Enhanced electrochemical performance of cobalt oxide nanocube intercalated reduced graphene oxide for supercapacitor application

Arshid Numan, Navaneethan Duraisamy*, Fatin Saiha Omar, Y. K. Mahipal, K. Ramesh and S. Ramesh*
Centre for Ionics University of Malaya, Department of Physics, Faculty of Science, University of Malaya, Lembah Pantai, 50603 Kuala Lumpur, Malaysia. E-mail: rameshtsubra@gmail.com; naveennanoenergy@gmail.com; Fax: +60-3-7967-4146; Tel: +60-3-7967-4391

Received 4th January 2016 , Accepted 23rd March 2016

First published on 24th March 2016


Abstract

We investigated different molar concentrations of cobalt precursor intercalated reduced graphene oxide (rGO) as possible electrode materials for supercapacitors. Cobalt oxide (Co3O4) nanocubes intercalated reduced graphene oxides (rGO) were synthesized via a facile hydrothermal method. It has been found that the Co3O4 particles with a cubical shape are decorated on rGO matrix with an average size of ∼45 nm. The structural crystallinity of rGO–Co3O4 composites was examined by X-ray diffraction (XRD). Raman spectroscopy confirmed the successful reduction of GO to rGO and effective interaction between Co3O4 and the rGO matrix. The electrochemical performances of rGO–Co3O4 electrodes were examined using cyclic voltammetry and charge–discharge techniques. The maximum specific capacitance (278 F g−1) is observed at current density of 200 mA g−1 in the C2 electrode resulting from effective ion transfer and less particle aggregation of Co3O4 on the rGO matrix than in the other electrodes. C2 exhibits good rate capability and excellent long-term cyclic stability of 91.6% for 2000 cycles. The enhanced electrochemical performance may result from uniform intercalation of cobalt oxide over the rGO. These results suggest that the Co3O4 intercalated rGO matrix could play a role in improved energy storage capability.


1. Introduction

The imminent depletion of fossil fuels and undesirable consequences of air pollution are alarming significant environmental concerns. These issues have motivated researchers to develop green, sustainable, and highly efficient alternative energy resources (solar energy, wind energy, etc.) and energy storages (batteries, supercapacitors).1 Among the various types of energy devices, the supercapacitor (SC, also known as an electrochemical capacitor) has emerged as an attractive energy storage device because of its extremely high power density, rapid dynamics of charge propagation, excellent cyclic retention, good energy density, and minimum charge separation compared with conventional capacitors and batteries.2,3

Generally, SCs can be classified into three different types according to the charge storage phenomenon: the electric double layer capacitor (EDLC), the pseudocapacitor, and the hybrid supercapacitor. In EDLC, the charges are stored in the electrode and electrolyte interfaces, whereas in the pseudocapacitor, charges are stored electrostatically via reversible adsorption of ions from the electrolyte into the electrode surface, resulting in higher specific capacitance than that of EDLC.4–6 Hybrid supercapacitors consist of asymmetric EDLC and pseudo (battery)-type electrode materials, hybridizing the advantages of EDLC and pseudocapacitors.

Usually, pseudocapacitors employ noble/transition metal oxides or conducting polymers as electrode materials.7 From the library of possible transition metal oxides, cobalt oxide (Co3O4) has emerged as a leader. It has high surface to volume ratio, simple preparation method, outstanding chemical durability, a promising ratio of surface atoms, and diverse morphology.8 However, the practical applications of such metal oxide-based pseudocapacitors may still be limited because of sluggish electron kinetics and rapid capacity decay arising from highly corrosive electrolyte and easily damaged morphologies of the materials during the faradic reactions.9

Many approaches have been explored for obtaining desirable electrode materials with different structures, to obtain high power and energy density.10–13 Incorporation of transition metal oxides with carbon materials can enhance electrochemical performance because of the enormous surface area and high electrical conductivity. Herein, graphene (as a two-dimensional (2D) monolayer of carbon atoms with hexagonal honeycomb lattice structure) has attracted tremendous attention because of its unique physicochemical properties, especially high surface area (2630 m2 g−1), ballistic conductivity (106 S cm−1), good electronic configuration,14 and a wide electrochemical window, suggesting it as a potential candidate for good electrochemical performance.15–17 The synergistic effect of rGO with cobalt oxide nanocomposite can enhance electrochemical redox activity, reversibility, electrical conductivity, and cycling charge discharge performance.18,19

Various techniques have been investigated for fabrication of cobalt oxide nanostructures, such as an unaided nano-casting method for preparation of Co3O4 nanowires,20 a surfactant-aided approach for fabrication of Co3O4 nanoboxes,21 low temperature synthesis of single-crystalline Co3O4 nanorods on silicon substrate,22 thermal oxidative decomposition growth of Co3O4 nanorods,23 nanotubes,24 nanowalls,25 gamma radiolysis-assisted formation of Co3O4 nanoparticles.26 Besides these, a hydrothermal method also has been employed to synthesize different structural morphologies.27 Of all these techniques, hydrothermal is the most preferred wet chemical route to fabricate nanostructured metal oxides because of its simplicity and the ability to tune various synthesis parameters (pH, temperature, time, and concentration of regents or precursors).28 Therefore, in this work, we report synthesis of a cobalt oxide nanocube intercalated rGO matrix by a facile one-pot hydrothermal route for energy storage applications.29

2. Experimental methods

2.1. Materials

Graphite flakes were purchased from Asbury Inc. (USA). Potassium permanganate (KMnO4, >99%), sulfuric acid (H2SO4 ∼98%), hydrochloric acid (HCl ∼35%), phosphoric acid (H3PO4 ∼85%), and ammonia solution (NH3, 25%) were purchased from R & M Chemicals. Cobalt acetate tetrahydrate (Co(CH3COO)2·4H2O), absolute ethanol (C2H6O ∼99.8%), and hydrogen peroxide (H2O2, 35%) were obtained from Sigma-Aldrich and Merck. All experiments were performed using doubly distilled water.

2.2. Synthesis of rGO–Co3O4 composites

In a typical synthesis of rGO–Co3O4 composite, 20 mL of freshly prepared30 exfoliated GO solution (1 mg mL−1) was added dropwise to 25 mL of ethanol under stirring. Ten milliliters of 0.25 mmol Co(CH3COO)2·4H2O was added drop by drop to the solution under vigorous stirring at 60 °C. Then, 15 mL of ammonia (6%) solution was added slowly to the mixture under constant stirring. Subsequently, the complete reaction mixture was transferred to a 100 mL Teflon lined stainless steel autoclave and subjected to hydrothermal treatment at 150 °C for 5 h. Then, the obtained precipitates of rGO–Co3O4 nanocomposites were washed with DI water and ethanol several times followed by drying at 90 °C in a hot air oven. The entire experiment was repeated to synthesize rGO–Co3O4 composites with other mole ratios (0.5, 0.75, and 1 mmol) of Co(CH3COO)2·4H2O. The synthesized hybrid Co3O4 nanocubes/rGO composites with different molar ratios of cobalt precursor 0.25, 0.5, 0.75, and 1 mmol were denoted as C1, C2, C3, and C4, respectively.

2.3. Electrochemical measurements

The working electrode was fabricated by mixing electroactive materials, polyvinylidene fluoride (PVDF), and carbon black with mass percentage ratio of 75[thin space (1/6-em)]:[thin space (1/6-em)]15[thin space (1/6-em)]:[thin space (1/6-em)]10 in 1-methyl-2-pyrrolidone (NMP) medium under ultrasonication. The viscous slurry was subsequently drop-casted and compressed on nickel foam (area of 1 cm2). The mass loading of prepared working electrodes (C1, C2, C3, and C4) without carbon black and PVDF was ∼5.1 mg. The working electrodes were immersed in 1 M KOH aqueous solution for 2 h before taking the electrochemical measurements, carried out in a three-electrode cell. The active material-coated Ni foam was used as the working electrode, with platinum wire as the counter electrode and Ag/AgCl as the reference electrode in a 1 M KOH aqueous solution at room temperature. Cyclic voltammetry (CV), electrochemical impedance spectroscopy (EIS), and galvanostatic charge discharge were performed with Gamry (Reference 3000 instrument). CV measurements of synthesized samples were performed in the potential range of 0 to 0.55 V at different scan rates of 1, 3, 5, 10, 20, 30, 40, and 50 mV s−1. EIS measurements were recorded in a frequency range from 0.01 Hz to 100 kHz under AC amplitude of 5 mV at open circuit voltage.

2.4. Characterization techniques

The structural crystallinity of rGO–Co3O4 nanocomposite was studied using Philips X'pert X-ray diffractometer with copper Kα radiation (λ = 1.5418 nm) at a scan rate of 0.02 degree per second. Raman spectra were obtained using the Renishaw inVia 2000 system green laser emitting at 514 nm. The surface morphologies of nanocomposites were studied using field emission scanning electron microscopy (JEOL JSM-7600F) and high-resolution transmission electron microscopy (JEOL JEM-2100F).

3. Results and discussion

3.1. Mechanism of Co3O4 intercalated rGO matrix

A schematic representation of formation of the rGO–Co3O4 nanocubes is shown in Fig. 1. The Co3O4 intercalated reduced graphene oxide was synthesized via a hydrothermal method. In the first step, GO solution (containing functional groups, such as hydroxyl (–OH), epoxy (C–O–C), carboxyl (–COOH), and carbonyl group) was sonicated for 1 hour to obtain exfoliated graphene oxide sheets. In the second step, addition of Co(CH3COO)2 in exfoliated GO solution led to adsorption of Co2+ ions on graphene oxide because of the electrostatic force of attraction between Co2+ ions and oxygen-based functional groups, resulting in the local creation of a bridge as Co–O–C bonding.31,32 In the third step, the reaction medium was changed to basic conditions using a NH3 solution followed by hydrothermal process. During the hydrothermal process, the GO–Co(OH)2 transformed into rGO–Co3O4 nanocubes. Restacking of rGO sheets was controlled by incorporation of Co3O4 nanocubes during reduction of GO to rGO.18
image file: c6ra00160b-f1.tif
Fig. 1 Schematic illustration of steps for the formation of rGO–Co3O4 hybrid: (a) exfoliated graphene oxide (GO) sheets, (b) intercalation of cobalt oxide, (c) ​cobalt nanocubes intercalated reduced graphene oxide (rGO).

3.2. X-ray diffraction analysis

The crystalline structures of synthesized GO, rGO, Co3O4, and rGO–Co3O4 were studied by recording the XRD patterns (Fig. 2). GO (Fig. 2a) showed a sharp intensity peak at 2θ value of 10° corresponding to the (001) plane.33,34 The sharpness of the peak indicated a larger interlayer distance because of the presence of functional groups on the graphene basal plane.18 Under the hydrothermal process, the diffraction peak of GO at 2θ value of 10° disappeared and two new broader peaks were generated at 2θ values of 26.8° and 42.7°, which correspond to the (002) and (100) planes of rGO (Fig. 2b), respectively.33 The two diffraction peaks observed revealed the disordered stacking of graphene sheets, indicating efficient transformation from GO to rGO.34,35 Fig. 2c and d shows the XRD patterns of pure Co3O4 and rGO–Co3O4 composite. The diffraction pattern of pure Co3O4 exhibited strong crystalline peaks at 2θ = 31.2°, 36.8°, 44.7°, 55.5°, 59.2°, and 65.1° corresponding to the (220), (311), (400), (422), (511), and (440) planes of a face centered cubic structure of nanostructured Co3O4 (JCPDS card no. 42-1467). The rGO–Co3O4 composite showed the diffraction peaks of Co3O4, which was intercalated with rGO matrix. No other characteristic peaks were observed, suggesting the purity of the Co3O4 embedded rGO matrix.36
image file: c6ra00160b-f2.tif
Fig. 2 XRD patterns for (a) GO, (b) rGO, (c) Co3O4, and (d) rGO–Co3O4.

3.3. Raman analysis

Raman spectroscopy is a suitable and non-destructive tool to monitor the functional groups and structural defects in materials, especially carbon-based graphene/graphene oxide/reduced graphene oxides. Fig. 3a–d shows the Raman spectra of GO, rGO, and the Co3O4–rGO composite. The spectra of GO and rGO (Fig. 3a and b) exhibited well-referred D and G bands at 1355 cm−1 and 1588 cm−1, respectively. The D band is ascribed to the lattice defect induced phonon mode. The G band corresponds to the in-plane vibrations of C–C atoms and a doubly degenerated phonon mode (E2g symmetry) at the Brillouin zone center.37–39 In general, there was a slight increase in ID/IG ratio from 0.95 to 1.36, suggesting formation of partially ordered crystal structures and decreased size of in-plane sp2 domains during reduction of GO to rGO.38 However, in the rGO–Co3O4 hybrid, the ID/IG ratio was decreased to 0.90, because of decreasing the sp2 domain size of carbon atoms and the reduction of sp3 to sp2 carbon.40 Additionally, defects were observed in the rGO matrix because of the removal of functional groups from GO layers.32 This confirmed successful reduction of GO to rGO. The 2D band or second order zone boundary phonon is extremely important for differentiation between monolayer and multilayer sheets in graphene-based material. Here, the 2D band was observed at 2901 cm−1 in GO, and retained at the same position in the rGO–Co3O4 composite (Fig. 3d).41,42 The inset of Fig. 3d shows intensity peaks at ∼194, 482, 525, 615, and 686 cm−1 corresponding to the B1g, Eg, F2g, F2g, and A1g modes of Co3O4, respectively.
image file: c6ra00160b-f3.tif
Fig. 3 Raman spectra for (a) GO, (b) rGO, (c) Co3O4, and (d) the rGO–Co3O4 composite.

3.4. Surface morphology

Surface morphologies of synthesized pure Co3O4 and rGO–Co3O4 composite were investigated via FE-SEM and HR-TEM analyses. Fig. 4a and b shows the clear distinction between pure cobalt oxide and composite of cobalt oxide with rGO matrix (C2). The FE-SEM image (Fig. 4a) revealed the cubical surface morphology of Co3O4 with particle aggregation, which led to reduction of the electrochemical surface area. This can be overcome by introducing rGO matrix as shown in Fig. 4b. The rGO matrix provided a platform to anchor cobalt oxide nanocubes, which were well distributed and intercalated with rGO sheets (C2 composite). However, C3 and C4 composite exhibited Co3O4 aggregation on rGO matrix because of a high concentration of cobalt precursor. Particle aggregation was increased with increasing concentration of cobalt precursor, as clearly shown in Fig. S1a and b.
image file: c6ra00160b-f4.tif
Fig. 4 Surface morphology: (a) FE-SEM image of pure Co3O4, (b) FE-SEM image of rGO–Co3O4 composite, (c) HR-TEM images of pure Co3O4, (d) HR-TEM image of rGO–Co3O4 composite.

Fig. 4c and d shows the HR-TEM image of pure Co3O4 and rGO–Co3O4 composite (C2). Fig. 4c shows the cubical shape of pure Co3O4 with an average particle size of ∼45 nm. The cubical shape of Co3O4 was well distributed on transparent rGO matrix (Fig. 4d). This result suggested that the rGO matrix provides a platform for improved electrochemical performance over that of pure Co3O4 because of effective intercalation and less particle aggregation of Co3O4 on rGO matrix.

3.5. Electrochemical performance

3.5.1. Cyclic voltammetry (CV). The electrochemical performance of pure Co3O4 and rGO–Co3O4 electrodes was evaluated using CV analysis. Cyclic voltammograms were performed for C1, C2, C3, and C4 with a potential window from 0 to +0.55 V at various scan rates. The measured currents were normalized with the electrode mass.

Fig. 5a–d shows the CV curves of C1, C2, C3, and C4 at different scan rates. Herein, the well-defined redox peaks show the pseudocapacitive behavior of composites, resulting from interaction of hydroxyl ions with the respective electrode.43 While increasing the scan rates, redox peak intensities increase with a slight peak shift towards higher potential, evidence of fast redox reactions occurring at the interface between the active material and electrolyte.44 Generally, the rate capability was mainly dependent on three processes: (i) the diffusion of electrolyte ions, (ii) the adsorption of ions on the electrode surface, and (iii) the charge transfer in the electrode. However, all of these processes were relatively slow at higher scan rate, leading to a reduction in specific capacitance.45 In addition, the characteristic shape of the CV curve was not significantly changed, which was an indication that the rGO–Co3O4 composites have outstanding rate capability.46


image file: c6ra00160b-f5.tif
Fig. 5 CV curves of (a) C1, (b) C2, (c) C3, and (d) C4 electrodes at different scan rates in 1 M KOH electrolyte.

The charge storage mechanism of Co3O4 for a pseudocapacitor electrode in alkaline medium can be represented as a simple OH entering reaction, described in the following equation.34

 
image file: c6ra00160b-t1.tif(1)

The CV curves of pure Co3O4 nanocubes at different scan rates are depicted in Fig. S2. Well-defined redox peaks were not observed, perhaps because of aggregation of Co3O4 nanocubes resulting in reduction of the active electrochemical surface area. In addition, with increasing scan rates, the cathodic and anodic peaks rapidly shift to the low and high potentials, respectively. This is mainly because of diffusion of OH ions having less time to interact with respective electrode at high scan rates.47,48 However, in the composites (rGO–Co3O4 nanocubes), the rGO matrix provides a highly conductive platform for the Co3O4 nanocubes because of the uniform dispersion and intercalation of the nanocubes on the matrix. This leads to an increase in electrochemical surface area by enhancing the active sites. A comparison of CV results of pure Co3O4 and composite (rGO–Co3O4) is shown in Fig. S3, which suggests that the composite has well-defined redox peaks at lower potential than that of pure Co3O4 because of non-aggregation of Co3O4 nanocubes on the rGO matrix. Therefore, we confirmed that the rGO matrix not only provided a highly conductive platform for the Co3O4 nanocubes, but also facilitated non-aggregated Co3O4 nanocubes.

3.5.2. Galvanostatic charge discharge study. The cyclic stability of Co3O4 decorated rGO matrix was examined by galvanostatic charge–discharge analysis. Fig. 6 shows the galvanostatic charge/discharge curve of rGO–Co3O4 composite electrodes at various current densities (from 200 to 500 mA g−1) under a potential range from 0 to 0.45 V (vs. Ag/AgCl).
image file: c6ra00160b-f6.tif
Fig. 6 Galvanostatic charge/discharge curves for (a) C1, (b) C2, (c) C3, and (d) C4.

The specific capacitance can be calculated by the following equation:

 
image file: c6ra00160b-t2.tif(2)
where “I” is the discharge current (A), “t” is the discharge time (s), “m” is the mass of active material (g), and V is the discharge potential (V). The obtained specific capacitances of 216, 278, 136, and 123 F g−1 correspond to C1, C2, C3, and C4, respectively, at the current density of 200 mA g−1.

Fig. 7a shows that specific capacitance varies with respect to molar concentration. Lower specific capacitance was observed in pure cobalt oxide, from aggregation of Co3O4 nanocubes, resulting in poor electrochemical performance. Introducing the rGO caused the specific capacitance to increase significantly from 118.8 F g−1 to 216 F g−1 (C1), as shown in Fig. 7a. The improved specific capacitance resulted from high dispersion and intercalation of Co3O4 on rGO matrix. Further increases in molar concentration of cobalt precursor resulted in the specific capacitances attaining a maximum value of 278 F g−1 (C2). However, beyond this addition of cobalt concentration, the specific capacitance began to decrease (C3 and C4) because of aggregation of cobalt oxide nanocubes on the rGO, resulting in limiting of OH ion diffusion into the electrode surface.


image file: c6ra00160b-f7.tif
Fig. 7 (a) Variation in specific capacitance at different current densities for C1, C2, C3, and C4 samples. (b) Comparison of SC with respect to molar concentration of Co3O4 precursor for C1, C2, C3, and C4 samples.

Fig. 7b shows the specific capacitance versus current density plot. The values of specific capacitance decrease with increasing current density because of asynchronous movement of electric charges with respect to current rates. At higher current densities, the diffusion/migration of charges through the electrodes was very slow.49

3.5.3. Electrochemical impedance spectroscopy. To further elucidate the origin of high electrochemical performance, electrochemical impedance spectroscopy (EIS) was carried out to examine rGO–Co3O4 composites. The equivalent series resistance (ESR) was obtained from the intersection point of the curves with the axis of real impedance. The difference in ESR of electrodes is attributed to the different conductance of electrode materials. ESRs of the rGO–Co3O4 electrodes (Fig. 8) were found to be 1.25, 1.3, 1.37, and 1.57 Ω corresponding to C1, C2, C3, and C4, respectively. The ESR of rGO–Co3O4 composites increased with increasing cobalt precursor concentration, indicating poor conductivity of material at high molar concentration. This may be because of aggregation of cobalt oxide and weak interaction of Co3O4 with the rGO matrix. At the high frequency region, the observed larger semicircle indicated poor electrical conductivity because of high interfacial charge transfer resistance.50
image file: c6ra00160b-f8.tif
Fig. 8 Nyquist plots for C1, C2, C3, and C4. Inset: magnified view of Nyquist plot at higher frequencies.

The long-term cycling stability test is a key factor to evaluate electrodes for practical applications.51 Long-term cycling performance over 2000 cycles of electrode material was carried out by repeating the charge/discharge test at a current density of 3 A g−1. The cyclic stability of composites C1, C2, C3, and C4 corresponds to 86%, 91.6%, 94%, and 97.5% (Fig. 9), respectively. These results confirmed that the cyclic stability increased with increasing molar concentration of cobalt precursor. However, the specific capacitance was decreased beyond some limitation of molar concentration because of particle aggregation. The maximum specific capacitance (278 F g−1) with longer cyclic stability (91.6% after 2000 cycles) was observed at C2 as compared with other electrodes. Therefore, these results indicate that C2 is a promising candidate for high performance energy storage systems.


image file: c6ra00160b-f9.tif
Fig. 9 Variation of specific capacitance as a function of cycle number from 5th to 2000th.

4. Conclusion

In this work, the effect of different molar concentrations of cobalt precursor-based cobalt oxide–reduced graphene oxide was studied on electrochemical performance. Different molar concentrations of Co3O4 intercalated rGO were synthesized using a hydrothermal method. Surface morphology of the composite materials revealed uniform distribution of Co3O4 on the rGO matrix with an average particle size of ∼45 nm. The maximum specific capacitance (278 F g−1) was observed in C2, with lower charge transfer resistance than that of other electrodes because of effective ion transfer and fewer particle aggregations on rGO matrix. Moreover, C2 exhibited long-term cyclic stability of 91.6% over 2000 cycles with good reversibility. These results demonstrate that controlling the molar concentration of cobalt precursor-based Co3O4 with rGO composite has potential in development of electrode materials for energy storage applications.

Acknowledgements

This work was supported by the High Impact Research Grant (H-21001-F000046) from Ministry of Education, Malaysia and University of Malaya Research Grant (RP025A-14AFR).

References

  1. C. Yu, J. Yang, C. Zhao, X. Fan, G. Wang and J. Qiu, Nanoscale, 2014, 6, 3097 RSC.
  2. J. K. Change, C. H. Huang, W. T. Tsai, M. J. Deng and I. W. Sun, J. Power Sources, 2008, 179, 435–440 CrossRef.
  3. L. L. Zhang, R. Zhaou and X. S. Zhao, J. Mater. Chem., 2010, 20, 5983–5992 RSC.
  4. J. J. Yoo, K. Balakrishnan, J. Huang, V. Meunier, B. G. Sumpter, A. Srivastava, M. Conway, A. L. M. Reddy, J. Yu, R. Vajtai and P. M. Ajayan, Nano Lett., 2011, 11, 1423–1427 CrossRef CAS PubMed.
  5. Y. Zhu, S. Murali, M. D. Stoller, K. J. Ganesh, W. Cai, P. J. Ferreira, A. Pirkle, R. M. Wallace, K. A. Cychosz, M. Thommes, D. Su, E. A. Stach and R. S. Ruoff, Science, 2011, 332, 1537–1541 CrossRef CAS PubMed.
  6. X. Fan, C. Yu, J. Yang, Z. Ling and J. Qiu, Carbon, 2014, 7, 130–141 CrossRef.
  7. K. R. Prasad, K. Koga and N. Miura, Chem. Mater., 2004, 16, 1845–1847 CrossRef CAS.
  8. M. Jafarian, M. G. Mahjani, H. Heli and F. Gobal, Electrochim. Acta, 2003, 48, 3423–3429 CrossRef CAS.
  9. Y. Huang, J. Liang and Y. Chen, Small, 2012, 8, 1805–1834 CrossRef CAS PubMed.
  10. M. N. Hyder, S. W. Lee, F. C. Cebeci, D. J. Schmidt, Y. Shao-Horn and P. T. Hammond, ACS Nano, 2011, 5, 8552–8561 CrossRef CAS PubMed.
  11. A. Izadi-Najafabadi, Adv. Mater., 2010, 22, E235–E241 CrossRef CAS PubMed.
  12. P. C. Chen, G. Shen, Y. Shi, H. Chen and C. Zhou, ACS Nano, 2010, 4, 4403–4411 CrossRef CAS PubMed.
  13. X. Tian, M. Shi, X. Xu, M. Yan, L. Xu, A. M. Khan, C. Han, L. He and L. Mai, Adv. Mater., 2015, 27, 7476–7482 CrossRef CAS PubMed.
  14. H. Wang, L. Zhi, K. Liu, L. Dang, Z. Liu, Z. Lei, C. Yu and J. Qiu, Adv. Funct. Mater., 2015, 25, 5420–5427 CrossRef CAS.
  15. Y. Zhu, S. Murali, W. Cai, X. Li, J. W. Suk, J. R. Potts and R. S. Ruoff, Adv. Mater., 2010, 22, 3906–3924 CrossRef CAS PubMed.
  16. H. Chang and H. Wu, Energy Environ. Sci., 2013, 6, 3483 CAS.
  17. J. Yang, C. Yu, X. Fan, S. Liang, S. Li, H. Huang, Z. Ling, C. Hao and J. Qiu, Energy Environ. Sci., 2016 10.1039/c5ee03633j.
  18. Z. S. Wu, Y. Sun, Y. Z. Tan, S. Yang, X. Feng and K. Mullen, J. Am. Chem. Soc., 2012, 134, 19532–19535 CrossRef CAS PubMed.
  19. C. Xiang, M. Li, M. Zhi, A. Manivannan and N. Wu, J. Power Sources, 2013, 226, 65–70 CrossRef CAS.
  20. Z. Wang, X. Chen, M. Zhang and Y. Qian, Solid State Sci., 2005, 7, 13–15 CrossRef CAS.
  21. T. He, D. Chen, X. Jiao and Y. Wang, Adv. Mater., 2006, 8, 1078–1082 CrossRef.
  22. L. He, Z. Li and Z. Zhang, Nanotechnology, 2008, 15, 155606 CrossRef PubMed.
  23. E. L. Salabas, A. Rumplecker, F. Kleitz, F. Radu and F. Schüth, Nano Lett., 2006, 12, 2977–2981 CrossRef PubMed.
  24. T. Li, S. Yang, L. Huang, B. Gu and Y. Du, Nanotechnology, 2004, 15, 1479–1482 CrossRef CAS.
  25. T. Yu, Y. W. Zhu, X. J. Xu, Z. X. Shen, P. Chen, C. T. Lim and C. H. Sow, Adv. Mater., 2005, 13, 1595–1598 CrossRef.
  26. L. M. Alrehaily, J. M. Joseph, M. C. Biesinger, D. A. Guzonasc and J. C. Wren, Phys. Chem. Chem. Phys., 2013, 15, 1014–1024 RSC.
  27. X. Liu, Q. Long, C. Jiang, B. Zhan, C. Li, S. Liu and X. Dong, Nanoscale, 2013, 14, 6525–6529 RSC.
  28. M. M. Rahmana, J. Z. Wanga, X. L. Deng, Y. Li and H. K. Liu, Electrochim. Acta, 2009, 55, 504–510 CrossRef.
  29. M. M. Lencka, M. Oledzka and R. E. Riman, Chem. Mater., 2000, 12, 1323–1330 CrossRef CAS.
  30. N. M. Huang, H. N. Lim, C. H. Chia, M. A. Yarmo and M. R. Muhamad, Int. J. Nanomed., 2011, 6, 3443–3448 CrossRef CAS PubMed.
  31. R. Kumar, H. J. Kim, S. Park, A. Srivastava and I. Oh, Carbon, 2014, 79, 92–202 CrossRef.
  32. K. H. An, W. S. Kim, Y. S. Park, J. Mi Moon, D. J. Bae, S. C. Lim, Y. S. Lee and Y. H. Lee, Adv. Funct. Mater., 2001, 11, 387–392 CrossRef CAS.
  33. C. M. Chen, Q. Zhang, M.-G. Yang, C.-H. Huang, Y.-G. Yang and M.-Z. Wang, Carbon, 2012, 50, 3572–3584 CrossRef CAS.
  34. M. M. Shahid, A. Pandikumar, A. M. Golsheikh, N. M. Huang and H. N. Lim, RSC Adv., 2014, 4, 62793–62801 RSC.
  35. R. Kanemoto, A. Anas, Y. Matsumoto, R. Ueji, T. Itoh, Y. Baba, S. Nakanishi, M. Ishikawa and V. Biju, J. Phys. Chem. C, 2008, 112, 8184–8191 CAS.
  36. Z.-S. Wu, W. Ren, L. Wen, L. Gao, J. Zhao, Z. Chen, G. Zhou, F. Li and H.-M. Cheng, ACS Nano, 2010, 4, 3187–3194 CrossRef CAS PubMed.
  37. A. C. Ferrari, Solid State Commun., 2007, 143, 47–57 CrossRef CAS.
  38. D. Naumenko, V. Snitka, B. Snopok, S. Arpiainen and H. Lipsanen, Nanotechnology, 2012, 23, 465703 CrossRef PubMed.
  39. S. S. Li, J. N. Zheng, X. Ma, Y. Y. Hu, A. J. Wang, J. R. Chen and J. J. Feng, Nanoscale, 2014, 6, 5708–5713 RSC.
  40. Y. Kim, Y. Noh, E. J. Lim, S. Lee, S. M. Choi and W. B. Kim, J. Mater. Chem. A, 2014, 2, 6976 CAS.
  41. G. T. S. How, A. Pandikumar, H. N. Ming and L. H. Ngee, Sci. Rep., 2014, 4, 5044 CAS.
  42. S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas, A. Kleinhammes, Y. Jia, Y. Wu, S. T. Nguyen and R. S. Ruoff, Carbon, 2007, 45, 1558–1565 CrossRef CAS.
  43. Q. Zheng, B. Zhang, X. Lin, X. Shen, N. Yousefi, Z. D. Huang, Z. Li and J. K. Kim, J. Mater. Chem., 2012, 22, 25072 RSC.
  44. N. Duraisamy, A. Numan, K. Ramesh, K. H. Choi, S. Ramesh and S. Ramesh, Mater. Lett., 2015, 161, 694–697 CrossRef CAS.
  45. J. Li, W. Zhao, F. Huang, A. Manivannan and N. Wu, Nanoscale, 2011, 3, 5103 RSC.
  46. H. W. Wang, Z. Ai Hu, Y. Q. Chang, Y. L. Chen, Z. Y. Zhang, Y. Y. Yang and H. Y. Wu, Mater. Chem. Phys., 2011, 130, 672–679 CrossRef CAS.
  47. C. Xu, B. Li, H. Du, F. Kang and Y. Zeng, J. Power Sources, 2008, 180, 664–670 CrossRef CAS.
  48. V. Subramanian, H. Zhu, R. Vajtai, P. M. Ajayan and B. Wei, J. Phys. Chem. B, 2005, 109, 20207–20214 CrossRef CAS PubMed.
  49. Q. Fu, F. Tietz and D. Stöver, J. Electrochem. Soc., 2006, 153, D74–D83 CrossRef CAS.
  50. A. Pendashteh, M. Mousavi and M. S. Rahmanifar, Electrochim. Acta, 2013, 88, 347–357 CrossRef CAS.
  51. S. Yoon, E. Kang, J. K. Kim, C. W. Lee and J. Lee, Chem. Commun., 2011, 47, 1021–1030 RSC.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra00160b

This journal is © The Royal Society of Chemistry 2016