Skip to main content
ACS Omega logoLink to ACS Omega
. 2021 Sep 23;6(39):25269–25276. doi: 10.1021/acsomega.1c03050

High-Capacity Anode Material for Lithium-Ion Batteries with a Core–Shell NiFe2O4/Reduced Graphene Oxide Heterostructure

Chang Liu 1, Tong Zhang 1, Lixin Cao 1, Kun Luo 1,*
PMCID: PMC8495711  PMID: 34632186

Abstract

graphic file with name ao1c03050_0007.jpg

A novel composite consisting of transition-metal oxide and reduced graphene oxide (rGO) has been designed as a highly promising anode material for lithium-ion batteries (LIBs). The anode material for LIBs exhibits high-rate capability, outstanding stability, and nontoxicity. The structural characterization techniques, such as X-ray diffraction, Raman spectra, and transmission electron microscopy, indicate that the material adopts a unique core–shell structure with NiFe2O4 nanoparticles situated in the center and an rGO layer coated on the surface of NiFe2O4 particles (denoted as NiFe2O4/rGO). The NiFe2O4/rGO material with a core–shell structure exhibits an excellent electrochemical performance, which shows a capacity of 1183 mA h g–1 in the first cycle and maintains an average capacity of ∼1150 mA h g–1 after 900 cycles at a current density of 500 mA g–1. This work provides a broad field of vision for the application of transition-metal-oxide materials in electrodes of lithium-ion batteries, which is of great significance for further development of lithium-ion batteries with excellent performance.

1. Introduction

As human beings increasingly rely on various types of energy-driven equipment, researchers have paid more attention to the evolution of energy storage and conversion devices.14 Lithium-ion batteries (LIBs) are one of the most competitive systems because of their high theoretical capacity, security, low cost, nontoxicity, etc.5,6 Although LIBs with a graphite anode that have been commercialized show low operating voltage and outstanding cycling stability, a low theoretical capacity of about 372 mA h g–1 of graphite is far below the demand of large facilities.79 Therefore, exploring a new anode material with higher capacity and better structural stability is a priority.

Transition-metal oxides (TMOs) have been proved to be very promising anode materials because they can deliver reversible capacities almost 3 times higher than the specific capacity of graphite materials, as well as for their low cost.10 However, transition-metal-oxide materials exhibit poor rate capability and drastic capacity fading because of low conductivity and large volume expansion during cycling. Various strategies have often been applied to solve these problems. The particle size of TMOs can be controlled through a series of chemical methods, promoting electrochemical processes and maintaining structural integrity.1114 Carbon materials can also be introduced to fabricate composites, which can buffer the volume expansion of TMOs during the cycling and increase electronic conductivity.1517 Graphene oxide (GO)/MoS3 has been prepared through a facile method that shows the highest specific capacity of 685 mA h g–1 after 1000 cycles at 2 A g–1.18 In addition, materials with nanostructures such as a hollow structure or a core–shell structure through material engineering can achieve excellent electrochemical performances. For example, core–shell porous nanocubic Mn2O3/TiO2, which was synthesized as a high-performance anode for LIBs, achieved a cycling capacity of 936 mA h g–1 after 100 cycles at a charging/discharging rate of 200 mA g–1.19

Among various TMO anode materials, NiFe2O4 has been considered as an appealing electrode material for LIBs because of high theoretical capacity (915 mA h g–1). Previous studies have reported a reversible capacity of ∼900 mA h g–1 exhibited by the NiFe2O4 anode material.20,21 The Coulombic efficiency is as low as 34.8 and 98.5% for the 1st and 30th cycles, respectively.22 However, similar to other TMOs, the NiFe2O4 anode material also exhibits poor electronic conductivity and fast capacity fading during cycling.23 To overcome these insufficiencies, structural and morphological modifications on oxide materials were carried out by researchers.2426 A hollow NiFe2O4 material exhibits a stable discharge capacity of 840–850 mA h g–1 at a current density of 200 mA g–1.24 NiFe2O4 nanofibers exhibit an outstanding capacity of 1000 mA h g–1 after 100 cycles with a high Coulombic efficiency of 100%.25 The sandwich-structured graphene–NiFe2O4–carbon nanocomposite has been prepared via a hydrothermal method, followed by carbon coating, which exhibits an exceptional cycle stability of 1195 mA h g–1 after 200 cycles at 500 mA g–1.26

In this study, we successfully synthesized a NiFe2O4/reduced graphene oxide (rGO) composite with a core–shell structure via a facile chemical process. The as-prepared NiFe2O4 material adopts a particle size of 18–20 nm, which is beneficial for the participation of NiFe2O4 in electrochemical reactions. The rGO network is considered to serve as a conductor of electrons and provide pathways for Li-ion diffusion to improve the electronic and ionic conductivities. More importantly, rGO on the surface of NiFe2O4 could act as a protective layer to form a unique core–shell structure, which can ease the volume changes of NiFe2O4 during cycling. Therefore, as an anode material for LIBs, the NiFe2O4/rGO composite exhibits improved lithium-storage performances. After 900 cycles, the NiFe2O4/rGO composite still exhibits an average capacity of ∼1150 mA h g–1 at 500 mA g–1. The findings provide wide insights into the utilization of transition-metal-oxide materials for electrodes in lithium-ion batteries, which is of vital importance to further developing LIBs with excellent performances.

2. Results and Discussion

The fabrication method of the NiFe2O4/rGO composite utilized in this study is schematically illustrated in Figure 1. Transition-metal nitrate and citric acid were mixed and heated to obtain NiFe2O4. NiFe2O4 and reduced graphene oxide (rGO) were then placed in ethanol for an ultrasonic process before a hydrothermal reaction. The NiFe2O4/rGO composite with a core–shell structure was prepared through a self-assembly treatment using a hydrothermal method.

Figure 1.

Figure 1

Schematic illustration of the synthesis process of NiFe2O4/rGO.

The structural information of the NiFe2O4/rGO material was characterized by X-ray diffraction (XRD). Figure 2a shows the XRD patterns of NiFe2O4 and the NiFe2O4/rGO composite. In the representative XRD pattern, all of the diffraction peaks can be well indexed to the cubic inverse spinel-type NiFe2O4 (JCPDS card No. 54-0964). These characteristic diffraction peaks of NiFe2O4 at 18.428, 35.701, 37.320, 37.630, 43.383, 53.814, 57.293, and 63.021° correspond to the crystal planes of (111), (220), (311), (222), (400), (422), (511), and (440). The formation of a pure compound can be confirmed as no other phases or impurity peaks could be found. In addition, the XRD pattern of the NiFe2O4/rGO composite is similar to that of NiFe2O4 and all of the characteristic peaks of NiFe2O4 are well preserved, indicating that the NiFe2O4 phase was maintained intact in the chemical reaction where the rGO material was deposited on the surface of NiFe2O4. A new broad peak appears at about 26° for the NiFe2O4/rGO composite, which corresponds to the (002) characteristic peak of rGO. The structure of NiFe2O4/rGO was also determined by Raman spectroscopy, as shown in Figure 2b. The Raman spectrum of the rGO exhibits two peaks at 1346 and 1582 cm–1, which represent the D and G modes of amorphous carbon, respectively. The D band corresponds to the presence of defects, and the G band indicates ideal graphitic sp2 carbons.2729 The ratio of D and G band intensities (ID/IG) is considered to indicate the degree of disorder in a carbon structure. As shown in Figure 2b, the ID/IG ratio of NiFe2O4/rGO is 1.10. This result indicates that rGO has a lower degree of graphitization, which is conducive to the intercalation and diffusion of Li+ ions.2931 On comparing the Raman spectra of NiFe2O4 and NiFe2O4/rGO, there are some other peaks recognized for NiFe2O4/rGO in the range of 400–700 cm–1, which were attributed to the spinel structure of NiFe2O4.32,33 The T2g(1) mode at 191 cm–1 corresponds to the translational drive of the tetrahedron, in which the Ni/Fe metal cations are tetrahedrally coordinated with oxygen atoms. At 319 cm–1, the Eg mode is related to the symmetric bending of the oxygen anions with respect to the Ni/Fe metal cations.31 The T2g(2) mode at 478 cm–1 is considered to be the asymmetric stretching of the Fe/Ni and O bond at octahedral sites, while the T2g(3) mode at 546 cm–1 is the result of the asymmetric bending of oxygen anions with tetrahedral and octahedral metal cations.32 The A1g mode at 606 cm–1 belongs to the symmetric stretching of oxygen anions along Ni–O and Fe–O bonds, as shown Figure S1.34,35

Figure 2.

Figure 2

XRD patterns of NiFe2O4/rGO (black line) and NiFe2O4 (red line) (a), and Raman spectra of NiFe2O4/rGO (red line) and rGO (black line) (b).

The X-ray photoelectron spectroscopy (XPS) of NiFe2O4/rGO was carried out to study the chemical states of transition-metal cations, as shown in Figure 3. The Ni 2p spectrum shows two characteristic peaks at 855.1 and 872.5 eV and two satellites peaks at 862.0 and 878.0 eV owing to Ni 2p1/2 and Ni 2p3/2,36 which imply that Ni exists as Ni2+.37 The characteristic peaks of the Fe 2p spectrum at 713.2 and 726.9 eV are attributed to Fe 2p3/2 and Fe 2p1/2, and there are two satellites peaks at 718.2 and 733.1 eV.36,38 These signs confirm the existence of Fe3+ ions in the composite. The results of the XPS spectra correspond well to the observations in the XRD pattern of the composite, indicating the formation of the NiFe2O4 phase. The thermogravimetric analysis (TGA) curve for NiFe2O4/rGO is shown in Figure S2, which is tested in the range of 50–600 °C at the rate of 5 °C/min in air. Graphene burns at a temperature between 400 and 600 °C in air.14 Therefore, the NiFe2O4/rGO nanocomposite is composed of ∼10 wt % rGO and ∼90 wt % NiFe2O4.

Figure 3.

Figure 3

Fe 2p high-resolution XPS spectra of NiFe2O4/rGO (a), and Ni 2p high-resolution XPS spectra of NiFe2O4/rGO (b).

The morphology of the NiFe2O4/rGO composite has been determined by transmission electron microscopy (TEM). Figures S3a–c show the TEM images of NiFe2O4. The NiFe2O4 nanoparticles show an average size of 20 nm. Such a small particle renders more contact area between the anode and the electrolyte, thus improving the participation of NiFe2O4 in electrochemical process. The selected area electron diffraction (SAED) pattern of NiFe2O4 (Figure S3d) shows a ring-like diffraction pattern, revealing a polycrystalline structure of NiFe2O4. Figure 4 shows the HRTEM images of NiFe2O4/rGO at low and high magnifications. Compared with Figure 4c, it is obviously shown in Figure 4d that the NiFe2O4 particles are embedded in rGO and there is a shell with a thickness of ∼5 nm coating on the surface of NiFe2O4 particles, confirming the formation of a core–shell-structured NiFe2O4/rGO. Figure 4e exhibits a ring-like diffraction pattern, which could reveal the polycrystalline structure of NiFe2O4/rGO. As shown in Figure 4f, all small grains show a single-crystal structure, with obvious crystal lattice fringes. The crystal plane spacing is 0.294 nm, corresponding to the (220) plane of spinel NiFe2O4. The element distribution in the core–shell structure of NiFe2O4/rGO is also demonstrated by the energy-dispersive X-ray spectroscopy (EDS), as shown in Figure 4. It can be definitely observed from the distribution of Ni, Fe, O, and C elements that Ni, Fe, and O atoms are uniformly concentrated in the center of the particles, confirming the existence of the NiFe2O4 phase, whereas the C atoms are dominantly situated on the circumference of the particles, which undoubtedly indicates the formation of a heterogeneous core–shell NiFe2O4/rGO structure observed in Figure 4d.

Figure 4.

Figure 4

TEM images of NiFe2O4/rGO (a, b, and d); TEM image of NiFe2O4 (c); SAED pattern of NiFe2O4/rGO (e); HRTEM image of NiFe2O4/rGO (f); from top to bottom: elemental mapping images showing the elemental distribution of Ni, Fe, O, and C and the overlapped image (g).

To explore the electrochemical performances of core–shell NiFe2O4/rGO, a cyclic voltammetry (CV) test was conducted. The CV curves of the initial four cycles are shown in Figure 5a, which has been tested at a scan rate of 0.2 mV s–1 at the potential window of 0.01–3.00 V vs Li/Li+. The cathodic peak of the first cycle started at 0.75 V, which is related to the formation of a solid–electrolyte interface (SEI) film and the reduction of NiFe2O4 during the first discharging process (eq 1).39 The weak wide peak at about 1.5 V may be the result of insertion of Li+ ions into the NiFe2O4/rGO electrode. The two broad anodic peaks at 1.6 and 1.8 V are attributed to the formation of NiO and Fe2O3 according to eqs 2 and 3 in the subsequent anodic scans, respectively.27,40,41 After one cycle, all reduction peaks moved slightly toward a higher potential, implying that there was a loss of capacity because of the formation of the irreversible phase in the first cycle.42 Meanwhile, the CV curves of NiFe2O4/rGO in the second and third cycles almost overlap, indicating the excellent reversibility of the electrochemical reactions.

2. 1
2. 2
2. 3

Figure 5.

Figure 5

(a) Cyclic voltammetry curves of the NiFe2O4/rGO electrode. Scan rate: 0.2 mV s–1; voltage range: 0.01–3.00 V. (b) Galvanostatic charge/discharge profiles of the NiFe2O4/rGO electrode. Voltage range: 0.01–3.00 V; current density: 500 mA g–1. (c) Nyquist plots of the NiFe2O4 (red) and NiFe2O4/rGO (black) electrodes. (d) CV curves of the NiFe2O4/rGO electrode at different scan rates. (e) Rate capabilities of NiFe2O4/rGO. (f) Nyquist plots of NiFe2O4/rGO at the 1st and 20th cycles. (g) Charge/discharge capacity and Coulombic efficiency of the NiFe2O4/rGO electrode at 500 mA g–1.

The charge/discharge curves of NiFe2O4/rGO were obtained between 0.01 and 3.00 V at 500 mA g–1. As shown in Figure 5b, the initial charge–discharge curves exhibit two obvious plateaus at 1.70 and 0.75 V, which correspond well to the above CV results. The discharge plateau was replaced by a slope from 1.00 to 0.80 V in the subsequent cycle, corresponding to the irreversible reaction such as the formation of an SEI film. Furthermore, the charge/discharge platform is well maintained with the same shape after 50 cycles, indicating the outstanding reversibility of electrochemical reaction.

Figure 5c shows electrochemical impedance spectra of NiFe2O4 and NiFe2O4/rGO. All of the Nyquist plots consist of a slope line and a semicircle. The slope line is related to the Li+ diffusion in the fibrous framework (Zw), and the semicircle corresponds to the charge-transfer resistance (Rct).43 The fitted results show that the Rct (54 Ω) of the NiFe2O4/rGO electrode is much lower than that of the NiFe2O4 electrode (154 Ω) and the diffusion coefficient of NiFe2O4/rGO (1.42 × 10–12 cm2 s–1) is higher than that of pure NiFe2O4 (6.64 × 10–13 cm2 s–1), which indicates that the electronic conductivity and ion mobility of the composite can be improved by coating rGO on NiFe2O4. Figure 5g shows the cycling performance of NiFe2O4/rGO at 500 mA g–1. The NiFe2O4/rGO anode shows a first discharge capacity of 1183 mA h g–1, which is much larger than the reversible capacity (∼900 mA h g–1) reported previously.20,21 The excess capacity may be related to the formation of SEI at the electrode/electrolyte interface caused by electrolyte reduction. The exposed vacancy defects of rGO also render more Li storage, resulting in extra capacity contribution.44 In addition, excess oxygen in the anode could also react with some lithium, leading to extra capacity.28,46 The first charge capacity is 842 mA h g–1 with a low Coulombic efficiency of 71.2%. The loss of capacity during charge process can be attributed to the irreversible decomposition of the SEI and the incomplete conversion reaction.46,45

However, there is a drastic capacity fading of NiFe2O4/rGO during the initial 85 cycles due to the practical structural destruction caused by electrolytic grinding during the cycle, which was observed previously in the NiFe2O4-based anode material.28,40 In subsequent cycles, the reversible capacity of the NiFe2O4 electrode gradually increases, which may be attributed to the formation of the SEI film and activation of NiFe2O4/rGO during the lithium insertion/extraction process.46 After 900 cycles, a relatively stable reversible capacity of ∼1150 mA h g–1 could be achieved with the Coulombic efficiency close to 100%, indicating the stabilization of the SEI film between the electrode and the electrolyte.40,46Figure S5 exhibits the cycling performance of NiFe2O4. A stable capacity of 105 mA h g–1 has been observed after a fast capacity fading. Compared to the NiFe2O4 electrode, the NiFe2O4/rGO electrode exhibits improved cycling performance. The enhanced cycling stability of NiFe2O4/rGO could be attributed to the following two reasons: (1) rGO as a shell layer can adapt to the volume change during the cycling process and alleviate the pulverization and aggregation of NiFe2O4; (2) the rGO network can effectively increase the electronic conductivity of NiFe2O4, facilitating the rapid transfer of electrons.47,48

Figure 5d shows the CV diagrams at different scan rates (0.2 to 1.8 mV s–1). With increase in the scan rate, the intensity of peaks at 0.75 and 1.70 V increases. In addition, the oxidation peak moves toward a higher voltage and the reduction peak moves toward a lower voltage with increase of the scan rate, indicating the higher electrochemical polarization for NiFe2O4/rGO. The Nyquist plots of the NiFe2O4/rGO nanoparticles after different cycles (the 1st and the 20th) are shown in Figure 5f. On comparing the Nyquist plots of the NiFe2O4/rGO composite after cycles, the Rct value increases from 17 Ω for the initial electrode to 54 Ω for the 20th cycle. The diffusion coefficients of NiFe2O4/rGO (initial) and NiFe2O4/rGO (after 20 cycles) are 1.42 × 10–12 and 5.24 × 10–11 cm2 s–1, respectively. This drastic change may be attributed to the activation of electrode materials and enhanced transport properties of NiFe2O4/rGO. In combination with the results of the cycling performance, this trend in resistance is attributed to the formation of the SEI layer and the activation of NiFe2O4.

To further study the rate capacity of the NiFe2O4/rGO electrode, the electrode has been evaluated at different rates, as shown in Figure 5e. At a rate of 0.1 A g–1, the discharge capacity reaches ∼1745.8 mA h g–1. At high rates of 0.2, 0.5, 1, and 2 A g–1, the discharge capacities still remain at values of 804.2, 553.2, 376.3, and 272.4 mA h g–1, respectively. The discharge capacity of 467.9 mA h g–1 was recovered when the rate was returned to 0.1 A g–1. These results show that the NiFe2O4/rGO composite with a core–shell structure has great potential as a high-rate anode material for LIBs.

3. Conclusions

In conclusion, the NiFe2O4/rGO composite with a unique core–shell structure has been successfully prepared through a facile chemical process. The NiFe2O4/rGO electrode exhibits excellent cycling and rate performances as an anode material for LIBs. The rGO coating layer on the surface can not only benefit for the increase of electronic conductivity but also suppress the volume change during the charge/discharge processes. The introduction of the rGO layer onto NiFe2O4 provides an exceptional specific capacity of ∼1150 mA h g–1 after 900 cycles at 500 mA g–1 with a high Coulombic efficiency of 100%, which is significantly improved when compared with that of NiFe2O4 nanoparticles. The EIS measurements show that the charge-transfer resistance of core–shell NiFe2O4/rGO is smaller than that of the NiFe2O4 nanoparticles. The excellent lithium-storage properties of the core–shell NiFe2O4/rGO electrode may be ascribed to the core–shell structure and the small size of NiFe2O4 particles, which can effectively improve the electronic conductivity of the electrodes and accommodate the change of volume during the lithium-ion insertion/extraction process. Thus, the NiFe2O4/rGO composite has great potential in the development of Li-ion batteries.

4. Experimental Section

4.1. Material Synthesis

In the preparation process, 4.08 g of Fe(NO3)3·9H2O, 1.10 g of Ni(Ac)2·9H2O, and 3.04 g of anhydrous citric acid were dissolved in 20 ml of anhydrous ethanol solution and stirred at room temperature for 0.5 h. Then, the reaction mixture was continuously stirred for 2 h at 60 °C and then heated at 110 °C for 24 h to obtain a homogeneous mixture. NiFe2O4 was obtained by placing the mixture in a muffle oven at 5 °C min–1 to 450 °C for 5 h. Reduced graphene oxide (rGO) purchased from Alab (Shanghai) Chemical Technology Company (thickness: <1 nm; diam: 1–100 μm) was placed in 20 ml ethanol according to the proportion of ultrasonic for 1 h. The mixture was placed in a Teflon high-pressure reactor for hydrothermal reaction at 180 °C for 12 h. The product was washed with ethanol three times and dried in 60 °C for 24 h.

4.2. Material Characterization

The microstructures and morphologies of the materials were determined using a transmission electron microscope (TEM, JEM, 2011F). The crystal structures were evaluated using a powder X-ray diffraction (XRD) meter (Smart Lab 9KW) with Cu Kα radiation. Raman spectroscopy was performed using a Thermo DXR Microscope with 532 nm excitation laser wavelength. PHI 5000 Versa Probe, ULVAC-PHI with an Al Kα X-ray source (1486.6 eV), was used to measure X-ray photoelectron spectroscopy (XPS). The TGA curves were obtained using a SHIMADZU DTG-60H thermo balance in air with a 5 °C min–1 heating rate from 50 to 600 °C.

4.3. Electrochemical Measurements

Active materials, super P, and poly(vinylidene fluoride) (PVDF) (weight ratio: 8:1:1) were mixed in NMP to obtain a slurry. Then, the slurry was coated onto the Cu foil and dried at 80 °C for 12 h in a vacuum oven. The electrolyte was 1.0 M solution of LiPF6 in the mixture of EC/DEC (1:1 by volume). The cells were assembled in a glovebox filled with Ar. The galvanostatic charge/discharge tests were conducted on a LANHE electrochemical workstation with a potential window of 0.01–3.00 V (vs Li+/Li). Cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) measurements were conducted at the EC-Lab electrochemical workstation. CV was tested at a scan rate of 0.2 mV s–1 in a voltage range of 0.01–3.00 V. The CV of NiFe2O4/rGO at various scan rates was also tested in a voltage range of 0.01–3.00 V. In the EIS measurement, the amplitude of the AC signal is 0.2 mV and the frequency is between 100 kHz and 0.01 Hz.

Acknowledgments

This study was supported by the Key Laboratory of Advanced Energy Materials Chemistry, Renewable Energy Conversion and Storage Center, College of Chemistry, Nankai University.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.1c03050.

  • Raman spectra of NiFe2O4 (black line) and NiFe2O4/rGO (red line) (Figure S1); TGA curve for NiFe2O4/rGO (Figure S2); TEM images of NiFe2O4 (a, b, and c) and SAED pattern of NiFe2O4 (d) (Figure S3); EDS spectrum of NiFe2O4/rGO (Figure S4); and charge/discharge capacity and Coulombic efficiency of the NiFe2O4 electrode (Figure S5). The calculation of lithium-ion diffusion coefficients has been listed on page S7 (PDF)

The authors declare no competing financial interest.

Supplementary Material

ao1c03050_si_001.pdf (256.9KB, pdf)

References

  1. Kang K.; Meng Y. S.; Bréger J.; Grey C. P.; Ceder G. Electrodes with High Power and High Capacity for Rechargeable Lithium Batteries. Science 2006, 311, 977. 10.1126/science.1122152. [DOI] [PubMed] [Google Scholar]
  2. Roy P.; Srivastava S. K. Nanostructured anode materials for lithium ion batteries. J. Mater. Chem. A 2015, 3, 2454–2484. 10.1039/C4TA04980B. [DOI] [Google Scholar]
  3. Tarascon J. M.; Armand M. Issues and challenges facing rechargeable lithium batteries. Nature 2001, 414, 359–367. 10.1038/35104644. [DOI] [PubMed] [Google Scholar]
  4. Song S. C.; Zuo D. C.; An C. S.; Zhang X. H.; Li J. H.; He Z. J.; Li Y. J.; Zheng J. C. Self-assembled GeOX/Ti3C2TX Composites as Promising Anode Materials for Lithium Ion Batteries. Inorg. Chem. 2020, 59, 4711–4719. 10.1021/acs.inorgchem.9b03784. [DOI] [PubMed] [Google Scholar]
  5. Cheng X.-B.; Zhang R.; Zhao C.-Z.; Zhang Q. Toward Safe Lithium Metal Anode in Rechargeable Batteries: A Review. Chem. Rev. 2017, 117, 10403–10473. 10.1021/acs.chemrev.7b00115. [DOI] [PubMed] [Google Scholar]
  6. Liu J.; Bao Z.; Cui Y.; Dufek E. J.; Goodenough J. B.; Khalifah P.; Li Q.; Liaw B. Y.; Liu P.; Manthiram A.; Meng Y. S.; Subramanian V. R.; Toney M. F.; Viswanathan V. V.; Whittingham M. S.; Xiao J.; Xu W.; Yang J.; Yang X.-Q.; Zhang J.-G. Pathways for practical high-energy long-cycling lithium metal batteries. Nat. Energy 2019, 4, 180–186. 10.1038/s41560-019-0338-x. [DOI] [Google Scholar]
  7. Armand M.; Tarascon J. M. Building better batteries. Nature 2008, 451, 652–657. 10.1038/451652a. [DOI] [PubMed] [Google Scholar]
  8. Ding B.; Cai Z.; Ahsan Z.; Ma Y.; Zhang S.; Song G.; Yuan C.; Yang W.; Wen C. A Review of Metal Silicides for Lithium-Ion Battery Anode Application. Acta Metall. Sin. 2021, 34, 291–308. 10.1007/s40195-020-01095-z. [DOI] [Google Scholar]
  9. Zheng C.; Zeng L.; Wang M.; Zheng H.; Wei M. Synthesis of hierarchical ZnV2O4 microspheres and its electrochemical properties. CrystEngComm 2014, 16, 10309–10313. 10.1039/C4CE01445F. [DOI] [Google Scholar]
  10. Poizot P.; Laruelle S.; Grugeon S.; Dupont L.; Tarascon J. M. Nano-sized transition-metal oxides as negative-electrode materials for lithium-ion batteries. Nature 2000, 407, 496–499. 10.1038/35035045. [DOI] [PubMed] [Google Scholar]
  11. Hao Q.; Lei D.; Yin X.; Zhang M.; Liu S.; Li Q.; Chen L.; Wang T. 3-D mesoporous nano/micro-structured Fe3O4/C as a superior anode material for lithium-ion batteries. J. Solid State Electrochem. 2011, 15, 2563–2569. 10.1007/s10008-010-1232-4. [DOI] [Google Scholar]
  12. Zhu K.; Zhang Y.; Qiu H.; Meng Y.; Gao Y.; Meng X.; Gao Z.; Chen G.; Wei Y. Hierarchical Fe3O4 microsphere/reduced graphene oxide composites as a capable anode for lithium-ion batteries with remarkable cycling performance. J. Alloys Compd. 2016, 675, 399–406. 10.1016/j.jallcom.2016.02.214. [DOI] [Google Scholar]
  13. Yue Y.; Liang H. Micro- and Nano-Structured Vanadium Pentoxide (V2O5) for Electrodes of Lithium-Ion Batteries. Adv. Energy Mater. 2017, 7, 1602545 10.1002/aenm.201602545. [DOI] [Google Scholar]
  14. Lu S. J.; Wang Z. T.; Zhang X. H.; He Z. J.; Tong H.; Li Y. J.; Zheng J. C. In Situ-Formed Hollow Cobalt Sulfide Wrapped by Reduced Graphene Oxide as an Anode for High-Performance Lithium-Ion Batteries. ACS Appl. Mater. Interfaces 2020, 12, 2671–2678. 10.1021/acsami.9b18931. [DOI] [PubMed] [Google Scholar]
  15. Liu L.; Yu Y.; Yan C.; Li K.; Zheng Z. Wearable energy-dense and power-dense supercapacitor yarns enabled by scalable graphene–metallic textile composite electrodes. Nat. Commun. 2015, 6, 7260 10.1038/ncomms8260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Lian C.; Wang Z.; Lin R.; Wang D.; Chen C.; Li Y. An efficientfficient, controllable and facile two-step synthesis strategy: Fe3O4@RGO composites with various Fe3O4 nanoparticles and their supercapacitance properties. Nano Res. 2017, 10, 3303–3313. 10.1007/s12274-017-1543-8. [DOI] [Google Scholar]
  17. Tang L. B.; Zhang B.; An C. S.; Li H.; Xiao B.; Li J. H.; He Z. J.; Zheng J. C. Ultrahigh-Rate Behavior Anode Materials of MoSe2 Nanosheets Anchored on Dual-Heteroatoms Functionalized Graphene for Sodium-Ion Batteries. Inorg. Chem. 2019, 58, 8169–8178. 10.1021/acs.inorgchem.9b00971. [DOI] [PubMed] [Google Scholar]
  18. Zhou Y.; Li Y.; Wang Q.; Wang Q.; Du R.; Zhang M.; Sun X.; Zhang X.; Kang L.; Jiang F. Ultrasmall MoS3 Loaded GO Nanocomposites as High-Rate and Long-Cycle-Life Anode Materials for Lithium- and Sodium-Ion Batteries. ChemElectroChem 2019, 6, 3113–3119. 10.1002/celc.201900756. [DOI] [Google Scholar]
  19. Chen X. Q.; Lin H. B.; Zheng X. W.; Cai X.; Xia P.; Zhu Y. M.; Li X. P.; Li W. S. Fabrication of core–shell porous nanocubic Mn2O3@TiO2 as a high-performance anode for lithium ion batteries. J. Mater. Chem. A 2015, 3, 18198–18206. 10.1039/C5TA04238K. [DOI] [Google Scholar]
  20. Alcántara R.; Jaraba M.; Lavela P.; Tirado J. L.; Jumas J. C.; Olivier-Fourcade J. Changes in oxidation state and magnetic order of iron atoms during the electrochemical reaction of lithium with NiFe2O4. Electrochem. Commun. 2003, 5, 16–21. 10.1016/S1388-2481(02)00520-9. [DOI] [Google Scholar]
  21. Park G. D.; Cho J. S.; Kang Y. C. Multiphase and Double-Layer NiFe2O4@NiO-Hollow-Nanosphere-Decorated Reduced Graphene Oxide Composite Powders Prepared by Spray Pyrolysis Applying Nanoscale Kirkendall Diffusion. ACS Appl. Mater. Interfaces 2015, 7, 16842–16849. 10.1021/acsami.5b04891. [DOI] [PubMed] [Google Scholar]
  22. Liu H.; Zhu H.; Yang H. A low temperature synthesis of nanocrystalline spinel NiFe2O4 and its electrochemical performance as anode of lithium-ion batteries. Mater. Res. Bull. 2013, 48, 1587–1592. 10.1016/j.materresbull.2012.12.072. [DOI] [Google Scholar]
  23. Lavela P.; Tirado J. L. CoFe2O4 and NiFe2O4 synthesized by sol–gel procedures for their use as anode materials for Li ion batteries. J. Power Sources 2007, 172, 379–387. 10.1016/j.jpowsour.2007.07.055. [DOI] [Google Scholar]
  24. Li S.; Li A.; Zhang R.; He Y.; Zhai Y.; Xu L. Hierarchical porous metal ferrite ball-in-ball hollow spheres: General synthesis, formation mechanism, and high performance as anode materials for Li-ion batteries. Nano Res. 2014, 7, 1116–1127. 10.1007/s12274-014-0474-3. [DOI] [Google Scholar]
  25. Cherian C. T.; Sundaramurthy J.; Reddy M. V.; Suresh Kumar P.; Mani K.; Pliszka D.; Sow C. H.; Ramakrishna S.; Chowdari B. V. R. Morphologically Robust NiFe2O4 Nanofibers as High Capacity Li-Ion Battery Anode Material. ACS Appl. Mater. Interfaces 2013, 5, 9957–9963. 10.1021/am401779p. [DOI] [PubMed] [Google Scholar]
  26. Heidari E. K.; Zhang B.; Sohi M. H.; Ataie A.; Kim J.-K. Sandwich-structured graphene–NiFe2O4–carbon nanocomposite anodes with exceptional electrochemical performance for Li ion batteries. J. Mater. Chem. A 2014, 2, 8314–8322. 10.1039/c4ta00507d. [DOI] [Google Scholar]
  27. Dong T.; Wang G.; Yang P. Electrospun NiFe2O4@C fibers as high-performance anode for lithium-ion batteries. Diamond Relat. Mater. 2017, 73, 210–217. 10.1016/j.diamond.2016.09.024. [DOI] [Google Scholar]
  28. Choi M.; Koppala S. K.; Yoon D.; Hwang J.; Kim S. M.; Kim J. A route to synthesis molybdenum disulfide-reduced graphene oxide (MoS2-RGO) composites using supercritical methanol and their enhanced electrochemical performance for Li-ion batteries. J. Power Sources 2016, 309, 202–211. 10.1016/j.jpowsour.2016.01.081. [DOI] [Google Scholar]
  29. Li F.; Zhang L.; Li J.; Lin X.; Li X.; Fang Y.; Huang J.; Li W.; Tian M.; Jin J.; Li R. Synthesis of Cu–MoS2/rGO hybrid as non-noble metal electrocatalysts for the hydrogen evolution reaction. J. Power Sources 2015, 292, 15–22. 10.1016/j.jpowsour.2015.04.173. [DOI] [Google Scholar]
  30. Xu Z. L.; Zhang B.; Zhou Z. Q.; Abouali S.; Akbari Garakani M.; Huang J.; Huang J. Q.; Kim J. K. Carbon nanofibers containing Si nanoparticles and graphene-covered Ni for high performance anodes in Li ion batteries. RSC Adv. 2014, 4, 22359–22366. 10.1039/C4RA03066D. [DOI] [Google Scholar]
  31. Liu Y.; Zhang N.; Yu C.; Jiao L.; Chen J. MnFe2O4@C Nanofibers as High-Performance Anode for Sodium-Ion Batteries. Nano Lett. 2016, 16, 3321–3328. 10.1021/acs.nanolett.6b00942. [DOI] [PubMed] [Google Scholar]
  32. Bakhshi H.; Mohammad Azari M.; Shokuhfar A. A facile approach for obtaining NiFe2O4@C core-shell nanoparticles and their magnetic properties assessment. Diamond Relat. Mater. 2020, 110, 108159 10.1016/j.diamond.2020.108159. [DOI] [Google Scholar]
  33. D’Ippolito V.; Andreozzi G. B.; Bersani D.; Lottici P. P. Raman fingerprint of chromate, aluminate and ferrite spinels. J. Raman Spectrosc. 2015, 46, 1255–1264. 10.1002/jrs.4764. [DOI] [Google Scholar]
  34. Paswan S. K.; Kumari S.; Kar M.; Singh A.; Pathak H.; Borah J. P.; Kumar L. Optimization of structure-property relationships in nickel ferrite nanoparticles annealed at different temperature. J. Phys. Chem. Solids 2021, 151, 109928 10.1016/j.jpcs.2020.109928. [DOI] [Google Scholar]
  35. Sousa M. H.; Tourinho F. A.; Rubim J. C. Use of Raman micro-spectroscopy in the characterization of MIIFe2O4 (M = Fe, Zn) electric double layer ferrofluids. J. Raman Spectrosc. 2000, 31, 185–191. . [DOI] [Google Scholar]
  36. Fu M.; Qiu Z.; Chen W.; Lin Y.; Xin H.; Yang B.; Fan H.; Zhu C.; Xu J. NiFe2O4 porous nanorods/graphene composites as high-performance anode materials for lithium-ion batteries. Electrochim. Acta 2017, 248, 292–298. 10.1016/j.electacta.2017.07.027. [DOI] [Google Scholar]
  37. Zhao J.; Li X.; Cui G.; Sun X. Highly-active oxygen evolution electrocatalyzed by an Fe-doped NiCr2O4 nanoparticle film. Chem. Commun. 2018, 54, 5462–5465. 10.1039/C8CC02568A. [DOI] [PubMed] [Google Scholar]
  38. Chen Q.; Wang R.; Lu F.; Kuang X.; Tong Y.; Lu X. Boosting the Oxygen Evolution Reaction Activity of NiFe2O4 Nanosheets by Phosphate Ion Functionalization. ACS Omega 2019, 4, 3493–3499. 10.1021/acsomega.8b03081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Huang X. H.; Tu J. P.; Zhang B.; Zhang C. Q.; Li Y.; Yuan Y. F.; Wu H. M. Electrochemical properties of NiO–Ni nanocomposite as anode material for lithium ion batteries. J. Power Sources 2006, 161, 541–544. 10.1016/j.jpowsour.2006.03.039. [DOI] [Google Scholar]
  40. Gao X.; Wang J.; Zhang D.; Nie K.; Ma Y.; Zhong J.; Sun X. Hollow NiFe2O4 nanospheres on carbon nanorods as a highly efficient anode material for lithium ion batteries. J. Mater. Chem. A 2017, 5, 5007–5012. 10.1039/C6TA11058D. [DOI] [Google Scholar]
  41. Zhu J.; Yang D.; Rui X.; Sim D.; Yu H.; Hoster H. E.; Ajayan P. M.; Yan Q. Facile Preparation of Ordered Porous Graphene–Metal Oxide@C Binder-Free Electrodes with High Li Storage Performance. Small 2013, 9, 3390–3397. 10.1002/smll.201300755. [DOI] [PubMed] [Google Scholar]
  42. Cho J. S.; Hong Y. J.; Kang Y. C. Design and Synthesis of Bubble-Nanorod-Structured Fe2O3–Carbon Nanofibers as Advanced Anode Material for Li-Ion Batteries. ACS Nano 2015, 9, 4026–4035. 10.1021/acsnano.5b00088. [DOI] [PubMed] [Google Scholar]
  43. Fu Y.; Wan Y.; Xia H.; Wang X. Nickel ferrite–graphene heteroarchitectures: Toward high-performance anode materials for lithium-ion batteries. J. Power Sources 2012, 213, 338–342. 10.1016/j.jpowsour.2012.04.039. [DOI] [Google Scholar]
  44. Zheng T.; Gozdz A. S.; Amatucci G. G. Reactivity of the Solid Electrolyte Interface on Carbon Electrodes at Elevated Temperatures. J. Electrochem. Soc. 1999, 146, 4014–4018. 10.1149/1.1392585. [DOI] [Google Scholar]
  45. Chen C. H.; Hwang B. J.; Do J. S.; Weng J. H.; Venkateswarlu M.; Cheng M. Y.; Santhanam R.; Ragavendran K.; Lee J. F.; Chen J. M.; Liu D. G. An understanding of anomalous capacity of nano-sized CoO anode materials for advanced Li-ion battery. Electrochem. Commun. 2010, 12, 496–498. 10.1016/j.elecom.2010.01.031. [DOI] [Google Scholar]
  46. Yuan F.-W.; Yang H.-J.; Tuan H.-Y. Alkanethiol-Passivated Ge Nanowires as High-Performance Anode Materials for Lithium-Ion Batteries: The Role of Chemical Surface Functionalization. ACS Nano 2012, 6, 9932–9942. 10.1021/nn303519g. [DOI] [PubMed] [Google Scholar]
  47. Zhang Y.; Cao W.; Cai Y.; Shu J.; Cao M. Rational design of NiFe2O4–rGO by tuning the compositional chemistry and its enhanced performance for a Li-ion battery anode. Inorg. Chem. Front. 2019, 6, 961–968. 10.1039/C9QI00055K. [DOI] [Google Scholar]
  48. Chen Y.; Zhu J.; Qu B.; Lu B.; Xu Z. Graphene improving lithium-ion battery performance by construction of NiCo2O4/graphene hybrid nanosheet arrays. Nano Energy 2014, 3, 88–94. 10.1016/j.nanoen.2013.10.008. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao1c03050_si_001.pdf (256.9KB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES