Spatial confinement of free-standing graphene sponge enables excellent stability of conversion-type Fe2O3 anode for sodium storage

Jun Dong Senyuan Tan Sunbin Yang Yalong Jiang Ruxing Wang Jian Ao Zilun Chen Chaohai Zhang Qinyou An Xiaoxing Zhang

Citation:  Jun Dong, Senyuan Tan, Sunbin Yang, Yalong Jiang, Ruxing Wang, Jian Ao, Zilun Chen, Chaohai Zhang, Qinyou An, Xiaoxing Zhang. Spatial confinement of free-standing graphene sponge enables excellent stability of conversion-type Fe2O3 anode for sodium storage[J]. Chinese Chemical Letters, 2025, 36(3): 110010. doi: 10.1016/j.cclet.2024.110010 shu

Spatial confinement of free-standing graphene sponge enables excellent stability of conversion-type Fe2O3 anode for sodium storage

English

  • Na-ion batteries (NIBs) are recognized as one of the most promising energy storage technologies for green-grid systems due to the high abundance of sodium resources [1-4]. However, compared to the advanced lithium-ion batteries (LIBs) (approaching 300 Wh/kg), the energy density of NIBs needs to be further increased [5]. One of the main factors that affect their development is the lack of affordable and high-performance anode materials. Conversion-type anode materials (MaXb; M = Ni, Co, Fe, Mn, Cu, none, etc.; X = O, S, Se, Te, P, N, etc.) undergo multi-electron reactions, presenting high theoretical capacity [6-10]. Meanwhile, they are usually low-cost, environmentally friendly, and non-toxicity. Hence, conversion-type anode materials are considered very promising for obtaining NIBs with high energy density. One of the key obstacles restricting their developments is the poor cycling stability. During the intercalation of Na+ ions, the atoms in MaXb structure have been rearranged, and MaXb is converted into nano-sized M particles and NacX phases [7,11]. The above molecular-level reorganization process results in huge volume changes and finally severe pulverization of the active materials, leading to an inferior electrochemical performance of MaXb. In response to the above problems, the conventional optimization strategies include nano-sized structures design, doping and adding electronic conductivity additives [6,12]. Spatial confinement effect enabled by constructing a fast electronic/ionic network around the MaXb materials is promising for achieving a long cycling life. This network could help to withstand volume expansion of the sodiated phases while improving electron/ion transport kinetics.

    When it comes to constructing a robust 3D electronic/ionic network, constructing 3D graphene framework (GF) as a conductive network is an effective strategy to improve the electrochemical performance of sulfur-based cathode [13]. The 3DGF is constructed by conjugated and interlocked graphene nanosheets, forming a continuous and hierarchical porous network. It can withstand the volume expansion of sodiated phases and provide an efficient ion transport path. Besides, the 3DGF displays a high conductivity of up to ~1400 S/m, which benefits to improve the electronic conductivity of composite materials [13]. Moreover, the strong mechanical strength of 3DGF can make a freestanding electrode without inactive binders. Considering the above advantages, 3DGF can be applied as a robust ionic/electronic conductive network in conversion-type anode materials.

    Fe2O3 anode presents a theoretical capacity of up to 1007 mAh/g and is one of the most promising conversion-type anodes of NIBs [14]. Herein, a self-supporting 3D graphene sponge decorated with Fe2O3 nanocubes (rGO@Fe2O3) was designed. It should be emphasized that the carbon composite is a promising strategy to optimize properties [15-18]. In the reported work, graphene mainly presents a two-dimensional (2D) sheet structure. However, we constructed a three-dimensional (3D) hierarchical porous network structure by cross-linking the 2D graphene sheet, in which the Fe2O3 nanocubes are encapsulated by graphene. The 3DGF provides a high-porosity structure, and Fe2O3 nanocubes are well dispersed in 3DGF without agglomeration. The 3DFG can not only act as a structural framework to buffer the volume expansion of Fe2O3 nanocubes but also serve as an ionic/electronic transport network. In particular, the rGO@Fe2O3 sponge can be directly acted as a self-supporting electrode without other inactive components such as binders and conductive additives. Subsequently, the electrochemical performance test, morphology evolution, and reaction kinetics analysis of the materials were carried out. It is found that sodiated Fe2O3 nanocubes in rGO@Fe2O3 sponge have not yet pulverized. The rGO@Fe2O3 sponge delivers excellent electrochemical performance. Finally, the assembled full NIBs display high power and energy density. This work unveils the potential of achieving NIBs with exceptional longevity through constructing a 3D ionic/electronic conductive network decorated with a nanosized conversion-type anode.

    The high-porosity rGO@Fe2O3 sponge was synthesized by self-assembly and calcination processes (Fig. 1a). First, graphene oxides decorated with Fe2O3 nanocubes (GO@Fe2O3) hydrogel were obtained by self-assembly under hydrothermal conditions. Then, the 3D rGO@Fe2O3 sponge was prepared by further annealing in the N2 atmosphere. The pure Fe2O3 nanoparticles were also synthesized for comparison. XRD patterns of the Fe2O3 and rGO@Fe2O3 are shown in Fig. S1a (Supporting information), and the phase is α-Fe2O3 (ICCD 01–084–0308). In the Raman spectra of rGO@Fe2O3, all peaks are assigned to rGO and Fe2O3 (Fig. S1b in Supporting information). The obvious peaks between 200 and 300 cm−1 are derived from the Fe2O3 [19]. The graphitic carbon (G band, 1590 cm−1) and the disordered carbon (D band 1351 cm−1) are also observed [20]. The measured ID/IG of around 1.08 indicates the presence of defects in rGO which help nanoparticles adhere to the surface of rGO [20]. The XPS survey spectrum displays Fe, O, C and N elements in the rGO@Fe2O3 sponge (Fig. S2a in Supporting information). Additionally, the Fe 2p XPS spectrum can be fitted into an one doublet Fe 2p1/2 (724.9 eV) and Fe 2p3/2 (711.3 eV) [15,21], indicating the existence of Fe2O3, coincident with the result of XRD pattern (Fig. S2b in Supporting information). In addition, the XPS spectrum of C 1 s can be fitted into three peaks at 284.7, 285.8, and 288.9 eV, ascribing to the C-C, O-C=O, and C-O (Fig. S2c in Supporting information). The results indicate the existence of oxygen-containing groups on the surface of rGO which could anchor Fe2O3 nanocubes by the Fe-O-C bond, restricting its growth and making the nanoparticles dispersed evenly.

    Figure 1

    Figure 1.  Schematic of the fabrication and structural characterization of rGO@Fe2O3 sponge. (a) Schematic illustration for the synthesis of rGO@Fe2O3 sponge. (b, c) SEM and TEM images of rGO@Fe2O3 sponge. (d, e) Nitrogen adsorption and desorption isotherm and pore-size distribution of rGO@Fe2O3 sponge.

    The morphology of Fe2O3 and rGO@Fe2O3 sponge was further investigated. The average size of pure Fe2O3 nanocubes is about 140 nm (Fig. S3 in Supporting information). As for the rGO@Fe2O3 sponge, it is found that Fe2O3 nanocubes are dispersed on the rGO surface without agglomeration (Figs. 1b and c), that is conducive to expose more active sites and further obtain excellent electrochemical performance. The Fe2O3 nanocubes exhibit an average size of about 100 nm. The high-resolution TEM (HRTEM) image reveals that Fe2O3 nanocubes are in close contact with the rGO sheets with a thickness of less than 5 nm (Fig. S4a in Supporting information), beneficial to fast electron transport. The distinct lattice fringes present a d-spacing of 0.36 nm and 0.27 nm, matching to the (012) plane and (104) plane of the α-Fe2O3 (Fig. S4b in Supporting information). The diffraction rings mean the polycrystalline characteristic (inset Fig. S4b). Besides, the distribution of Fe, O, C, and N elements is in good agreement with the rGO@Fe2O3 morphology (Fig. S5 in Supporting information). Meanwhile, the nitrogen element is attributed to nitrogen doping in rGO, consistent with the XPS analysis results (Fig. S2a). The 3D porous structures of the rGO@Fe2O3 are investigated by nitrogen adsorption-desorption isotherms (Fig. 1d) and pore-size distributions (Fig. 1e). The Brunauer–Emmett–Teller specific surface area of rGO@Fe2O3 is 31.87 m2/g. The pores of rGO@Fe2O3 range mostly between 2 nm and 100 nm based on the Barrett-Joyner-Halenda model. Interestingly, the well-defined mesopores (~3 nm/40 nm) and macropores (~60 nm) were obtained, presenting a hierarchical porous network, favorable for electrolyte accessibility and rapid Na-ion diffusion (Fig. 1e).

    To evaluate the Na+-ion storage performance, the electrode slices were first prepared, with sodium metal as a counter electrode. The rGO@Fe2O3 sponge can be directly used as electrode slices without adding additional binders and conductive carbon black (Fig. S6 in Supporting information), further achieving batteries with higher gravimetric energy density. Based on the CHN elemental analysis, the carbon content of rGO@Fe2O3 is around 56% (Table S1 in Supporting information). The gravimetric capacity of the rGO@Fe2O3 sponge is calculated based on the quality of the entire electrode. The initial discharge/charge capacities are 1683/859 mAh/g with a first Coulombic efficiency of 51% at 0.1 A/g in the voltage range of 0.01–3 V vs. Na+/Na (Fig. 2a). The capacity loss may ascribe to the irreversible consumption of sodium ions by the functional groups on the graphene sheet and the formation of solid electrolyte interface (SEI) film. After initial cycle, rGO@Fe2O3 electrode exhibits high reversibility.

    Figure 2

    Figure 2.  Electrochemical performances of rGO@Fe2O3 sponge and Fe2O3. (a) The galvanostatic charge-discharge curves of rGO@Fe2O3 sponge at 0.1 A/g. (b, c) CV curves of rGO@Fe2O3 and Fe2O3 at 0.2 mV/s at first three cycles. (d) Long-term cycles at 1 A/g of rGO@Fe2O3. (e) Cycling performance comparison of current work to the state of the art. (f) Rate performance of rGO@Fe2O3 and Fe2O3 at different current densities. (g) Rate performance comparison of current work to the state of the art.

    Figs. 2b and c display the cyclic voltammetry (CV) curves. The irreversible cathodic peaks at the initial cycle correspond to the SEI formation. Compared with Fe2O3 anode, the first-loop CV curves of rGO@Fe2O3 have more irreversible peaks, which may be due to the irreversible redox reaction of the functional groups on the surface of rGO, and this may also be the reason for the low initial Coulomb efficiency. In the second cycle, there are peak 1 and 2 (0.76 V, 0.01 V) on the reduction side and peak 3 and 4 (0.3 V, 1.5 V) on the oxidation side. The peak 1 indicates the Na+ insertion process into Fe2O3, corresponding to the Na2Fe2O3 formation [11]. The peak 2 can be attributed to the phase conversion from NaxFe2O3 to Na2O and Fe. During Na+ extraction, the peak 3 is assigned to the oxidation of Fe, resulting in the formation of Na2Fe2O3. The peak 4 at 1.5 vs. Na+/Na is attributed to the desodiation of Na2Fe2O3. Consequently, the reversible conversion reaction can be described as followed: Fe2O3 + 6Na++ 6e↔ 2Fe + 3Na2O [11,22]. The CV curves of pure Fe2O3 anode show a similar reaction process. The critical point is that the two anodic peaks (peak 3’ and peak 4’) of the pure Fe2O3 anode shift toward higher potential side than that of the rGO@Fe2O3, resulting a larger overpotential. The lower overpotential of the rGO@Fe2O3 anode indicates that the 3D conductive network framework can effectively reduce its impedance of desodiation reaction and improve the reaction kinetics.

    The cycling performance of the rGO@Fe2O3 sponge was also evaluated (Fig. 2d). Compared with Fe2O3, the rGO@Fe2O3 sponge shows higher capacity. After 2000 cycles at 1 A/g, the discharge capacity of rGO@Fe2O3 sponge is 327 mAh/g, which maintains 63.4% of the original capacity. The specific capacity of rGO@Fe2O3 decreases within 100 cycles and then stabilizes, and the corresponding polarization also increases first and then remains unchanged (Fig. S7a in Supporting information). These results suggest that the fading of capacity may be related to the increase of polarization. In addition, the 3D network structure can provide abundant active interfacial sites that may decrease during the cycling process, leading to the capacity decay. Fig. 2f shows the rate capability of rGO@Fe2O3 sponge and Fe2O3 at various current densities. The reversible capacity of Fe2O3 only exhibits 50 mAh/g at 4 A/g. In contrast, the reversible capacity of rGO@Fe2O3 anode is 271 and 162 mAh/g at 4 A/g and 20 A/g. The corresponding galvanostatic charge-discharge curves maintain the inclined-line shape at different current densities, indicating a fast capacitive behavior (Fig. S7b in Supporting information). The comparisons of cycling performance and rate capacity between previous reports and current work are shown in Figs. 2e and g [14-16,20,21,23-42]. It demonstrates that the Fe2O3 anode in this work has an obvious improvement in rate and cycling performance with the help of a 3D electronic/ionic network. The optimized performance may be attributed to its unique structure. 3D rGO conductive framework with high porosity and resilience facilitates the transport of the electrons/ions and accommodates the volume expansion of the Fe2O3 nanocubes during the desodiation/sodiation.

    The kinetics process of Na+ ion diffusion was studied by the galvanostatic intermittent titration technique (GITT) method [43-45]. The electrode was activated after three cycles to reach the equilibrium state for the GITT test. The time required for rGO@Fe2O3 anode to complete GITT discharge or charge test is as long as 25 h, much higher than the result of pure Fe2O3 anode (10 h, Fig. 3a). The Na+ diffusion coefficients (DNa) for rGO@Fe2O3 anode at the initial states of discharging and charging are initially high at about 9.5 × 10−8 and 3.6 × 10−8 cm2/s (Fig. 3b and Fig. S8 in Supporting information). The calculated average DNa+ over the entire discharge or charge process are about 2.12 × 10−8 and 3.19 × 10−8 cm2/s, significantly higher than those of pure Fe2O3 (1.34 × 10−8 and 9.7 × 10−9 cm2/s). In the electrochemical impedance spectrum (EIS) measurements, the rGO@Fe2O3 anode at a fully charged state presents a smaller semicircle than that of Fe2O3 (Fig. 3c), indicating a smaller charge-transfer resistance. The fast Na+ diffusion kinetics and low charge-transfer resistance of the rGO@Fe2O3 anode could be attributed to a unique 3D ionic/electronic conductive network.

    Figure 3

    Figure 3.  Electrochemical kinetic analysis. (a) GITT profiles of discharge and charge processes of rGO@Fe2O3 and Fe2O3 measured after three cycles at the current density of 0.02 A/g. (b, c) Na+ diffusion coefficients as a function of the states of charging process and Nyquist plots of rGO@ Fe2O3 and Fe2O3 at fully charged state. (d) CV curves of rGO@Fe2O3 at the scan rates ranging from 0.2 mV/s to 1 mV/s. (e)The normalized peak currents versus scan rate to determine the b value for anodic and cathodic peaks of rGO@Fe2O3. (f) The proportion of diffusion−controlled and capacitive contributions to the total charge storage at different scant rates.

    The CV measurements were also carried out to study reaction kinetics during sodiation/desodiation process. When increasing the scan rates, the cathodic peak of the rGO@Fe2O3 electrode still exists and only slightly shifts (Fig. 3d and Fig. S9a in Supporting information). However, the cathodic peak of the Fe2O3 electrode disappeared (Figs. S9b and c in Supporting information). This result shows that rGO@Fe2O3 exhibits good charge transfer and ion diffusion ability during fast charge and discharge, effectively reducing the polarization of the electrode and improving its rate capacity. Then, as proposed by Dunn et al., a semi-quantitative analysis is conducted. The peak current (ip) obeys the power law, as described in Eq. 1 [46]:

    (1)

    where a and b are constants, and b-value is obtained by linearly fitting log(i) and log(v). The b = 0.5 and b = 1 respectively indicates a semi-infinite diffusion process and a surface-controlled capacitive kinetics. Fig. 3e displays the fitting results, and the b-values of rGO@Fe2O3 (0.75/0.8) are close to 1, indicating an apparent capacitive charge storage process.

    The quantitative determination of the capacitive contribution to the total capacity was conducted according to the following Eq. 2 [47]:

    (2)

    where the current response i(V) at a specific potential could be divided into capacitive contribution (k2v) and diffusion-controlled part (k1v1/2). Fig. 3f shows the proportion of capacitive contribution at different scan rates, with all the capacitive contributions greater than 50%. The capacitive contribution for the rGO@Fe2O3 is 75.7% at 1 mV/s (Fig. S9d in Supporting information). These results demonstrate that capacitive contributions dominate the total capacity.

    The SEM and TEM were used to investigate the morphology changes of Fe2O3 particles during sodiation/desodiation. Figs. 4a and b present the morphological changes of Fe2O3 nanocubes in the rGO@Fe2O3 sponge from the fresh to the fully discharged state. The metallic iron is observed at the fully discharged state (Fig. S10 in Supporting information). The nanocubes remain well dispersed on the rGO sheet without agglomeration and maintain an intact morphology. Besides, the nanocube size in the rGO@Fe2O3 sponge exhibits an increasing tendency (Fig. 4c). The primary size is in the range of 80–140 nm in the initial state, which changes to that in the range of 90–270 nm after deep discharge. The average size-distribution increases from the initial 112.18 nm to 186.5 nm. After desodiation, the lattice fringes of Fe2O3 recovered, but the lattice fringes of Fe still present (Figs. S11a-c in Supporting information). This partially irreversible transformation process affects the first Coulomb efficiency of the electrode. After desodiation, the particles would shrink, with a uniform distribution of 144.92 nm (Fig. S11d in Supporting information). The above results indicate that the sodiated Fe2O3 nanocubes in the rGO@Fe2O3 sponge maintain an intact morphology and do not pulverized. In comparison, the nanocubes in bare Fe2O3 are clearly deformed at fully discharged state (Figs. 4d and e). As shown in Fig. 4f, the primary size is in the range of 90–190 nm in the initial state, which changes to in the range of 30–140 nm after deep discharge. The uniform distribution decreases from the initial 142.27 nm to 79.78 nm, meaning that more small particles appear. Thus, bare Fe2O3 nanocubes without the protection of rGO sheets undergo a significant pulverization process. The above results indicate that the unique 3DGF effectively buffers the volume expansion of particles and inhibits their pulverization.

    Figure 4

    Figure 4.  Morphological changes characterization of Fe2O3 particles. TEM images of rGO@Fe2O3 sponge (a) at fresh states and (b) fully discharged states for Na-ion storage. (c) The size distribution of Fe2O3 nanocubes in rGO@ Fe2O3 sponge at fresh and fully discharged state. SEM images of pure Fe2O3 particles (d) at fresh states and (e) fully discharged states. (f) The size distribution of bare Fe2O3 particles at fresh and fully discharged states. (g) Schematic of structure changes of rGO@Fe2O3 and Fe2O3 after sodiation.

    Based on the above discussion, the 3DGF with a continuous and hierarchical porous network enables spatial confinement to achieve outstanding results, which is visually demonstrated for the first time by the results of SEM and TEM in Fig. 4. The 3DGF not only acts as a structural framework to suppresses particle pulverization but also serves as an ionic/electronic conductive network to improve the electrochemical reaction kinetics (Fig. 4g). As a result, the rGO@Fe2O3 displays an excellent rate capability and cycling stability.

    It has been found that rGO@Fe2O3 in this work exhibits a superior specific capacity, which is favorable for obtaining full batteries with higher energy density. The coin-type Na3V2(PO4)3@C//rGO@Fe2O3 NIB was assembled. The anode material was pre-activated first (Fig. S12 in Supporting information). The test parameters, such as energy density, power density and specific capacity, are calculated based on the total mass of the cathode and anode materials. It delivers a specific capacity of 115 mAh/g at 0.1 A/g (Figs. S13a and b in Supporting information). At a high specific current of 4 A/g and 20 A/g, the specific capacity is 43.7 mAh/g and 30 mAh/g. As shown in Fig. S13c (Supporting information), the Ragone plots of Na3V2(PO4)3@C//rGO@Fe2O3 NIB exhibit a maximum energy density of 265.3 Wh/kg. A high energy density of 53.7 Wh/kg could still be obtained after 36 s of discharging. Fig. S13d (Supporting information) shows the long-term cycling performance of the full cell. After 200 cycles, the specific capacity of the full cell is 32 mAh/g.

    In summary, a self-supporting 3D graphene network decorated with Fe2O3 nanocubes is rationally crafted, which can be directly used as an electrode without additional binders and conductive additives. Such unique hierarchical porous architecture consisted of rGO sheets enables spatial confinement to achieve outstanding results, which can not only accommodate volume changes of sodiated Fe2O3 nanocubes, but also ensure favorable transport kinetics of electrons and Na+. Consequently, the high DNa of rGO@Fe2O3 during the entire discharging/charging process is about 2.12 × 10−8/3.19 × 10−8 cm2/s, and the Na-ion storage in rGO@Fe2O3 sponge exhibits capacitive behavior. The rGO@Fe2O3 sponge anode delivers superior cycling performance (327.32 mAh/g after 2000 cycles) and superior rate capability (162 mAh/g at 20 A/g) for Na+ storage, which is better than the most reported results. This work provides a universal strategy for constructing a 3D conductive network to optimize the electrochemical performance of conversion-type electrodes.

    The authors declare that they have no known competing finan- cial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Jun Dong: Writing – review & editing, Writing – original draft, Visualization, Methodology, Investigation, Funding acquisition, Data curation, Conceptualization. Senyuan Tan: Writing – review & editing, Validation, Investigation, Data curation. Sunbin Yang: Methodology, Data curation. Yalong Jiang: Writing – review & editing, Supervision, Resources, Project administration, Funding acquisition, Conceptualization. Ruxing Wang: Validation, Resources, Project administration, Funding acquisition. Jian Ao: Validation. Zilun Chen: Validation. Chaohai Zhang: Validation. Qinyou An: Writing – review & editing, Supervision, Resources, Project administration, Funding acquisition, Conceptualization. Xiaoxing Zhang: Writing – review & editing, Supervision, Resources, Project administration, Funding acquisition, Conceptualization.

    This work is supported by National Natural Science Foundation of China (Nos. 52307239, 52102300, 52207234), the Natural Science Foundation of Hubei Province (Nos. 2022CFB1003, 2021CFA025).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110010.


    1. [1]

      Q. Wei, X. Chang, D. Butts, et al., Nat. Commun. 14 (2023) 7. doi: 10.1038/s41467-022-35617-3

    2. [2]

      C. Vaalma, D. Buchholz, M. Weil, et al., Nat. Rev. Mater. 3 (2018) 18013. doi: 10.1038/natrevmats.2018.13

    3. [3]

      L. Kitsu Iglesias, E.N. Antonio, T.D. Martinez, et al., Adv. Energy Mater. 13 (2023) 2302171. doi: 10.1002/aenm.202302171

    4. [4]

      T. Long, P. Chen, B. Feng, et al., Chin. Chem. Lett. 35 (2024) 109267. doi: 10.1016/j.cclet.2023.109267

    5. [5]

      Z. Li, G. Zhou, S. Li, et al., Joule 7 (2023) 2609–2621. doi: 10.1016/j.joule.2023.10.003

    6. [6]

      S. Qiao, Q. Zhou, M. Ma, et al., ACS Nano 17 (2023) 11220–11252. doi: 10.1021/acsnano.3c02892

    7. [7]

      Q. Li, H. Li, Q. Xia, et al., Nat. Mater. 20 (2021) 76–83. doi: 10.1038/s41563-020-0756-y

    8. [8]

      M. Zhang, X. Liu, J. Gu, et al., Chin. Chem. Lett. 34 (2023) 108471. doi: 10.1016/j.cclet.2023.108471

    9. [9]

      F. Kong, Y. Ge, S. Tao, et al., Chin. Chem. Lett. 35 (2023) 108552. doi: 10.1016/j.cclet.2023.108552

    10. [10]

      A. Yin, L. Yang, Z. Zhuang, et al., Energy Storage. 2 (2020) e132. doi: 10.1002/est2.132

    11. [11]

      B. Philippe, M. Valvo, F. Lindgren, et al., Chem. Mater. 26 (2014) 5028–5041. doi: 10.1021/cm5021367

    12. [12]

      N.T. Aristote, K. Zou, A. Di, et al., Chin. Chem. Lett. 33 (2022) 730–742. doi: 10.1016/j.cclet.2021.08.049

    13. [13]

      B. Papandrea, X. Xu, Y. Xu, et al., Nano Res. 9 (2016) 240–248. doi: 10.1007/s12274-016-1005-1

    14. [14]

      T. Li, A. Qin, L. Yang, et al., ACS Appl. Mater. Interfaces 9 (2017) 19900–19907. doi: 10.1021/acsami.7b04407

    15. [15]

      Y. Zhang, Q. Wang, K. Zhu, et al., Chem. Eng. J. 428 (2022) 131204. doi: 10.1016/j.cej.2021.131204

    16. [16]

      C. Zheng, X. Xu, Q. Lin, et al., Carbon N Y 176 (2021) 31–38. doi: 10.1016/j.carbon.2021.01.122

    17. [17]

      P. Ma, Z. Zhuang, J. Cao, et al., ACS Appl. Energy Mater. 5 (2022) 6417–6422. doi: 10.1021/acsaem.2c00860

    18. [18]

      Z. Zhuang, L. Yang, B. Ju, et al., Energy Storage 2 (2020) e109. doi: 10.1002/est2.109

    19. [19]

      A. Kormanyos, E. Kecsenovity, A. Honarfar, et al., Adv. Funct. Mater. 30 (2020) 2002124. doi: 10.1002/adfm.202002124

    20. [20]

      L. Shi, Y. Li, F. Zeng, et al., Chem. Eng. J. 356 (2019) 107–116. doi: 10.1016/j.cej.2018.09.018

    21. [21]

      M. Chen, D. Niu, J. Mao, et al., ACS Appl. Energy Mater. 4 (2021) 5888–5896. doi: 10.1021/acsaem.1c00691

    22. [22]

      H. Usui, Y. Domi, E. Iwama, et al., Mater. Chem. Phys. 272 (2021) 125023. doi: 10.1016/j.matchemphys.2021.125023

    23. [23]

      Q. Wang, Y. Ma, L. Liu, et al., Nanomaterials 10 (2020) 782. doi: 10.3390/nano10040782

    24. [24]

      Y. Chen, X. Yuan, C. Yang, et al., J. Alloys Compd. 777 (2019) 127–134. doi: 10.1145/3294109.3295645

    25. [25]

      T. Hou, X. Sun, D. Xie, et al., Chem. Eur. J. 24 (2018) 14786–14793. doi: 10.1002/chem.201802916

    26. [26]

      M. Fiore, G. Longoni, S. Santangelo, et al., Electrochim. Acta 269 (2018) 367–377. doi: 10.1016/j.electacta.2018.02.161

    27. [27]

      Z. Qiang, Y. Chen, B. Gurkan, et al., Carbon N Y 116 (2017) 286–293. doi: 10.1016/j.carbon.2017.01.093

    28. [28]

      G.D. Park, J.S. Cho, J.K. Lee, et al., Sci. Rep. 6 (2016) 22432. doi: 10.1038/srep22432

    29. [29]

      H. Li, L. Xu, H. Sitinamaluwa, et al., Compos. Commun. 1 (2016) 48–53. doi: 10.1038/nmat4465

    30. [30]

      D. Li, J. Zhou, X. Chen, et al., ACS Appl. Mater. Interfaces 8 (2016) 30899–30907. doi: 10.1021/acsami.6b09444

    31. [31]

      Z. Zhang, Y. Wang, S. Chou, et al., J. Power Sources 280 (2015) 107–113. doi: 10.1504/IJCNDS.2015.070290

    32. [32]

      Z. Wu, Y. Zhong, J. Liu, et al., J. Mater. Chem. A 3 (2015) 10092–10099. doi: 10.1039/C5TA01334H

    33. [33]

      Y. Fu, Q. Wei, X. Wang, et al., J. Mater. Chem. A 3 (2015) 13807–13818. doi: 10.1039/C5TA02994E

    34. [34]

      S. Wang, W. Wang, P. Zhan, et al., ChemElectroChem 1 (2014) 1636–1639. doi: 10.1002/celc.201402208

    35. [35]

      M. Valvo, F. Lindgren, U. Lafont, et al., J. Power Sources 245 (2014) 967–978. doi: 10.1016/j.jpowsour.2013.06.159

    36. [36]

      C. Li, H. Su, K. Zhang, et al., J. Alloys Compd. 858 (2021) 157714. doi: 10.1016/j.jallcom.2020.157714

    37. [37]

      Y. Chen, Z. Guo, B. Jian, et al., Nanomaterials 9 (2019) 1770. doi: 10.3390/nano9121770

    38. [38]

      S. Meng, D. Zhao, L. Wu, et al., J. Alloys Compd. 737 (2018) 130–135. doi: 10.1016/j.jallcom.2017.12.077

    39. [39]

      T. Guo, H. Liao, P. Ge, et al., Mater. Chem. Phys. 216 (2018) 58–63. doi: 10.1016/j.matchemphys.2018.05.054

    40. [40]

      M. Li, C. Ma, Q.C. Zhu, et al., Dalton Trans. 46 (2017) 5025–5032. doi: 10.1039/C7DT00540G

    41. [41]

      B. Sun, S.J. Bao, J.Le Xie, et al., RSC Adv. 4 (2014) 36815. doi: 10.1039/C4RA04686B

    42. [42]

      Z. Jian, B. Zhao, P. Liu, et al., Chem. Commun. 50 (2014) 1215–1217. doi: 10.1039/C3CC47977C

    43. [43]

      Y. Shen, Y. Jiang, Z. Yang, et al., Adv. Sci. 9 (2022) e2104504. doi: 10.1002/advs.202104504

    44. [44]

      Z. Zhuang, F. Zhang, D. Gandla, et al., ACS Appl. Mater. Interfaces 15 (2023) 38530–38539. doi: 10.1021/acsami.3c08766

    45. [45]

      Z. Zhuang, F. Zhang, Y. Zhou, et al., Materials Today Energy 30 (2022) 101192. doi: 10.1016/j.mtener.2022.101192

    46. [46]

      V. Augustyn, J. Come, M.A. Lowe, et al., Nat. Mater. 12 (2013) 518–522. doi: 10.1038/nmat3601

    47. [47]

      V. Augustyn, P. Simon and B. Dunn, Energy Environ. Sci. 7 (2014) 1597. doi: 10.1039/c3ee44164d

  • Figure 1  Schematic of the fabrication and structural characterization of rGO@Fe2O3 sponge. (a) Schematic illustration for the synthesis of rGO@Fe2O3 sponge. (b, c) SEM and TEM images of rGO@Fe2O3 sponge. (d, e) Nitrogen adsorption and desorption isotherm and pore-size distribution of rGO@Fe2O3 sponge.

    Figure 2  Electrochemical performances of rGO@Fe2O3 sponge and Fe2O3. (a) The galvanostatic charge-discharge curves of rGO@Fe2O3 sponge at 0.1 A/g. (b, c) CV curves of rGO@Fe2O3 and Fe2O3 at 0.2 mV/s at first three cycles. (d) Long-term cycles at 1 A/g of rGO@Fe2O3. (e) Cycling performance comparison of current work to the state of the art. (f) Rate performance of rGO@Fe2O3 and Fe2O3 at different current densities. (g) Rate performance comparison of current work to the state of the art.

    Figure 3  Electrochemical kinetic analysis. (a) GITT profiles of discharge and charge processes of rGO@Fe2O3 and Fe2O3 measured after three cycles at the current density of 0.02 A/g. (b, c) Na+ diffusion coefficients as a function of the states of charging process and Nyquist plots of rGO@ Fe2O3 and Fe2O3 at fully charged state. (d) CV curves of rGO@Fe2O3 at the scan rates ranging from 0.2 mV/s to 1 mV/s. (e)The normalized peak currents versus scan rate to determine the b value for anodic and cathodic peaks of rGO@Fe2O3. (f) The proportion of diffusion−controlled and capacitive contributions to the total charge storage at different scant rates.

    Figure 4  Morphological changes characterization of Fe2O3 particles. TEM images of rGO@Fe2O3 sponge (a) at fresh states and (b) fully discharged states for Na-ion storage. (c) The size distribution of Fe2O3 nanocubes in rGO@ Fe2O3 sponge at fresh and fully discharged state. SEM images of pure Fe2O3 particles (d) at fresh states and (e) fully discharged states. (f) The size distribution of bare Fe2O3 particles at fresh and fully discharged states. (g) Schematic of structure changes of rGO@Fe2O3 and Fe2O3 after sodiation.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  114
  • HTML全文浏览量:  7
文章相关
  • 发布日期:  2025-03-15
  • 收稿日期:  2024-04-17
  • 接受日期:  2024-05-13
  • 修回日期:  2024-04-25
  • 网络出版日期:  2024-05-14
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章