Man Jianga,
Jing-Li Zhangb,
Fei Qiaob,
Rui-Ying Zhangb,
Ling-Bao Xing*b,
Jin Zhoub,
Hongyou Cuib and
Shuping Zhuo*b
aResources and Environmental Engineering, Shandong University of Technology, Zibo 255049, P. R. China
bSchool of Chemical Engineering, Shandong University of Technology, Zibo 255049, P. R. China. E-mail: lbxing@sdut.edu.cn; zhuosp_academic@yahoo.com
First published on 11th May 2016
Self-assembled three-dimensional (3D) reduced graphene hydrogels (RGHs) were fabricated by the facile chemical reduction of a graphene oxide (GO) dispersion with ammonia using acetaldehyde oxime as reducing and doping agent. The chemical reduction of GO was confirmed by Fourier transform infrared spectroscopy (FT-IR), X-ray photoelectron spectroscopy (XPS), X-ray diffraction (XRD), and Raman spectroscopy. The hierarchical porosity and structures of the resulting RGHs can be demonstrated by field emission scanning electron microscopy (FESEM) and N2 sorption experiments. Benefiting from the developed porosity with micro-meso hierarchical pore texture, the specific capacitance of RGHs exhibited high specific capacitances of 230.4, 155.3, 234.2, 155.1 and 191.8 F g−1 at 0.3 A g−1 for RGHs-1, RGHs-2, RGHs-5, RGHs-10 and RGHs-15 in 6 M KOH electrolyte, respectively. More importantly, the RGHs maintained high capacitances of 167.1, 110.4, 142.3, 106.9 and 142.3 F g−1 (the retention rates are 72.5, 71.1, 60.8, 68.9 and 74.2% for RGHs-1, RGHs-2, RGHs-5, RGHs-10 and RGHs-15) at a very high current density of 20 A g−1, indicating good electrochemical stability and a high degree of reversibility in the repetitive charge/discharge cycling test.
Among the many materials used in supercapacitors, graphene has attracted considerable interest since its multiple appealing features such as high specific surface area, excellent electrical conductivity, and extraordinary chemical/electrochemical stability and mechanical flexibility, which have shown potential applications in electrochemical energy storage devices such as supercapacitors,5–15 batteries,16–20 fuel cells.21–24 However, the obtained two-dimensional (2D) graphene tend to interact with each other to form irreversible aggregates or overlapping to graphitic structure due to noncovalent interactions such as π–π stacking and hydrophobic interactions between the intersheet of graphene, resulting in a dramatic decrease of the surface area.25–36 Consequently, the intrinsic properties of the obtained graphene sheets can not be exploited thoroughly, thus inefficient utilization of graphene layers for electrochemical energy storage. In order to fully utilize the high intrinsic surface area and further explore the new functions of graphene, self-assembly of nanoscale graphene into monolithic macroscopic materials with 3D porous networks can largely translate the properties of individual graphene into the resulting macrostructures. The formation of 3D graphene network can effectively prevent graphene from restacking and retain the high specific surface area that is necessary for high specific capacitance. Additionally, with a highly interconnected graphene network for excellent electron transport and interpenetrated porous network for rapid ion transport, the 3D graphene macrostructures are ideally suited for supercapacitor electrodes. The methods to prepare 3D structures especially reduced graphene hydrogels or aerogels generally include hydrothermal reduction, hydrothermal combined with chemical reduction, chemical reduction, N-doped hydrothermal reduction, solvothermal reduction, electrochemical reduction and so on,25–36 in which chemical reduction by using various reducing agents, such as ferrous ions,37 NaHSO3,38 mercaptoacetic acid,39 HI,40,41 ammonium sulfide,42 hydrazine,43,44 phenolic acids,45 polydopamine,46 hydroquinones,47 paraphenylene diamine,48,49 L-phenylalanine,50 sodium ascorbate,51–54 thiourea dioxide,55 thiocarbohydrazide56 and so on, has attracted extensive attention due to its lower temperature and shorter time in the process of preparation in contrast to other methods.
As low-toxicity, high efficiency and versatile reagents, oximes were used as oxygen scavenger in boiler water usually. Guo and coworkers have firstly prepared a highly stable graphene suspension by using dimethyl ketoxime (DMKO) as reductant, in which DMKO could be hydrolyzed in aqueous solution and slowly generates hydroxylammonium and acetone under alkaline condition at elevated temperature.57 Therefore, GO was reduced into graphene by in situ generated hydroxylammonium, which was an appropriate reductant for the GO reduction in aqueous suspension as reported recently.58,59 As a same reducing oxime, acetaldehyde oxime can also be hydrolyzed to hydroxylammonium and acetaldehyde, which has been also used to be oxygen scavenger in boiler water. Based on the above analysis, it was reasonably considered that the strong reducing capacity of acetaldehyde oxime in alkaline solution was attributed to the synergic effect of hydroxylammonium and acetaldehyde after hydrolysis. In present work, self-assembled 3D reduced graphene hydrogels (RGHs) were prepared in a facile chemical reduction process in GO suspension with acetaldehyde oxime as reducing agents. In the preparation process of RGHs (Scheme 1), acetaldehyde oxime and ammonia are added into the GO suspension under sonication to get a brown opaque colloidal solution. Then, the mixtures are placed in the oil bath at 90 °C without stirring for 4 h to get cylindrical RGHs. Benefiting from the abundant micro-meso hierarchical pore texture, the supercapacitors based on RGHs exhibited a high specific capacitance of 230.4, 155.3, 234.2, 155.1 and 191.8 F g−1 at 0.3 A g−1 for RGHs-1, RGHs-2, RGHs-5, RGHs-10 and RGHs-15 in 6 M KOH electrolyte, respectively. More importantly, the RGHs can maintain a high capacitance up to 167.1, 110.4, 142.3, 106.9 and 142.3 F g−1 at a very high current density of 20 A g−1, the retention rate are 72.5, 71.1, 60.8, 68.9 and 74.2% for RGHs-1, RGHs-2, RGHs-5, RGHs-10 and RGHs-15, respectively. It also showed that the electrode based on RGHs has good stability and high reversibility in the charge/discharge cycling test. The specific capacitance of the supercapacitor were maintained at 199.3, 133.1, 167.2, 127.4 and 163.9 F g−1 (capacitance retention ∼95.3, 91.5, 91.3, 88.9 and 91.2%) after 5000 cycles at 1 A g−1.
The specific capacitance is calculated by the following equation:
Cm = (It)/(ΔVm) |
Sample | SBET (m2 g−1) | VT (cm3 g−1) | Vmeso (cm3 g−1) | Vmicro (cm3 g−1) | D (nm) |
---|---|---|---|---|---|
RGHs-1 | 152 | 0.21 | 0.19 | 0.02 | 3.7 |
RGHs-2 | 82 | 0.14 | 0.10 | 0.04 | 5.3 |
RGHs-5 | 139 | 0.22 | 0.20 | 0.02 | 4.4 |
RGHs-10 | 86 | 0.22 | 0.21 | 0.01 | 7.1 |
RGHs-15 | 132 | 0.68 | 0.61 | 0.07 | 16.4 |
Field emission scanning electron microscopy (FESEM) was applied to observe the interior microstructures of the obtained RGHs. As observed from the SEM images shown in Fig. 2 and S1–S4,† a typical 3D network with pore structures formed by cross-linking of graphene sheets are quite uniform in the large scale, and the hierarchical pore with the wide size distribution ranges from micropores to mesopores, which is in agreement with N2 sorption experiment. As described in the introduction section, the 3D porous structures can be formed through the partial overlapping or aggregating of flexible graphene sheets via π–π stacking and hydrophobic interactions in the reduction process by using acetaldehyde oxime in an aqueous suspension of GO with ammonia.
The XRD patterns of GO and RGHs are shown in Fig. 3a. The feature diffraction peak of GO appeared at 9.43°, giving an interlayer spacing (d-spacing) of 9.37 Å. This interlayer space was much higher than that of pristine graphite (3.4 Å) as a result of the introduction of oxygenated functional groups on graphene sheets. For the resulting RGHs (Fig. 3a), the peak located at 9.43° disappeared, while small broad bumps near 23.82, 25.16, 25.35, 23.68 and 25.76° appeared which could be attributed to an interlayer spacing of 3.73, 3.54, 3.51, 3.75 and 3.46 for RGHs-1, RGHs-2, RGHs-5, RGHs-10 and RGHs-15, indicating the stacking of graphene sheets during the process of chemical reduction of GO. The broad XRD peaks of the resulting freeze-dried RGHs samples can be explained by the exfoliation of the layered GO and destruction of the regular stacks of graphite or graphite oxide. The structural changes after the reduction process are further reflected in the Raman spectra of GO and as-prepared RGHs. In Raman spectra as shown in Fig. 3b, the intensity ratios of the well-documented D bands (the A1g symmetry mode) in the vicinity of 1345 cm−1 and G bands (the E2g mode of the sp2 carbon atoms) in the vicinity of 1596 cm−1 of RGHs were enhanced after the chemical reduction process, compared with that of GO, indicating the improvement of the disordered graphene sheets. The intensity ratio of ID/IG is 0.89 in GO55,56,60–64 and increases to 1.08, 1.00, 1.06, 1.02 and 1.06 for RGHs-1, RGHs-2, RGHs-5, RGHs-10 and RGHs-15, demonstrating that the chemical reduction altered the structure of GO. FT-IR spectra is employed to further prove the removal of oxygenated groups by acetaldehyde oxime with ammonia (Fig. S5†). The broad and intense band at 3000–3500 cm−1 related to hydroxyl groups and peak at 1728 cm−1 related to carbonyl and carboxyl groups decrease, while a new peak at 1665 cm−1 attributed to the stretching and bending vibrations from CC appears, corresponding to the remaining sp2 character from the unoxidized graphitic domains. Furthermore, new peaks at 1548 and 1100 cm−1 corresponding to the C
N and C–N bending vibration appear, implying the doping of nitrogen in the reducing process. The element compositions of the prepared samples are determined by XPS measurement (Fig. S6†). Based on the XPS results, nitrogen doping of RGHs is further confirmed, which are mainly from the hydrolysed component of hydroxylammonium. The N content of RGHs is in the range of 2.46–5.13 atm%, while the O content of RGHs decreases from 29.06 to 8.28, 9.39 and 13.01 atm% for RGHs-1, RGHs-2 and RGHs-5, respectively. Because the XPS analysis only give the relative atomic ratio on the surface of RGHs, we further employ elemental analysis to analyse the C, H, O and N in RGHs. As shown in Table S1,† the C, H and O contents increase dramatically after reduction of acetaldehyde oxime, but have no obvious changes with the increase of the addition of acetaldehyde oxime. However, the N content increases obviously from 7.65 to 8.37 mass% with increasing of acetaldehyde oxime. This can be attributed to the increase of the reduction agents of hydrolyzed hydroxylammonium in the reduction process.
High resolution XPS measurements were further performed to analyze atom binding states of the prepared materials (Fig. 4), which showed the C 1s, O 1s and N 1s deconvolution spectra of RGHs-1 (a and b), RGHs-2 (c and d) and RGHs-5 (e and f). In case of C 1s of RGHs-1, four peaks at 284.4, 286.3, 287.4 and 288.8 eV corresponding to CC/C–C in sp2-hybridized domains, C–O (epoxy and hydroxyl), C
O (carbonyl) and O
C–O (carboxyl) groups were observed. Besides, the appearance of the new characteristic peak corresponding to C–N bond (285.2 eV) indicates the successful doping of nitrogen. The N 1s of RGHs-1 are determined by 398.1, 399.6 and 401.2 eV (Fig. S7†), which can be attributed to pyridinic N, pyrrolic N and graphitic N, respectively. In case of O 1s of RGHs-1, three peaks of C–O (carbonyl), C
O (epoxy and hydroxyl) and O
C–O (carboxyl) groups are determined by the peaks at 531.5, 532.7 and 533.6 eV, respectively. Compared to that of GO,55,56,60–64 the peaks of C–O, C
O and O
C–O of RGHs became much weaker, indicating the removal of oxygenated groups. Moreover, the C 1s, N 1s and O 1s spectra in RGHs-2 and RGHs-5 also gave the same phenomenon and results as RGHs-1. All the high-resolution XPS spectra demonstrate the effective removal of oxygen-containing groups and doping of nitrogen by acetaldehyde oxime with ammonia in the reduction process compared to GO.
![]() | ||
Fig. 4 High-resolution XPS spectra of C 1s and O 1s peaks for RGHs-1 (a and b), RGHs-2 (c and d) and RGHs-5 (e and f). |
The electrochemical performance of the obtained RGHs as electrode materials for supercapacitors were evaluated by cyclic voltammograms (CV) and galvanostatic charge/discharge test in a three-electrode system (Fig. 5). Fig. 5a shows the CV curves of RGHs-1, RGHs-2, RGHs-5, RGHs-10 and RGHs-15 at scan rate of 10 mV s−1. The CV curves at different scan rates from 5 to 150 mV s−1 (Fig. S8†) of RGHs all show rectangular-like shape with obvious current enlargement spread over a wide range of −0.4–0 V, indicating excellent capacitive behavior and the presence of pseudo-capacitance effect, which could be attributed to the complex and overlapped redox reactions introduced by N, O-doped species. These results can be further confirmed by galvanostatic charge/discharge measurements. As shown in Fig. 5b and S9,† the galvanostatic charge/discharge curves at different current density exhibit nearly triangular shapes with a small deviation from linearity, implying good capacitive behaviors originating from the combination of electric double layer capacitance and pseudocapacitance. It is considered that the galvanostatic charge/discharge measurement is a more accurate technique to determine the specific capacitance of electrode materials, especially for the one involving pseudocapacitance.
![]() | ||
Fig. 5 (a) Cyclic voltammograms of the supercapacitor based on RGHs at 10 mV s−1; and (b) galvanostatic charge/discharge curves of RGHs at 0.3 A g−1. |
The specific capacitance of the RGHs electrodes are evaluated to be 230.4, 155.3, 234.2, 155.1 and 191.8 F g−1 at a discharge current density of 0.3 A g−1 for RGHs-1, RGHs-2, RGHs-5, RGHs-10 and RGHs-15, respectively (Fig. 5b). As shown in Fig. S9† and 6a–e, the specific capacitances of RGHs are further investigated with the increasing of charge/discharge current density from 0.3 to 20 A g−1. It can be seen that the specific capacitances decrease at high charging/discharging current density. This is a common feature of real supercapacitor since there is no enough time for the electrolyte ions to diffusion into the entire pore surface, especially at higher charge current density. However, RGHs-1, RGHs-2, RGHs-5, RGHs-10 and RGHs-15 can still maintain the specific capacitances of 167.1, 110.4, 142.3, 106.9 and 142.3 F g−1 at a very high current density of 20 A g−1, the retention rate are 72.5, 71.1, 60.8, 68.9 and 74.2%, respectively (Fig. 6f). The excellent performance of RGHs can be attributed to the good combination of the hierarchical porous structure increasing the effective surface area and surface accessibility and the nitrogen-containing functional groups inducing the additional pseudocapacitance as well as improving the wettability of the material.
EIS data in a typical Nyquist plots for RGHs at frequencies ranging from 10 kHz to 1 mHz are given in Fig. 7a. For ideal porous electrodes, Nyquist plot is a vertical straight line perpendicular to the horizontal coordinate. However, for real porous electrodes, Nyquist plot can be separated into 3 parts: high frequency region, middle frequency region, low frequency region. At very high frequencies, the imaginary part (Z′) of the impedance is near to zero and the real part of resistance (Z′′) is derived from the electrolyte and the contact between the electrode and the current collector (Rs). The uncompleted semicircle loop at high frequency demonstrates the charge transfer resistance (Rct) at the interface between the electrolyte and electrode which is related to pseudocapacitive performance. The 45° slope region at middle frequency can be attributed to the ions diffusion/transport from the electrolyte to the pore on the surface of samples. In the low-frequency region, the Nyquist plot is a straight line, representing the dominance of ideal double-layer charge/discharge behaviors. The more vertical the line is, the more ideal the capacitor is. It is shown that all RGHs have almost vertical lines, indicating ideal capacitive behavior. This ideal capacitive behavior should also be ascribed to the micro-meso hierarchical pore texture of RGHs that favors fast ionic diffusion. Fig. 7b gives the cycle durability of RGHs electrodes investigated by galvanostatic charge/discharge measurement at a current density of 1 A g−1 for 5000 cycles. It is clear that the specific capacitance still remains at 199.3, 133.1, 167.2, 127.4 and 163.9 F g−1 after 5000 cycles at 1 A g−1, which are ∼95.3, 91.5, 91.3, 88.9 and 91.2% of the initial capacitance value after 5000 cycles, indicating the as-prepared RGHs display an outstanding cycling stability as electrode materials for supercapacitors.
![]() | ||
Fig. 7 (a) Electrochemical impedance spectra; and (b) long-term cycle test of RGHs measured at a current density of 1 A g−1 within the potential range from 0 to −0.9 V. |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra04348h |
This journal is © The Royal Society of Chemistry 2016 |