Next Article in Journal
Facile Synthesis of ZnO/WO3 Nanocomposite Porous Films for High-Performance Gas Sensing of Multiple VOCs
Previous Article in Journal
H2 Plasma and PMA Effects on PEALD-Al2O3 Films with Different O2 Plasma Exposure Times for CIS Passivation Layers
Previous Article in Special Issue
Bimetallic MOFs-Derived Hollow Carbon Spheres Assembled by Sheets for Sodium-Ion Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Coupling of Mn2O3 with Heteroatom-Doped Reduced Graphene Oxide Aerogels with Improved Electrochemical Performances for Sodium-Ion Batteries

by
Nor Fazila Mahamad Yusoff
1,
Nurul Hayati Idris
1,*,
Muhamad Faiz Md Din
2,
Siti Rohana Majid
3,
Noor Aniza Harun
4 and
Lukman Noerochim
5,*
1
Energy Storage Research Group, Faculty of Ocean Engineering Technology and Informatics, Universiti Malaysia Terengganu, Kuala Nerus 21300, Terengganu, Malaysia
2
Department of Electrical and Electronic Engineering, Faculty of Engineering, National Defence University of Malaysia, Kem Sungai Besi, Kuala Lumpur 57000, Malaysia
3
Center for Ionics University of Malaya, Department of Physics, Faculty of Science, University of Malaya, Kuala Lumpur 50603, Malaysia
4
Advance Nano Materials (ANOMA) Research Group, Faculty of Science and Marine Environment, Universiti Malaysia Terengganu, Kuala Nerus 21300, Terengganu, Malaysia
5
Department of Materials and Metallurgical Engineering, Institut Teknologi Sepuluh Nopember, Surabaya 60111, Indonesia
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(4), 732; https://doi.org/10.3390/nano13040732
Submission received: 29 December 2022 / Revised: 7 February 2023 / Accepted: 13 February 2023 / Published: 15 February 2023
(This article belongs to the Special Issue Carbon Nanomaterials for Electrochemical Energy Storage)

Abstract

:
Currently, efforts to address the energy needs of large-scale power applications have expedited the development of sodium–ion (Na–ion) batteries. Transition-metal oxides, including Mn2O3, are promising for low-cost, eco-friendly energy storage/conversion. Due to its high theoretical capacity, Mn2O3 is worth exploring as an anode material for Na-ion batteries; however, its actual application is constrained by low electrical conductivity and capacity fading. Herein, we attempt to overcome the problems related to Mn2O3 with heteroatom-doped reduced graphene oxide (rGO) aerogels synthesised via the hydrothermal method with a subsequent freeze-drying process. The cubic Mn2O3 particles with an average size of 0.5–1.5 µm are distributed to both sides of heteroatom-doped rGO aerogels layers. Results indicate that heteroatom-doped rGO aerogels may serve as an efficient ion transport channel for electrolyte ion transport in Mn2O3. After 100 cycles, the electrodes retained their capacities of 242, 325, and 277 mAh g−1, for Mn2O3/rGO, Mn2O3/nitrogen-rGO, and Mn2O3/nitrogen, sulphur-rGO aerogels, respectively. Doping Mn2O3 with heteroatom-doped rGO aerogels increased its electrical conductivity and buffered volume change during charge/discharge, resulting in high capacity and stable cycling performance. The synergistic effects of heteroatom doping and the three-dimensional porous structure network of rGO aerogels are responsible for their excellent electrochemical performances.

1. Introduction

Sodium-ion (Na-ion) batteries have good prospects as a replacement for lithium-ion (Li-ion) batteries used in electronic devices due to their low cost, vast quantity of sodium resources, and high energy density batteries, which are comparable to those of Li-ion energy storage devices [1,2,3]. Na-ion energy storage device deployment remains hampered by issues with rate performance, Coulombic efficiency, and cycle stability, among other factors. The main reason for this is that Na+ has a larger radius (0.102 nm) than Li+ (0.076 nm), resulting in slower reaction kinetics for the sodiation/desodiation of Na+ [4]. Ideally, anode materials should have microscopic internal structures that can store Na+ and enable a reversible process for sodiation/desodiation during charge/discharge. Additionally, they should be inexpensive, simple to prepare, and compatible with electrolytes. As a result, the primary challenge for Na-ion battery devices is to combine excellent anode materials with optimal electrolytes.
Because of their abundant source and high energy density, transition-metal oxides are promising anode materials for Na–ion batteries [5,6,7]. Due to their high theoretical capacity, low cost, and eco-friendliness, manganese oxides have garnered a great deal of attention [8,9]. Unfortunately, like other transition-metal oxides, their low conductivity and severe volume expansion during the charge/discharge process lead to poor electrochemical performances [10,11], limiting their practical application despite their aforementioned benefits. Several works have been reported on MnO2 [12,13] and Mn3O4 [14,15]; however, their electrochemical performances as an anode for Na-ion batteries are unsatisfactory. Due to its high theoretical specific capacity (1018 mAh g−1), Mn2O3 is thought to be one of the most promising candidates among them. However, Mn2O3 has low reversibility, poor cyclic stability, and low Coulombic efficiency as a result of the volume pulverisation during the charge and discharge process, the disintegration of its structure, and its low conductivity, which limited its applications [11,16]. Significant efforts [17,18] have been made to date to resolve the aforementioned issues. A good strategy is to use nanostructured Mn2O3 particles where the internal void and smaller particles could provide a large electrode–electrolyte contact area while also shortening ion and electron transport distances, buffering the volume change [19,20]. Another approach is to create a nanocomposite with carbon-based materials to improve the electrical conductivity between the transition-metal oxide particles while also acting as a buffer to accommodate volume changes during cycling to ensure structure stability [6,21].
Reduced graphene oxide (rGO) aerogel has been widely used to improve the electrochemical performance of the Na–ion batteries by shortening the transportation path for Na+ ions through its porous aerogel structure and also to sustain the volume change [22,23]. The open channel in the porous structure of the three-dimensional (3D) rGO aerogel provides a substantial area for ion transport. This structure, which has either macropores, mesopores, or micropores, can be created simply via freeze-drying. Unfortunately, due to excessive graphene sheet stacking, graphene has a large irreversible capacity, low Coulombic efficiency, rapid capacity fading, and a low specific capacity of 372 mAh g−1 [24,25,26]. Doping heteroatoms such as nitrogen (N) and/or sulphur (S) could alter the physicochemical properties of rGO aerogel. It has been demonstrated that N-doping could increase the specific capacity and electronic conductivity of carbon-based materials [27,28,29,30] due to the doping defect sites and functionalized groups. S-doping, conversely, could expand the distance between graphene layers, thereby favouring the sodiation/desodiation processes [31,32,33]. The combination of these two heteroatoms, namely nitrogen and sulphur (N,S)-doping, may thus improve the electrochemical properties of Na-ion batteries through a synergistic effect. Our previous study [34] have shown the improved electrochemical performances of Mn3O4/heteroatom-doped rGO aerogel as anode in Na-ion batteries. In this work, the Mn2O3/heteroatom-doped rGO aerogel was synthesised (Scheme 1), and its electrochemical performances were investigated. The ammonia (NH3) and L-Cysteine served as the precursor of the N and N,S, respectively. During heating, the graphite oxide (GO) layers self-assembled into 3D networks due to interactions, hydrogen bonding, coordination and electrostatic interactions, while simultaneously being reduced to rGO and introducing the nitrogen and/or sulphur to the rGO layers. The heteroatom doping on the porous structure of rGO aerogel provides further assistance for Na-ion migration, and the 3D-interconnected rGO layers will provide the conductive path that is advantageous for electron transportation. Our findings may contribute to the existing body of knowledge and aid in the development of this material in the future.
The Mn2O3/heteroatom-doped rGO aerogel demonstrates an impressive discharge capacity of up to 100 cycles with remarkable Coulombic efficiencies of up to 99% in comparison with the undoped rGO. The strategy adopted here is useful in enhancing the electrode’s capacity and cycling stability. Additionally, the presence of rGO aerogel might improve conductivity and enable quick ion transport inside the electrode. The enhanced specific capacity and cyclability demonstrated by the Mn2O3/rGO and Mn2O3/heteroatom-doped rGO aerogels in our work provide new information for other researchers attempting to forecast the potential for this nanocomposite to be used in Na-ion batteries.

2. Materials and Methods

2.1. Chemicals

Graphite powders (99.99%), sulphuric acid (H2SO4), potassium permanganate (KMnO4) (99.0%), hydrogen peroxide (H2O2; 30%), ammonia solution (NH3), L-Cysteine, carbon black (>99.5%), polyvinylidene fluoride (PVDF), N-methyl-2-pyrrolidone (NMP), sodium metal (99.9% trace metal base), sodium perchlorate (NaClO4; 98%), propylene carbonate (PC; 99.7%), and fluoroethylene carbonate (FEC; 99%) were purchased from Sigma–Aldrich, St. Louis, MO, USA. D-(+)-glucose (C6H12O6) was obtained from Merck Millipore, and hydrochloric acid (HCl; 35.0%) was received from Alfa Aesar. All chemicals were used without further purification.

2.2. Synthesis of Mn2O3/rGO and Mn2O3/Heteroatom-Doped rGO Aerogels

GO was synthesised according to the modified Hummer method [35]. Mn2O3 was prepared using the hydrothermal method and then thermally decomposed, as previously reported [17]. To synthesise Mn2O3/rGO aerogel, as prepared, GO (90 mg) was dispersed in 18 mL deionised (DI) water, and then, Mn2O3 was added into the above solution. After that, the solution was transferred to a stainless-steel autoclave with a capacity of 125 mL and heated at 180 °C for 12 h. The obtained precipitates (Mn2O3/rGO hydrogel) were then freeze-dried to obtain the Mn2O3/rGO aerogel. A similar procedure was adopted to synthesise the Mn2O3/heteroatom-doped rGO aerogel, where 4 mL NH3 as the N source was added to the suspension and 45 mg L-Cysteine was added as a nitrogen/sulphur source. The collected powders were denoted as Mn2O3/N-rGO and Mn2O3/N,S-rGO aerogels.

2.3. Physical Characterisation

X-ray diffraction (XRD; Rigaku Miniflex II, Tokyo, Japan) was used to determine the crystal structure. The amount of Mn2O3 in the nanocomposite was quantified using a thermogravimetric analyser (TGA; Perkin Elmer, Waltham, MA, USA) at a heating rate of 10 min−1 in air. A scanning electron microscope (SEM; JSM-6360L; JEOL, Akishima, Tokyo, Japan) and transmission electron microscopy (TEM; TECNAI G2 20 S-TWIN FEI Company, Lincoln, NE, USA) were used to examine the morphology of the samples. Fourier transform infrared (FTIR) measurement was conducted using an IR Tracer-100 from Shimadzu (Kyoto, Japan). Raman spectroscopy (Renishaw, Gloucestershire, UK) was used to collect Raman spectra. X-ray photoelectron spectroscopy (XPS) was used to inspect the chemical states of the elements present in the composite powders using an Axis Ultra DLD XPS (Kratos, Manchester, UK), and the obtained spectra were fitted using CASA software.

2.4. Electrochemical Characterisation

The active materials, carbon black and PVDF (weight ratio of 75:20:5), were dissolved in NMP and then coated on a copper foil and left to dry at 100 °C overnight. The fabrication of CR2032 coin-type cells was conducted in an argon-filled glovebox (Unilab, MBRAUN, Germany, H2O, O2 < 0.1 ppm). The active material electrode was sandwiched with sodium metal (Sigma–Aldrich, 99.9%) and separated using a glass fibre (GF/D; Whatman). The electrolyte used was 1 M NaClO4 in PC with the addition of 5 wt.% FEC. A Neware battery analyser was used to investigate the galvanostatic charge/discharge behaviour, and cyclic voltammetry (CV) was conducted using the CHI 700E between 0.01 and 3.0 V (vs. Na+/Na).

3. Results and Discussion

The TGA was conducted in the air to estimate the amount of Mn2O3 in the nanocomposites (Figure 1). The weight loss from ambient until 100 °C was mainly due to the evaporation of the physisorbed water, and then the weight loss up to 400 °C occurred due to the decomposition of the hydroxyl and carboxyl groups [36] present in the rGO. Whereas, for the N-rGO and N,S-rGO aerogels, the weight loss between 150 °C and 600 °C resulted from the decomposition of nitrogen and/or sulphur-containing functional groups, respectively, and was also related to the decarboxylation and elimination of hydroxyl functionalities [37]. The rGO is completely decomposed after 600 °C [38]. Therefore, the amounts of Mn2O3 in Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels were estimated to be 15, 16, and 19 wt%, respectively.
The XRD was used to investigate the structure of the Mn2O3 and its nanocomposites (Figure 2). No impurity phases were observed, and the Mn2O3 can be indexed to the cubic structure of Mn2O3 (JCPDS no. 11061). The presence of Mn2O3 in the nanocomposites indicates that the Mn2O3 was successfully embedded into the rGO layer structure. The broad diffraction peak appeared at 25° and was ascribed to the layered and hexagonal structure of the (002) lattice plane of the graphite-like structure [39]. The intensity of the diffraction peaks of the Mn2O3 was decreased in the Mn2O3-containing samples due to a large amount of rGO aerogel, as confirmed via the TGA.
The morphology of all samples was viewed using SEM (Figure 3) and TEM (Figure 4). Figure 3a–f shows that SEM images of the crumpled and wrinkled structure of rGO layers can be observed in all rGO aerogel-containing samples. The porous structure of rGO aerogel was interconnected through physical cross-linking to form a networked structure with pores smaller than 1 µm. The porous structure could act as an effective channel to shorten the Na–ion diffusion pathway for fast kinetics of Na–ion transport and provide a cushion to accommodate the volume changes during the charge/discharge processes [40]. Figure 3b,d,f show that the Mn2O3 cubic microstructures of 0.5–1.5 µm are distributed on the surface of the rGO layers, which are aggregates of individual Mn2O3. This suggests that Mn2O3 cubic structures and rGO aerogel layers assembled efficiently during hydrothermal. Figure 3g shows that the pristine Mn2O3 possesses a cubic morphology with a size ranging from 1.1 to 1.9 µm. Furthermore, the Mn2O3 is anchored on the surface of the heteroatom-doped rGO aerogel layers and could act as a spacer to prevent the rGO aerogel layers from restacking [41,42].
As shown in Figure 4, further investigation of the Mn2O3 anchored on the heteroatom-doped rGO aerogel layers is conducted using TEM. Typical crumpled rGO sheets formed a 3D framework (Figure 4a,d,g) aligned with the observations from the SEM images. The overlapping thin lamellar structures with clear edges and curved layers were evident in the heteroatom-doped rGO aerogels. The structural flaws caused by heteroatom doping resulted in a more wrinkled surface in the N-rGO (Figure 4d) and N,S-rGO (Figure 4g) aerogels. The TEM images (Figure 4b,e,h) further revealed the Mn2O3 were surrounded by a few rGO aerogel layers and the diameter of individual Mn2O3 is measured to be 13–17 nm based on the histogram (inset). The Mn2O3 is less transparent in comparison to its surroundings, demonstrating its porous nature. Previously, we have reported that the Mn2O3 exhibited a porous structure [17]. The high-resolution TEM (Figure 4c,f,i) demonstrated a clear interlayer d-spacing of the cubic phase of Mn2O3 of 0.25 nm in the (222) plane and 0.38 nm corresponding to the (211) plane.
To support the presence of heteroatoms in the nanocomposites, Figure 5 shows the FTIR spectra of Mn2O3, Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels. The apparent peaks at 1730 cm−1 (C=O stretching vibration), 1363 cm−1 (O–H deformation), and 1215 cm−1 (C–O stretching vibration) indicate that GO is successfully reduced and the corresponding product is rGO [43,44]. Another FTIR peak at 1440 cm−1 attributed to the C–O stretching deformation and at 1569 cm−1 related to the aromatic skeletal v (C=C) stretching vibration of rGO aerogel is also observed [45]. The existence of S atoms on the N,S-rGO and Mn2O3/N,S-rGO aerogels was established via the appearance of peaks at 1096 cm−1 (C=S) and 2451 cm−1 (S–H stretching vibration) [46]. Furthermore, the C–N stretching vibration located at 1422 cm−1 is an indicator that nitrogen is successfully doped in the nitrogen-containing nanocomposites. The FTIR spectra of Mn2O3 also appeared in the Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels, in which the Mn–O stretching vibration appeared at 541 and 675 cm−1 [47]. Hence, the FTIR results indicate that there is a Mn–O, C–N, C=S and N–H bonding and thus confirmed that Mn2O3 nanoparticles were incorporated into the rGO aerogel layers.
Raman spectra for all samples are also recorded (Figure 6). Two typical characteristic peaks of carbon are observed in all samples: the D band at 1340 cm−1, which reflects the defect and disorder of carbon in graphitic layers, and the G band at 1590 cm−1, which associates with the characteristics of the E2g mode of sp2 hybridised carbon [48,49]. The intensity ratio between the D and G bands (ID/IG) can be used to assess surface defects. According to this figure, the ID/IG of rGO aerogel is 1.33. After nitrogen/sulphur doping, the ID/IG of N-rGO and N,S-rGO aerogels were calculated to be 0.95 and 0.98, respectively. After nitrogen and/or sulphur doping, rGO layer defects were established, and the ID/IG ratio was altered via the introduction of cubic Mn2O3 into graphene layers. By contrast, the calculated intensity ratios for Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels were 1.35, 1.00, and 1.02, respectively. Hence, the band located at ~340 cm−1 could be ascribed to the asymmetric stretch of bridge oxygen species (Mn–O–Mn), in correspondence to that obtained for bulk Mn2O3 particles [50].
Figure 7 depicts the survey scan of the heteroatom-doped rGO aerogels and the core-level spectra of O 1s, C 1s, N 1s, S 2p, and Mn 2p for Mn2O3/N-rGO and Mn2O3/N,S-rGO aerogels. The presence of nitrogen and nitrogen–sulphur is visible in the survey spectra of Mn2O3/N-rGO and Mn2O3/N,S-rGO aerogels, respectively (Figure 7a,b). Furthermore, the O 1s, C 1s, N 1s, Mn2p3/2, and Mn 2p1/2 peaks are located at 532, 284, 399, 641, and 653 eV and are found in the survey spectrum. Figure 7b exhibits the S 2p peak located at 164 eV. Figure 7c,d shows that the C 1s region can be resolved into three peaks. The peak at 284 eV corresponds to the C–C signal, and 287 eV is related to the C=O, whereas, the peak located at 285.2 eV is related to the C–O and C–N signal for Mn2O3/N-rGO aerogel and C–O, C–N and C–S at 285.3 eV for the Mn2O3/N,S-rGO aerogel. Figure 7e,f shows the O 1s deconvoluted XPS spectra, and the peaks located at 531, 533, and 535 eV corresponded to C–O–Mn, C=O, and surface absorbed oxygen (O–C–O/C–OH), respectively. Moreover, one peak that appeared at 529 eV, as depicted in Figure 7e, might be related to Mn–O. In the N 1s spectra (Figure 7g,h), three fitted peaks can be assigned to pyridinic N (397 eV), pyrrolic N (399 eV), and graphitic N (401 eV) doped in graphene. It has also been reported that the existence of pyridinic- and pyrrolic-type nitrogens can facilitate the transfer of Na-ion and electrons, improving the rate capability of the electrodes [51,52,53]. Additionally, the N concentration in Mn2O3/N-rGO aerogel was 15.54%, whereas, for the Mn2O3/N,S-rGO aerogel, the concentration of N and S is 3.36% and 3.41%, respectively. The high-resolution Mn 2p (Figure 7i,j) can be fitted into Mn 2p3/2 (641 eV) and Mn 2p1/2 (653 eV). The observed spin-orbit splitting represented by the difference in binding energy values of these two levels is 12 eV, in line with the previous report [54]. Moreover, the S 2p can be fitted into three components of S p3/2, S p1/2 and SOx centred at 163.7, 164.8 and 168.4 eV (Figure 7k), thus confirming the introduction of S atoms into disordered graphene lattice [55].
The electrochemical activity of Mn2O3 and their nanocomposites is evaluated using CV (Figure 8) at a scan rate of 0.1 mV s−1 over 0.01–3.00 V. All samples show similar reduction and oxidation peaks in their CV curves. In the first cycle (Figure 8a–d), the main cathodic peak at 0.8–0.9 V attributes to the reduction of Mn2O3 to Mn−O and the formation of Na2O and solid electrolyte interphase (SEI) layers over the active materials, which can be expressed as Equation (1) [17,56].
Mn2O3 + 2Na+ + 2e → 2MnO + Na2O
The broad peak at 0.01–0.4 V may be ascribed to the Na–ion insertion, the MnO to Mn irreversible conversion reaction (Equation (2)), and the formation of SEI on the electrode’s surface.
MnO + 2Na+ + 2e → Mn + Na2O
During the second to fifth cycles, the slight shift of the reduction peak at higher potential and the overlapping CV curves indicate the electrochemical reversibility of Na-ions. Na-ion batteries exhibited wider and weaker CV peaks in comparison to Li-ion batteries since Na-ions have a larger ionic radius than Li-ion and move between graphitic carbon layers more slowly than Li-ion [57,58]. Figure 8d shows that the Mn2O3/N,S-rGO aerogel exhibited an anodic peak at 1.95 V, which may be due to the sulphur incorporated into the N,S-rGO aerogel [59]. The following reaction (Equation (3)) describes the reversible mechanism during the sodiation/desodiation
Mn2O3 + 6Na+ + 6e ↔ 2Mn + 3Na2O
Figure 9 represents the typical charge and discharge profiles of the samples for several cycles at 0.1 A g−1. The long discharge plateau observed at the first cycle could be due to the formation of the SEI layer, which leads to the low initial Coulombic efficiency. The subsequent cycle shows no distinct plateau, which is consistent with the CV curves.
Further evaluation of the electrochemical performances of the Mn2O3, Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogel electrodes was conducted, and Figure 10 shows the results. Figure 10a depicts that the Mn2O3/N-rGO aerogel exhibited the highest irreversible discharge capacity of 2507 mAh g−1, followed by Mn2O3/N,S-rGO aerogel (1677 mAh g−1), Mn2O3/rGO (828 mAh g−1), and pristine Mn2O3 (544 mAh g−1) at a current density of 0.1 A g−1. The Mn2O3/N-rGO aerogel delivers a reversible discharge capacity of 594 mAh g−1 during the second cycle and gradually decreases to 325 mAh g−1 after 100 cycles with 87% capacity retention. For Mn2O3/N,S-rGO aerogel, the reversible discharge capacity decreased from 448 mAh g−1 (second cycle) to 279 mAh g−1 (100th cycle) with an 83% retention rate. These results indicate that the heteroatom doping on the rGO aerogel layers could tailor the physicochemical properties of the rGO, which in turn enhanced the discharge capacity and cyclability of the Mn2O3. The advantages of anchoring the Mn2O3 on the rGO layers can also be seen in the electrochemical properties of Mn2O3/rGO aerogel versus pristine Mn2O3. In the second cycle, the Mn2O3/rGO aerogel exhibits a reversible discharge capacity of 305 mAh g−1 and gradually reaches 242 mAh g−1 after 100 cycles with a capacity retention of 71%. Whereas, for pristine Mn2O3, the reversible discharge capacity at the second cycle was 294 mAh g−1 and decreased to 157 mAh g−1 after 100 cycles with 71% capacity retention. During the initial few cycles, the porous structure of rGO aerogel promoted the formation of excessive SEI layers, resulting in a lower initial Coulombic efficiency [60]. After a few cycles, the Coulombic efficiencies for all electrodes reached 90% due to the relatively stable SEI that formed during the cycling. Additionally, the Mn2O3/N-rGO aerogel (Figure 10b) has better rate capability compared with other electrodes with discharge capacities of 260, 222, 201, 177, 232, and 272 mAh g−1 at current densities of 0.2, 0.4, 0.6, 0.8, 1.0, and returning 0.2 A g−1, respectively. Interestingly, other electrodes also show good rate capability performance when executed at different current densities, but with slightly lower discharge capacity values compared with Mn2O3/N-rGO aerogel.
The overall electrochemical performances of Mn2O3/heteroatom-doped rGO aerogels reported here are superior to those reported previously (Table 1). It is believed the pronounced electrochemical performances of Mn2O3/heteroatom-doped rGO aerogels can be attributed to the networked structure of the nanocomposites, which have a variety of favourable properties. First, the porous networking of Mn2O3 and rGO aerogel may shorten the diffusion path for Na-ions and facilitate the electrolyte’s penetration into the inner surface of the active materials. The interconnected rGO networks in this structure offer an additional active site for Na-ion storage and also offer a large area for reversible reactions [42,61,62]. Second, the 3D rGO aerogel networks improved the mechanical integrity of the electrode by accommodating the strain and stress during the sodiation/desodiation processes. Consequently, structure destruction of the active materials can be avoided while maintaining the cyclability of the electrode. The aggregation of Mn2O3 could also be inhibited [63,64]. Third, the electron transfer ability of the Mn2O3 was enhanced in combination with the heteroatom-doped rGO aerogel. The electrical conductivity of the active materials is greatly improved by decreasing the charge-transfer resistance, which in turn facilitates the strong electronic interconnection of the active materials to ensure rapid charge transfers within the electrode [65]. Lastly, doping nitrogen and co-doping nitrogen/sulphur improve the physicochemical characteristics of the rGO [28]. The addition of heteroatoms to rGO aerogels may improve their electrical conductivity and electrochemical activity by fostering charge transfer between adjacent carbon atoms [45,66,67]. The strategies of doping heteroatoms into rGO aerogels could be ideal platforms to tackle the problems related to the use of transition-metal oxide as anode materials for Na-ion batteries, especially electrode pulverisation during cycling. Several challenges related to Mn2O3 have to be overcome, and modifications and development of the Mn2O3 with nanocomposites are necessary to achieve the desired electrochemical properties for Na storage for practical applications.

4. Conclusions

In summary, the integration of Mn2O3 with the 3D framework of heteroatom-doped rGO aerogel was demonstrated to be an effective strategy to mitigate problems such as volume changes during the charge/discharge process, mechanical degradation, and capacity fading. The cubic Mn2O3 particles (with an average size of 0.5–1.5 µm) obtained via thermal decomposition of MnCO3 are anchored in porous and crumpled rGO sheets of micrometre dimension. This architecture ensures closed contact between the electrolyte and the active materials that enable efficient electron transport from each active Mn2O3 cubic particle to the current collector via the rGO aerogel. The aerogel structure produced heteroatom-doped rGO layers with strong interconnections, and the Mn2O3 particles distributed on the rGO layers prevented the restacking of the rGO layers. A strong synergistic effect between Mn2O3 and heteroatom-doped rGO in the nanocomposite becomes more apparent with cycling and contributes to the excellent cycling performance of the electrode. After 100 cycles, the capacities of the Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels were retained at 242, 325, and 277 mAh g−1, respectively. In comparison to Mn2O3/rGO, the N- and N,S-doped rGO aerogels with Mn2O3 displayed higher specific capacities, excellent cyclability, and rate capability. Thus, Mn2O3 anchored on the heteroatom-doped rGO aerogel could effectively improve the Na storage capacity of Mn2O3 and provide a useful strategy for producing high-yield anode materials for practical Na-ion batteries.

Author Contributions

N.F.M.Y.: conceptualization, methodology, formal analysis, writing—original draught, visualization. N.H.I.: conceptualization, validation, investigation, writing—review and editing, supervision. M.F.M.D.: methodology, validation, writing—review and editing, visualisation. S.R.M.: Methodology, Validation, Writing—Review & Editing. N.A.H.: methodology, validation, writing—review and editing. L.N.: methodology, validation, and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The Fundamental Research Grant Scheme (FRGS; FRGS/1/2018/STG07/UMT/02/3) (VOT 59520) provided funding for this research.

Data Availability Statement

Not applicable.

Acknowledgments

The Ministry of Higher Education Malaysia is acknowledged for financial assistance provided by the Fundamental Research Grant Scheme (FRGS; FRGS/1/2018/STG07/UMT/02/3).

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. 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]
  2. Zhao, X.; Wang, H.-E.; Yang, Y.; Neale, Z.G.; Massé, R.C.; Cao, J.; Cai, W.; Sui, J.; Cao, G. Reversible and fast Na-ion storage in MoO2/MoSe2 heterostructures for high energy-high power Na-ion capacitors. Energy Storage Mater. 2018, 12, 241–251. [Google Scholar] [CrossRef]
  3. Cheng, T.; Xu, J.; Tan, Z.; Ye, J.; Tao, Z.; Du, Z.; Wu, Y.; Wu, S.; Ji, H.; Yu, Y. A spray-freezing approach to reduced graphene oxide/MoS2 hybrids for superior energy storage. Energy Storage Mater. 2018, 10, 282–290. [Google Scholar] [CrossRef]
  4. Hong, S.Y.; Kim, Y.; Park, Y.; Choi, A.; Choi, N.S.; Lee, K.T. Charge carriers in rechargeable batteries: Na ions vs. Li ions. Energy Environ. Sci. 2013, 6, 2067–2081. [Google Scholar] [CrossRef]
  5. Cao, K.; Jiao, L.; Liu, Y.; Liu, H.; Wang, Y.; Yuan, H. Ultra-high capacity lithium-ion batteries with hierarchical CoO nanowire clusters as binder free electrodes. Adv. Funct. Mater. 2015, 25, 1082–1089. [Google Scholar] [CrossRef]
  6. Zhang, N.; Han, X.; Liu, Y.; Hu, X.; Zhao, Q.; Chen, J. 3D porous γ-Fe2O3@C nanocomposite as high-performance anode material of Na-ion batteries. Adv. Energy Mater. 2015, 5, 1401123. [Google Scholar] [CrossRef]
  7. Chen, X.Q.; Lin, H.B.; Zheng, X.W.; Cai, X.; Xia, P.; Zhu, Y.M.; Li, X.P.; Li, W.S. Fabrication of core–shell porous nanocubic Mn2O3@TiO2 as a high-performance anode for lithium ion batteries. J. Mater. Chem. A 2015, 3, 18198–18206. [Google Scholar] [CrossRef]
  8. Cheng, C.; Huang, Y.; Wang, N.; Jiang, T.; Hu, S.; Zheng, B.; Yuan, H.; Xiao, D. Facile fabrication of Mn2O3 nanoparticle-assembled hierarchical hollow spheres and their sensing for hydrogen peroxide. ACS Appl. Mater. Interfaces 2015, 7, 9526–9533. [Google Scholar] [CrossRef] [PubMed]
  9. Huang, S.Z.; Jin, J.; Cai, Y.; Li, Y.; Tan, H.Y.; Wang, H.E.; Van Tendeloo, G.; Su, B.L. Engineering single crystalline Mn3O4 nano-octahedra with exposed highly active {011} facets for high performance lithium ion batteries. Nanoscale 2014, 6, 6819–6827. [Google Scholar] [CrossRef] [PubMed]
  10. Su, H.; Xu, Y.-F.; Feng, S.-C.; Wu, Z.-G.; Sun, X.-P.; Shen, C.-H.; Wang, J.-Q.; Li, J.-T.; Huang, L.; Sun, S.-G. Hierarchical Mn2O3 hollow microspheres as anode material of lithium ion battery and its conversion reaction mechanism investigated by XANES. ACS Appl. Mater. Interfaces 2015, 7, 8488–8494. [Google Scholar] [CrossRef] [PubMed]
  11. Bai, Z.; Zhang, Y.; Zhang, Y.; Guo, C.; Tang, B.; Sun, D. MOFs-derived porous Mn2O3 as high-performance anode material for Li-ion battery. J. Mater. Chem. A 2015, 3, 5266–5269. [Google Scholar] [CrossRef]
  12. Zhang, Z.; Zhao, X.; Li, J. Facile synthesis of nanostructured MnO2 as anode materials for sodium-ion batteries. ChemNanoMat 2016, 2, 196–200. [Google Scholar] [CrossRef]
  13. Weng, Y.-T.; Huang, T.Y.; Lim, C.H.; Shao, P.S.; Hy, S.; Kuo, C.Y.; Cheng, J.H.; Hwang, B.J.; Lee, J.F.; Wu, N.L. An unexpected large capacity of ultrafine manganese oxide as a sodium-ion battery anode. Nanoscale 2015, 7, 20075–20081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Jiang, Y.; Hu, M.; Zhang, D.; Yuan, T.; Sun, W.; Xu, B.; Yan, M. Transition metal oxides for high performance sodium ion battery anodes. Nano Energy 2014, 5, 60–66. [Google Scholar] [CrossRef] [Green Version]
  15. Mahamad, Y.N.F.; Idris, N.H.; Md, D.M.F.; Majid, S.R.; Harun, N.A.; Rahman, M.M. Electrochemical sodiation/desodiation into Mn3O4 nanoparticles. ACS Omega 2020, 5, 29158–29167. [Google Scholar] [CrossRef]
  16. Alagar, S.; Madhuvilakku, R.; Mariappan, R.; Piraman, S. Nano-architectured porous Mn2O3 spheres/cubes vs rGO for asymmetric supercapacitors applications in novel solid-state electrolyte. J. Power Sources 2019, 441, 227181–227193. [Google Scholar] [CrossRef]
  17. Yusoff, N.F.M.; Idris, N.H.; Din, M.F.M.; Majid, S.R.; Harun, N.A.; Rahman, M.M. Investigation on the electrochemical performances of Mn2O3 as a potential anode for Na-ion batteries. Sci. Rep. 2020, 10, 9207–9217. [Google Scholar] [CrossRef]
  18. Huang, S.-Z.; Jin, J.; Cai, Y.; Li, Y.; Deng, Z.; Zeng, J.-Y.; Liu, J.; Wang, C.; Hasan, T.; Su, B.-L. Three-dimensional (3D) bicontinuous hierarchically porous Mn2O3 single crystals for high performance lithium-ion batteries. Sci. Rep. 2015, 5, 14686–14698. [Google Scholar] [CrossRef] [Green Version]
  19. Zhao, Y.; Li, X.; Yan, B.; Li, D.; Lawes, S.; Sun, X. Significant impact of 2D graphene nanosheets on large volume change tin-based anodes in lithium-ion batteries: A review. J. Power Sources 2015, 274, 869–884. [Google Scholar] [CrossRef]
  20. Jiang, X.; Yu, W.; Wang, H.; Xu, H.; Liu, X.; Ding, Y. Enhancing the performance of MnO by double carbon modification for advanced lithium-ion battery anodes. J. Mater. Chem. A 2016, 4, 920–925. [Google Scholar] [CrossRef]
  21. Bai, Z.; Zhang, Y.; Zhang, Y.; Guo, C.; Tang, B. A large-scale, green route to synthesize of leaf-like mesoporous CuO as high-performance anode materials for lithium ion batteries. Electrochim. Acta 2015, 159, 29–34. [Google Scholar] [CrossRef]
  22. Li, Y.; Ye, D.; Liu, W.; Shi, B.; Guo, R.; Pei, H.; Xie, J. A three-dimensional core-shell nanostructured composite of polypyrrole wrapped MnO2/reduced graphene oxide/carbon nanotube for high performance lithium ion batteries. J. Colloid Interface Sci. 2017, 493, 241–248. [Google Scholar] [CrossRef] [PubMed]
  23. Hao, Q.; Liu, B.; Ye, J.; Xu, C. Well encapsulated Mn3O4 octahedra in graphene nanosheets with much enhanced Li-storage performances. J. Colloid Interface Sci. 2017, 504, 603–610. [Google Scholar] [CrossRef] [PubMed]
  24. Wu, Z.S.; Zhou, G.; Yin, L.C.; Ren, W.; Li, F.; Cheng, H.M. Graphene/metal oxide composite electrode materials for energy storage. Nano Energy 2012, 1, 107–131. [Google Scholar] [CrossRef]
  25. Buqa, H.; Goers, D.; Holzapfel, M.; Spahr, M.E.; Novák, P. High rate capability of graphite negative electrodes for lithium-ion batteries. J. Electrochem. Soc. 2005, 152, A474–A481. [Google Scholar] [CrossRef]
  26. Chen, X.; Zhang, Y. The main problems and solutions in practical application of anode materials for sodium ion batteries and the latest research progress. Int. J. Energy Res. 2021, 45, 9753–9779. [Google Scholar] [CrossRef]
  27. Wang, Z.; Qie, L.; Yuan, L.; Zhang, W.; Hu, X.; Huang, Y. Functionalized N-doped interconnected carbon nanofibers as an anode material for sodium-ion storage with excellent performance. Carbon 2013, 55, 328–334. [Google Scholar] [CrossRef]
  28. Jing, M.; Chen, Z.; Li, Z.; Li, F.; Chen, M.; Zhou, M.; He, B.; Chen, L.; Hou, Z.; Chen, X. Facile synthesis of ZnS/N,S co-doped carbon composite from zinc metal complex for high-performance sodium-ion batteries. ACS Appl. Mater. Interfaces 2018, 10, 704–712. [Google Scholar] [CrossRef]
  29. Yang, F.; Zhang, Z.; Du, K.; Zhao, X.; Chen, W.; Lai, Y.; Li, J. Dopamine derived nitrogen-doped carbon sheets as anode materials for high-performance sodium ion batteries. Carbon 2015, 91, 88–95. [Google Scholar] [CrossRef]
  30. Yue, X.; Huang, N.; Jiang, Z.; Tian, X.; Wang, Z.; Hao, X.; Jiang, Z.J. Nitrogen-rich graphene hollow microspheres as anode materials for sodium-ion batteries with super-high cycling and rate performance. Carbon 2018, 130, 574–583. [Google Scholar] [CrossRef]
  31. Deng, X.; Xie, K.; Li, L.; Zhou, W.; Sunarso, J.; Shao, Z. Scalable synthesis of self-standing sulfur-doped flexible graphene films as recyclable anode materials for low-cost sodium-ion batteries. Carbon 2016, 107, 67–73. [Google Scholar] [CrossRef]
  32. Wang, X.; Li, G.; Hassan, F.M.; Li, J.; Fan, X.; Batmaz, R.; Xiao, X.; Chen, Z. Sulfur covalently bonded graphene with large capacity and high rate for high-performance sodium-ion batteries anodes. Nano Energy 2015, 15, 746–754. [Google Scholar] [CrossRef]
  33. Quan, B.; Jin, A.; Yu, S.H.; Kang, S.M.; Jeong, J.; Abruña, H.D.; Jin, L.; Piao, Y.; Sung, Y.E. Solvothermal-derived S-doped graphene as an anode material for sodium-ion batteries. Adv. Sci. 2018, 5, 1700880–1700913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Mahamad Yusoff, N.F.; Idris, N.H.; Md Din, M.F.; Majid, S.R.; Harun, N.A. Enhanced electrochemical performances of Mn3O4/heteroatom-doped reduced graphene oxide aerogels as an anode for sodium-ion batteries. Nanomaterials 2022, 12, 3569. [Google Scholar] [CrossRef]
  35. Hummers, J.W.S.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  36. Hu, N.; Gao, R.; Wang, Y.; Wang, Y.; Chai, J.; Yang, Z.; Kong, E.S.W.; Zhang, Y. The preparation and characterization of non-covalently functionalized graphene. J. Nanosci. Nanotechnol. 2012, 12, 99–104. [Google Scholar] [CrossRef] [Green Version]
  37. Chen, J.; Yao, B.; Li, C.; Shi, G. An improved Hummers method for eco-friendly synthesis of graphene oxide. Carbon 2013, 64, 225–229. [Google Scholar] [CrossRef]
  38. Zhang, L.; Shi, G. Preparation of highly conductive graphene hydrogels for fabricating supercapacitors with high rate capability. J. Phys. Chem. C 2011, 115, 17206–17212. [Google Scholar] [CrossRef]
  39. Ghosh, A.; Ghosh, S.; Seshadhri, G.M.; Ramaprabhu, S. Green synthesis of nitrogen-doped self-assembled porous carbon-metal oxide composite towards energy and environmental applications. Sci. Rep. 2019, 9, 5187. [Google Scholar] [CrossRef] [Green Version]
  40. Kang, J.-H.; Chen, J.-S. Using ethylenediamine to prepare three dimensional nitrogen-doped graphene aerogel/sulfur composite for lithium-sulfur batteries. Diam. Relat. Mater. 2018, 88, 222–229. [Google Scholar] [CrossRef]
  41. Li, Y.; Zhao, Y.; Ma, C.; Shi, J.; Zhao, Y. Highly monodispersed graphene/SnO2 hybrid nanosheets as bifunctional anode materials of Li-ion and Na-ion batteries. J. Alloys Compd. 2020, 821, 153506–153516. [Google Scholar] [CrossRef]
  42. Zhang, L.; Ge, D.; Geng, H.; Zheng, J.; Cao, X.; Gu, H. Synthesis of porous Mn2O3 embedded in reduced graphene oxide as advanced anode materials for lithium storage. New J. Chem. 2017, 41, 7102–7107. [Google Scholar] [CrossRef]
  43. Ouyang, Z.; Lei, Y.; Chen, Y.; Zhang, Z.; Jiang, Z.; Hu, J.; Lin, Y. Preparation and specific capacitance properties of sulfur, nitrogen co-doped graphene quantum dots. Nanoscale Res. Lett. 2019, 14, 1–9. [Google Scholar] [CrossRef]
  44. Meng, J.; Cao, Y.; Suo, Y.; Liu, Y.; Zhang, J.; Zheng, X. Facile fabrication of 3D SiO2@graphene aerogel composites as anode material for lithium ion batteries. Electrochim. Acta 2015, 176, 1001–1009. [Google Scholar] [CrossRef]
  45. Wu, Z.-S.; Winter, A.; Chen, L.; Sun, Y.; Turchanin, A.; Feng, X.; Müllen, K. Three-dimensional nitrogen and boron co-doped graphene for high-performance all-solid-state supercapacitors. Adv. Mater. 2012, 24, 5130–5135. [Google Scholar] [CrossRef]
  46. Rochman, R.A.; Wahyuningsih, S.; Ramelan, A.H.; Hanif, Q.A. Preparation of nitrogen and sulphur Co-doped reduced graphene oxide (rGO-NS) using N and S heteroatom of thiourea. IOP Conf. Ser. Mater. Sci. Eng. 2019, 509, 012119–012127. [Google Scholar] [CrossRef]
  47. Ding, Y.; Hou, C.; Li, B.; Lei, Y. Sensitive hydrazine detection using a porous Mn2O3 nanofibers-based sensor. Electroanalysis 2011, 23, 1245–1251. [Google Scholar] [CrossRef]
  48. Jiang, Z.; Li, Y.; Han, C.; He, Z.; Ma, W.; Meng, W.; Jiang, Y.; Dai, L.; Wang, L. Superior lithium storage performance of hierarchical N-doped carbon encapsulated NaTi2(PO4)3 microflower. Ceram. Int. 2020, 46, 1954–1961. [Google Scholar] [CrossRef]
  49. Wall, M. The Raman spectroscopy of graphene and the determination of layer thickness. Thermo Sci 2011, 5, 1–5. [Google Scholar]
  50. Rahaman, H.; Laha, R.M.; Maiti, D.K.; Ghosh, S.K. Fabrication of Mn2O3 nanorods: An efficient catalyst for selective transformation of alcohols to aldehydes. RSC Adv. 2015, 5, 33923–33929. [Google Scholar] [CrossRef]
  51. Wang, T.; Wang, L.; Wu, D.; Xia, W.; Zhao, H.; Jia, D. Hydrothermal synthesis of nitrogen-doped graphene hydrogels using amino acids with different acidities as doping agents. J. Mater. Chem. A 2014, 2, 8352–8361. [Google Scholar] [CrossRef]
  52. Li, D.; Chen, H.; Liu, G.; Wei, M.; Ding, L.-X.; Wang, S.; Wang, H. Porous nitrogen doped carbon sphere as high performance anode of sodium-ion battery. Carbon 2015, 94, 888–894. [Google Scholar] [CrossRef]
  53. Zou, L.; Lai, Y.; Hu, H.; Wang, M.; Zhang, K.; Zhang, P.; Fang, J.; Li, J. N/S co-doped 3D porous carbon nanosheet networks enhancing anode performance of sodium-ion batteries. Chem. A Eur. J. 2017, 23, 14261–14266. [Google Scholar] [CrossRef]
  54. Gopi, C.V.V.M.; Vinodh, R.; Sambasivam, S.; Obaidat, I.M.; Singh, S.; Kim, H.J. Co9S8-Ni3S2/CuMn2O4-NiMn2O4 and MnFe2O4-ZnFe2O4/graphene as binder-free cathode and anode materials for high energy density supercapacitors. Chem. Eng. J. 2020, 381, 122640. [Google Scholar] [CrossRef]
  55. Li, W.; Zhou, M.; Li, H.; Wang, K.; Cheng, S.; Jiang, K. A high performance sulfur-doped disordered carbon anode for sodium ion batteries. Energy Environ. Sci. 2015, 8, 2916–2921. [Google Scholar] [CrossRef]
  56. Wang, B.; Li, F.; Wang, X.; Wang, G.; Wang, H.; Bai, J. Mn3O4 nanotubes encapsulated by porous graphene sheets with enhanced electrochemical properties for lithium/sodium-ion batteries. Chem. Eng. J. 2019, 364, 57–69. [Google Scholar] [CrossRef]
  57. Valvo, M.; Philippe, B.; Lindgren, F.; Tai, C.-W.; Edström, K. Insight into the processes controlling the electrochemical reactions of nanostructured iron oxide electrodes in Li-and Na-half cells. Electrochim. Acta 2016, 194, 74–83. [Google Scholar] [CrossRef] [Green Version]
  58. Wang, X.; Liu, X.; Wang, G.; Xia, Y.; Wang, H. One-dimensional hybrid nanocomposite of high-density monodispersed Fe3O4 nanoparticles and carbon nanotubes for high-capacity storage of lithium and sodium. J. Mater. Chem. A 2016, 4, 18532–18542. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Zhao, Y.; Bakenov, Z.; Tuiyebayeva, M.; Konarov, A.; Chen, P. Synthesis of hierarchical porous sulfur/polypyrrole/multiwalled carbon nanotube composite cathode for lithium batteries. Electrochim. Acta 2014, 143, 49–55. [Google Scholar] [CrossRef]
  60. Ye, J.; Zhao, H.; Song, W.; Wang, N.; Kang, M.; Li, Z. Enhanced electronic conductivity and sodium-ion adsorption in N/S co-doped ordered mesoporous carbon for high-performance sodium-ion battery anode. J. Power Sources 2019, 412, 606–614. [Google Scholar] [CrossRef]
  61. Kandula, S.; Bae, J.; Cho, J.; Son, J.G. Gram-scale synthesis of rGO wrapped porous α-Fe2O3 as an advanced anode material for Na-ion batteries with superior cyclic stability. Compos. Part B Eng. 2021, 220, 108995–109005. [Google Scholar] [CrossRef]
  62. Liu, Q.; Hu, Z.; Zou, C.; Jin, H.; Wang, S.; Li, L. Structural engineering of electrode materials to boost high-performance sodium-ion batteries. Cell Rep. Phys. Sci. 2021, 2, 100551–100579. [Google Scholar] [CrossRef]
  63. Li, M.; Zhu, K.; Zhao, H.; Meng, Z. Recent progress on graphene-based nanocomposites for electrochemical sodium-ion storage. Nanomaterials 2022, 12, 2837. [Google Scholar] [CrossRef]
  64. Xie, X.; Chen, S.; Sun, B.; Wang, C.; Wang, G. 3D networked tin oxide/graphene aerogel with a hierarchically porous architecture for high-rate performance sodium-ion batteries. ChemSusChem 2015, 8, 2948–2955. [Google Scholar] [CrossRef] [PubMed]
  65. Sun, J.; Lee, H.-W.; Pasta, M.; Yuan, H.; Zheng, G.; Sun, Y.; Li, Y.; Cui, Y. A phosphorene–graphene hybrid material as a high-capacity anode for sodium-ion batteries. Nat. Nanotechnol. 2015, 10, 980–985. [Google Scholar] [CrossRef]
  66. Wu, Z.-S.; Ren, W.; Xu, L.; Li, F.; Cheng, H.-M. Doped graphene sheets as anode materials with superhigh rate and large capacity for lithium ion batteries. ACS Nano 2011, 5, 5463–5471. [Google Scholar] [CrossRef] [PubMed]
  67. Jeong, H.M.; Lee, J.W.; Shin, W.H.; Choi, Y.J.; Shin, H.J.; Kang, J.K.; Choi, J.W. Nitrogen-doped graphene for high-performance ultracapacitors and the importance of nitrogen-doped sites at basal planes. Nano Lett. 2011, 11, 2472–2477. [Google Scholar] [CrossRef]
  68. Li, F.; Ma, J.; Ren, H.; Wang, H.; Wang, G. Fabrication of MnO nanowires implanted in graphene as an advanced anode material for sodium-ion batteries. Mater. Lett. 2017, 206, 132–135. [Google Scholar] [CrossRef]
  69. Zhang, X.; Zhu, G.; Yan, D.; Lu, T.; Pan, L. MnO@C nanorods derived from metal-organic frameworks as anode for superiorly stable and long-life sodium-ion batteries. J. Alloys Compd. 2017, 710, 575–580. [Google Scholar] [CrossRef]
  70. Qin, Y.; Jiang, Z.; Guo, L.; Huang, J.; Jiang, Z.J. Controlled thermal oxidation derived Mn3O4 encapsulated in nitrogen doped carbon as an anode for lithium/sodium ion batteries with enhanced performance. Chem. Eng. J. 2021, 406, 126894–126906. [Google Scholar] [CrossRef]
Scheme 1. The schematic representation for the synthesis of Mn2O3/rGO aerogels and Mn2O3/heteroatom-doped rGO aerogels.
Scheme 1. The schematic representation for the synthesis of Mn2O3/rGO aerogels and Mn2O3/heteroatom-doped rGO aerogels.
Nanomaterials 13 00732 sch001
Figure 1. TGA curves of rGO, N-rGO, N,S-rGO, Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels.
Figure 1. TGA curves of rGO, N-rGO, N,S-rGO, Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels.
Nanomaterials 13 00732 g001
Figure 2. XRD patterns of Mn2O3, Mn2O3/rGO, and Mn2O3/heteroatom-doped rGO aerogels.
Figure 2. XRD patterns of Mn2O3, Mn2O3/rGO, and Mn2O3/heteroatom-doped rGO aerogels.
Nanomaterials 13 00732 g002
Figure 3. SEM images of (a) rGO aerogel, (b) Mn2O3/rGO aerogel, (c) N-rGO aerogel, (d) Mn2O3/N-rGO aerogel, (e) N,S-rGO aerogel, (f) Mn2O3/N,S-rGO aerogel, and (g) pristine Mn2O3.
Figure 3. SEM images of (a) rGO aerogel, (b) Mn2O3/rGO aerogel, (c) N-rGO aerogel, (d) Mn2O3/N-rGO aerogel, (e) N,S-rGO aerogel, (f) Mn2O3/N,S-rGO aerogel, and (g) pristine Mn2O3.
Nanomaterials 13 00732 g003
Figure 4. TEM images of (a) rGO aerogel, (b) Mn2O3/rGO aerogel (inset: particle size distribution), (d) N-rGO aerogel, (e) Mn2O3/N-rGO aerogel (inset: particle size distribution), (g) N,S-rGO aerogel, (h) Mn2O3/N,S-rGO aerogel (inset: particle size distribution), and high-resolution TEM images of (c) Mn2O3/rGO aerogel, (f) Mn2O3/N-rGO aerogel, and (i) Mn2O3/N,S-rGO aerogel.
Figure 4. TEM images of (a) rGO aerogel, (b) Mn2O3/rGO aerogel (inset: particle size distribution), (d) N-rGO aerogel, (e) Mn2O3/N-rGO aerogel (inset: particle size distribution), (g) N,S-rGO aerogel, (h) Mn2O3/N,S-rGO aerogel (inset: particle size distribution), and high-resolution TEM images of (c) Mn2O3/rGO aerogel, (f) Mn2O3/N-rGO aerogel, and (i) Mn2O3/N,S-rGO aerogel.
Nanomaterials 13 00732 g004
Figure 5. FTIR spectra of Mn2O3, rGO, N-rGO, N,S-rGO, Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels.
Figure 5. FTIR spectra of Mn2O3, rGO, N-rGO, N,S-rGO, Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels.
Nanomaterials 13 00732 g005
Figure 6. Raman spectra of Mn2O3, rGO, N-rGO, N,S-rGO, Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels.
Figure 6. Raman spectra of Mn2O3, rGO, N-rGO, N,S-rGO, Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels.
Nanomaterials 13 00732 g006
Figure 7. XPS spectra of the survey scan (a,b), C 1s (c,d), O 1s (e,f), N 1s (g,h), Mn 2p (i,j) for the Mn2O3/N-rGO and Mn2O3/N,S-rGO aerogels, respectively, and (k) S 2p spectra of the Mn2O3/N,S-rGO aerogel.
Figure 7. XPS spectra of the survey scan (a,b), C 1s (c,d), O 1s (e,f), N 1s (g,h), Mn 2p (i,j) for the Mn2O3/N-rGO and Mn2O3/N,S-rGO aerogels, respectively, and (k) S 2p spectra of the Mn2O3/N,S-rGO aerogel.
Nanomaterials 13 00732 g007
Figure 8. CV curves of (a) Mn2O3, (b) Mn2O3/rGO aerogel, (c) Mn2O3/N-rGO aerogel, and (d) Mn2O3/N,S-rGO aerogel at a scanning rate of 0.1 mV s−1.
Figure 8. CV curves of (a) Mn2O3, (b) Mn2O3/rGO aerogel, (c) Mn2O3/N-rGO aerogel, and (d) Mn2O3/N,S-rGO aerogel at a scanning rate of 0.1 mV s−1.
Nanomaterials 13 00732 g008
Figure 9. Charge and discharge profiles of (a) Mn2O3, (b) Mn2O3/rGO aerogel, (c) Mn2O3/N-rGO aerogel, and (d) Mn2O3/N,S-rGO aerogel at 0.1 A g−1.
Figure 9. Charge and discharge profiles of (a) Mn2O3, (b) Mn2O3/rGO aerogel, (c) Mn2O3/N-rGO aerogel, and (d) Mn2O3/N,S-rGO aerogel at 0.1 A g−1.
Nanomaterials 13 00732 g009
Figure 10. (a) Cyclability and Coulombic efficiencies and (b) rate capabilities of Mn2O3, Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels.
Figure 10. (a) Cyclability and Coulombic efficiencies and (b) rate capabilities of Mn2O3, Mn2O3/rGO, Mn2O3/N-rGO, and Mn2O3/N,S-rGO aerogels.
Nanomaterials 13 00732 g010
Table 1. Comparison of the electrochemical performances of the Mn-based oxide anode materials for Na-ion batteries.
Table 1. Comparison of the electrochemical performances of the Mn-based oxide anode materials for Na-ion batteries.
Mn-Based Anode MaterialsInitial Discharge Capacity (mAh g−1)Discharge Capacity at 2nd Cycle (mAh g−1)Discharge Capacity after 100 Cycles (mAh g−1)Reference
MnO/graphene553316191[68]
MnO/C560-260[69]
Mn3O4/pGS700-233[56]
Mn3O4/NC710.8-249.5[70]
Mn3O4/N-rGO aerogel1950470374[35]
Mn3O4/N,S-rGO aerogel884425281[35]
Mn2O3/N-rGO aerogel2507594325This work
Mn2O3/N,S-rGO aerogel1677448279This work
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

Mahamad Yusoff, N.F.; Idris, N.H.; Md Din, M.F.; Majid, S.R.; Harun, N.A.; Noerochim, L. Coupling of Mn2O3 with Heteroatom-Doped Reduced Graphene Oxide Aerogels with Improved Electrochemical Performances for Sodium-Ion Batteries. Nanomaterials 2023, 13, 732. https://doi.org/10.3390/nano13040732

AMA Style

Mahamad Yusoff NF, Idris NH, Md Din MF, Majid SR, Harun NA, Noerochim L. Coupling of Mn2O3 with Heteroatom-Doped Reduced Graphene Oxide Aerogels with Improved Electrochemical Performances for Sodium-Ion Batteries. Nanomaterials. 2023; 13(4):732. https://doi.org/10.3390/nano13040732

Chicago/Turabian Style

Mahamad Yusoff, Nor Fazila, Nurul Hayati Idris, Muhamad Faiz Md Din, Siti Rohana Majid, Noor Aniza Harun, and Lukman Noerochim. 2023. "Coupling of Mn2O3 with Heteroatom-Doped Reduced Graphene Oxide Aerogels with Improved Electrochemical Performances for Sodium-Ion Batteries" Nanomaterials 13, no. 4: 732. https://doi.org/10.3390/nano13040732

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