Next Article in Journal
Self-Filtering Monochromatic Infrared Detectors Based on Bi2Se3 (Sb2Te3)/Silicon Heterojunctions
Previous Article in Journal
Comparison of Different Approaches to Surface Functionalization of Biodegradable Polycaprolactone Scaffolds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

N-Doped Modified Graphene/Fe2O3 Nanocomposites as High-Performance Anode Material for Sodium Ion Storage

Material and Energy School, Guangdong University of Technology, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2019, 9(12), 1770; https://doi.org/10.3390/nano9121770
Submission received: 17 November 2019 / Revised: 8 December 2019 / Accepted: 10 December 2019 / Published: 12 December 2019

Abstract

:
Sodium-ion storage devices have received widespread attention because of their abundant sodium resources, low cost and high energy density, which verges on lithium-ion storage devices. Electrochemical redox reactions of metal oxides offer a new approach to construct high-capacity anodes for sodium-ion storage devices. However, the poor rate performance, low Coulombic efficiency, and undesirable cycle stability of the redox conversion anodes remain a huge challenge for the practical application of sodium ion energy storage devices due to sluggish kinetics and irreversible structural change of most conversion anodes during cycling. Herein, a nitrogen-doping graphene/Fe2O3 (N-GF-300) composite material was successfully prepared as a sodium-ion storage anode for sodium ion batteries and sodium ion supercapacitors through a water bath and an annealing process, where Fe2O3 nanoparticles with a homogenous size of about 30 nm were uniformly anchored on the graphene nanosheets. The nitrogen-doping graphene structure enhanced the connection between Fe2O3 nanoparticles with graphene nanosheets to improve electrical conductivity and buffer the volume change of the material for high capacity and stable cycle performance. The N-GF-300 anode material delivered a high reversible discharge capacity of 638 mAh g−1 at a current density of 0.1 A g−1 and retained 428.3 mAh g−1 at 0.5 A g−1 after 100 cycles, indicating a strong cyclability of the SIBs. The asymmetrical N-GF-300//graphene SIC exhibited a high energy density and power density with 58 Wh kg−1 at 1365 W kg−1 in organic solution. The experimental results from this work clearly illustrate that the nitrogen-doping graphene/Fe2O3 composite material N-GF-300 is a potential feasibility for sodium-ion storage devices, which further reveals that the nitrogen doping approach is an effective technique for modifying carbon matrix composites for high reaction kinetics during cycles in sodium-ion storage devices and even other electrochemical storage devices.

Graphical Abstract

1. Introduction

With the rapid development of portable mobile electronic devices and electric vehicles, the importance of high-performance electrical energy storage (EES) devices is becoming increasingly prominent. The EES devices based on sodium ion such as sodium ion batteries (SIBs) and sodium ion supercapacitors (SICs) have attracted great attention due to their low cost, the rich abundance of sodium resource, and their high energy density, which approaches that of lithium-ion energy storage devices [1,2,3]. However, there are still some obstacles hindering the practical application of sodium ion energy storage devices, such as poor rate performance, low coulombic efficiency, and undesirable cycle stability. The foremost reason is the bigger radius of Na+ (0.102 nm) compared to Li+ (0.076 nm), which leads to low reaction kinetics for Na+ insertion/extraction from the anode materials [4,5,6]. Meanwhile, excellent anode materials should have suitable microscopic internal structures that can accommodate the lager Na+ and enable reversible insertion/extraction electrochemical behavior, and should also be compatible with electrolytes, cost-efficient, and easy to prepare. Therefore, the exploitation of prominent anode materials, synchronously coupling with optimum electrolytes, is the main challenge for sodium ion energy storage devices.
For Na+-storage negative electrode, alloy-based materials such as transition metal oxides (TMOs) and transition metal sulfides (TMSs) have been attracting increased interest because of their tailored morphology and high capacity. Among them, Fe2O3 is considered to be one of the most promising candidates due to their high theoretical specific capacity (1007 mAh g−1), abundance in earth, easy availability, low cost, and environmental friendliness [7,8]. Compared with bulk electrode materials, nanoscaled materials provide larger active surface area and shorter ion transport distances, which can significantly improve ion diffusion efficiency and enhance rate capability [9]. Valvo et al. investigated nanostructured Fe2O3 in composite-coated electrodes for the first time as a potentially viable negative electrode for SIBs with a capacity of 250 mAh g−1 after 60 cycles at 130 mA g−1 [10]. Nevertheless, iron oxides in nano size suffer from low electron conductivity and the defect of volume expansion and contraction during the charge/discharge processes, finally leading to low capacity and poor cycling performance [11,12,13]. To resolve this issue of volume change, a good strategy is to coat conductive carbon materials with Fe2O3 to form multicomponent composite materials. Then, graphene is considered to be the preferred carbon material because of its special two-dimensional structure, the excellent electrical conductivity and large theoretical specific surface area (2630 m2 g−1) [1,14]. The graphene nanosheets can provide enough active surface to anchor nano-Fe2O3 particles and offer an extremely conductive framework to promote electron transport [15]. Thus, a series of nanostructured Fe2O3/graphene composites, for example Fe2O3@GNS (graphene nanosheet), Fe2O3/r-GONRAs (reduced graphene oxide nanorod arrays), Fe2O3/N-GNS (nitrogen-doped graphene nanosheet), a-Fe2O3/pBC-N (nitrogen-doped pyrolyzed bacterial cellulose), γ-Fe2O3/PCF (porous carbon fibers), Fe2O3@C@MoS2@C and so on, were synthesized by various methods as anode electrode materials for SIBs with diverse electrochemical properties listing in Table 1. Among these, Li et al. synthesized Fe2O3/rGO nanocomposites, delivered 500 mAh g−1 after 100 cycles at a current density of 50 mA g−1, which was attributed to the single crystal structure of the Fe2O3 particles, and could maintain the structural integrity of the electrode and the electrochemical reversibility of the conversion process [15]. Chen’s group prepared 3D (three-dimensional) porous γ-Fe2O3@Carbon nanocomposites with ultrasmall Fe2O3 nanoparticles, stable C matrix, and 3D porous structure, providing a capacity of 317 mAh g−1 even at a current density of 8 A g−1 [16].
In this paper, nitrogen atoms were introduced to fabricate N-doping graphene/Fe2O3 (GF) nanocomposites as Na+ storage anode materials for SIBs and SISs. Fe2O3 nanocrystalline particles with a size of about 50 nm are uniformly anchored onto graphene nanosheets to obtain the GF sample through a one-step water bath synthesis method. Then, in order to further reduce graphene oxide for high conductivity, GF-300 was prepared. Finally, nitrogen atoms were doped into the graphene nanosheets by a further heat treatment in ammonia gas, and the obtained sample was noted as N-GF-300. The nitrogen-doped graphene-based framework shown in Figure 1 resulted in some advantages in this work, as follows: 1) enhancing the Fe2O3-graphene binding through the large electron affinity of nitrogen; 2) improving the nanocomposites of electrical conductivity to accelerate charge transport; 3) increasing reactive sites to improve energy conversion efficiency; and 4) producing the defects into the graphene nanosheets for high capacity [27].
Benefitting from the superior structural advantages, the final sample of N-GF-300 was proved to have excellent Na+ storage properties as the anode material for SIBs and SICs. The relative SIBs delivered a reversible discharge capacity of 638 mAh g−1 at 0.1 A g−1 and retained 300 mAh g−1 at 0.5 A g−1 after 100 cycles. Meanwhile, the asymmetrical N-GF-300//graphene SIC (the N-GF-300 sample was as the anode and graphene as the cathode material) exhibited a high energy density and power density with 58 Wh kg−1 at 1365 W kg−1 in 1 M NaClO4 organic solution. Furthermore, the electrochemical impedance spectroscopy (EIS) was utilized to seriously analyze the reason nitrogen doping into N-GF-300 could enhance the electrochemical performance, and it was finally revealed that nitrogen doping into the carbon-based electrode material is an effective method to improve the electrochemical properties for electrode materials.

2. Experimental Section

2.1. Synthesis of the Material

Firstly, the graphene/Fe2O3 (GF) nanocomposite was synthesized by a water bath process shown in Figure 1. In a typical synthesis procedure, 100 mL FeCl3·6H2O (0.901 g) solution was slowly added to 60 mL graphene oxide (2 mg mL−1) dispersion solution in a 250 mL round-bottom flask. After stirring for 2.5 h, 0.096 mL hydrazine hydrate (85 wt.%) was added as a reductant to the above solution; then, the mixed solution was subjected to a water bath at 80 °C for 24 h. The whole low-temperature hydrothermal process was kept under stirring. The product of the low-temperature hydrothermal reaction was treated by a series of procedures like vacuum filtering, washing and freeze-drying, and then the sample GF nanocomposite could be obtained. The GF-300 sample was prepared by annealing GF at 300 °C for 2 h under an argon atmosphere. The N-GF-300 was finally obtained via heating GF-300 to dope nitrogen atoms at 600 °C for 2 h in an ammonia atmosphere.

2.2. Structural Characterization of the Material

The morphology of the prepared material was observed by using a Hitachi SU8010 field emission scanning electron microscope (FESEM) (Hitachi, Japan). High-resolution transmission electron microscope (HRTEM) and mapping of the material were obtained by using a JEM-2100F field emission transmission electron microscope (JEOL, Japan). The Raman spectrum was collected using a Horiba LabRAM HR800 Raman spectrometer (HORIBA, France) with a 532 nm laser wavelength under a grating of 1800 gr/mm. X-ray diffraction (XRD) was measured by using a Bruker D8 Advance X-ray diffractometer (BRUKER, Germany) with a copper target. To test the content of Fe2O3 in the composite, thermogravimetric analysis (TGA) was recorded on a Thermogravimetric and synchronous thermal-infrared spectroscopy (Mettler-Toledo, USA) combined system in air from 10 to 800 °C at a heating rate of 10 °C min−1. X-ray photoelectron spectroscopy (XPS) was conducted on PHI 5000 VersaProbe II in situ X-ray photoelectron spectrometer (Ulvac-Phi, Japan). The specific surface area and pore size distribution curves were measured by the Brunauer Emmett Teller (BET) and density functional theory (DFT) method on a Micromeritics ASAP 2460 system (MICROMERITICS, USA), respectively.

2.3. Electrochemical Measurements

The electrochemical properties of the electrode material were measured by assembling them into a button cell. The electrode was produced by placing the mixture, which consisted of 60% material, 30% superconducting carbon black, and 10% PTFE (polytetrafluoroethylene) binder, onto a nickel foil and drying at 120 °C for 10 h under vacuum. The button cell (2032 type) was assembled in an argon glove box with water and oxygen content below 0.1 using the previously prepared electrode material as the working electrode, sodium as the auxiliary electrode and the counter electrode. The electrolyte used was 1 M sodium trifluomethanesulfonate (NaCF3SO3) in diglyme = 100 vol%, 1M NaClO4 in EC (ethylene carbonate): DMC (dimethyl carbonate): EMC (ethyl methyl carbonate) = 1:1:1 vol% with 5% FEC and 1 M NaPF6 in EC:DMC:EMC = 1:1:1 vol% with 5% FEC. The galvanostatic discharge/charge test was performed on a Land cycler (Wuhan Kingnuo Electronic Co., Wuhan, China). Cyclic voltammetry (CV) was tested on an CHI660E (CH Instruments, Shanghai, China) electrochemical workstation at a certain scan rate between 0.005 V and 3 V. The electrochemical impedance spectroscopy (EIS) was measured on the CHI660E electrochemical workstation with an alternating current (AC) voltage of 5 mV amplitude from 100 kHz to 100 mHz. For sodium ion capacitor devices, the cathode was made with a mixture of 60% graphene, 30% superconducting carbon black, and 10% PTFE and then placed on foamed nickel. For the anode, it was the same as the anode material used in the above assembled sodium ion battery, followed by assembly into a device in a glove box. The electrolyte used was 1 M sodium perchlorate (NaClO4) in EC:DMC:EMC = 1:1:1 Vol% with 5% FEC. The assembled devices were tested for EIS, CV, galvanostatic charge and discharge (GCD), etc., and the test parameters are the same as before. The specific energy density and power density of the sodium ion capacitor were calculated by the following formula [28]:
E   ( Wh   kg 1 ) = 0.5 C × Δ V 2 3.6
P   ( W   kg 1 ) = E × 3600 Δ t
where C is the specific capacitances of the supercapacitor cell, △V (V) is the voltage window during discharge process after the IR drop, △t (s) is the discharge time, E (Wh kg−1) is the specific energy density and P (W kg−1) is the power density.

3. Results and Discussion

Figure 1 shows a sketch diagram of the synthesis process of the N-GF-300 nanocomposite through a water bath and an annealing process in NH3 atmosphere. The water bath method with low temperature is a feasible approach for synthesizing the gram level above for nanomaterials with high production and low energy consumption.
The morphology of GF, GF-300, N-GF-300 was measured by FESEM, as shown in Figure 2. In the SEM image of N-GF-300 (Figure 2c), Fe2O3 nanoparticles with a homogenous size of about 30 nm were uniformly anchored on the graphene nanosheets, revealing a better connection between Fe2O3 nanoparticles with graphene nanosheets compared to GF and GF-300. The nanostructured Fe2O3 particles in the N-GF-300 composite are tightly coated by the N-doped graphene sheet matrix, which facilitates the sodium ions rapidly inserted and extracted to improve the electronic conductivity. Most importantly, in terms of size, Fe2O3 particles in N-GF-300 are smaller than those in GF and GF-300 because of the greater number of active sites and the heterogeneous nucleation caused by N-doped graphene sheets [29]. The isotherm diagrams in Figure 2d clearly show the type IV adsorption isotherm with a distinct hysteresis loop indicating the mesoporous structure of these materials. The annealing at 300 °C for GF-300 with lower specific surface area (SSA) (97.9 m2 g−1) revealed the removal of unstable microporous structures from GF (102.1 m2 g−1), and the N-doping process further improved the SSA to 102.1 m2 g−1 with richer mesopores. As a whole, the low SSA and abundant mesoporous structure of N-GF-300 can effectively reduce the side reactions of electrolyte and increase ion transport channels. The XRD patterns were measured to investigate the crystal structure of the as-prepared GF, GF-300 and N-GF-300 materials, and all characteristic peaks were in good agreement with the standard card of Fe2O3 (JCPDS card No. 33-0664) (Figure 2e). The distinct peak of graphene is inconspicuous from the XRD pattern, which is attributed to the low graphene sheet content and the high removal of oxygen inserted into the graphite layer by reduction [30,31]. Raman spectroscopy was conducted to characterize the structural features of the graphene in GF, GF-300 and N-GF-300 composites. As shown in the Raman spectra (Figure 2f), the two main peaks located at 1394 and 1592 cm−1 were attributed to be D and G band of graphene, respectively [32]. The front band (D band) was attributed to the graphite lattice vibration of the Raman-active E2g2 mode, while the back band (G band) was attributed to the A1g mode, which comes from certain types of disorders and defects in the graphite crystal structure. The ID/IG (the intensity ratio of the D and G bands) ratios are presented in increasing order of GF, GF-300 and N-GF-300, and are 1.28, 1.37 and 1.38, respectively. The larger number of the ID/IG demonstrates the more amorphous structure of the carbon-based materials [33]. Therefore, the N-GF-300 composite has a more disordered microstructure than GF-300 and GF, and this result indicates that the nitrogen atoms had been successfully doped and the defects had been produced in the structure of the graphene sheets in N-GF-300, changing the perfect six-carbon ring structure [34].
HRTEM and mapping images of N-GF-300 were measured in order to further study the microstructure and the distribution of chemical elements as shown in Figure 3a–d, respectively. In Figure 3a,b, the Fe2O3 and graphene phases can be clearly distinguished. In addition, in Figure 3c,d, the mapping of N-GF-300 confirmed that the elements of, C, O, Fe and N were uniformly distributed in the composite. The HRTEM image (Figure 3e) shows that lattice fringes of the composite can be observed with a pitch of (012). XPS measurements were performed to further understand the surface chemistry of N-GF-300 and reveal that the weight content of N is up to 3.66%, which is a moderate doping level. As shown in Figure 3g, the two major peaks of Fe 2p located at 711.2 and 724.9 eV correspond to Fe 2p2/3 and Fe 2p1/2, respectively, which is consistent with Fe2O3 previously reported [18,35]. More importantly, the XPS spectra manifest the strong interface interaction that exists between Fe2O3 and N-doping graphene nanosheets in the composite of N-GF-300. As seen from Figure 3f, the C1s spectrum of N-GF-300 is mainly composed of six kinds of bonds: C–C/C=C (284.6 eV), C–N (285.6 eV), C–O (286.5 eV), C=O (287.8 eV) and O=C–O (289.4 eV) [7,18,21,36]. The N1s spectrum is made up of three peaks in Figure 3h, located at 398.6, 399.8 and 401 eV, which correspond to pyridinic-N, pyrrolic-N and graphitic-N, respectively [37,38].
The contents of three N-doping types as pyridinic-N, pyrrolic-N and graphitic-N (Figure 4b) in N-GF-300, GF-300, GF samples were compared in Figure 4a. The peak strength and peak area of pyridinic-N in N-GF-300 is much bigger than that of GF-300 and GF, and the result indicates that the nitrogen doping carried out by ammonia gas was beneficial in increasing the content of pyridinic-N. Notably, pyridinic-N was proved to enhance the conduction of electrons and facilitate diffusion of sodium ions [39]; therefore, N-GF-300 is more beneficial to improving the electrochemical performance for SIBs and SISs.
The electrochemical performance of N-GF-300 as the anode material of SIBs and SISs was expressed by CV, GCD and EIS. In the SIBs, three different electrolytes, including 1.0 M NaClO4, 1.0 M NaPF6 in EC:DMC:EMC = 1:1:1 vol% with 5.0% FEC, respectively, and 1.0 M NaCF3SO3 in diglyme were investigated by GCD, as shown in Figure 5a. Compared to 1.0 M NaClO4 and 1.0 M NaPF6 in EC:DMC:EMC = 1:1:1 vol% with 5.0% FEC, the SIBs with the ether-based electrolyte of 1.0 M NaCF3SO3 in diglyme delivered the highest discharge capacity of 378.4 mAh g−1 at a current of 0.5 A g−1. The reason for this is attributed a special solid electrolyte interphase (SEI) formed in the ether-based electrolyte [40], and the details of this will be discussed in our future work. Therefore, the studies on the electrochemical performance of SIBs carried out later were based on the ether electrolytes. Figure 5b shows the first three cycles of the CV curve for the N-GF-300 electrode measured at a scan rate of 0.05 mVs−1 at a range of 0.005–3 V. According to previous reports, the insertion and extraction of sodium ions in Fe2O3 is based on the following reaction [5,10,16,17]:
F e 2 O 3 + 6 N a + + 6 e 2 F e 0 + 3 N a 2 O
In the first CV curve from Figure 5b, there are two cathode peaks appearing at 0.5 and 0.89 V, respectively, due to the formation of the SEI film and the produce of NaxFe2O3 from the insertion of sodium ions into the Fe2O3. In the second CV curve, the cathode peaks at 0.63 and 0.94 V are accompanied by anode peaks at 0.6 and 1.6 V, which correspond to the reversible redox reaction of Fe3+ ↔Fe2+ and Fe2+ ↔Fe0 [5,16]. Figure 5c provides a comparison of the CV curves of the three electrode materials of GF, GF-300 and N-GF-300 at a scan rate of 1 mV s−1. Obviously, the CV curve area of the N-GF-300 material is larger than the other two, corresponding to the highest capacitance of 703 F g−1 compared to 532.5 F g−1 of GF-300 and 300.5 F g−1 of GF. Figure 5d is the rate performance of the three samples at different current densities. The discharge specific capacities of N-GF-300 at 0.1, 0.2, 0.5, 1, and 2 A g−1 are 638.4, 565.6, 464.9, 380.7 and 297.4 mAh g−1, respectively. When the current density returned to 0.1 and 0.2 A g−1, the discharge specific capacities of N-GF-300 could also return to the start level with 632.2 and 520.2 mAh g−1, respectively. In general, compared to GF and GF-300, the rate performance of N-GF-300 is superior to GF and GF-300, and more obvious at the higher the current density, which is attributed to the increasing charge transport of N-GF-300 by doping nitrogen atoms into the graphene sheets. The N-GF-300 electrode exhibits stable cycling performance and high Coulombic efficiency, as shown in Figure 5e. After 100 cycles at a current density of 0.5 A g−1, the discharge capacity of N-GF-300 could be maintained at up to 88.8%, while only 53% and 50.5% of the discharge capacity could be retained for GF-300 and GF respectively.
Electrochemical impedance spectroscopy (EIS) was utilized to further discover the internal energy storage mechanism of N-GF-300 as the anodes in SIBs. Figure 5f provides the EIS measurements of three composite electrode materials. The Nyquist diagram consists of a semicircle that intersects the graph and the accompanying line portion. The semicircular portion (in the high frequency region) corresponds to the electron transfer limiting process, and the straight portion (low frequency region) represents the diffusion limiting process. In the faster electron transfer process, the AC impedance spectrum contains only the straight portion, while the slower electron transfer process is characterized by a large semicircular region. The diameter of the semicircle is equal to the electron transfer resistance (Rct). The intercept of the semicircle on the real axis (Zre) to the electrolyte resistance Re [41,42]. According to the circuit diagram [43], the fitting results are shown in Table 2. The solution resistance of N-GF-300 and GF-300 is similar to that of GF for using the same electrolyte, but the electron transfer resistance of N-GF-300 (Rct = 18.24 Ω) is obviously smaller than that of GF (Rct = 52.31 Ω) and GF-300 (Rct = 40.17 Ω), indicating that the doping of N atoms is advantageous to enhance the electrical conductivity and electrochemical activity of N-GF-300 during the electrochemical process for the anode of SIBs [34].
Except for the anode material of the SIBs, N-GF-300 is also a highly appropriate anode material for SICs. In this work, an asymmetric SIC (Graphene//N-GF-300) comprised of graphene as the cathode material and N-GF-300 as the anode material was investigated by a series of electrochemical measurements, as shown in Figure 6, and a symmetric SIC N-GF-300//N-GF-300 was assembled for comparison.
The CV curves of the SICs Graphene//N-GF-300 and N-GF-300//N-GF-300 in Figure 6a maintained a typical quasi-rectangular shape. At a scan rate of 20 mV s−1, the specific capacitance of the symmetric SIC N-GF-300//N-GF-300 was 130 F g−1, while the Graphene//N-GF-300 supercapacitor had a much higher specific capacitance of 228 F g−1 due to the asymmetric configuration. Figure 6b shows the rate performance of the Graphene//N-GF-300 and N-GF-300//N-GF-300 capacitors at current densities of 1, 2, 5, 8, 10, 1 and 2 A g−1. The results show that the electrochemical performance of Graphene//N-GF-300 is significantly better than that of N-GF-300//N-GF-300. The energy density and power density of the SICs were calculated according to Equations (1) and (2), as shown in the Ragone plot (Figure 6c). The Graphene//N-GF-300 asymmetric capacitors exhibit higher energy density and power density compared to the N-GF-300//N-GF-300 symmetric capacitors. The energy density of the asymmetric capacitor reached 58 Wh kg−1 (1365 W kg−1) at a current density of 1 A g−1. Even with the current density reaching up to 10 A g−1, the energy density of the asymmetric capacitor Graphene//N-GF-300 still retained 28.1 Wh kg−1 (10,116 W kg−1).
Herein, the EIS measurement was used to further verify the superiority of the Graphene//N-GF-300 capacitor by analyzing internal resistance value of SICs. Figure 6d shows the Nyquist plots of the Graphene//N-GF-300 and N-GF-300//N-GF-300 devices, and the corresponding experimental impedance data and fitted data consisted of a loop in the high frequency region and a sloped line in the low frequency region. The equivalent circuit for fitting is Re(Q(RctZw))Cl and its schematic diagram is inset in Figure 6d. Here is a brief overview of the meaning of the five equivalent components: Re is the equivalent series resistance, Rct is the charge transfer resistance, Zw is the Warburg impedance, Q is the constant phase element (CPE), and C1 is the finite capacitance. Re includes the resistance of the electrolyte, the inherent resistance of the electrode material, and the resistance at the electrode material/collector interface. Rct with respect to the diameter of the semicircle in the high frequency region indicates the dynamic resistance of charge transfer when the Faraday reaction occurs. Zw is related to the slope of the 45° portion of the Nyquist plot in the intermediate frequency region, which is the impedance of the diffusion control process in the electrolyte. Cl is a finite capacitor. The electric double layer at the electrode/electrolyte boundary is usually equal to the electric double layer capacitor, but this deviates from the almost pure electrode layer due to the dispersion effect. The equivalent element denoted this part as CPE(Q). Its impedance is defined as:
ZQ = [Y0(jw)n]−1
leading to
Z Q = [ Y 0 ( w ) n ] 1 cos ( n π 2 )
Z Q = [ Y 0 ( w ) n ] 1 sin ( n π 2 )
where the dimension unit of Y0 is 1 S cm−2 s−n, which is a constant independent of frequency, and n is an exponent with the value between −1 and 1. When n = −1, the CPE is a pure inducer; when n = 0, the CPE is a pure resistor; when n = 1, the CPE is a pure capacitor; when 0 < |n|< 1, the CPE is Q [33]. Table 3 lists the value of each element in the equivalent circuit calculated by the complex nonlinear least-squares (CNLS) method. As can be seen from Table 3, Graphene//N-GF-300 is obviously better than N-GF-300//N-GF-300. Figure 6e shows the cycle performance of Graphene//N-GF-300 capacitors at current densities of 5 A g−1. In addition, the two Graphene//N-GF-300 capacitor devices in series can illuminate a red LED lamp (minimum operating voltage is 3 V), as shown in Figure 6f.

4. Conclusions

In summary, a nitrogen-doping graphene/Fe2O3 composite material N-GF-300 was successfully prepared as Na+ storage anode for SIBs and SICs through a water bath and an annealing process. The nitrogen-doping graphene-based framework in N-GF-300 enhanced Fe2O3-graphene binding through the high electron affinity of nitrogen to improve electrical conductivity and accelerate charge transport. Meanwhile, the reactive sites and defects increased in the graphene nanosheets during the nitrogen atom doping process, which could improve energy conversion efficiency at high capacity. Finally, the excellent electrochemical performance of the relative SIBs was indicated, with high capacity, good rate capability and stable cycling performance. A reversible discharge capacity of 638 mAh g−1 was delivered at the current density of 0.1 A g−1, and 428.3 mAh g−1 was retained at 0.5 A g−1 after 100 cycles. The asymmetrical N-GF-300//graphene SIC exhibited a high energy density and power density with 58 Wh kg−1 at 1365 W kg−1 in organic solution.

Author Contributions

Conceptualization, C.Z.; Formal analysis, Y.C. and C.Z.; Funding acquisition, C.Z.; Investigation, Y.C.; Methodology, Y.C.; Project administration, C.Z.; Resources, C.Z. and H.Z.; Supervision, C.Z.; Validation, Y.C., Z.G. and B.J.; Visualization, Y.C.; Writing—original draft, Y.C. and C.Z.; Writing—review & editing, Z.G. and C.Z.

Funding

This research is funded by the Science and Technology Project of Guangzhou (201804010397).

Conflicts of Interest

The authors declare no conflict of interest.

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.; et al. 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. 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]
  6. Yang, H.; Xu, R.; Yao, Y.; Ye, S.; Zhou, X.; Yu, Y. Multicore–Shell Bi@N-doped Carbon Nanospheres for High Power Density and Long Cycle Life Sodium- and Potassium-Ion Anodes. Adv. Funct. Mater. 2019, 29, 1809195. [Google Scholar] [CrossRef]
  7. Meng, S.; Zhao, D.L.; Wu, L.L.; Ding, Z.W.; Cheng, X.W.; Hu, T. Fe2O3/nitrogen-doped graphene nanosheet nanocomposites as anode materials for sodium-ion batteries with enhanced electrochemical performance. J. Alloys Compd. 2018, 737, 130–135. [Google Scholar] [CrossRef]
  8. Yang, Q.; Bi, R.; Yung, K.; Pecht, M. Electrochemically reduced graphene oxides/nanostructured iron oxides as binder-free electrodes for supercapacitors. Electrochim. Acta 2017, 231, 125–134. [Google Scholar] [CrossRef]
  9. Wang, K.X.; Li, X.H.; Chen, J.S. Surface and interface engineering of electrode materials for lithium-ion batteries. Adv. Mater. 2015, 27, 527–545. [Google Scholar] [CrossRef]
  10. Valvo, M.; Lindgren, F.; Lafont, U.; Björefors, F.; Edström, K. Towards more sustainable negative electrodes in Na-ion batteries via nanostructured iron oxide. J. Power Sources 2014, 245, 967–978. [Google Scholar] [CrossRef]
  11. Cherian, C.T.; Sundaramurthy, J.; Kalaivani, M.; Ragupathy, P.; Kumar, P.S.; Thavasi, V.; Reddy, M.V.; Sow, C.H.; Mhaisalkar, S.G.; Ramakrishna, S.; et al. Electrospun α-Fe2O3 nanorods as a stable, high capacity anode material for Li-ion batteries. J. Mater. Chem. 2012, 22, 12198–12204. [Google Scholar] [CrossRef]
  12. Zhang, G.; Shi, Y.; Wang, H.; Jiang, L.; Yu, X.; Jing, S.; Xing, S.; Tsiakaras, P. A facile route to achieve ultrafine Fe2O3 nanorods anchored on graphene oxide for application in lithium-ion battery. J. Power Sources 2019, 416, 118–124. [Google Scholar] [CrossRef]
  13. Li, Y.; Huang, Y.; Zheng, Y.; Huang, R.; Yao, J. Facile and efficient synthesis of α-Fe2O3 nanocrystals by glucose-assisted thermal decomposition method and its application in lithium ion batteries. J. Power Sources 2019, 416, 62–71. [Google Scholar] [CrossRef]
  14. Sharifi, T.; Gracia-Espino, E.; Barzegar, H.R.; Jia, X.; Nitze, F.; Hu, G.; Nordblad, P.; Tai, C.W.; Wagberg, T. Formation of nitrogen-doped graphene nanoscrolls by adsorption of magnetic gamma-Fe2O3 nanoparticles. Nat. Commun. 2013, 4, 1–9. [Google Scholar] [CrossRef]
  15. Li, T.; Qin, A.; Yang, L.; Chen, J.; Wang, Q.F.; Zhang, D.H.; Yang, H.X. 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]
  16. 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]
  17. Jian, Z.; Zhao, B.; Liu, P.; Li, F.; Zheng, M.; Chen, M.; Shi, Y.; Zhou, H. Fe2O3 nanocrystals anchored onto graphene nanosheets as the anode material for low-cost sodium-ion batteries. Chem. Commun. 2014, 50, 1215–1217. [Google Scholar] [CrossRef]
  18. 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]
  19. Li, H.; Xu, L.; Sitinamaluwa, H.; Wasalathilake, K.; Yan, C. Coating Fe2O3 with graphene oxide for high-performance sodium-ion battery anode. Compos. Commun. 2016, 1, 48–53. [Google Scholar] [CrossRef] [Green Version]
  20. Liu, X.; Chen, T.; Chu, H.; Niu, L.; Sun, Z.; Pan, L.; Sun, C.Q. Fe2O3-reduced graphene oxide composites synthesized via microwave-assisted method for sodium ion batteries. Electrochim. Acta 2015, 166, 12–16. [Google Scholar] [CrossRef]
  21. Meng, J.; Suo, Y.; Li, J.; Zheng, G.; Liu, Y.; Zhang, J.; Zheng, X. Nitrogen-doped graphene aerogels as anode materials for lithium-ion battery: Assembly and electrochemical properties. Mater. Lett. 2015, 160, 392–396. [Google Scholar] [CrossRef]
  22. Shi, L.; Li, Y.; Zeng, F.; Ran, S.; Dong, C.; Leu, S.Y.; Boles, S.T.; Lam, K.H. In situ growth of amorphous Fe2O3 on 3D interconnected nitrogen-doped carbon nanofibers as high-performance anode materials for sodium-ion batteries. Chem. Eng. J. 2019, 356, 107–116. [Google Scholar] [CrossRef]
  23. Chen, Y.; Yuan, X.; Yang, C.; Lian, Y.; Razzaq, A.A.; Shah, R.; Guo, J.; Zhao, X.; Peng, Y.; Deng, Z. γ-Fe2O3 nanoparticles embedded in porous carbon fibers as binder-free anodes for high-performance lithium and sodium ion batteries. J. Alloys Compd. 2019, 777, 127–134. [Google Scholar] [CrossRef]
  24. Xu, D.; Chen, W.; Zheng, M.; Huang, X.; Fang, Y.; Yu, X. Nanoflakes assembled hydrangea-like Fe2O3@C@MoS2@C nanocomposite as high performance anode materials for lithium/sodium ion batteries. Electrochim. Acta 2018, 265, 419–429. [Google Scholar] [CrossRef]
  25. Guo, T.; Liao, H.; Ge, P.; Zhang, Y.; Tian, Y.; Hong, W.; Shi, Z.; Shao, C.; Hou, H.; Ji, X. Fe2O3 embedded in the nitrogen-doped carbon matrix with strong C–O–Fe oxygen-bridge bonds for enhanced sodium storages. Mater. Chem. Phys. 2018, 216, 58–63. [Google Scholar] [CrossRef]
  26. Wu, X.; Chen, W.; Key, J.; Wu, W. One-pot solvothermal synthesis of fern leaf-like α-Fe2O3@C/graphene from ferrocene with enhanced lithium and sodium storage properties. Powder Technol. 2018, 323, 424–432. [Google Scholar] [CrossRef]
  27. Lv, R.; Cui, T.; Jun, M.S.; Zhang, Q.; Cao, A.; Su, D.S.; Zhang, Z.; Yoon, S.H.; Miyawaki, J.; Mochida, I.; et al. Open-Ended, N-Doped Carbon Nanotube-Graphene Hybrid Nanostructures as High-Performance Catalyst Support. Adv. Funct. Mater. 2011, 21, 999–1006. [Google Scholar] [CrossRef]
  28. Zhang, J.; Zhao, H.; Li, J.; Jin, H.; Yu, X.; Lei, Y.; Wang, S. In Situ Encapsulation of Iron Complex Nanoparticles into Biomass-Derived Heteroatom-Enriched Carbon Nanotubes for High-Performance Supercapacitors. Adv. Energy Mater. 2019, 9, 1803221. [Google Scholar] [CrossRef]
  29. Meng, X.; Xu, Y.; Sun, X.; Xiong, L.; Wang, Q. Nitrogen-doped graphene assists Fe2O3 in enhancing electrochemical performance. J. Power Sources 2016, 326, 389–396. [Google Scholar] [CrossRef]
  30. Hao, D.; Xuefen, C.; Liangdong, Q.; Xiaohui, Z. Fabrication, Characterization and Properties of Superparamagnetic Reduced Graphene Oxide/Fe3O4 Hollow Sphere Nanocomposites. Rare Metal Mat. Eng. 2016, 45, 1669–1673. [Google Scholar] [CrossRef] [Green Version]
  31. Li, J.; Qin, W.; Xie, J.; Lin, R.; Wang, Z.; Pan, L.; Mai, W. Rational design of MoS2-reduced graphene oxide sponges as free-standing anodes for sodium-ion batteries. Chem. Eng. J. 2018, 332, 260–266. [Google Scholar] [CrossRef]
  32. Zhang, F.; Alhajji, E.; Lei, Y.; Kurra, N.; Alshareef, H.N. Highly Doped 3D Graphene Na-Ion Battery Anode by Laser Scribing Polyimide Films in Nitrogen Ambient. Adv. Energy Mater. 2018, 8, 1800353. [Google Scholar] [CrossRef]
  33. Zheng, C.; Yoshio, M.; Qi, L.; Wang, H. A 4 V-electrochemical capacitor using electrode and electrolyte materials free of metals. J. Power Sources 2014, 260, 19–26. [Google Scholar] [CrossRef]
  34. 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]
  35. Yu, Z.; Li, H.; Lu, J.; Zhang, X.; Liu, N.; Zhang, X. Hydrothermal synthesis of Fe2O3/graphene nanocomposite for selective determination of ascorbic acid in the presence of uric acid. Electrochim. Acta 2015, 158, 264–270. [Google Scholar] [CrossRef]
  36. Liu, Z.; Zhang, H.; Yang, Q.; Chen, Y. Graphene/V2O5 hybrid electrode for an asymmetric supercapacitor with high energy density in an organic electrolyte. Electrochim. Acta 2018, 287, 149–157. [Google Scholar] [CrossRef]
  37. Nair, A.K.; Elizabeth, I.; Gopukumar, S.; Thomas, S.; Kala, M.S.; 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]
  38. Li, X.; Geng, D.; Zhang, Y.; Meng, X.; Li, R.; Sun, X. Superior cycle stability of nitrogen-doped graphene nanosheets as anodes for lithium ion batteries. Electrochem. Commun. 2011, 13, 822–825. [Google Scholar] [CrossRef]
  39. Li, Z.; Wu, G.; Deng, S.; Wang, S.; Wang, Y.; Zhou, J.; Liu, S.; Wu, W.; Wu, M. Combination of uniform SnO2 nanocrystals with nitrogen doped graphene for high-performance lithium-ion batteries anode. Chem. Eng. J. 2016, 283, 1435–1442. [Google Scholar] [CrossRef]
  40. Zhang, J.; Wang, D.W.; Lv, W.; Zhang, S.; Liang, Q.; Zheng, D.; Kang, F.; Yang, Q.H. Achieving superb sodium storage performance on carbon anodes through an ether-derived solid electrolyte interphase. Energy Environ. Sci. 2017, 10, 370–376. [Google Scholar] [CrossRef]
  41. Wang, S.; Li, L.; Shao, Y.; Zhang, L.; Li, Y.; Wu, Y.; Hao, X. Transition-Metal Oxynitride: A Facile Strategy for Improving Electrochemical Capacitor Storage. Adv. Mater. 2019, 31, 1806088. [Google Scholar] [CrossRef] [PubMed]
  42. Ren, H.; Yu, R.; Qi, J.; Zhang, L.; Jin, Q.; Wang, D. Hollow Multishelled Heterostructured Anatase/TiO2 (B) with Superior Rate Capability and Cycling Performance. Adv. Mater. 2019, 31, 1805754. [Google Scholar] [CrossRef] [PubMed]
  43. Sun, H.; Mei, L.; Liang, J.; Zhao, Z.; Lee, C.; Fei, H.; Ding, M.; Lau, J.; Li, M.; Wang, C.; et al. Three-dimensional holey-graphene/niobia composite architectures for ultrahigh-rate energy storage. Science 2017, 356, 599–604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. N-GF-300 synthesis diagram, wherein white: H atom; blue: C atom; red: N atom; green: O atom; blue purple: Fe2O3.
Figure 1. N-GF-300 synthesis diagram, wherein white: H atom; blue: C atom; red: N atom; green: O atom; blue purple: Fe2O3.
Nanomaterials 09 01770 g001
Figure 2. FESEM images of (a) GF; FESEM images of (b) GF-300; FESEM image of (c) N-GF-300; (d) Nitrogen adsorption/desorption isotherms; (e) XRD patterns of GF, GF-300 and N-GF-300; (f) Raman spectra of GF, GF-300 and N-GF-300.
Figure 2. FESEM images of (a) GF; FESEM images of (b) GF-300; FESEM image of (c) N-GF-300; (d) Nitrogen adsorption/desorption isotherms; (e) XRD patterns of GF, GF-300 and N-GF-300; (f) Raman spectra of GF, GF-300 and N-GF-300.
Nanomaterials 09 01770 g002
Figure 3. TEM image indicating a Fe2O3 sphere covered by graphene: (a) N-GF-300; TEM image indicating a Fe2O3 sphere covered by graphene: (b) N-GF-300; (c,d) Distribution of C, N, O and Fe in N-GF-300; (e) HRTEM image showing the lattice fringes; XPS spectra of N-GF-300: (f) C1s, (g) Fe2p, (h) N1s.
Figure 3. TEM image indicating a Fe2O3 sphere covered by graphene: (a) N-GF-300; TEM image indicating a Fe2O3 sphere covered by graphene: (b) N-GF-300; (c,d) Distribution of C, N, O and Fe in N-GF-300; (e) HRTEM image showing the lattice fringes; XPS spectra of N-GF-300: (f) C1s, (g) Fe2p, (h) N1s.
Nanomaterials 09 01770 g003
Figure 4. (a) N1s spectra of GF, GF-300 and N-GF-300, and (b) N-doped graphene framework, with corresponding configurations of N-containing functional groups, wherein green: pyridinic N, red: pyrrolic N, blue: graphitic N.
Figure 4. (a) N1s spectra of GF, GF-300 and N-GF-300, and (b) N-doped graphene framework, with corresponding configurations of N-containing functional groups, wherein green: pyridinic N, red: pyrrolic N, blue: graphitic N.
Nanomaterials 09 01770 g004
Figure 5. (a) Initial galvanostatic charge-discharge curves of different electrolytes for GF at 0.5A g−1 current density; (b) CV curves of the first three cycles for N-GF-300 between 0.005–3 V at scan rate of 0.05 mV s−1; (c) CV curves of GF, GF-300, N-GF-300 electrode at scan rate of 1 mV s−1; (d) Rate performance at different current densities; (e) cycle performance at 0.5 A g−1; (f) EIS of GF, GF-300 and N-GF-300 electrode.
Figure 5. (a) Initial galvanostatic charge-discharge curves of different electrolytes for GF at 0.5A g−1 current density; (b) CV curves of the first three cycles for N-GF-300 between 0.005–3 V at scan rate of 0.05 mV s−1; (c) CV curves of GF, GF-300, N-GF-300 electrode at scan rate of 1 mV s−1; (d) Rate performance at different current densities; (e) cycle performance at 0.5 A g−1; (f) EIS of GF, GF-300 and N-GF-300 electrode.
Nanomaterials 09 01770 g005
Figure 6. (a) CV curves of Graphene//N-GF-300, N-GF-300//N-GF-300 device at scan rate of 20 mV/s; (b) rate performance of Graphene//N-GF-300, N-GF-300//N-GF-300 device at different current densities; (c) Ragone plots of Graphene//N-GF-300, N-GF-300//N-GF-300 device at different current densities; (d) EIS of our prepared device-based Graphene//N-GF-300 and N-GF-300//N-GF-300; (e) cycling performance of Graphene//N-GF-300 device measured at a scan rate of 5 A g−1; (f) digital photograph of an electronic LED powered by two flexible devices in series.
Figure 6. (a) CV curves of Graphene//N-GF-300, N-GF-300//N-GF-300 device at scan rate of 20 mV/s; (b) rate performance of Graphene//N-GF-300, N-GF-300//N-GF-300 device at different current densities; (c) Ragone plots of Graphene//N-GF-300, N-GF-300//N-GF-300 device at different current densities; (d) EIS of our prepared device-based Graphene//N-GF-300 and N-GF-300//N-GF-300; (e) cycling performance of Graphene//N-GF-300 device measured at a scan rate of 5 A g−1; (f) digital photograph of an electronic LED powered by two flexible devices in series.
Nanomaterials 09 01770 g006
Table 1. Progress in the synthesis and electrochemical properties of Fe2O3 composites anode material for SIBs. (The definitions for the electrolyte abbreviations in Table 1 are listed here: EC (ethylene carbonate), DEC (diethyl carbonate), DMC (dimethyl carbonate), PC (propylene carbonate) and FEC (fluoroethylene carbonate))
Table 1. Progress in the synthesis and electrochemical properties of Fe2O3 composites anode material for SIBs. (The definitions for the electrolyte abbreviations in Table 1 are listed here: EC (ethylene carbonate), DEC (diethyl carbonate), DMC (dimethyl carbonate), PC (propylene carbonate) and FEC (fluoroethylene carbonate))
Electrode MaterialsSynthesis MethodVoltage Range (vs. Na/Na+)ElectrolyteCycling DataRate CapabilityRef.
Fe2O3@GNSnanocasting technology0.05–31M NaPF6 in EC:DEC=1:1250/200th/0.5 A g−1110/2 A g−1[17]
Fe2O3/r-GONRAshydrothermal0.01–31M NaPF6 in EC:DEC=1:1 with 5% FEC~332/300th/0.2 A g−192/1.6 C[5]
Fe2O3@GNSice bath0.01–2.51M NaSO3CF3 in diglyme110/500th/2 A g−1194 /2 A g−1[18]
Fe2O3/GOfreezing-dry0.01–2.51M NaClO4 in EC:DMC=1:1420/100th/0.1 C90/10 C[19]
Fe2O3-RGOmicrowave assisted0.005–31M NaClO4 in EC:PC=1:1289/50th/0.05 A g−132.8/2 A g−1[20]
Fe2O3/N-GNShydrothermal0.01–31M NaClO4 in EC:DEC=1:1306/50th/0.05 A g−1132/1 A g−1[7]
Fe2O3/rGOsolvothermal0.01–31M LiPF6 in EC:DEC=1:1 with 5.0% FEC~500/100th/0.05 A g−1216/2 A g−1[15]
3Dporousγ-Fe2O3@Caerosol spray pyrolysis technique0.04–31M NaClO4 in EC:DEC=1:1358/1400th/2 A g−1317/8 A g−1[16]
a-Fe2O3/rGOmicrowave hydrothermal0.05–31M NaClO4 or LiPF6 in EC:DEC=1:1310/150th/0.1 A g−177/2 A g−1[21]
a-Fe2O3/pBC-Nin situ growth0.01–31 M NaPF6 in EC:PC=1:1408/350/0.1 A g−1183/3 A g−1[22]
γ-Fe2O3/PCF-0.01–31M NaSO3CF3 in tetraglyme290/50th/0.1 C230/2 C[23]
Fe2O3@C@MoS2@Chydrothermal0.1–31M NaClO4 in EC:DEC=1:1498/200/0.1 A g−1150/2 A g−1[24]
Fe2O3@NC-0.01–31mol/L NaClO4 in PC with 5% FEC155.3/200th/4 A g−1167.8/4 A g−1[25]
α-Fe2O3@C/rGOsolvothermal0.01–31mol/L NaClO4 in PC with 2% FEC--[26]
N-GF-300water bath0.005–31M NaSO3CF3 in diglyme428.3/100th/0.5 A g−1300/2 A g−1this work
Table 2. EIS Fitting results of N-GF-300, GF-300 and GF electrode before cycling.
Table 2. EIS Fitting results of N-GF-300, GF-300 and GF electrode before cycling.
ReRctCPE1-T/ΩCPE1-P/ΩW1-R/ΩW1-T/ΩW1-P/Ω
N-GF-30012.7118.240.00039980.63550.63550.31240.4290
GF-30012.2240.170.0010240.46209.4370.29490.4201
GF10.2552.310.00092050.428312.830.46460.4391
Table 3. The value of Re, C (= Q, when n =1), Q(Y0, n), Rct, Zw(Y0) and C1 calculated by CNLS using the fitting EIS data based on the relative equivalent circuit.
Table 3. The value of Re, C (= Q, when n =1), Q(Y0, n), Rct, Zw(Y0) and C1 calculated by CNLS using the fitting EIS data based on the relative equivalent circuit.
Impedance ParameterRe (Ω)Rct (Ω)Zw (Y0)C1 (F)nC (F, n = 1)
Graphene//N-GF-3004.3587.5410.039870.047180.75370.000118
N-GF-300//NGF-3004.14214.090.0270.01230.73480.0001265

Share and Cite

MDPI and ACS Style

Chen, Y.; Guo, Z.; Jian, B.; Zheng, C.; Zhang, H. N-Doped Modified Graphene/Fe2O3 Nanocomposites as High-Performance Anode Material for Sodium Ion Storage. Nanomaterials 2019, 9, 1770. https://doi.org/10.3390/nano9121770

AMA Style

Chen Y, Guo Z, Jian B, Zheng C, Zhang H. N-Doped Modified Graphene/Fe2O3 Nanocomposites as High-Performance Anode Material for Sodium Ion Storage. Nanomaterials. 2019; 9(12):1770. https://doi.org/10.3390/nano9121770

Chicago/Turabian Style

Chen, Yaowu, Zhu Guo, Bangquan Jian, Cheng Zheng, and Haiyan Zhang. 2019. "N-Doped Modified Graphene/Fe2O3 Nanocomposites as High-Performance Anode Material for Sodium Ion Storage" Nanomaterials 9, no. 12: 1770. https://doi.org/10.3390/nano9121770

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