Xiujiao Gaoa,
Haipeng Lva,
Zhihong Lia,
Qunjie Xua,
Haimei Liu*a,
Yonggang Wang*b and
Yongyao Xiab
aShanghai Key Laboratory of Materials Protection and Advanced Materials in Electric Power, College of Environmental and Chemical Engineering, Shanghai University of Electric Power, Shanghai 200090, China. E-mail: liuhm@shiep.edu.cn
bDepartment of Chemistry, Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Institute of New Energy, Fudan University, Shanghai 200433, China. E-mail: ygwang@fudan.edu.cn
First published on 2nd November 2016
Developing low cost, highly efficient, stable and earth-abundant electrode materials for supercapacitors is critical for energy storage devices. In this work, a two-dimensional (2D) nickel-iron layered double hydroxide (Ni–Fe LDH) hybrid with 2D graphene was constructed into a three-dimensional (3D) aerogel by a facile one-step hydrothermal process assisted by freeze-drying treatment. Compared with the 2D structure, the 3D hybrid aerogel shows several advantages, including a unique porous framework, a multidimensional electron transport pathway and excellent electrical conductivity. When used as a supercapacitor electrode, benefiting from the above characteristics, the Ni–Fe LDH/graphene hybrid aerogel (Ni–Fe LDH/GHA) displays a high capacitance of 1196 F g−1 at 1 A g−1 and outstanding cycling stability with a capacitance retention of 80% after 2000 cycles. Furthermore, an asymmetric supercapacitor device with Ni–Fe LDH/GHA and active carbon as the positive and negative electrodes, respectively, achieved an energy density of 17.6 W h kg−1 at a power density of 650 W kg−1 and excellent long-term cycle stability (a specific capacitance of 91 F g−1 at 1 A g−1 after 3500 cycles with 87.2% retention). This work demonstrates that low cost, high performance Ni–Fe LDH/GHA has great potential for practical applications as a positive electrode in supercapacitors.
As a typical class of two-dimensional (2D) inorganic layered pseudocapacitor materials, layered double hydroxide (LDH) have received extensive attention in the electrochemistry field in recent years due to their easy-exchangeability, wide metal ion tunability and plentiful supply of redox active sites. Furthermore, their easily controlled facile chemical modification, low cost and lower environment pollution satisfies the requirements for sustainable development. In addition, due to the pre-intercalation of water molecules and the rapid expansion and contraction of their flexible structures, the 2D pseudocapacitor nanomaterials can store energy between nanosheets by a fast ion adsorption mechanism. Thus, the nature of 2D nanomaterials favours fast ionic transport through 2D channels that are free to expand and contract. Due to the above mentioned merits, over a long period of time, LDH-based pseudocapacitor materials have been extensively investigated, for instance, recent research has shown that Ni–Al,8 Ni–Co,9 Co–Al10 and Ni–Co–Al LDH11 as SC positive electrode materials have advantages because mixed-metal hydroxide compounds have superior electrochemical performances than single metal compositions.12,13
As a classic electroactive material, Ni–Al LDHs, formed by Al3+ partly replacing the Ni2+ of α-Ni(OH)2, have demonstrated both improved cycle lives and high rate capacitance. However, the electrochemical characteristics of Ni–Al LDH are still largely limited because of their poor electric conductivity and aggregation issues. Moreover, during electrochemical testing, it is always observed that Al3+ ions dissolve in alkaline electrolyte in the form of Al(OH)4− or AlO2−, which again drives the transformation of Ni–Al LDHs into the low capacitance of β-Ni(OH)2.14–17 To solve these issues, many other transition metal-based materials have been explored as replacements, because they are earth-abundant and stable relatively.18,19 Recently, the material of Ni–Fe LDH is a hot topic for applications in various fields, such as non-precious metal catalysts in oxygen20 and hydrogen evolution reactions,21 dye sorption22 and so on. As expected, in Ni–Fe LDH, it can be considered that the Fe3+ ions partly replace the Ni2+ of α-Ni(OH)2 to form a LDH structure and Fe3+ ions thus enhance the electrical conductivity of NiOOH.15,17 Very recently, a flower-like Ni–Fe LDH/carbon black composite was prepared and it exhibited good performance when serving as an alkaline battery cathode.14 However, as a pseudocapacitor-based electrode material, the Ni–Fe LDH is rarely studied.
Alternatively, building various carbon-based hybrid materials is an effective solution to improve the poor cycling lifespan of electrode materials, which results from their poor conductivity and structural stability. Among carbon-based hybrid materials, graphene-based functional materials have the priority owing to their large surface area, good conductivity and excellent structural stability. Generally, interconnected pores of three-dimensional graphene aim to prevent the restacking of graphene sheets. Compared with 2D graphene, the graphene aerogels offer more advantages, sparked by their unique architecture that integrates the advantages of multidimensional electron transport pathways.23 Therefore, how to avoid the self-stacking of graphene and combine graphene with redox materials to improve the electrochemical performance of EDLCs remains an important scientific challenge.
In this work, a facile hydrothermal method was developed for the fabrication of a hybrid graphene aerogel based on Ni–Fe LDH nanosheets, assisted by freeze-drying treatment. This novel aerogel SC electrode material, in which Ni–Fe LDHs are grown and anchored on a 3D graphene interconnected nanosheet skeleton to create a unique porous structure, is expected to provides the open channels for electrolyte transportation, and also allow most of the Ni–Fe LDH sheets to be exposed to the electrolyte. Moreover, the interconnected mesopore channels can provide a more favorable path to shorten electrolyte diffusion pathways and facilitate ion transport. Therefore, the electrochemical performance of the Ni–Fe LDH/GHA can be significantly improved.
For a two electrode asymmetric SC system, Ni–Fe LDH/GHA served as the positive electrode and active carbon (AC) served as the negative electrode. The cyclic voltammetry (CV), galvanostatic charge and discharge (GCD) measurements were carried out using a CHI660E electrochemical workstation and cycle life studies were measured by an Autolab electrochemical workstation in a 6 M KOH aqueous solution. All of the experiments were performed at room temperature with N2 protection.
The specific capacitance (Cs) values of the working electrodes were calculated in the three-electrode system from charge–discharge profiles using the following equation:
Cs = (I × Δt)/(ΔV × m) | (1) |
For the two electrode asymmetric system, the following equation was used to calculate Cs:
C = I × Δt/M × ΔV | (2) |
The energy and power density of the SC device were calculated using the following equations:25
E = 1/2 × C × ΔV2 | (3) |
P = E/Δt | (4) |
For comparison, bare graphene aerogels (GAs) and Ni–Fe LDH were also prepared. The morphology of the GAs, pure Ni–Fe LDH and Ni–Fe LDH/GHA composites were directly proved by the SEM images displayed in Fig. 2a–d. Among them, Fig. 2a is the image of GAs and the 3D network structure can be clearly observed. Such a 3D network structure provides hierarchical porous features with a wide scale distribution. However, the pure Ni–Fe LDH exhibits a compacted lamellar structure and a somewhat aggregated phenomenon in Fig. 2b. In order to prevent the LDH nanoflakes from self-aggregation and the loss of effective active area, it is necessary to construct a robust and rigid structure to support them. It is noteworthy that a panoramic SEM image of the as-fabricated 3D Ni–Fe LDH/GHA (Fig. 2c) shows an interconnected and porous 3D architecture. Moreover, the Ni–Fe LDH nanosheets occupy most of the available surface of the graphene, giving much higher loadings of LDH nanosheets in the hybrid aerogels. It is clearly seen that most of the Ni–Fe LDHs in the composite material have regular hexagon shapes and are several tens of nanometers in thickness in Fig. 2d. These Ni–Fe LDH nanosheets are interlacedly grown on the graphene skeleton surface, which prevent the agglomeration of LDH. That is important for effective electrolyte transport and active-site accessibility.28 The element mapping images of the Ni–Fe LDH/GHA sample illustrate the distribution of very elements in Fig. 2e. It is clear that Ni, Fe and C were overlapped with the same shape, demonstrating that Ni–Fe LDH was distributed evenly on the surface of graphene. From all above, it can be observed that the morphology of Ni–Fe LDH/GHA is a uniformly distributed 3D structure.
The above points were strongly confirmed by the following results. In order to further determine the porous structure and check the surface area of various samples, N2 adsorption desorption measurements were carried out (Fig. S2†) and it is verified that the GAs, pure Ni–Fe LDH and Ni–Fe LDH/GHA have specific surface areas of 113.6, 37.2 and 101.3 m2 g−1, respectively. Even with amounts of Ni–Fe LDH nanosheets grown on the surface of graphene, the specific surface area of the hybrid aerogels was still retained well. Compared with pure LDHs, the 3D structure of hybrid aerogels has higher BET surface area which would allow the materials and electrolyte to have good contact. The porosity of materials is highly important for superior performance as electrodes for SCs. Therefore, BJH analysis was carried out to measure the pore size and distribution of the Ni–Fe LDHs/GHA. Its average pore diameters were mainly distributed a narrow size between 3.6 nm and 10.3 nm, which are classified as mesopores. The above evaluation confirms that the 3D sample has a porous nanostructure. It is clear that the Ni–Fe LDH/GHA almost has as high as a specific surface area as the GAs and these mesopores will be favourable for improvement of electrochemical performance because of the benefit to fast ion transfer.
The crystal structures of the obtained GAs, pure Ni–Fe LDHs and Ni–Fe LDH/GHA were investigated by XRD, as displayed in Fig. 3. The powder XRD pattern of the composite (Fig. 3a and b) exhibited the typical signature of hexagonal lattices with diffraction peaks which are attributed to the (003), (006), (012), (015), (018), (110) and (013) planes with R3m rhombohedral symmetry (JCPDS card 22-0700),29 similar to the diffraction patterns of the α-Ni(OH)2 phase, as reported respectively.26 As shown in Fig. 3c, the GO was turned into GHAs after the hydrothermal reduction, the typical peak at 2θ = 25.2, its broad feature indicates the distortion from ordered arrangement of graphene sheets along their stacking direction.
XRD was used to identify the phases. Further information about the microstructure of the samples was obtained from TEM. Fig. 4a and b show TEM images of the GAs, it is observed that the graphene sheets stacked densely to form a wrinkled shell surface and graphene nanosheets interconnected to form a porous structure. The TEM images of Ni–Fe LDH with different magnifications reveal an uniform nano-size distribution of Ni–Fe LDH particles in Fig. 4c and d. It can be clearly observed that the LDH is in nearly hexagonal platelets with a lateral size. Fig. 4e and f further reveal that the 3D structure of the hybrid aerogels was constructed with graphene sheets and multiple overlapping Ni–Fe LDHs nanosheets with an interlayer separation of 0.26 nm, in agreement with its (012) plane.30 A typical HRTEM image of the hybrid proves the intimate interfacial contact among two components (Fig. 4f). Ni–Fe LDH platelets are randomly laid or vertical grown on the graphene nanosheets and the Ni–Fe LDH platelets deposited on the graphene sheets still have hexagonal shapes.
![]() | ||
Fig. 4 TEM and HRTEM images of various samples: (a and b) GAs; (c and d) Ni–Fe LDH; (e and f) Ni–Fe LDH/GHA. |
Raman spectroscopy is a non-destructive method for characterising graphitic materials, therefore, to further characterise and determine the ordered and disordered crystal structures of the as-prepared samples. Raman spectroscopy of the 3D Ni–Fe LDH/GHA, pure Ni–Fe LDHs, GA and GO were measured and shown in Fig. S3.† For Ni–Fe LDHs (Fig. S3b†), besides the peaks at 477 and 684 cm−1, which are ascribed to NiFe-LDHs,22,23 Raman shift samples at 1344 and 1583 cm−1 verify the formation of GAs. The peak at 1344 cm−1 is ascribed to disordered carbon which is induced by the graphitic structure and functional groups attached to the surface of the graphene sheets, while the peak at 1583 cm−1 is ascribed to graphite in-plane vibrations,23 which are induced by sp2-bond carbon atoms in a 2D hexagonal graphitic layer. The ID/IG ratios for GAs and hybrid aerogel are 1.3, respectively, demonstrating a decrease in the average size of the sp2 domains due to the partial removal of oxygen functional groups. Raman spectroscopy of the 3D Ni–Fe LDH/GHA (Fig. S3a†) reveals that the D-band and G-band located at around 1344 and 1583 cm−1, respectively, exhibit a ratio of integrated peak intensities.
In order to further identify the chemical state of the C, Ni and Fe elements, high-resolution XPS spectra of the Ni–Fe LDH/GHA and the GO composite were analysed (Fig. 5) and the full XPS spectra of various as-prepared samples were investigated in Fig. S4.† Four components of carbon atoms in different functional groups are indicated in Fig. 5b. The peaks at 284.6, 285.8, 287.4 and 288.7 eV are designated as CC, C–O, C
O and –COO, respectively.23 The C elements are compared in GO and the hybrid aerogel in Fig. 5a and b, it is noticeable that the hybrid aerogel exhibits the same oxygen functionalities in the C 1s spectrum as GO. However, compared with GO, the intensities of the peaks for hybrid aerogel centred at 284.6, 286.3, 288.4 and 289.3 eV, especially for the intensity of the peak centred at 286.3 eV, are much smaller, whereas for the peak at 284.6 eV, the intensity is higher than that observed for GO. In order to further identify the chemical state of the Ni and Fe elements, the high-resolution XPS spectra of the Ni–Fe LDHs and Ni–Fe LDH/GHA composites were analysed. The distance of Ni 2p1/2 and Ni 2p3/2 is 17.6 eV, moreover, the Ni 2p core level spectrum of the Ni–Fe LDHs (Fig. 5c) shows the Ni 2p1/2 (∼874.0 eV) and Ni 2p3/2 (∼856.3 eV) spin–orbit doublets accompanied by satellite peaks (∼880.2 and ∼861.2 eV), suggesting the divalent Ni2+ state.23,26,27 The Fe 2p core-level spectrum of the Ni–Fe LDHs (Fig. 5d) presented Fe 2p1/2 (∼726.0 eV) and Fe 2p3/2 (∼712.7 eV) spin–orbit peaks, indicating the existence of Fe3+ species.31,32
![]() | ||
Fig. 5 High-resolution XPS spectra of (a) C 1s (GO), (b) C 1s, (c) Ni 2p, (d) Fe 2p of Ni–Fe LDH/GHA. |
The attractive morphologies of the Ni–Fe LDH/GHA show promise for energy storage applications. Hence, the electrochemical behaviour of the Ni–Fe LDH and Ni–Fe LDH/GHA nanostructures were further characterised by CV (Fig. 6a and b) in a three electrode system using a 6 M KOH aqueous electrolyte. In a potential range of 0.0 to 0.50 V, CV curves show typical pseudocapacitor behaviour at different sweep rates. A pair of redox peaks in CV curves due to the faradaic reaction of redox reaction of Ni2+ ↔ Ni3+, which refers to LDH-Ni(OH)2 + OH ↔ LDH-NiOOH + H2O + e−.33 Notably, the Ni–Fe LDH/GHA electrode demonstrated a larger redox current density than the Ni–Fe LDH, implying more efficient utilisation of the electroactive components, owing to the 3D structure of the Ni–Fe LDHs with graphene nanoparticles. The potential of oxidation peaks presented a tendency to positively shift while the reduction peaks moved somewhat which suggested a more facile conversion of Ni3+ to Ni2+ than the reverse process for Ni–Fe LDH materials under fast charge/discharge conditions.
The galvanostatic charging/discharging profiles for the pure Ni–Fe LDH and Ni–Fe LDH/GHA electrodes at different current densities are shown in Fig. 6c and d. Compared with the corresponding pure Ni–Fe LDHs, the hybrid aerogel electrode exhibits excellent specific capacitances (1196 F g−1) at 1 A g−1 and excellent retention rate even at a rather high current density (72% capacitance retention at 10 A g−1). The specific capacitances of both samples gradually decrease with the increase of the current density. Apparently, the capacitances of the 3D hybrid material are much higher than pure Ni–Fe LDH at various constant current densities, especially the high rate performance in Fig. 6e. This performance is superior to many of the similar single active materials of LDHs, such as Ni–Al LDH,34 Co–Al LDH10 and Co–Fe LDH35 electrode materials (Table S1†36–39). However, it is noted that when compared with some two or more active components of LDHs, for example, Ni–Co LDH, Ni–Mn LDH and Ni–Co–Al LDH, the present Ni–Fe LDH/GHA do not represent the best performance, it is also the aim to optimise its performance in the future. To further evaluate the durability of our material, the cycling stability of the SC was tested through a cyclic charge/discharge process at a current density of 10 A g−1. As exhibited in Fig. 6f, it is observed that the specific capacity of the Ni–Fe LDH/GHA decreases slightly from 698 to 558 F g−1 after 2000 cycles, with an excellent capacity retention of 80% of initial value. As validated by SEM and TEM investigation (Fig. S5†), after a prolonged charge–discharge process, the regular hexagon shapes structure of Ni–Fe LDH is still maintained and without apparent damage. The 3D graphene structure also can be seen and to support the LDH to release the mechanical stress during charge–discharge process, and therefore it is beneficial in maintaining the stability of the electrode structure.
To further investigate the samples for practical application, two electrode SCs device Ni–Fe LDH/GHA//AC were fabricated, in which pressed Ni foam substrate was used as the current collector, employing AC as negative electrode and Ni–Fe LDH/GHA as the positive electrode (Fig. S6†). It illustrates the CV curves of the Ni–Fe LDH/GHA//AC asymmetric SCs at various sweep rates from 5 to 50 mV s−1 with a potential window of 1.3 V in Fig. 7a. The CV curves of the device exhibit an efficient charge storage behaviour, which still retain a primary shape even at a sweep rate of 50 mV s−1. To evaluate the electrochemical performance further, GCD tests were conducted at various current densities from 0.5 to 5 A g−1, as shown in Fig. 7b. It is noted that the asymmetric SCs deliver a specific capacitance of 91 F g−1 at 0.5 A g−1 and still maintain 40 F g−1 at 5 A g−1. For SCs, energy density and power density are two key parameters in evaluating the performances of SCs. Fig. 7c displays the relation between energy and power density of the device calculated by eqn (3) and (4). The energy density can reach as highly as 21.3 W h kg−1 at a power density of 356 W kg−1, and still maintains an energy density of 9.3 W h kg−1 at a power density of 3451 W kg−1. Compared with other LDH materials, the performance of the Ni–Fe/GHA material in alkaline supercapacitor is beneficial to the development of the LDH supercapacitors (more details in Table S2, ESI†).40–44 The long term cycling performance is an important criterion for SCs. Fig. 7d display the stability of the device. It is obtained that the capacitance retention of the Ni–Fe LDH/GHA//AC device is 87.2% after 3500 cycles. The GCD curve of the device at 1 A g−1 after 3500 cycles (Fig. S5d†) verifies the stability of the device. All the results of this work compare favourably with earlier related reports.34–39
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra19495h |
This journal is © The Royal Society of Chemistry 2016 |