Khadija
Kouchi
a,
Marwa
Tayoury
a,
Abdelwahed
Chari
a,
Loubna
Hdidou
a,
Zakaria
Chchiyai
b,
Khadija
El kamouny
c,
Youssef
Tamraoui
a,
Bouchaib
Manoun
ab,
Jones
Alami
a and
Mouad
Dahbi
*a
aMaterials Science, Energy, and Nano-engineering Department, Mohammed VI Polytechnic University, Lot 660-Hay Moulay Rachid, 43150, Ben Guerir, Morocco. E-mail: Mouad.dahbi@um6p.ma
bHassan First University, FST Settat, Rayonnement-Matière et Instrumentation, S3M, 26000, Settat, Morocco
cGreen Tech Institute Department, Mohammed VI Polytechnic University UM6P, Ben Guerir, Morocco
First published on 15th February 2024
Lithium-ion batteries (LIBs) have gained considerable attention from the scientific community due to their outstanding properties, such as high energy density, low self-discharge, and environmental sustainability. Among the prominent candidates for anode materials in next-generation LIBs are the spinel ferrites, represented by the MFe2O4 series, which offer exceptional theoretical capacities, excellent reversibility, cost-effectiveness, and eco-friendliness. In the scope of this study, Ni0.5Mg0.5Fe1.7Mn0.3O4 nanoparticles were synthesized using a sol–gel synthesis method and subsequently coated with a carbon layer to further enhance their electrochemical performance. TEM images confirmed the presence of the carbon coating layer on the Ni0.5Mg0.5Fe1.7Mn0.3O4/C composite. The analysis of the measured X-ray diffraction (XRD) and Raman spectroscopy results confirmed the formation of nanocrystalline Ni0.5Mg0.5Fe1.7Mn0.3O4 before coating and amorphous carbon in the Ni0.5Mg0.5Fe1.7Mn0.3O4/C after the coating. The Ni0.5Mg0.5Fe1.7Mn0.3O4 anode material exhibited a much higher specific capacity than the traditional graphite material, with initial discharge/charge capacities of 1275 and 874 mA h g−1, respectively, at a 100 mA g−1 current density and a first coulombic efficiency of 68.54%. The long-term cycling test showed a slight capacity fading, retaining approximately 85% of its initial capacity after 75 cycles. Notably, the carbon-coating layer greatly enhanced the stability and slightly increased the capacity of the as-prepared Ni0.5Mg0.5Fe1.7Mn0.3O4. The first discharge/charge capacities of Ni0.5Mg0.5Fe1.7Mn0.3O4/C at 100 mA g−1 current density reached 1032 and 723 mA h g−1, respectively, and a first coulombic efficiency of 70.06%, with an increase of discharge/charge capacities to 826.6 and 806.2 mA h g−1, respectively, after 75 cycles (with a capacity retention of 89.7%), and a high-rate capability of 372 mA h g−1 at 2C. Additionally, a full cell was designed using a Ni0.5Mg0.5Fe1.7Mn0.3O4/C anode and an NMC811 cathode. The output voltage was about 2.8 V, with a high initial specific capacity of 755 mA hg−1 at 0.125C, a high rate-capability of 448 mA h
g−1 at 2C, and a high-capacity retention of 91% after 30 cycles at 2C. The carbon coating layer on Ni0.5Mg0.5Fe1.7Mn0.3O4 nanoparticles played a crucial role in the excellent electrochemical performance, providing conducting, buffering, and protective effects.
To overcome the shortcomings associated with graphite and silicon-based anodes, spinel Li4Ti5O12 is another anode material that has been used for LIBs. Its discharge potential is about 1.5 V, which makes it safer than the commonly used graphite anode because the problem of lithium plating can be largely avoided.12 The Li4Ti5O12 demonstrates outstanding stability and specific capacity at high current densities. However, its theoretical capacity is limited to 175 mA h g−1.13 As a result, more research is necessary to develop new anode materials that offer improved safety, low cost, high theoretical capacity, and easy synthesis processes.
Spinel ferrite oxides, including ZnFe2O4,14 CoFe2O4,15 NiFe2O4,16 CuFe2O4,17 MgFe2O4,18etc., have gained considerable interest as potential anode materials for use in lithium-ion batteries (LIBs). This is largely because they have significantly higher theoretical capacities compared to graphite and Li4Ti5O12 anode materials.19,20
For instance, CaFe2O4, CoFe2O4, NiFe2O4, and CuFe2O4 exhibited higher specific capacities than the traditional graphite (372 mA h g−1). ZnFe2O4 can also be considered as a good anode material candidate owing to its elevated theoretical specific capacity of around 1000 mA h g−1.21
However, the main problems of spinel oxide anode materials are their low electrical conductivity and considerable volume expansion.22 To overcome these issues, carbon coating has been reported as a suitable coating material that can effectively enhance the capacity retention and cycling stability of spinel materials.17 Moreover, it participates in reducing the volume change during the discharging/charging operations and facilitates the Li-ion diffusion into the metal oxide structure.13
For example, Jin et al. have successfully prepared CuFe2O4/C hollow spheres through a hydrothermal growth method based on polymers, followed by calcination. After the 70th cycle at a current density of 100 mA cm−2, the anodic material demonstrated a specific capacity of 550 mA h g−1, which is significantly greater than the initial specific capacity of the CuFe2O4 hollow spheres (∼120 mA h g−1).23 In addition, Deng et al. reported a significant enhancement in the electrochemical properties of ZnFe2O4/C hollow spheres. They found that the ZnFe2O4 sample modified with carbon exhibited a specific capacity of 841 mA h g−1 after 30 cycles, along with a high-rate capability.24
To our knowledge, no studies have reported on the electrochemical performances of nanocrystalline Ni0.5Mg0.5Fe1.7Mn0.3O4 as anode materials for LIBs. This study details the preparation of Ni0.5Mg0.5Fe1.7Mn0.3O4 and Ni0.5Mg0.5Fe1.7Mn0.3O4/C nanoparticles, aiming to explore their potential as LIB anodes. We enhanced the cycling stability and rate capability of Ni0.5Mg0.5Fe1.7Mn0.3O4 anode through carbon coating. The nanocrystalline powders were characterized using X-ray diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), and Raman spectroscopy. The anode materials’ performance for LIBs was evaluated using electrochemical tests in a lithium half-cell, including cyclic voltammetry, galvanostatic charge–discharge, and electrochemical impedance spectroscopy (EIS) tests. Additionally, the carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C anode's performance for LIBs was specifically evaluated in the full-cell configuration.
In this synthesis process, the reagents reacted to produce the Ni0.5Mg0.5Fe1.7Mn0.3O4 spinel oxide, as depicted by the subsequent chemical equation:
0.5 Ni(NO3)2·6H2O + 0.5 Mg(NO3)2·6H2O + 1.7 FeCl3·6H2O + 0.3 MnCl2·4H2O → Ni0.5Mg0.5Fe1.7Mn0.3O4 + 2.85 Cl2↑ + 2 NO2↑ + 17.4 H2O↑ |
Carbon coating was performed by mixing the as-synthesized Ni0.5Mg0.5Fe1.7Mn0.3O4 material with sucrose powder as the carbon source (Ni0.5Mg0.5Fe1.7Mn0.3O4:
glucose = 95
:
5, which was the percentage) using acetone solution. Subsequently, the composite underwent a calcination process at a temperature of 600 °C for a duration of 5 hours, while being exposed to an argon environment to get the Ni0.5Mg0.5Fe1.7Mn0.3O4/C sample.
The galvanostatic discharge/charge and long-term cycling tests on the newly investigated uncoated and coated spinel oxide were carried out at a current density of 100 mA g−1 (100 mA g−1 = 0.125C) over the 0.01–3.00 V voltage range. Rate capability analysis was conducted at various C-rates, including 0.03C, 0.06C, 0.125C, 0.25C, 0.625C, 1.25C and 2C. Cyclic voltammetry curves of uncoated electrode Ni0.5Mg0.5Fe1.7Mn0.3O4 were obtained by varying the potential between 0.01 V and 3.0 V and utilizing a scan rate of 1 mV s−1. Electrochemical impedance spectra (EIS) were characterized over a frequency range of 100 kHz to 0.01 Hz. The electrochemical testing of the full cell was carried out at a current density of 100 mA g−1, within a voltage range of 1–4 V. For this, we used an NMC811 electrode, sourced from Argonne National Laboratory, as the positive electrode, and our synthesized carbon-coated spinel oxide Ni0.5Mg0.5Fe1.7Mn0.3O4/C as the negative electrode. The full cell, comprising NMC811 vs. Ni0.5Mg0.5Fe1.7Mn0.3O4/C, was assembled using the same procedure as for the half-cells of Ni0.5Mg0.5Fe1.7Mn0.3O4 and Ni0.5Mg0.5Fe1.7Mn0.3O4/C negative electrodes. However, in the full-cell configuration, the NMC811 electrode was used instead of the lithium metal disc. The N/P (negative/positive) mass ratio was maintained at approximately 0.84. A discharge/charge rate of 0.125C and 2C was applied, corresponding to a specific current of 100 mA g−1 and 1600 mA g−1 for the negative electrode material. All potentials were referenced to the Li/Li+ electrode. It should be noted that discharging and charging were regarded as lithiation and delithiation processes in this study. All electrochemical tests were conducted at ambient temperature utilizing a multi-channel potentiostat (MPG-2, Bio-Logic).
![]() | ||
Fig. 1 Room temperature XRD patterns of the synthesized Ni0.5Mg0.5Fe1.7Mn0.3O4 and Ni0.5Mg0.5Fe1.7Mn0.3O4/C materials. |
The X-ray diffraction (XRD) pattern of the carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C composite did not reveal any diffraction peaks for carbon, indicating that the carbon produced from glucose decomposition may be amorphous.29 In addition, there is no apparent difference between the XRD patterns of uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 and carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C materials. Therefore, the XRD results confirm the formation of the Ni0.5Mg0.5Fe1.7Mn0.3O4/C composite.
To conduct a more comprehensive examination of the structural characteristics, the X-ray diffraction (XRD) pattern of the synthesized Ni0.5Mg0.5Fe1.7Mn0.3O4 spinel ferrite was subjected to refinement using the Rietveld method. Nevertheless, the process of structural refinement was conducted using the cubic structure and the Fdm space group. Fig. 2(a) presents the Rietveld refinement plot for the synthesized Ni0.5Mg0.5Fe1.7Mn0.3O4 spinel. The red circles represent the experimental X-ray diffraction (XRD) pattern, the black solid lines depict the theoretical XRD pattern, and the blue solid line corresponds to the difference curves. The vertical green bars indicate the positions of various Bragg reflections. Fig. 2(a) presented in this study demonstrates a strong correspondence between the observed X-ray diffraction (XRD) pattern and the calculated XRD pattern. This agreement provides evidence of the successful structural refinement achieved for the synthesized Ni0.5Mg0.5Fe1.7Mn0.3O4 spinel ferrite. Furthermore, the evaluation of the quality of structural refinement was performed by considering several R-factors, such as the weighted profile factor (Rwp), profile factor (Rp), expected factor (Rexp), and Chi-squared value (χ2). Table S1 (ESI†) summarizes the findings from the powder X-ray diffraction analysis of the produced Ni0.5Mg0.5Fe1.7Mn0.3O4 spinel oxide using Rietveld refinement. The results include structural parameters, Chi-squared factor (χ2), and various reliability R-factors. Based on the obtained R-factors values, it can be concluded that the synthesized Ni0.5Mg0.5Fe1.7Mn0.3O4 spinel oxide crystallizes in the cubic Fd
m structure, characterized by a lattice parameter of a = 8.36215 Å.
Fig. 2(b) illustrates the crystal structure of the Ni0.5Mg0.5Fe1.7Mn0.3O4 sample, which exhibits cubic Fdm symmetry. The analysis demonstrates that the synthesized material exhibits a crystal structure that is characterized by the connectivity of tetrahedra and octahedra sites, which are composed of oxygen anions. Based on the established framework, it becomes apparent that there are three distinct and easily discernible classifications of sites, denoted as 8a, 16d, and 32e sites. Notably, the 8a and 16d sites can be correlated with the tetrahedral and octahedral sites, respectively. The cations Ni2+ and Mg2+ are present in a ratio of 1/8 within the tetrahedral sites formed by the anions O2−. Conversely, the cations Fe3+/Mn3+ are situated in half of the octahedral sites formed by the anions O2−. Furthermore, it should be noted that all the sites in the 32e structure are currently occupied by the O2− anions. Hence, it is evident within the provided crystal structure that a substantial fraction of the tetrahedral sites (7/8) and half of the octahedral sites are vacant, thereby providing vacancies for the incorporation of lithium ions in the lacunae of the unit cell.
To further investigate the crystal structure of the synthesized Ni0.5Mg0.5Fe1.7Mn0.3O4 material and to confirm the presence of a carbon layer in the Ni0.5Mg0.5Fe1.7Mn0.3O4/C nanoparticles, Raman scattering spectroscopy was employed. Fig. 3 shows the Raman spectra of the uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 and carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C materials. The Raman spectra of the coated material exhibit two distinct peaks at wavenumbers of approximately 1328.21 and 1582.06 cm−1, which correspond to the typical D and G bands for carbon, respectively.30,31 The D band is associated with the A1g phonon of sp3 carbon atoms at the edge and disordered carbon, while the G band is ascribed to the in-plane vibration of sp2 carbon atoms in carbon atomic rings or long-chain carbon. The ID/IG ratio is a common evaluation value for carbon materials, indicating the degree of disorder and its association with electronic conductivity.32,33 The calculated ID/IG value of 0.97 suggests a significant degree of disorder within the carbon structure of the as-prepared Ni0.5Mg0.5Fe1.7Mn0.3O4/C composite, indicating amorphous carbon and agreeing with the XRD results.
![]() | ||
Fig. 3 Raman scattering curves collected at room temperature of the uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 and carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C samples. |
The Raman spectra of the uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 material exhibited six peaks between 100 and 700 cm−1, which is also observed in the carbon-coated material.
It's pertinent to remember that, in the group theory analysis, the AB2O4-type spinels exhibiting cubic Fdm symmetry are predicted to have five active Raman modes, which may be defined as A1g, Eg, and 3T2g.34 The Raman scattering spectra of the produced material, obtained at room temperature, are depicted in Fig. 3. The remarkable crystallinity of the Ni0.5Mg0.5Fe1.7Mn0.3O4 sample is confirmed by the Raman spectra presented in Fig. 3. A total of six distinct Raman bands were detected at the respective wavenumbers of 189, 323, 473, 523, 650, and 691 cm−1. The detected Raman bands, have distinct characteristics that align with the crystal structure of MgFe2O4, namely the cubic spinel configuration. This crystal structure is commonly associated with the Fd
m space group.35 Based on prior research, it has been shown that the Raman bands detected at a wavenumber of 189 cm−1 can be attributed to the T2g(3) mode, which signifies the vibrations of the localized lattice within the octahedral sub-lattice.28 Nevertheless, it has been observed that two separate Raman bands have been found within the frequency range of 300–500 cm−1. The bands seen in this study are associated with the Eg and T2g(2) vibrational modes. Specifically, these bands are allocated to the symmetric and asymmetric bending vibrations of the A/B–O tetrahedron.23 In contrast, the three Raman peaks detected at elevated wavenumbers, namely at 523, 650, and 691 cm−1, are ascribed to the asymmetric and symmetric stretching vibrations of oxygen atoms that are linked to metal ions situated within tetrahedral sites. The vibrational modes are designated as T2g(1) for asymmetric stretching vibrations and A1g for symmetric stretching vibrations, respectively. The Raman bands observed at a frequency greater than 600 cm−1 are attributed to the vibrational modes of the A-site. The A1g mode exhibits a distinct separation into two distinct modes at about 650 and 691 cm−1, which can be attributed to the presence of various ions, specifically Mg2+ and Ni2+, within the tetrahedral site.36 This observation confirms the presence of a pure cubic spinel phase in the material that was synthesized.
The size and morphology of particles significantly impact the material's electrochemical performance. Creating active materials at the nanoscale is a highly efficient approach to enhancing electrode kinetics. Nanoparticles provide a larger contact area between the electrode and electrolyte, shorten the pathway for Li+ ion diffusion and electron movement, and increase the number of electrochemically active sites.37
The morphology of the as-prepared Ni0.5Mg0.5Fe1.7Mn0.3O4 sample was investigated by SEM analysis, as illustrated in Fig. 4(a)–(c). The prepared material demonstrates the presence of spherical particles, with an average diameter of around 217 nm. Besides, EDS elemental mapping of Ni0.5Mg0.5Fe1.7Mn0.3O4 is illustrated in Fig. 4(d)–(h) in which the signals corresponding to O, Mg, Ni, Fe, and Mn elements are detected with well-distributed elements and concentrated on the particles without obvious element segregation. Elemental analysis of the synthesized material was conducted using EDS spectroscopy. The surface elemental composition of the Ni0.5Mg0.5Fe1.7Mn0.3O4 material was determined using EDX analysis. Fig. S1, which can be found in the ESI,† provides evidence of the presence of nickel (Ni), magnesium (Mg), iron (Fe), manganese (Mn), and oxygen (O) in the as-synthesized sample, thereby confirming its purity.
![]() | ||
Fig. 4 (a)–(c) SEM images of Ni0.5Mg0.5Fe1.7Mn0.3O4 material. (d)–(h) EDS mapping images of the O, Ni, Mg, Fe, and Mn elements for Ni0.5Mg0.5Fe1.7Mn0.3O4 material. |
TEM analysis was used to observe the inner microstructure of the uncoated and carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4 materials. The images are presented in Fig. 5 and Fig. S2 (ESI†). Fig. S2 (ESI†) shows that there is no carbon layer on the surface of the uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 nanoparticles. However, Fig. 5(a)–(e) reveal amorphous carbon coating layers on the surface of the carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C nanoparticles. The homogeneity of various elements in the carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C powder was examined through high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) along with corresponding elemental mapping technique. The images produced by HAADF-STEM and energy-dispersive spectroscopy (EDS) mapping of Ni0.5Mg0.5Fe1.7Mn0.3O4/C, are presented in Fig. 5(f)–(l). These images reveal a consistent distribution of C, O, Ni, Mg, Fe, and Mn elements within the carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C nanoparticles.
![]() | ||
Fig. 5 (a)–(e) TEM images of carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C, (f) STEM image, and (g)–(l) corresponding elemental mapping images of the Ni0.5Mg0.5Fe1.7Mn0.3O4/C sample. |
The electrochemical performance of Ni0.5Mg0.5Fe1.7Mn0.3O4 and Ni0.5Mg0.5Fe1.7Mn0.3O4/C were evaluated by cyclic voltammetry and galvanostatic charge/discharge experiments at various current densities. These tests were conducted at room temperature, within a potential range ranging from 0.01 V to 3.0 V. Fig. 6 illustrates the cyclic voltammetry profile of the Ni0.5Mg0.5Fe1.7Mn0.3O4 electrode within the voltage range of 0.01–3.0 V, with a scan rate of 1 mV s−1. The initial cycle of the Ni0.5Mg0.5Fe1.7Mn0.3O4 sample exhibits notable distinctions compared to subsequent cycles. This disparity can be attributed to the cathodic lithiation process during the first cycle, wherein two cathodic peaks are observed. A pronounced reduction peak is observed at a potential of 0.01 V (versus Li+/Li), which is succeeded by a subsequent reduction peak occurring at 0.74 V (versus Li+/Li). These peaks are likely a result of the incorporation of Li+ ions into the Ni0.5Mg0.5Fe1.7Mn0.3O4 material through a series of multistep reactions, these reactions could be assigned to the cations reduced to metallic status (Fe3+/Fe0, Mn3+/Mn0, and Ni2+/Ni0 without Mg2+ reduction to Mg0 due to the high bond energy of MgO38) with the formation of SEI films and the generation of amorphous Li2O (eqn (1)).36,39 However, in this case, it can be observed that the cathodic peak at 0.01 V is indicative of the distinctive behavior associated with the insertion of Li+ ions into the Ni0.5Mg0.5Fe1.7Mn0.3O4 material.40 The initial anodic peak observed at 1.8 V can be ascribed to the oxidation processes of iron (Fe) to ferric ions (Fe3+), nickel (Ni) to nickel ions (Ni2+), and manganese (Mn) to manganese ions (Mn2+), facilitated by the creation of iron(III) oxide Fe2O3 (eqn (3)), MnO (eqn (4)), and NiO (eqn (2)) phases with the decomposition of the Li2O phase.7,25 Additionally, the pronounced cathodic peak recorded at a potential of 0.01 V (vs. Li+/Li) during the initial cathodic peak could be ascribed to the formation of the solid–electrolyte interphase (SEI) layer, which arises from the decomposition of the electrolyte. This phenomenon is widely recognized for its significant contribution to the substantial reduction in capacity observed during the initial discharge cycle.41 In the subsequent four CV cycles, they have similar cathodic and anodic peaks at 0.65 V and 1.9 V, respectively, The cathodic peaks for the subsequent four cycles could be linked to the reduction of NiO, Fe2O3, and MnO to Ni, Fe, and Mn, respectively. The following anodic process might be attributed to the oxidation of metallic nickel (Ni) to divalent ions (Ni2+), metallic iron (Fe) to ferric ions (Fe3+), and metallic manganese (Mn) to divalent ions (Mn2+), indicating similar kinetics during the charge and discharge of the Ni0.5Mg0.5Fe1.7Mn0.3O4 material.
![]() | ||
Fig. 6 Cyclic voltammetry profile in the voltage range of 0.01–3.0 V at a scan rate of 1 mV s−1 of the Ni0.5Mg0.5Fe1.7Mn0.3O4 electrode. |
Based on the literature,25,26,42,43 and the analysis that has been previously discussed, the electrochemical reactions of the Ni0.5Mg0.5Fe1.7Mn0.3O4 compound can be described as follows:
Ni0.5Mg0.5Fe1.7Mn0.3O4 + 7.1Li+ + 7.1e− → 0.5Ni + 0.5MgO + 1.7Fe + 0.3Mn + 3.55Li2O | (1) |
0.5Ni + 0.5Li2O ↔ 0.5NiO + Li+ + 1e− | (2) |
(1.7/2)Fe2O3 + 5.1Li+ + 5.1e− ↔ 1.7Fe + 2.55Li2O | (3) |
0.3MnO + 0.6Li+ + 0.6e− ↔ 0.3Mn + 0.3Li2O | (4) |
To improve the cycling stability of Ni0.5Mg0.5Fe1.7Mn0.3O4 material, a carbon coating process was adopted. The carbon coating was demonstrated to improve the metal oxide's electrochemical stability during long-term cycling. The carbon prevents the metal oxide particle agglomeration and maintains the structure's stability. Additionally, it participates in reducing the volume change during discharge/charge processes and facilitates the Li+ ion diffusion into the metal oxide's structure.28,44 The charge–discharge profile of both uncoated and coated anode materials is depicted in Fig. 7(a) and (b). The data were obtained by conducting experiments within the voltage range of 0.01–3 V (vs. Li/Li+) and at a current density of 100 mA g−1 for LIBs. The initial discharge and charge capacities for the uncoated and coated materials were determined to be 1275/874 and 1032/723 mA h g−1, respectively, accompanied by initial coulombic efficiencies (ICE) of 68.54% and 70.06%, respectively. Consequently, the first coulombic efficiency of the Ni0.5Mg0.5Fe1.7Mn0.3O4/C anode is slightly higher than that of the uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 anode. This improvement may be attributed to the carbon layer, which protects the electrode materials from direct contact with the liquid electrolyte.45 In addition, it is important to note that the initial capacity reduction during the first cycle is mainly due to the formation of a solid electrolyte interphase (SEI) layer on the electrode surface, accompanied by the generation of the inactive magnesium oxide phase (MgO).46 In the 75th cycle, the charge capacities of uncoated and carbon-coated anode materials were 744 mA h g−1 and 806 mA h g−1 respectively. From the long-term cycling, the coated material shows an increase in capacity after 75 cycles compared to the uncoated material which shows a decrease in capacity, as indicated in Fig. 7(c). Moreover, the carbon-coated material displayed a capacity retention of approximately 89.7%, while reaching nearly 100% coulombic efficiency in the 75th cycle. The details of the electrochemical properties of these two electrodes are shown in Table S2 (ESI†). The excellent cycle stability of the Ni0.5Mg0.5Fe1.7Mn0.3O4 material after carbon coating can be attributed to the stable solid electrolyte interphase (SEI) formed between the electrode and the electrolyte. Additionally, the improvement in the cycling stability of the coated material, compared to the uncoated sample, is a result of enhanced mechanical stability and stress resistance due to volume changes during the charge/discharge process. The coating limits the material's volume expansion by limiting its capacity; indeed, the anode did not reach its maximum capacity. Consequently, the volume change during lithiation is reduced, leading to improved stability.47 As a result, both the electrochemical performance and cycling stability of the electrode are enhanced.
To better understand the advantages of carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4 nanoparticles in lithium-based energy storage, we compared the performance of uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 and carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C rate capability in terms of Li+ insertion/extraction (Fig. 8(a)). The two electrodes were cycled at different current rates of 0.03C, 0.06C, 0.125C, 0.25C, 0.625C, 1.25C and 2C, corresponding to 25, 50, 100, 200, 500, 1000 and 1600 mA g−1. The coated electrode cell delivers average charge capacities of 720, 778, 760, 731, 690, 651, and 372 mA h g−1. At the highest current density of 1600 mA g−1, Ni0.5Mg0.5Fe1.7Mn0.3O4/C delivers 372 mA h g−1 instead of the limited capacity of 192 mA h g−1 for Ni0.5Mg0.5Fe1.7Mn0.3O4. More importantly, when the current density was returned to 25 mA g−1, a large irreversible capacity of 818 mA h g−1 was recovered, showing a strong tolerance to the rapid insertion/extraction of Li+ ions. To further evaluate the long-term cycling performance at high current density, uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4, and carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C electrodes were charged and discharged for 100 cycles at 2C for comparison. As shown in Fig. 8(b), the uncoated material Ni0.5Mg0.5Fe1.7Mn0.3O4 experienced an apparent capacity fading after 50 cycles and its capacity retention reached only 18.24% after 100 cycles, while the Ni0.5Mg0.5Fe1.7Mn0.3O4/C still retained 91.17% of the initial discharge capacity after 100 cycles, reflecting the unstable structure of the uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 and the stable structure of the carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C. The details of the electrochemical properties of these two electrodes are shown in Table S2 (ESI†).
To understand the carbon coating effect in more detail, electrochemical impedance spectroscopy (EIS) tests were performed before and after 100 cycles at a high current rate of 2C (100 mA g−1 = 0.125C). These tests were performed in the frequency range of 100 kHz to 0.01 Hz using cells with uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 and carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C electrodes (Fig. 8(c) and (d)). From Fig. 8(c) and (d), the observed impedance spectra showed a semicircle in the high-frequency region, corresponding to the charge transfer resistance due to the transport of lithium ions across the electrode/electrolyte interface, and an inclined straight line in the low-frequency region, corresponding to the Warburg impedance due to the diffusion of Li+ ions into the electrode materials.48Fig. 8(c) and (d) show the equivalent circuit model, consisting of the contact resistance (Rs), surface film resistance (Rf), charge transfer resistance (Rct) and Warburg impedance (W4); the fitting parameters are summarized in Table S3 (ESI†).49 The Rtotal (where Rtotal = Rs + Rf + Rct) can be closely monitored to explore the origin of the electrochemical properties of LiB cells in practice. A small Rtotal indicates improved cycling performance and rate capability of LiB. From the equivalent circuit model, the fitted Rct parameter for cells before cycling are 18646 and 10
237 Ω for uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 and carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C electrodes, respectively. The lower Rct value of the carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C electrode suggests a higher charge diffusion efficiency in this novel nanostructure.50 Meanwhile, Rct decreased to 2145 and 400 Ω after 100 cycles at a 2C rate for uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 and carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C electrodes, respectively. It is observed that the Rct values after 100 cycles were lower than those of the cell before cycling, indicating a lower charge transfer resistance, which suggests an improvement in electron transport during the repeated lithiation and delithiation processes. For Rtotal, the results indicate that the carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C electrode has a significantly lower total resistance after cycling compared to the uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 and carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C electrodes before cycling, indicating improved stability and rate capability performance. Thus, the EIS result shows that the Ni0.5Mg0.5Fe1.7Mn0.3O4/C composite electrode has higher electrical conductivity compared to the uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 material, resulting in stable and higher reversible capacity.
The Ni0.5Mg0.5Fe1.7Mn0.3O4/C carbon-coated electrode exhibits excellent cycling performance and rate capability compared to the uncoated Ni0.5Mg0.5Fe1.7Mn0.3O4 electrode, which is attributed to the protective carbon coating layer which protects the Ni0.5Mg0.5Fe1.7Mn0.3O4 nanoparticles from direct contact with the electrolyte, thereby maintaining the structural integrity of the nanoparticles during the lithiation–delithiation process. Furthermore, the carbon layers not only mitigate the volume expansion of Ni0.5Mg0.5Fe1.7Mn0.3O4 nanoparticles but also maintain stable electronic conductivity during the lithiation–delithiation process impedance text ref. 51.
To evaluate the performance of the carbon-coated Ni0.5Mg0.5Fe1.7Mn0.3O4/C material in a complete battery system, a full cell was designed using an NMC811 cathode (Fig. 9(a) and (b)). The full cell's output voltage was 2.8 V at 0.125C current rate, within the voltage window of 1–4 V, matching the voltage difference between the NMC811 cathode and Ni0.5Mg0.5Fe1.7Mn0.3O4/C anode. Furthermore, in Fig. 9(a), the charge/discharge cycles of the full cell at a 0.125C current rate are illustrated within the voltage range of 1–4 V. The full cell exhibited a capacity of 755 mA h g−1 with a coulombic efficiency of 65% during the first cycle, based on the mass of the Ni0.5Mg0.5Fe1.7Mn0.3O4/C anode. Subsequently, after 4 cycles, the capacity decreased to 633 mA h g−1, with an increase in the coulombic efficiency to 97%. The cycling performance of the full cell was investigated at 2C within the voltage window of 1–4 V (Fig. 9(b)). After 30 cycles, a capacity of 448 mA h g−1 was achieved with a capacity retention of 91% and a coulombic efficiency of 98%. The low capacity of the Ni0.5Mg0.5Fe1.7Mn0.3O4/C electrode in the full cell is mainly due to its low first coulombic efficiency, which did not exceed 65%. The low first coulombic efficiency of the Ni0.5Mg0.5Fe1.7Mn0.3O4/C anode may be attributed to various factors, such as the inadequate balance between the positive and negative electrodes, electrolyte decomposition during cycling, and other factors.
Fig. 10 shows the reversible capacity, the average potential, and the estimated energy density of anode materials for lithium full cells. In this study, we demonstrate that the Ni0.5Mg0.5Fe1.7Mn0.3O4/C composite material displays a reversible capacity above 800 mA h g−1. This value is notably 2–3 times greater than the reversible capacity observed in the graphite anode material,35 and also shows a high specific capacity compared to the spinel Li4Ti5O12,52 TiO2,53 and Li3VO4,54 anode materials. Furthermore, the energy density of spinel oxide Ni0.5Mg0.5Fe1.7Mn0.3O4/C Li-ion cell is calculated to be 360 Wh kg−1 based on the capacity and the average potential of positive and negative electrodes with a relatively high operating voltage of 1.1 V versus Li/Li+ compared to the traditional graphite.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cp00182f |
This journal is © the Owner Societies 2024 |