Next Article in Journal
A Colloidal-Quantum-Dot Integrated U-Shape Micro-Light-Emitting-Diode and Its Photonic Characteristics
Next Article in Special Issue
Composites of Titanium–Molybdenum Mixed Oxides and Non-Traditional Carbon Materials: Innovative Supports for Platinum Electrocatalysts for Polymer Electrolyte Membrane Fuel Cells
Previous Article in Journal
Fabrication of WO3 Quantum Dots with Different Emitting Colors and Their Utilization in Luminescent Woods
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Amorphous Fe2O3 Anchored on N-Doped Graphene with Internal Micro-Channels as an Active and Durable Anode for Sodium-Ion Batteries

1
Collaborative Innovation Center for Nanomaterials & Devices, College of Physics, Qingdao University, Qingdao 266071, China
2
State Key Laboratory of Bio-Fibers and Eco-Textiles, Qingdao University, Qingdao 266071, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2024, 14(11), 937; https://doi.org/10.3390/nano14110937
Submission received: 28 April 2024 / Revised: 20 May 2024 / Accepted: 24 May 2024 / Published: 27 May 2024

Abstract

:
The reduced graphene oxide (rGO) exhibits outstanding electrical conductivity and a high specific surface area, making it a promising material for various applications. Fe2O3 is highly desirable due to its significant theoretical capacity and cost-effectiveness, high abundance, and environmental friendliness. However, the performance of these r-GO/Fe2O3 composite electrodes still needs to be further improved, especially in terms of cycle stability. The composite of Fe2O3 anchored on N-doped graphene with inside micro-channels (Fe2O3@N-GIMC) was used to be efficiently prepared. Because the inside channels can furnish extra transmission pathways and absorption websites and the interconnected structure can efficaciously forestall pulverization and aggregation of electrode materials. In addition, N do** is also beneficial to improve its electrochemical performance. Thus, it demonstrates exceptional sodium storage characteristics, including notable electrochemical activity, impressive initial Coulombic efficiency, and favorable rate performance. The optimized Fe2O3@N-GIMC indicates outstanding discharge capacity (573.5 mAh g−1 at 1 A g−1), significant rate performance (333.6 mAh g−1 at 8 A g−1), and stable long-term cycle durability (308.9 mAh g−1 after 1000 cycles at 1 A g−1, 200.8 mAh g−1 after 4000 cycles at 1 A g−1) as a sodium-ion battery anode. This presents a new approach for preparing graphene-based high-functional composites and lays a stable basis for further expanding its application field.

1. Introduction

As lithium-ion battery (LIB) performance advances, its extensive use in handheld electronic devices, electric cars, and other fields has a profound impact on the development of these fields, thus leading the change in living habits and social development. However, due to the low lithium content in the crust and the uneven distribution, the cost of LIBs has increased rapidly in recent years [1,2]. Therefore, the development of new cheap secondary batteries based on the abundant elements of the earth has become the focus of recent scientific research and industry attention. Among them, sodium-ion batteries (SIBs) based on widely available and low-cost sodium have been widely studied because of their electrochemical working principles similar to LIBs [3,4,5,6,7]. However, due to reasons such as the sluggish electrochemical reaction kinetics and unfavorable structure disintegration caused by the greater ionic radius and higher atomic mass, it is often challenging to secure high specific capacity and superior cycle reliability, thus affecting practical applications [8,9,10,11]. Therefore, it is imperative to identify appropriate anode materials for SIBs with excellent performance through materials’ structural tailoring or exploration of new electrode design principles.
Graphite is a reliable and commonly used commercial anode material for LIBs, with good cyclability and safety [12,13,14]. However, because the radius of Na+ is larger than the interlayer spacing (d-spacing) and for other reasons, it is not an ideal anode material for SIBs [15]. Due to the excellent properties such as good conductivity, high specific surface area, and wide electrochemical potential window, reduced graphene oxide (r-GO) has been extensively studied in the field of electrode materials for rechargeable batteries [15,16,17]. In addition to participating in electrochemical reactions as an active material, its excellent electrical conductivity helps to improve kinetics, and good mechanical properties can support the structural integrity of the electrodes during charge discharge cycles. Recent reports indicate that the electrochemical performance of rechargeable batteries based on r-GO strongly depends on its morphology and pore size distribution [3,15,18,19,20]. In our earlier works [21,22], an r-GO film featuring a network of interconnected internal micro-channels was effectively fabricated by integrating electrospinning and electrospray techniques. The findings demonstrate that the LIB utilizing an r-GO film as the anode material exhibits outstanding performance, characterized by a high specific capacity and exceptional cycle stability. This would be attributed to the presence of inner channels, which provide improved transmission pathways and absorption websites, in addition to the interconnected structure that efficaciously mitigates pulverization and aggregation of electrode substances. However, the sodium storage performance of the r-GO film is unsatisfactory, especially due to the low specific capacity [21].
In addition, heteroatom-do** and metal oxide/sulfide composites are also effective ways to enhance the sodium storage capabilities of r-GO-based materials. Because it can improve the electrical conductivity and promote Na+ adsorption and insertion, various heteroatoms have been doped into r-GO layers to regulate properties, such as N, P, B, S, and F [3,23,24,25,26,27,28]. Meanwhile, numerous metal oxides/sulfides have been extensively researched as promising electrode materials in SIBs due to their high theoretical specific capacity [29,30,31,32]. Among them, Fe2O3 is of significant interests because of its high theoretical capacity, low cost, high abundance, and environmental friendliness [33,34,35,36,37,38,39,40,41]. In these studies, some interesting work is to compound Fe2O3 with r-GO to improve the specific capacity of the material and solve the problem of rapid capacity decay due to the poor electrical conductivity of Fe2O3 [33,35,36,37]. However, the performance of these r-GO/Fe2O3 composite electrodes still needs to be further improved, especially in terms of cycle stability.
In this investigation, the combination film of nylon nanofibers, GO flakes, and Fe2O3 was synthesized using a combination of electrospinning and electrospray techniques. Subsequently, the reduction of GO sheets and the breakdown of nylon nanofibers took place through the process of thermal reduction. The successful formation of the Fe2O3@N-GIMC composite, where Fe2O3 is anchored on N-doped graphene with internal micro-channels (Fe2O3@N-GIMC), was achieved in this investigation. The outstanding performance of the material can be attributed to the increased availability of transmission pathways and absorption sites facilitated by the internal channels, as well as the interconnected architecture, which effectively mitigates the pulverization and aggregation of electrode materials. In addition, N do** is also beneficial to improve its electrochemical performance. It exhibits superior sodium storage properties such as high electrochemical activity, high initial Coulombic efficiency, and good rate performance. The optimized Fe2O3@N-GIMC indicates staggering discharge capacity (573.5 mAh g−1 at 1 A g−1), significant rate of achievement (333.6 mAh g−1 at 8 A g−1), and resilient long-term cycle robustness (308.9 mAh g−1 after 1000 cycles at 1 A g−1, 200.8 mAh g−1 after 4000 cycles at 1 A g−1) as a sodium-ion batteries anode. This presents a novel approach to fabricating high-performance composites using graphene and establishes a firm basis for broadening its range of applications.

2. Materials and Methods

2.1. Materials

A commercially available dispersion of monolayer graphene oxide (GO) in N, N-Dimethylformamide (DMF) solvent was obtained from Gaoxi Technology (Hangzhou, China), with a concentration of 10 mg g−1. Detailed technical indicators are shown in Table 1. Before use, the dispersion was reduced to one-third of its initial concentration. Then iron acetylacetonate was added into the prepared graphene dispersion. Formic acid (88%) and iron acetylacetone were purchased from China National Pharmaceutical Corporation (Bei**g, China) without further purification. Nylon PA66 (262.35 g mol−1) was purchased from Sigma-Aldrich (St. Louis, MO, USA). Then formic acid was used as the solvent to prepare a 16% concentration of nylon solution.

2.2. Preparation of the Combined Film of GO Flakes Containing Iron Acetylacetonate and Nylon Nanofibers

Nylon nanofiber film was prepared on aluminum foil by a traditional electrospinning process, which was used as the base material for subsequent experiments. Then, the prepared foundation was securely attached to the roller’s surface, and the nylon solution and the graphene dispersion doped with iron acetylacetonate loaded into different syringes. The infusion rate of the syringes was regulated by individual injection pumps. In contrast, the needle tips of the syringes were linked to a high-voltage power supply capable of producing a DC voltage. The spacing between the needle tip and the drum surface was adjusted to 5 cm and 10 cm, respectively. GO sheets doped with iron acetylacetonate were deposited on the substrate at a feeding rate of 25 μL·min−1 and a voltage of 14 kV. At the same time, nylon nanofibers were uniformly dispersed onto the substrate using the electrospinning technique at a rate of 25 μL·min−1 and a voltage of 10 kV, while the roller rotated at a speed of 200 rpm. By combining electrospinning and electrospray technology, GO sheets doped with iron acetylacetonate and nylon fiber combined films were prepared successfully. In addition, there is a nylon fiber film on the substrate, which is conducive to the combined film removed from the aluminum foil.

2.3. Preparation of Fe2O3@N-GIMC and Fe2O3@G

The combined film was removed from the aluminum foil and placed in a tubular furnace. It was thermally annealed under an argon-hydrogen gas mixture to reduce the GO sheet and eliminate the nylon nanofibers, while the iron acetylacetonate was thermally decomposed. During the thermal reduction process, the combined film was slowly heated from room temperature to 500 °C at a pace of 2 °C per minute and maintained at this temperature for 240 min. It was allowed to cool to room temperature and the film removed. After annealing, there was no nylon fiber in the combined film, so the composite of Fe2O3 anchored on N-doped graphene with internal microchannels (Fe2O3@N-GIMC) was prepared. If no nylon fiber was added in the preparation process, the composite of Fe2O3 anchored on graphene with no internal micro-channels (Fe2O3@G) could be prepared by repeating the spinning and thermal annealing procedures above. According to the mass ratio of iron acetylacetonate (m) and GO dispersion (n) in the precursor solution, the prepared composite material was labeled as Fe2O3@N-GIMC_m:n and the composite material prepared by not incorporating nylon fiber is labeled as Fe2O3@G_m:n.

2.4. Characterization

The morphologies of the graphene oxide (GO), reduced graphene oxide (rGO), Fe2O3@G, and Fe2O3@N-GIMC were analyzed employing a field emission scanning electron microscope (SEM) (FE-SEM, JEOL JSM-7800F, JEOL, Tokyo, Japan). Their compositions were determined via X-ray photoelectron spectroscopy (XPS) (Thermo Escalab 250, Thermo Fisher Scientific, Waltham, MA, USA). The structure was characterized by X-ray diffraction (XRD) (Japan, Smart Lab 3KW, Rigaku, Tokyo, Japan). The morphology was characterized by a Transmission Electron Microscope (TEM) (Japan Corporation, JEM 2100F, JEOL). The types and contents of micro-component elements were characterized by an Energy Dispersive Spectrometer (EDS) (Thermo Fishe, ESCALAB **, the size of Fe2O3 phases in Fe2O3@N-GIMC_1:2 is almost the same as Fe2O3@N-GIMC_1:4, and the Fe2O3 phases of both is uniformly dispersed. Still, the Fe2O3 phases density of 1:2 is greater on the graphene sheet per unit area. As the do** amount of iron acetylacetonate continues to increase, there is a serious aggregation of Fe2O3 phases in Fe2O3@N-GIMC_1:1, which also causes the size of Fe2O3 phases to grow to 210 nm.
X-ray Photoelectron Spectroscopy (XPS) measurements were meticulously employed to gain profound insights into the surface chemistry of Fe2O3@N-GIMC, effectively unveiling a noteworthy nitrogen weight content of up to 5.65%, which represents a moderate do** concentration (Figure 3a). As illustrated in Figure 3b, the C1s spectrum of GF is mainly composed of three kinds of bonds: C-C (284.8 eV), C-O-C (285.88 eV), and O-C=O (288.5 eV). This shows that the C element has been reduced [42]. To further prove the interaction between the keys, the O1s map was analyzed, as shown in Figure 3c. It is mainly composed of five types of keys: Fe-O (530.37 eV), C=O (531.35 eV), C-O-Fe (532.29 eV), C-O/O=C-O (533.53 eV), Fe-O (530.37 eV), which more indirectly proved the presence of iron trioxide in the composite material. The existence of C-O-Fe (532.29 eV) chemical bonds is compelling evidence for the robust interconnection between iron trioxide and graphene [37]. The Fe2p spectrum and the results of their fitting are displayed in Figure 3d. The spectrum revealed the presence of doublet Fe2p3/2 and Fe2p1/2 with binding energies of 711.25 eV and 724.7 eV, correspondingly. The Fe2p3/2 peak was also accompanied by a satellite peak positioned at 720.03 eV, which was indicative of α-Fe2O3 and aligned closely with the values reported in existing literature [48]. It indicates the presence of ferric oxide in r-GO, indirectly proving that the reaction was successful [37]. Most notably, XPS spectra demonstrate the robust interfacial connection between Fe2O3 and r-GO in the r-GO/Fe2O3 composite material. Figure 3e shows that the N1s spectrum consists of three peaks at 398.71 eV, 399.07 eV, and 400.96 eV, corresponding to pyridine-N, pyrrole-N, and graphite-N, respectively [49]. This shows that N do** of graphene sheets has been successfully achieved.
To investigate the sodium-ion insertion/extraction mechanisms in the Fe2O3@N-GIMC_1:2 electrode, cyclic voltammetry (CV) was employed, as depicted in Figure 4a. In the primary cycle, two distinct reduction peaks at around ~0.01 V and 0.85 V are associated with Na penetration into Fe2O3@N-GIMC_1:2 and the creation of the SEI layers, respectively. Following the vanishing of the peak at 0.85 V, indicative of the establishment of enduring solid electrolyte interphase (SEI) layers, enhanced cyclic durability is evidenced [50]. In the anodic CV curves of Fe2O3@N-GIMC_1:2, two discernible peaks at approximately 0.08 V and 1.37 V are observed, corresponding to the extraction of sodium from Fe2O3@N-GIMC_1:2 and the corrosion of certain reversible solid electrolyte interface components. In the second CV curve, the cathode peaks at 0.76 and 0.92 V are accompanied by anode peaks at 0.72 and 1.36 V, which correspond to the reversible redox reaction of Fe3+↔Fe2+, Fe2+↔Fe0, and Fe dissolution [51,52]. The cyclic voltammetry plots exhibit almost perfect overlap from the second cycle onwards, underscoring the exceptional electrochemical reversibility of the Fe2O3@N-GIMC_1:2 electrode.
The Fe2O3@N-GIMC_1:2 electrode demonstrates remarkable cycling stability and high Coulombic efficiency, as illustrated in Figure 4b. Following 1000 cycles at a current density of 1 A g−1, the discharge capacity retention reached 67%, surpassing the values of 65% and 48% observed for Fe2O3@N-GIMC_1:1 and Fe2O3@N-GIMC_1:4, respectively. This is because, since a small amount of Fe2O3 do** can only slightly improve the electrochemical performance, while excessive do** causes Fe2O3 agglomeration (this is confirmed in Figure 2g). So, the excellent electrochemical performance of Fe2O3@N-GIMC_1:2 should be attributed to the right amount of Fe2O3 do**. Moreover, after undergoing 1000 cycles at a current density of 1 A g−1, Fe2O3@N-GIMC_1:2 demonstrates an impressive discharge capacity retention of 67%, outperforming Fe2O3@G_1:2 which only maintains 53% of its initial discharge capacity. Similarly, after enduring 4500 cycles at the same current density, Fe2O3@N-GIMC_1:2 exhibits a remarkable discharge capacity retention of 51%, while Fe2O3@G_1:2 only manages to retain 45%. These results highlight the excellent long-term cycle stability of Fe2O3@N-GIMC_1:2. The Fe2O3@N-GIMC_1:2 demonstrates remarkable discharge capacity (573.5 mAh g−1 at 1 A g−1) as a sodium-ion batteries anode (Figure 4c), and that of Fe2O3@G_1:2 only 424.6 mAh g−1 (Figure 4d). This phenomenon can be attributed to the enhanced provision of ways of transmission and sites of absorption by the internal channels. At the same time, the interconnected architecture effectively reduces the pulverization and aggregation of electrode-related materials. Furthermore, N do** is also beneficial to improve its electrochemical performance. In Figure 4e, the rate performance of the two samples at various current densities is presented. The discharge-specific capacities of Fe2O3@N-GIMC_1:2 at current densities of 0.2, 0.5, 1, 2, 3, 5, and 8 A g−1 are measured at 440.6 mAh g−1, 430.3 mAh g−1, 408.0 mAh g−1, 389.7 mAh g−1, 374.1 mAh g−1, 354.3 mAh g−1, and 333.6 mAh g−1, respectively. Upon returning the current density to 0.2 A g−1, the discharge specific capacities of Fe2O3@N-GIMC_1:2 could revert to the initial level of 454.2 mAh g−1, demonstrating excellent tolerance to high currents. Overall, in comparison to Fe2O3@G_1:2, the rate performance of Fe2O3@N-GIMC_1:2 exhibits superiority over Fe2O3@G_1:2. Moreover, the Fe2O3@N-GIMC_1:2 electrode still maintains a high specific discharge capacity after a long cycle of 4000 cycles (Figure 4f). In conclusion, despite an equivalent Fe do** level in Fe2O3@N-GIMC_1:2 and Fe2O3@G_1:2, the Fe2O3@N-GIMC_1:2 electrode demonstrates superior initial coulombic efficiency, exceptional reversibility, improved rate performance and excellent cycling stability in sodium-ion batteries. On one side, the three-dimensional interconnected structure can effectively prevent fragmentation and clustering of electrode materials because of their exceptional structural integrity. It leads to fast ion transport and efficient use of active materials, thereby enhancing the speed performance and longevity. On the flip side, the multiple hollow channels can offer additional routes for ion transmission and sites for absorption of sodium ions, as well as reduce transmission distance, resulting in a higher capacity.
To comprehend the reaction speed as well as the reversible capabilities and extended cycle longevity of the Fe2O3@N-GIMC_1:2, cyclic voltammetry plots at different scanning speeds were gathered. (Figure 5a). The scan rates are increased from 0.2 to 3 mV s−1, similar outlines and slight peak shifts at different scan rates are illustrated by the excellent reversibility and small polarizations of Fe2O3@N-GIMC_1:2 in Figure 5a. Furthermore, based on the cyclic voltammetry plots, the correlation between the peak current (i) and the scan rate (v) complies with Equation (1):
i = avb
where a is a constant, and b can be ascertained by the gradients of log (i) versus log (v). If b equals 0.5, it signifies that the electrochemical process is governed by a diffusion-controlled mechanism; if b equals 1, the electrochemical process is dictated by surface-responsive capacity behavior. The b-values measured for the Fe2O3@N-GIMC_1:2 electrode (Figure 5b) is located within a range between 0.79 and 0.92, Their b values were all near 1, suggesting a surface capacitance-driven behavior, resulting in a swift Na+ intercalation/deintercalation process and excellent rate capability.
The internal energy storage mechanism of Fe2O3@N-GIMC_1:2 as the anode material for sodium-ion batteries (SIBs) was further elucidated through electrochemical impedance spectroscopy (EIS) analysis, as shown in Figure 5c. The Nyquist plot is characterized by the presence of a semicircular arc intersecting the graph, along with an accompanying linear segment. The semicircular region observed in the high-frequency domain of the Nyquist plot signifies the electron transfer-limited mechanism, while the linear portion evident in the low-frequency range denotes the diffusion-limited process. In the faster electron transport process, the AC impedance spectrum contains only a straight line portion, while the slower charge transport process has a larger semicircular region. The diameter of the semicircle corresponds to the charge transfer resistance (Rct), while the intercept (Zre) of the semicircle on the real axis represents the electrolyte resistance (Re). The fitting results are shown in Table 2 according to the equivalent circuit diagram. The Rct of the Fe2O3@N-GIMC_1:2 electrode is lower than Fe2O3@G_1:2 electrode. This suggests a more rapid charge transfer between the Fe2O3@N-GIMC_1:2 electrode surface. This shows that in the electrochemical process of the negative electrode of SIB, the channel of Fe2O3@N-GIMC_1:2 can shorten the ion diffusion distance and accelerate the ion diffusion, and the do** of N atom is conducive to improving the conductivity and electrochemical activity of Fe2O3@N-GIMC_1:2. Moreover, to reveal the detailed reaction kinetics for the Na+ intercalation/deintercalation processes, the galvanostatic intermittent titration technique (GITT) was performed (Figure 5d). The diffusion coefficient (D) of Fe2O3@N-GIMC_1:2 can be calculated from the GITT potential profiles using Fick’s second law with the following Equation (2):
D = 4 Π τ ( m B M b V M S ) 2 ( Δ E s Δ E t ) 2
where τ represents the duration of the current pulse; mB represents the mass loading of the electrode material; S represents the geometric area of the electrode; ∆Es is the quasi-thermodynamic equilibrium potential difference between before and after the current pulse; ∆Et is the potential difference during the current pulse; VM is the molar volume of the materials; and Mb is the molar mass of Fe2O3@N-GIMC_1:2 materials and the results are plotted as a function of electrochemical potential. As shown in Figure S2, Fe2O3@N-GIMC_1:2 delivers a high D-value at almost all potentials, resulting from reinforced effects of the abundant heterointerfaces, porous, and interconnected C skeleton.

4. Conclusions

In summary, the synthesis of Fe2O3 anchored onto nitrogen-doped graphene with internal micro-channels (Fe2O3@N-GIMC) was effectively achieved. The synthesized Fe2O3@N-GIMC_1:2 demonstrates remarkable discharge capacity (573.5 mAh g−1 at 1 A g−1), notable rate capability (333.6 mAh g−1 at 8 A g−1), and consistent long-term cycle stability (308.9 mAh g−1 after 1000 cycles at 1 A g−1, 200.8 mAh g−1 after 4000 cycles at 1 A g−1) as a sodium-ion battery anode. Following 1000 cycles under a current density of 1 A g−1, the Fe2O3@N-GIMC_1:2 composite retained 67% of its discharge capacity, while the Fe2O3@G_1:2 composite exhibited a lower capacity retention of 53%. This phenomenon can be ascribed to the enhanced availability of transmission pathways and absorption sites facilitated by the internal channels, alongside the interconnected architecture’s efficacy in mitigating electrode material pulverization and aggregation. In addition, N do** is also beneficial to improve its electrochemical performance. This offers a novel approach for producing graphene-based advanced composites and establishes a basis for broadening their applications.

Supplementary Materials

The following supporting information can be downloaded at: https://mdpi.longhoe.net/article/10.3390/nano14110937/s1, Figure S1: The Energy Dispersive Spectrometer (EDS) of Fe2O3@N-GIMC_1:4, Fe2O3@N-GIMC_1:2 and Fe2O3@N-GIMC_1:1; Figure S2: Diffusion coefficients of Fe2O3@N-GIMC_1:2.

Author Contributions

Conceptualization, W.H. and Y.L.; methodology, L.L. (Lin Li), W.H., and H.L.; investigation, L.L. (Lin Li), W.H., and X.Y.; data curation, L.L. (Lin Li), H.L., and L.L. (Linxin Liu); writing—original draft preparation, L.L. (Lin Li) and W.H.; writing—review and editing, L.L. (Lin Li) and W.H.; funding acquisition, Y.L. and W.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Shandong Provincial Natural Science Foundation (ZR2021MA029), the National Natural Science Foundation of China (11604173), and Shandong Province Introduction of Top Talents (Team) “One Thing, One Discussion” (DC1900013728).

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Goodenough, J.B.; Park, K.-S. The Li-Ion Rechargeable Battery: A Perspective. J. Am. Chem. Soc. 2013, 135, 1167–1176. [Google Scholar] [CrossRef] [PubMed]
  2. Pan, H.; Hu, Y.-S.; Chen, L. Room-temperature stationary sodium-ion batteries for large-scale electric energy storage†. Energy Environ. Sci. 2013, 6, 2338–2360. [Google Scholar] [CrossRef]
  3. Xu, J.; Wang, M.; Wickramaratne, N.P.; Jaroniec, M.; Dou, S.; Dai, L. High-Performance Sodium Ion Batteries Based on a 3D Anode from Nitrogen-Doped Graphene Foams. Adv. Mater. 2015, 27, 2042–2048. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, T.; Liu, Y.; Pan, L.; Lu, T.; Yao, Y.; Sun, Z.; Chua, D.H.C.; Chen, Q. Electrospun carbon nanofibers as anode materials for sodium ion batteries with excellent cycle performance. J. Mater. Chem. A 2014, 2, 4117–4121. [Google Scholar] [CrossRef]
  5. Ould, D.M.C.; Menkin, S.; O'Keefe, C.A.; Coowar, F.; Barker, J.; Grey, C.P.; Wright, D.S. Beyond the Norm: Synthesis and Electrochemical Study of High Concentrated NaPF6 Electrolytes. ECS Meet. Abstr. 2022; MA2022-01, 498. [Google Scholar]
  6. ** Endow TiO2 with Extraordinary Sodium Storage Performance. Adv. Mater. 2018, 30, 1801013. [Google Scholar] [CrossRef] [PubMed]
  7. Zhao, Y.; Wang, F.; Wang, C.; Wang, S.; Wang, C.; Zhao, Z.; Duan, L.; Liu, Y.; Wu, Y.; Li, W.; et al. Encapsulating highly crystallized mesoporous Fe3O4 in hollow N-doped carbon nanospheres for high-capacity long-life sodium-ion batteries. Nano Energy 2018, 56, 426–433. [Google Scholar] [CrossRef]
  8. **g, P.; Wang, Q.; Wang, B.; Gao, X.; Zhang, Y.; Wu, H. Encapsulating yolk-shell FeS2@carbon microboxes into interconnected graphene framework for ultrafast lithium/sodium storage. Carbon 2020, 159, 366–377. [Google Scholar] [CrossRef]
  9. Li, T.; Qin, A.; Yang, L.; Chen, J.; Wang, Q.; Zhang, D.; Yang, H. In Situ Grown Fe2O3 Single Crystallites on Reduced Graphene Oxide Nanosheets as High Performance Conversion Anode for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2017, 9, 19900–19907. [Google Scholar] [CrossRef] [PubMed]
  10. Cheng, D.; Ye, L.; Wei, A.; Xu, G.; Cao, Z.; Zhu, P.; Chen, Y. Constructing SnS/Fe2O3 heterostructure anchored on few-layered graphene as an ion-adsorption/diffusion enhancer for ultrafast and cycle-stable sodium storage. Chem. Eng. J. 2022, 457, 141243. [Google Scholar] [CrossRef]
  11. Zheng, C.; Xu, X.; Lin, Q.; Chen, Y.; Guo, Z.; Jian, B.; Li, N.; Zhang, H.; Lv, W. Confined growth of Fe2O3 nanoparticles by holey graphene for enhanced sodium-ion storage. Carbon 2021, 176, 31–38. [Google Scholar] [CrossRef]
  12. Zhang, Y.; Wang, Q.; Zhu, K.; Ye, K.; Wang, G.; Cao, D.; Yan, J. Edge sites-driven accelerated kinetics in ultrafine Fe2O3 nanocrystals anchored graphene for enhanced alkali metal ion storage. Chem. Eng. J. 2021, 428, 131204. [Google Scholar] [CrossRef]
  13. Li, D.; Zhou, J.; Chen, X.; Song, H. Amorphous Fe2O3/Graphene Composite Nanosheets with Enhanced Electrochemical Performance for Sodium-Ion Battery. ACS Appl. Mater. Interfaces 2016, 8, 30899–30907. [Google Scholar] [CrossRef] [PubMed]
  14. Sharifi, T.; Gracia-Espino, E.; Reza Barzegar, H.; Jia, X.; Nitze, F.; Hu, G.; Nordblad, P.; Tai, C.-W.; Wågberg, T. Formation of nitrogen-doped graphene nanoscrolls by adsorption of magnetic γ-Fe2O3 nanoparticles. Nat. Commun. 2013, 4, 2319. [Google Scholar] [CrossRef] [PubMed]
  15. Wang, L.; Wei, Z.; Mao, M.; Wang, H.; Li, Y.; Ma, J. Metal oxide/graphene composite anode materials for sodium-ion batteries. Energy Storage Mater. 2019, 16, 434–454. [Google Scholar] [CrossRef]
  16. Liu, Y.-T.; Zhang, P.; Sun, N.; Anasori, B.; Zhu, Q.-Z.; Liu, H.; Gogotsi, Y.; Xu, B. Self-Assembly of Transition Metal Oxide Nanostructures on MXene Nanosheets for Fast and Stable Lithium Storage. Adv. Mater. 2018, 30, 1707334. [Google Scholar] [CrossRef]
  17. Li, J.; Wang, Y.; Xu, W.; Wang, Y.; Zhang, B.; Luo, S.; Zhou, X.; Zhang, C.; Gu, X.; Hu, C. Porous Fe2O3 nanospheres anchored on activated carbon cloth for high-performance symmetric supercapacitors. Nano Energy 2018, 57, 379–387. [Google Scholar] [CrossRef]
  18. Yan, J.-X.; Leng, Y.-C.; Guo, Y.-N.; Wang, G.-Q.; Gong, H.; Guo, P.-Z.; Tan, P.-H.; Long, Y.-Z.; Liu, X.-L.; Han, W.-P. Highly Conductive Graphene Paper with Vertically Aligned Reduced Graphene Oxide Sheets Fabricated by Improved Electrospray Deposition Technique. ACS Appl. Mater. Interfaces 2019, 11, 10810–10817. [Google Scholar] [CrossRef]
  19. Gong, H.; Li, M.-F.; Yan, J.-X.; Lin, M.-L.; Liu, X.-L.; Sun, B.; Tan, P.-H.; Long, Y.-Z.; Han, W.-P. Highly conductive, flexible and functional multi-channel graphene microtube fabricated by electrospray deposition technique. J. Mater. Sci. 2019, 54, 14378–14387. [Google Scholar] [CrossRef]
  20. Han, W.; Wu, Y.; Gong, H.; Liu, L.; Yan, J.; Li, M.; Long, Y.; Shen, G. Reliable sensors based on graphene textile with negative resistance variation in three dimensions. Nano Res. 2021, 14, 2810–2818. [Google Scholar] [CrossRef]
  21. Yijun, W.; Linxin, L.; Sai, W.; Tong, Z.; Yun-Ze, L.; Wenpeng, H. Preparation of stretchable and porous graphene paper via functionalized with diol oligomer. J. Mater. Sci. Mater. Electron. 2023, 34, 636. [Google Scholar]
  22. Nai, J.; Kang, J.; Guo, L. Tailoring the shape of amorphous nanomaterials: Recent developments and applications. Sci. China Mater. 2015, 58, 44–59. [Google Scholar] [CrossRef]
  23. 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. [Google Scholar] [CrossRef] [PubMed]
  24. Cheng, Z.; Xu, J.; Zhong, H.; Yang, Y.; Chen, W. Self-assembly of α-Fe2O3 nanoparticles into rice ball shaped microstructures via a pyrolytic route. Superlattices Microstruct. 2010, 47, 253–258. [Google Scholar] [CrossRef]
  25. Nair, A.K.; Elizabeth, I.; Gopukumar, S.; Thomas, S.; M. S, K.; Kalarikkal, N. Nitrogen doped graphene—Silver nanowire hybrids: An excellent anode material for Lithium ion batteries. Appl. Surf. Sci. 2018, 428, 1119–1129. [Google Scholar] [CrossRef]
  26. Wang, G.; Shen, X.; Yao, J.; Park, J. Graphene nanosheets for enhanced lithium storage in lithium ion batteries. Carbon 2009, 47, 2049–2053. [Google Scholar] [CrossRef]
  27. Kong, D.; Cheng, C.; Wang, Y.; Liu, B.; Huang, Z.; Yang, H.Y. Seed-assisted growth of α-Fe2O3 nanorod arrays on reduced graphene oxide: A superior anode for high-performance Li-ion and Na-ion batteries†. J. Mater. Chem. A 2016, 4, 11800–11811. [Google Scholar] [CrossRef]
  28. Yang, Y.; Fu, W.; Lee, D.C.; Bell, C.; Drexler, M.; Ma, Z.F.; Magasinski, A.; Yushin, G.; Alamgir, F.M. Porous FeP/C composite nanofibers as high-performance anodes for Li-ion/Na-ion batteries. Mater. Today Energy 2020, 16, 100410. [Google Scholar] [CrossRef]
Figure 1. (a) Preparation of composite materials. The scanning electron microscope (SEM) of Fe2O3@G (b) and Fe2O3@N-GIMC (c,d); (e) the X-ray diffraction (XRD) of Fe2O3@N-GIMC and r-GO.
Figure 1. (a) Preparation of composite materials. The scanning electron microscope (SEM) of Fe2O3@G (b) and Fe2O3@N-GIMC (c,d); (e) the X-ray diffraction (XRD) of Fe2O3@N-GIMC and r-GO.
Nanomaterials 14 00937 g001
Figure 2. (ac) SEM image of Fe2O3@N-GIMC; (d) High Resolution Transmission Electron Microscope (HRTEM) image of Fe2O3@N-GIMC; (eg) Transmission Electron Microscope (TEM) images of Fe2O3@N-GIMC.
Figure 2. (ac) SEM image of Fe2O3@N-GIMC; (d) High Resolution Transmission Electron Microscope (HRTEM) image of Fe2O3@N-GIMC; (eg) Transmission Electron Microscope (TEM) images of Fe2O3@N-GIMC.
Nanomaterials 14 00937 g002
Figure 3. X-ray Photoelectron Spectroscopy (XPS) spectra of Fe2O3@N-GIMC. (a) XPS survey spectra; (b) C1s; (c) O1s; (d) Fe2p; (e) N1s.
Figure 3. X-ray Photoelectron Spectroscopy (XPS) spectra of Fe2O3@N-GIMC. (a) XPS survey spectra; (b) C1s; (c) O1s; (d) Fe2p; (e) N1s.
Nanomaterials 14 00937 g003
Figure 4. (a) CV plots at a scan rate of 0.1 mV s−1 for Fe2O3@N-GIMC_1:2; (b) cycle stability and electrochemical efficiency for Fe2O3@N-GIMC_1:4, Fe2O3@N-GIMC_1:2, Fe2O3@N-GIMC_1:1, and Fe2O3@G_1:2 cycled at a current of 1 A g−1. Initial five discharge–charge curves for Fe2O3@N-GIMC_1:2; (c) and Fe2O3@G_1:2; (d) at 1 A g−1; (e) rate capability of Fe2O3@N-GIMC_1:2 and Fe2O3@G_1:2; (f) cycling performance for Fe2O3@N-GIMC_1:2 cycled at a current of 1 A g−1.
Figure 4. (a) CV plots at a scan rate of 0.1 mV s−1 for Fe2O3@N-GIMC_1:2; (b) cycle stability and electrochemical efficiency for Fe2O3@N-GIMC_1:4, Fe2O3@N-GIMC_1:2, Fe2O3@N-GIMC_1:1, and Fe2O3@G_1:2 cycled at a current of 1 A g−1. Initial five discharge–charge curves for Fe2O3@N-GIMC_1:2; (c) and Fe2O3@G_1:2; (d) at 1 A g−1; (e) rate capability of Fe2O3@N-GIMC_1:2 and Fe2O3@G_1:2; (f) cycling performance for Fe2O3@N-GIMC_1:2 cycled at a current of 1 A g−1.
Nanomaterials 14 00937 g004
Figure 5. (a) CV curves of Fe2O3@N-GIMC_1:2 at various scan rates from 0.2 to 3 mV s−1; (b) log (i) versus log (v) plots; (c) electrochemical impedance spectroscopy (EIS) of the Fe2O3@N-GIMC_1:2 electrode and Fe2O3@G_1:2 electrode; (d) galvanostatic intermittent titration technique (GITT) potential profiles of Fe2O3@N-GIMC_1:2 and Fe2O3@G_1:2 electrode.
Figure 5. (a) CV curves of Fe2O3@N-GIMC_1:2 at various scan rates from 0.2 to 3 mV s−1; (b) log (i) versus log (v) plots; (c) electrochemical impedance spectroscopy (EIS) of the Fe2O3@N-GIMC_1:2 electrode and Fe2O3@G_1:2 electrode; (d) galvanostatic intermittent titration technique (GITT) potential profiles of Fe2O3@N-GIMC_1:2 and Fe2O3@G_1:2 electrode.
Nanomaterials 14 00937 g005
Table 1. Technical specifications of GO.
Table 1. Technical specifications of GO.
Model NumberSingle Layer RateTransverse Size DistributionParticle Size (D50)Viscosity
(cp, 10 mg/g)
GO2-DMF>99%20–30 μm1.9 ± 0.2 μm4800 ± 8000
Table 2. EIS fitting results of Fe2O3@N-GIMC_1:2 electrode and Fe2O3@G_1:2 electrode before cycling.
Table 2. EIS fitting results of Fe2O3@N-GIMC_1:2 electrode and Fe2O3@G_1:2 electrode before cycling.
Re/ΩRct/ΩCPE1-T/ΩCPE1-P/ΩW1-R/ΩW1-T/ΩW1-P/Ω
Fe2O3@N-GIMC_1:215.418420.0000000870.87118218.110.52
Fe2O3@G_1:2 14.120870.0000000450.64180317.200.49
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, L.; Li, H.; Liu, L.; Yan, X.; Long, Y.; Han, W. Amorphous Fe2O3 Anchored on N-Doped Graphene with Internal Micro-Channels as an Active and Durable Anode for Sodium-Ion Batteries. Nanomaterials 2024, 14, 937. https://doi.org/10.3390/nano14110937

AMA Style

Li L, Li H, Liu L, Yan X, Long Y, Han W. Amorphous Fe2O3 Anchored on N-Doped Graphene with Internal Micro-Channels as an Active and Durable Anode for Sodium-Ion Batteries. Nanomaterials. 2024; 14(11):937. https://doi.org/10.3390/nano14110937

Chicago/Turabian Style

Li, Lin, Hui Li, Linxin Liu, Xunchang Yan, Yunze Long, and Wenpeng Han. 2024. "Amorphous Fe2O3 Anchored on N-Doped Graphene with Internal Micro-Channels as an Active and Durable Anode for Sodium-Ion Batteries" Nanomaterials 14, no. 11: 937. https://doi.org/10.3390/nano14110937

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

Article Metrics

Back to TopTop