ordered N-Doped Carbon Coated Hollow Fe3O4 by Magnetic Self-Assembly for the Application of Li-Ion Battery. Chem. Eng. J. 2019, 356, 746–755. [Google Scholar] [CrossRef]
  40. Wang, B.; Zhang, X.; Liu, X.; Wang, G.; Wang, H.; Bai, J. Rational Design of Fe3O4@C Yolk-Shell Nanorods Constituting a Stable Anode for High-Performance Li/Na-Ion Batteries. J. Colloid Interface Sci. 2018, 528, 225–236. [Google Scholar] [CrossRef]
  41. Xu, X.; Liu, J.; Liu, Z.; Shen, J.; Hu, R.; Liu, J.; Ouyang, L.; Zhang, L.; Zhu, M. Robust Pitaya-Structured Pyrite as High Energy Density Cathode for High-Rate Lithium Batteries. ACS Nano 2017, 11, 9033–9040. [Google Scholar] [CrossRef]
  42. Liu, Z.; Yu, X.-Y.; Paik, U. Etching-in-a-Box: A Novel Strategy to Synthesize Unique Yolk-Shelled Fe3O4 @Carbon with an Ultralong Cycling Life for Lithium Storage. Adv. Energy Mater. 2016, 6, 1502318. [Google Scholar] [CrossRef]
  43. Guo, W.; Sun, W.; Lv, L.-P.; Kong, S.; Wang, Y. Microwave-Assisted Morphology Evolution of Fe-Based Metal–Organic Frameworks and Their Derived Fe2O3 Nanostructures for Li-Ion Storage. ACS Nano 2017, 11, 4198–4205. [Google Scholar] [CrossRef]
  44. Du, J.; Tang, Y.; Wang, Y.; Shi, P.; Fan, J.; Xu, Q.; Min, Y. A MOF-Derived Method to Construct Well-Arranged Porous Nanosheets for Lithium Ion Batteries. Dalton Trans. 2018, 47, 7571–7577. [Google Scholar] [CrossRef]
  45. Dai, R.; Sun, W.; Wang, Y. Ultrasmall Tin Nanodots Embedded in Nitrogen-Doped Mesoporous Carbon: Metal-Organic-Framework Derivation and Electrochemical Application as Highly Stable Anode for Lithium Ion Batteries. Electrochim. Acta 2016, 217, 123–131. [Google Scholar] [CrossRef]
  46. Wu, L.; Liu, Y.; Zhao, H.; Wang, Z.; Zhu, B.; Zhang, X.; He, P.; Liu, Y.; Yang, T. MOF-Derived Long Spindle-like Carbon-Coated Ternary Transition-Metal-Oxide Composite for Lithium Storage. ACS Omega 2022, 7, 16837–16846. [Google Scholar] [CrossRef]
  47. Ferrari, A.C.; Robertson, J. Raman spectroscopy in carbons: From nanotubes to diamond-Preface. Philos. Trans. R. Soc. Lond. Ser. Math. Phys. Eng. Sci. 2004, 362, 2269–2270. [Google Scholar] [CrossRef]
  48. Lv, L.-P.; Du, P.; Liu, P.; Li, X.; Wang, Y. Integrating Mixed Metallic Selenides/Nitrogen-Doped Carbon Heterostructures in One-Dimensional Carbon Fibers for Efficient Oxygen Reduction Electrocatalysis. ACS Sustain. Chem. Eng. 2020, 8, 8391–8401. [Google Scholar] [CrossRef]
  49. Duan, H.; Zhang, S.; Chen, Z.; Xu, A.; Zhan, S.; Wu, S. Self-Formed Channel Boosts Ultrafast Lithium Ion Storage in Fe3O4@Nitrogen-Doped Carbon Nanocapsule. ACS Appl. Mater. Interfaces 2020, 12, 527–537. [Google Scholar] [CrossRef]
  50. Xie, Y.; Qiu, Y.; Tian, L.; Liu, T.; Su, X. Ultrafine Hollow Fe3O4 Anode Material Modified with Reduced Graphene Oxides for High-Power Lithium-Ion Batteries. J. Alloys Compd. 2022, 894, 162384. [Google Scholar] [CrossRef]
  51. Pan, Q.; Ding, Y.; Yan, Z.; Cai, Y.; Zheng, F.; Huang, Y.; Wang, H.; Li, Q. Designed Synthesis of Fe3O4@NC Yolk-Shell Hollow Spheres as High Performance Anode Material for Lithium-Ion Batteries. J. Alloys Compd. 2020, 821, 153569. [Google Scholar] [CrossRef]
  52. Yuan, M.; Zhao, Y.; Sun, Z.; Bakenov, Z. Rational Design of MOFs-Derived Fe3O4@C Interwoven with Carbon Nanotubes as Sulfur Host for Advanced Lithium-sulfur Batteries. J. Electroanal. Chem. 2020, 877, 114608. [Google Scholar] [CrossRef]
  53. Kumar, R.; Singh, R.K.; Vaz, A.R.; Savu, R.; Moshkalev, S.A. Self-Assembled and One-Step Synthesis of Interconnected 3D Network of Fe3O4/Reduced Graphene Oxide Nanosheets Hybrid for High-Performance Supercapacitor Electrode. ACS Appl. Mater. Interfaces 2017, 9, 8880–8890. [Google Scholar] [CrossRef]
  54. Yi, Q.; Du, M.; Shen, B.; Ji, J.; Dong, C.; Xing, M.; Zhang, J. Hollow Fe3O4/Carbon with Surface Mesopores Derived from MOFs for Enhanced Lithium Storage Performance. Sci. Bull. 2020, 65, 233–242. [Google Scholar] [CrossRef]
  55. Ma, T.; Liu, X.; Sun, L.; Xu, Y.; Zheng, L.; Zhang, J. Strongly Coupled N-Doped Carbon/Fe3O4/N-Doped Carbon Hierarchical Micro/Nanostructures for Enhanced Lithium Storage Performance. J. Energy Chem. 2019, 34, 43–51. [Google Scholar] [CrossRef] [Green Version]
  56. Cong, B.; Hu, Y.; Sun, S.; Wang, Y.; Wang, B.; Kong, H.; Chen, G. Metal–Organic Framework Derived Amorphous VOx Coated Fe3O4/C Hierarchical Nanospindle as Anode Material for Superior Lithium-Ion Batteries. Nanoscale 2020, 12, 16901–16909. [Google Scholar] [CrossRef]
  57. Huang, J.; Dai, Q.; Cui, C.; Ren, H.; Lu, X.; Hong, Y.; Woo Joo, S. Cake-like Porous Fe3O4@C Nanocomposite as High-Performance Anode for Li-Ion Battery. J. Electroanal. Chem. 2022, 918, 116508. [Google Scholar] [CrossRef]
  58. Zhang, S.S.; Xu, K.; Jow, T.R. EIS Study on the Formation of Solid Electrolyte Interface in Li-Ion Battery. Electrochim. Acta 2006, 51, 1636–1640. [Google Scholar] [CrossRef]
  59. Moss, P.L.; Au, G.; Plichta, E.J.; Zheng, J.P. Investigation of Solid Electrolyte Interfacial Layer Development during Continuous Cycling Using Ac Impedance Spectra and Micro-Structural Analysis. J. Power Sources 2009, 189, 66–71. [Google Scholar] [CrossRef]
  60. Kang, L.; Zhang, M.; Zhang, J.; Liu, S.; Zhang, N.; Yao, W.; Ye, Y.; Luo, C.; Gong, Z.; Wang, C. Dual-defect surface engineering of bimetallic sulfide nanotubes towards flexible asymmetric solid-state supercapacitors. J. Mater. Chem. A 2020, 8, 24053–24064. [Google Scholar] [CrossRef]
  61. Liu, S.; Kang, L.; Zhang, J.; Jung, E.; Lee, S.; Jun, S.C. Structural engineering and surface modification of MOF-derived cobalt-based hybrid nanosheets for flexible solid-state supercapacitors. Energy Storage Mater. 2020, 32, 167–177. [Google Scholar] [CrossRef]
Scheme 1. Schematic illustration of the preparation of the Fe3O4@C nanocomposites.
Scheme 1. Schematic illustration of the preparation of the Fe3O4@C nanocomposites.
Batteries 09 00403 sch001
Figure 1. (a) XRD, (b) thermogravimetric analysis cures for Fe3O4@C composite.
Figure 1. (a) XRD, (b) thermogravimetric analysis cures for Fe3O4@C composite.
Batteries 09 00403 g001
Figure 2. (ac) TEM images of Fe3O4@C-773; (df) corresponding EDS element mapping signals of the C, Fe, and O elements. (g,h) TEM images of Fe3O4@C-873.
Figure 2. (ac) TEM images of Fe3O4@C-773; (df) corresponding EDS element mapping signals of the C, Fe, and O elements. (g,h) TEM images of Fe3O4@C-873.
Batteries 09 00403 g002
Figure 3. (a) Raman spectroscopy; (b,c): Isothermal N2-adsorption/desorption curves at 77 K and pore size distribution curve for Fe3O4@C-873 (b) and Fe3O4@C-773 (c), respectively.
Figure 3. (a) Raman spectroscopy; (b,c): Isothermal N2-adsorption/desorption curves at 77 K and pore size distribution curve for Fe3O4@C-873 (b) and Fe3O4@C-773 (c), respectively.
Batteries 09 00403 g003
Figure 4. (a,e) XPS spectrum, (b,f) Fe 2P XPS spectrum, (c,g): C 1s XPS spectrum, (d,h) O 1s XPS spectrum for Fe3O4@C-873 and Fe3O4@C-773, respectively.
Figure 4. (a,e) XPS spectrum, (b,f) Fe 2P XPS spectrum, (c,g): C 1s XPS spectrum, (d,h) O 1s XPS spectrum for Fe3O4@C-873 and Fe3O4@C-773, respectively.
Batteries 09 00403 g004
Figure 5. (a,c) cyclic voltammetry profiles of Fe3O4@C-873 and Fe3O4@C-773 at a scan rate of 0.1 mV s−1; (b,d) voltage capacity curves of Fe3O4@C-873 and Fe3O4@C-773 at different current density (0.2, 0.5, 1, 2 and 5 A g−1).
Figure 5. (a,c) cyclic voltammetry profiles of Fe3O4@C-873 and Fe3O4@C-773 at a scan rate of 0.1 mV s−1; (b,d) voltage capacity curves of Fe3O4@C-873 and Fe3O4@C-773 at different current density (0.2, 0.5, 1, 2 and 5 A g−1).
Batteries 09 00403 g005
Figure 6. Cycle performances of Fe3O4@C-873 (ac) and Fe3O4@C-773 (eg) at various current densities; rate performance of Fe3O4@C-873 (d) and Fe3O4@C-773 (h) at various current rates from 0.2 to 5.0 A g−1 in the potential range of 0.01~3.0 V (vs. Li/Li+).
Figure 6. Cycle performances of Fe3O4@C-873 (ac) and Fe3O4@C-773 (eg) at various current densities; rate performance of Fe3O4@C-873 (d) and Fe3O4@C-773 (h) at various current rates from 0.2 to 5.0 A g−1 in the potential range of 0.01~3.0 V (vs. Li/Li+).
Batteries 09 00403 g006
Figure 7. (a) electrochemical impedance spectra and (b) the corresponding equivalent circuit diagram.
Figure 7. (a) electrochemical impedance spectra and (b) the corresponding equivalent circuit diagram.
Batteries 09 00403 g007
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

Zhong, H.; Huang, W.; Wei, Y.; Yang, X.; Jiang, C.; Liu, H.; Zhang, W.; Liang, C.; Dai, L.; Xu, X. Facile Constructing Hierarchical Fe3O4@C Nanocomposites as Anode for Superior Lithium-Ion Storage. Batteries 2023, 9, 403. https://doi.org/10.3390/batteries9080403

AMA Style

Zhong H, Huang W, Wei Y, Yang X, Jiang C, Liu H, Zhang W, Liang C, Dai L, Xu X. Facile Constructing Hierarchical Fe3O4@C Nanocomposites as Anode for Superior Lithium-Ion Storage. Batteries. 2023; 9(8):403. https://doi.org/10.3390/batteries9080403

Chicago/Turabian Style

Zhong, Haichang, Wenlong Huang, Yukun Wei, Xin Yang, Chunhai Jiang, Hui Liu, Wenxian Zhang, Chu Liang, Leyang Dai, and Xijun Xu. 2023. "Facile Constructing Hierarchical Fe3O4@C Nanocomposites as Anode for Superior Lithium-Ion Storage" Batteries 9, no. 8: 403. https://doi.org/10.3390/batteries9080403

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop