You are on page 1of 6

View Online / Journal Homepage / Table of Contents for this issue

PAPER

www.rsc.org/materials | Journal of Materials Chemistry

Facile and controllable electrochemical reduction of graphene oxide and its applications
Yuyan Shao, Jun Wang, Mark Engelhard, Chongmin Wang and Yuehe Lin*
Received 2nd September 2009, Accepted 19th October 2009 First published as an Advance Article on the web 20th November 2009 DOI: 10.1039/b917975e
Downloaded by University Institute of Chemical Technology (UICT) on 23 November 2011 Published on 20 November 2009 on http://pubs.rsc.org | doi:10.1039/B917975E

Graphene oxide is electrochemically reduced which is called electrochemically reduced graphene oxide (ER-G). ER-G is characterized with scanning electron microscopy, transmission electron microscopy, X-ray photoelectron spectroscopy, and X-ray diffraction. The oxygen content is signicantly decreased and the sp2 carbon is restored after electrochemical reduction. ER-G exhibits much higher electrochemical capacitance and cycling durability than carbon nanotubes (CNTs) and chemically reduced graphene; the specic capacitance measured with cyclic voltammetry (20 mV s1) is 165, 86, and 100 F g1 for ER-G, CNTs, and chemically reduced graphene, respectively. The electrochemical reduction of oxygen and hydrogen peroxide are greatly enhanced on ER-G electrodes as compared with CNTs. ER-G has shown promising features for applications in energy storage, biosensors, and electrocatalysis.

1. Introduction
Graphene, a one-atom-thick sp2-bonded carbon sheet, has attracted strong scientic and technological interest1,2 since its discovery in 2004.3 It has shown great application potential in various elds, such as electronic devices,4 nanocomposite materials,5 sustainable energy storage and conversion (ultracapacitors,68 batteries,9 fuel cells,1014 solar cells15,16), and bioscience/biotechnologies1721 because of its unique physical and chemical properties (high surface area,1 excellent conductivity,1 and mechanical strength22). Producing graphene with desirable properties is the rst step to scientic research and applications, and this is still a signicant challenge.23 So far, several methods have been developed to produce graphene:2,23,24 (1) Micromechanical cleavage (scotch tape method).3 This method is widely used in many laboratories to obtain pristine perfect-structured graphene layer(s) for basic scientic research and for making proof-of-concept devices. However, it is not suitable for mass production. (2) Epitaxial graphene grown on SiC,25,26 ruthenium,27 Ni,28 and Cu.29 This is a potential mass-production method with the aim of producing graphene for electronics applications.4,24 (3) Chemical3033 or thermal reduction34 of graphite oxide (GO). (4) Bottom-up organic synthesis.35 The synthesis route from graphite oxides is considered to be the most economical way to the mass produce of graphene.2,23 However, it often employs hazardous chemicals (e.g., hydrazine) as reductants or rapid heat-treatment at high temperature. The green synthesis of graphene under mild conditions is preferred.33 Electrochemical methods are one promising green strategy for graphene synthesis, and several

research works have been reported.20,3640 In fact, most of the graphenes that are chemically or electrochemially synthesized contain other elements, such as oxygen, so it is more reasonable to call the graphene reduced graphene oxide. In this paper, we report the electrochemically controllable reduction of graphene oxide, and we found that the properties of ER-G are quite different from other nanostructured carbons (i.e., carbon nanotubes [CNTs]) and even different from graphene that is chemically reduced from GO. It exhibits greatly enhanced activity for the electrocatalytic reduction of O2 and H2O2, and much higher electrochemical capacitance for potential application in ultracapacitors. An ultracapacitor is a power device for electric energy storage, which is very important for developing renewable energy and electric vehicles.41,42 Oxygen reduction reaction (ORR) is important for fuel cells, metalair batteries, etc.43,44 Hydrogen peroxide is a general enzymatic product of oxidases and a substrate of peroxidases, which is important in biological processes and biosensor development.45 Therefore, this work seems promising for applications in energy storage, biosensors, and electrocatalysis.

2. Experimental
2.1 Graphite oxide preparation Graphite oxide (GO) was prepared with the modied Hummers method.31,46 Typically, 5.0 g graphite powder (<45 mm, SigmaAldrich) was added into 180 mL concentrated H2SO4 and stirred for 1 h in a hood. Then 60 mL fuming HNO3 was slowly added to the mixture under ice-cooling and stirring. After cooling down, 25 g KMnO4 was slowly added under ice-cooling and stirring. The mixed slurry was stirred at room temperature in a hood for 120 h. After that, 600 mL deionized (DI) water was slowly added into the reacted slurry and stirred for 2 h; then 30 mL 30% H2O2 was added, and the slurry immediately turned into a bright yellow solution with bubbling. The resultant solution was stirred for 2 h and then allowed to settle down for 24 h; after that, the
J. Mater. Chem., 2010, 20, 743748 | 743

Pacic Northwest National Laboratory, Richland, WA, 99352. E-mail: yuehe.lin@pnl.gov Electronic supplementary information (ESI) available: Images of electrochemically reduced graphene; cyclic voltammograms of capacitor charge/discharge curves, ORR and H2O2 reduction curves; XPS spectra. See DOI: 10.1039/b917975e

This journal is The Royal Society of Chemistry 2010

View Online

Downloaded by University Institute of Chemical Technology (UICT) on 23 November 2011 Published on 20 November 2009 on http://pubs.rsc.org | doi:10.1039/B917975E

supernatant was decanted. The resultant yellow slurry was centrifuged and then washed in 1000 mL DI water with 5 mL HCl (37%) and 3 mL 30% H2O2 added. After stirring for 2 h, the solution was centrifuged and then washed again. This process was repeated three times. After that, the yellow slurry was further washed with 500 mL DI water until the pH of the washing solution increased to neutral (6.5) (it took about 500 mL 12 washes). The remaining dark-yellow solid was dried under vacuum at 40  C for 48 h and ground to a ne powder. The dry process for GO must be carried out at low temperatures because it slowly decomposes (deoxygenates) above 60 to 80  C.47 2.2 Electrochemistry GO was rst exfoliated in Nanopure DI water. A total of 20 mg GO was mixed with 20 mL DI water and ultrasonicated for 1 h. The resultant homogeneous yellow-brown GO solution (1.0 mg mL1) can be stable for months because of the negatively charged surface of GO. The GO solution (20 mL) was applied onto prepolished glassy carbon/gold disk electrodes and allowed to dry in air, and then 10 mL 0.05 wt% Naon was cast and allowed to dry in air. The electrochemical reduction of GO was carried out with extended cyclic voltammetry (CV, 1.0  1.0 V vs. reversible hydrogen electrode [RHE], 50 mV s1) in 0.1 M Na2SO4 solution in a standard three-electrode cell with the Hg/Hg2SO4 and Pt foil as the reference and counter electrode, respectively (all electrode potentials are scaled to RHE except as specially stated). The capacitance behavior was characterized with CV and galvanostatic charge/discharge in 0.1 M Na2SO4. The cycling capacitance durability was studied with CV (0 to 1.1 V, 200 mV s1) in 0.1 M Na2SO4. The oxygen reduction was measured with a rotating disk electrode (Pine Instruments) in O2 saturated 0.5 M H2SO4 (1600 RPM, 10 mV s1). The electrochemical reduction of H2O2 was carried out in N2-saturated 5 mM H2O2 + 50 mM phosphate buffered saline (PBS) + 100 mM KCl (pH 7.4) (50 mV s1). Carbon nanotubes (Nanolab, Purity>95%) were tested in the same condition for comparison. Before use, CNTs were functionalized in an ultrasonication bath of concentrated H2SO4/ HNO3 for 4 h at room temperature and washed/repuried with DI water.48 2.3 Material characterization The X-ray photoelectron spectra (XPS) measurements were obtained on the Physical Electronics Quantum 2000 Scanning ESCA Microprobe using a focused monochromatic aluminium Ka X-ray (1486.6 eV) source and a spherical section analyzer. The X-ray beam used was a 100 W, 100-mm diameter beam rastered over a 1.3 0.2 mm area on the sample. Wide-scan data were collected using a pass energy of 117.4 eV, which produces a full width at half maximum (FWHM) of 1.6 eV for the Ag3d5/2 line. High-energy photoemission spectra were collected using a pass energy of 23.95 eV, which produces an FWHM of 0.80 eV for the Ag3d5/2 line. The binding-energy scale was calibrated using Cu2p3/2 at 932.62 0.05 eV and Au 4f at 83.96 0.05 eV. All samples were dried in vacuum before the XPS test. X-Ray diffraction (XRD) patterns were obtained using a Philips Xpert X-ray diffractometer with Cu Ka radiation at l 1.54
744 | J. Mater. Chem., 2010, 20, 743748

 Scanning electron microscopy (SEM) images were taken at A. FEI Company Helios Nanolab. Transmission electron microscopy (TEM) images were collected with holey carbon TEM grids on a JEOL JSM-2010 TEM microscope operated at 200 kV. ER-G Samples for XPS and XRD tests were prepared by applying ER-G ink onto clean substrates and were let to dry in vacuum at room temperature. ER-G ink was prepared by dispersing ER-G in 100% ethanol with ultrasonication.

Results

Fig. 1 shows the TEM and SEM images of ultrasonically exfoliated graphene oxide, which exhibits the typical wrinkle morphology of GO.49,50 From the TEM image, it can be seen that GO was exfoliated into single or very thin layers. GO was electrochemically reduced with extended potential cycling (1.0 V  1.0 V RHE, 50 mV s1) in 0.1 M Na2SO4. Fig. 2a shows the CVs recorded during the reduction process. The reduction current peak around 0.75 V is due to the electrochemical reduction of GO.51 It can be seen that the reduction peak rst increases and then decreases (to disappear) as the potential cycling proceeds. The redox current peaks around 0.50 V (Fig. 2a) increase with the potential cycling and then remain stable after a certain number of cycles (the specic capacitance calculated with CV6,52 keeps the same trend

Fig. 1 TEM (a) and SEM (b) images of graphene oxide.

This journal is The Royal Society of Chemistry 2010

View Online

Downloaded by University Institute of Chemical Technology (UICT) on 23 November 2011 Published on 20 November 2009 on http://pubs.rsc.org | doi:10.1039/B917975E

Fig. 3 Cyclic voltammograms (a, b) and galvanostatic charge/discharge curves (5 A g1) (c, d) of the electrochemically reduced graphene oxide (ER-G) and carbon nanotube (CNT) electrodes in 0.1 M Na2SO4.

Fig. 2 (a) Cyclic voltammograms of the electrochemical reduction of graphene oxide in 0.1 M Na2SO4 at 50 mV s1; (b) the specic capacitance changes with the reduction cycles, measured with cyclic voltammograms (0 to 0.9 V) at 20 mV s1.

[Fig. 2b]). The reduction degree of GO can be in situ monitored and controlled through the reduction peak (0.75 V) and the calculated specic capacitance. When the reduction peak (0.75 V) disappears, the electrochemical reduction of GO is considered to be completed (there are still some oxygen-containing groups on ER-G, but they are too stable under electrochemical conditions, to be discussed in the following sections). The reduction of GO can also be seen from the color of GO electrodes, which changes from yellow (before reduction) to black (after reduction) (Fig. S1). The electrochemical capacitance behavior of ER-G was characterized with CV and galvanostatic charge/discharge, and compared with CNTs (Fig. 3). The calculated specic capacitances are shown in Table 1. The measured specic capacitances depend on both the test methods and the materials. As expected, the faster charge/discharge produces lower measured capacitances.6,53 It can be seen that the specic capacitance of ER-G is always much higher than that of CNTs under the same test conditions; for example, the specic capacitance of ER-G measured with CV [20 mV s1] is 164.8 F g1, larger than the 85.8
Table 1 Specic capacitance (F g1) of ER-G and CNTs Galvanostatic charge/discharge/A g1 Test method ER-G CNT 5 150.4 83 20 118.4 69 Cyclic voltammogram/mV s1 20 164.8 85.8 100 136.8 81.6

Fig. 4 The specic capacitance of ER-G and CNTs changing with the charge/discharge cycle number.

F g1 for CNTs. This value is also much higher than the previously reported specic capacitance of chemically reduced graphene (100 F g1 tested at 20 mV s1 in 5.5 M KOH 6). The durability of materials is extremely important for applications in ultracapacitors.53 The cycling durability of ER-G and CNT electrodes is shown in Fig. 4. It can be seen that the specic capacitance of CNTs degrades with cycling, but increases slightly for ER-G in comparison with 10% degradation in only 1200 charge/discharge cycles for chemically reduced graphene.8 Therefore, ER-G is promising as an electrode material for ultracapacitors, which are becoming more and more important in energy storage.41,54 Fig. 5 shows the polarization curves of electrocatalytic reduction of oxygen (Fig. 5a) and hydrogen peroxide (Fig. 5b) on ER-G and CNT electrodes. The ER-G exhibits greatly enhanced catalytic activity towards the electrochemical reduction of O2 and H2O2. The onset potentials of ORR and H2O2 reduction are positively shifted by 250 mV and 150 mV on the ER-G compared with those on the CNT electrodes, respectively, the reduction currents are enhanced and a well-dened H2O2 reduction peak at 0.2 V (Ag/AgCl) is observed on ER-G (no well-dened H2O2 reduction peaks on chemically reduced graphene oxide19,55) (see also Fig. S4).
J. Mater. Chem., 2010, 20, 743748 | 745

This journal is The Royal Society of Chemistry 2010

View Online

Downloaded by University Institute of Chemical Technology (UICT) on 23 November 2011 Published on 20 November 2009 on http://pubs.rsc.org | doi:10.1039/B917975E

Fig. 5 (a) Oxygen reduction on ER-G and CNT electrodes in O2-saturated 0.5 M H2SO4 (10 mV s1, 1600 RPM); (b) H2O2 reduction on ER-G and CNT electrodes in N2-saturated 5 mM H2O2 + 10 mM PBS + 0.1 M KCl (50 mV s1).

Fig. 6 (a) C1s high-resolution XPS spectra of graphite, GO, and ER-G; (b) XRD patterns of graphite, GO, and ER-G; (c) XPS spectra of graphite, GO, ER-G, and CNTs (F is from Naon. Na at.% 4.7% and S at.% 0.6% for ER-G, implying that in addition to trace Na2SO4, a certain amount of Na+ ions are intercalated in ER-G).

GO before and after electrochemical reduction was characterized with XPS30 and XRD,50 and the results are shown in Fig. 6. Detailed information of the surface chemistry (oxygencontaining groups) is shown in Table 2. It can be seen that graphite exhibits a single sharp C1s XPS spectrum centered at 284.3 eV (mainly sp2 carbon, Fig. 6a). GO shows a typical C1s spectrum of heavily oxygenated carbon. ER-G shows a similar and relatively broad C1s XPS spectrum compared with graphite, indicating that partial sp2 carbon was restored after electrochemical reduction. There is an obvious hump (BE 285.0  289.0 eV) in the ER-G C1s spectrum, indicating that there are still oxygen-containing groups on ER-G. Fig. 6b shows the XRD patterns of graphite, GO, and ER-G. It can be seen that after oxidation, the sharp diffraction peak in graphite (2q 26.5 , corresponding to the interlayer distance d 0.336 nm) disappeared, and a new diffraction peak (2q 11.4 , d 0.777 nm) appeared in GO,49 indicating the complete oxidation of graphite. This is a prerequisite to successfully exfoliating GO and obtaining single-layer or few-layer graphene sheets.50 After the electrochemical reduction, the diffraction peak at 2q 26.5 does not reappear as seen in the graphene produced with other methods,50,57 while the XRD pattern of ER-G displays
Table 2 Fitted results (%) of high-resolution C1s XPS spectra of GO, ER-G and CNTs.30,56 COOR 289.1 eV 4.00 3.05 6.28 CO 287.8 eV 15.84 6.75 5.29 CO 286.5 eV 65.98 10.93 8.76 CC (defect/sp3) 285.15 eV 14.18 18.86 21.73 CC (sp2) 284.31 eV

an amorphous structure. This means that either all stacking of graphene layers in ER-G is lost, or any remaining stacking is disordered.50 Fig. 6c shows the full-scale XPS spectra of graphite, GO, ER-G, and CNTs. It can be seen that the relative intensity of oxygen signals, which indicates the oxygen content, is in the order of GO> ER-G > CNT > graphite. In ER-G XPS spectra, the signals of the elements F, Na, and S, in addition to carbon and oxygen, were observed. The element F is from Naon. Na and S are from Na2SO4 (S might also be from Naon). The XPS spectrum of CNTs is very clean, and no metal impurities were detected.

Discussion

Sample GO ER-G CNT

53.72 57.96

The electrochemical reduction of GO is an accumulative process. As is known, GO is insulating,46 so at the rst stage of the electrochemical reduction only GO in direct contact with a conducting gold disk/glassy carbon electrode can be electrochemically reduced on the condition that the electrolyte (ions) has access to it (the triple-phase boundary formation). More and more insulating GO turns to conducting graphene while the reduction process proceeds, which makes more GO accessible for electrochemical reduction and accounts for the increase of the reduction peak (0.75 V, Fig. 2a); and after a certain number of reduction cycles, the amount of GO on glass carbon/Au disk electrodes becomes less and less, which leads to the decrease and nal vanishing of the reduction peak. When the reduction peak (0.75 V) vanishes, the redox peaks (0.5 V) reach their highest point and keep stable with a certain number of further potential cycling. At this point, the electrochemically active surface area of ER-G also reaches the highest value and
This journal is The Royal Society of Chemistry 2010

746 | J. Mater. Chem., 2010, 20, 743748

View Online

remains constant. These are consistent with the trend of the calculated specic capacitance (double-layer non-Faradaic capacitance + the redox Faradaic capacitance) as shown in Fig. 2b. This also indicates that the redox couples (0.5 V) are electrochemically stable under such conditions. So the reduction of GO can be in situ monitored and controlled by its reduction peaks (0.75 V) and the redox current peaks (0.50 V). As is known, chemical reduction removes the oxygen from GO and produces hydrophobic graphene, which tends to restack because of the strong pp interaction between graphene sheets,30,58 which greatly reduces the electrochemically accessible surface area when assembled into the electrode. The electrochemical reduction removes partial oxygen from GO (as can be seen from the XPS spectra, Fig. 6 and Table 2) and also increases the hydrophobicity of graphene sheets. However the electrochemical reduction can only take place at the electrochemical triple-phase boundary (which is accessible to both electrons from the electrode and ions from the electrolyte) at very low potentials under which condition the positively charged Na+ ions (0.1 M Na2SO4 was used as an electrolyte) can be easily intercalated between graphene oxide layers.59 The full-scale XPS spectra (Fig. 6c) show the existence of F, Na+, and S in ER-G. The element F was from Naon, which was applied on the GO surface to stabilize GO on gold disk electrodes. Na+ and S were from Na2SO4 (S might also be from Naon), the supporting electrolyte in the experiment. The content of Na+ and S was calculated (through the high-resolution XPS spectra) to be 4.7 and 0.6%, respectively, which indicates that, in addition to trace Na2SO4 in ER-G, a certain amount of Na+ was intercalated in graphene oxide layers during the electrochemical reduction. The intercalation of Na+ in GO/ER-G prevents electrochemically reduced graphene sheets from re-stacking and provides the ER-G electrode with more electrochemically active surface area, which contributes to its high specic capacitance. The oxygen content is much higher for ER-G than CNTs and chemically reduced graphene, as calculated from XPS spectra (Fig. 6 and Table 2). This is also conrmed from the CVs of the electrodes: ER-G produces a much larger redox peak (0.5 V in Fig. 3a) than chemically reduced graphene (CVs in Ref. 6) and CNTs (Fig. 3b, and Fig. S3), which contributes to a large part of the capacitance for ER-G electrodes. It is generally known that carbon materials with oxygen-containing groups will degrade in capacitance with charge/discharge cycles,60 which is also observed for our CNT electrodes (Fig. 4) and previously reported chemically reduced graphene.8 However, the specic capacitance of ER-G (Fig. 4) slightly increases with the charge/discharge cycle, even though there is much higher oxygen content on ERG. This might be due to the electrochemical stabilization of oxygen-containing groups (the unstable functional groups have been reduced/removed in the extended CV during the GO reduction process). However the exact underlying mechanisms still need further investigation. CNTs have shown enhanced catalytic activity towards the electrochemical reduction of O26163 and H2O2,64 but metal impurities in CNTs might also attribute the catalytic effect for some previous works using unpuried CNTs.61,65,66 Our CNTs and ER-G do not contain such impurities (see XPS spectra in Fig. 6c). The onset potential of oxygen reduction on ER-G is about 0.6 V (RHE) as compared to 0 V (standard calomel
This journal is The Royal Society of Chemistry 2010

Downloaded by University Institute of Chemical Technology (UICT) on 23 November 2011 Published on 20 November 2009 on http://pubs.rsc.org | doi:10.1039/B917975E

electrode, SCE) on acid-washed CNTs in the literature61 (comparable with the CNTs used in this work). The onset potential of H2O2 reduction on ER-G is about 0.05 V (Ag/AgCl) much higher than that on CNTs (0.10 V, Ag/AgCl) in this work and the reported values in the literature.66 So ER-G exhibits much better catalytic activity towards oxygen reduction and H2O2 reduction than CNTs. The enhanced catalytic activity might be attributed to the abundant specic functional groups on ER-G. The XPS spectra of ER-G and CNTs are quite different: more C]O and CO groups on ER-G (Fig. S5 and Table 2). The CVs of ER-G and CNTs are also different with ER-G exhibiting enhanced redox current peaks (0.5 V) (Fig. 3ab and Fig. S3). The redox current peaks in the CVs are usually attributed to quinone-hydroquinone redox,56,67 which is consistent with XPS analysis. These functional groups can effectively catalyze the reduction of oxygen and hydrogen peroxide.68 Oxygen reduction is one key reaction in fuel cells and metalair batteries, and hydrogen peroxide is a general enzymatic product of oxidases and a substrate of peroxidases, which is important in biological processes and biosensor development.45 Therefore, the ndings of the enhanced electrocatalytic activity of ER-G indicates that ER-G is promising for applications in electrochemical energy conversion and storage and biosensors.

Conclusions
Graphene oxide is electrochemically reduced to produce graphene, which we call electrochemically reduced graphene oxide (ER-G). This reduction process can be in situ monitored and controlled. ER-G exhibits higher electrochemical capacitance and cycling durability than carbon nanotubes and chemically reduced graphene. The electrocatalytic activity of O2 reduction and H2O2 reduction is greatly enhanced on ER-G as compared to carbon nanotubes. Electrochemically reduced graphene oxide is a promising material for supercapacitors, fuel cells, metal/air batteries and biosensors.

Acknowledgements
This work was supported by a LDRD program at PNNL. Part of the research described in this paper was performed at the Environmental Molecular Sciences Laboratory, a national scienticuser facility sponsored by the U.S. Department of Energys (DOEs) Ofce of Biological and Environmental Research and located at the Pacic Northwest National Laboratory (PNNL). PNNL is operated for DOE by Battelle under Contract DEAC05-76L01830.

References
1 A. K. Geim and K. S. Novoselov, Nat. Mater., 2007, 6, 183191. 2 A. K. Geim, Science, 2009, 324, 15301534. 3 K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004, 306, 666669. 4 J. Hass, W. A. de Heer and E. H. Conrad, J. Phys.: Condens. Matter, 2008, 20, 323202. 5 T. Ramanathan, A. A. Abdala, S. Stankovich, D. A. Dikin, M. Herrera-Alonso, R. D. Piner, D. H. Adamson, H. C. Schniepp, X. Chen, R. S. Ruoff, S. T. Nguyen, I. A. Aksay, R. K. Prudhomme and L. C. Brinson, Nat. Nanotechnol., 2008, 3, 327331.

J. Mater. Chem., 2010, 20, 743748 | 747

View Online
6 M. D. Stoller, S. J. Park, Y. W. Zhu, J. H. An and R. S. Ruoff, Nano Lett., 2008, 8, 34983502. 7 D. W. Wang, F. Li, J. P. Zhao, W. C. Ren, Z. G. Chen, J. Tan, Z. S. Wu, I. Gentle, G. Q. Lu and H. M. Cheng, ACS Nano, 2009, 3, 17451752. 8 Y. Wang, Z. Q. Shi, Y. Huang, Y. F. Ma, C. Y. Wang, M. M. Chen and Y. S. Chen, J. Phys. Chem. C, 2009, 113, 1310313107. 9 E. Yoo, J. Kim, E. Hosono, H. Zhou, T. Kudo and I. Honma, Nano Lett., 2008, 8, 22772282. 10 B. Seger and P. V. Kamat, J. Phys. Chem. C, 2009, 113, 79907995. 11 R. Kou, Y. Y. Shao, D. H. Wang, M. H. Engelhard, J. H. Kwak, J. Wang, V. V. Viswanathan, C. M. Wang, Y. H. Lin, Y. Wang, I. A. Aksay and J. Liu, Electrochem. Commun., 2009, 11, 954957. 12 E. Yoo, T. Okata, T. Akita, M. Kohyama, J. Nakamura and I. Honma, Nano Lett., 2009, 9, 22552259. 13 Y. C. Si and E. T. Samulski, Chem. Mater., 2008, 20, 67926797. 14 Y. M. Li, L. H. Tang and J. H. Li, Electrochem. Commun., 2009, 11, 846849. 15 X. Wang, L. J. Zhi, N. Tsao, Z. Tomovic, J. L. Li and K. Mullen, Angew. Chem., Int. Ed., 2008, 47, 29902992. 16 J. B. Wu, H. A. Becerril, Z. N. Bao, Z. F. Liu, Y. S. Chen and P. Peumans, Appl. Phys. Lett., 2008, 92, 263302. 17 Z. Liu, J. T. Robinson, X. M. Sun and H. J. Dai, J. Am. Chem. Soc., 2008, 130, 1087610877. 18 H. Chen, M. B. Muller, K. J. Gilmore, G. G. Wallace and D. Li, Adv. Mater., 2008, 20, 35573561. 19 C. S. Shan, H. F. Yang, J. F. Song, D. X. Han, A. Ivaska and L. Niu, Anal. Chem., 2009, 81, 23782382. 20 Z. J. Wang, X. Z. Zhou, J. Zhang, F. Boey and H. Zhang, J. Phys. Chem. C, 2009, 113, 1407114075. 21 L. Tang, Y. Wang, Y. Li, H. Feng, J. Lu and J. Li, Adv. Funct. Mater., 2009, 19, 2782. 22 C. Lee, X. D. Wei, J. W. Kysar and J. Hone, Science, 2008, 321, 385 388. 23 S. Park and R. S. Ruoff, Nat. Nanotechnol., 2009, 4, 217224. 24 A. N. Obraztsov, Nat. Nanotechnol., 2009, 4, 212213. 25 W. A. de Heer, C. Berger, X. S. Wu, P. N. First, E. H. Conrad, X. B. Li, T. B. Li, M. Sprinkle, J. Hass, M. L. Sadowski, M. Potemski and G. Martinez, Solid State Commun., 2007, 143, 92100. 26 K. V. Emtsev, A. Bostwick, K. Horn, J. Jobst, G. L. Kellogg, L. Ley, J. L. McChesney, T. Ohta, S. A. Reshanov, J. Rohrl, E. Rotenberg, A. K. Schmid, D. Waldmann, H. B. Weber and T. Seyller, Nat. Mater., 2009, 8, 203207. 27 P. W. Sutter, J. I. Flege and E. A. Sutter, Nat. Mater., 2008, 7, 406 411. 28 K. S. Kim, Y. Zhao, H. Jang, S. Y. Lee, J. M. Kim, J. H. Ahn, P. Kim, J. Y. Choi and B. H. Hong, Nature, 2009, 457, 706710. 29 X. S. Li, W. W. Cai, J. H. An, S. Kim, J. Nah, D. X. Yang, R. Piner, A. Velamakanni, I. Jung, E. Tutuc, S. K. Banerjee, L. Colombo and R. S. Ruoff, Science, 2009, 324, 13121314. 30 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, 15581565. 31 S. Gilje, S. Han, M. Wang, K. L. Wang and R. B. Kaner, Nano Lett., 2007, 7, 33943398. 32 D. Li, M. B. Muller, S. Gilje, R. B. Kaner and G. G. Wallace, Nat. Nanotechnol., 2008, 3, 101105. 33 X. B. Fan, W. C. Peng, Y. Li, X. Y. Li, S. L. Wang, G. L. Zhang and F. B. Zhang, Adv. Mater., 2008, 20, 44904493. 34 H. C. Schniepp, J. L. Li, M. J. McAllister, H. Sai, M. Herrera-Alonso, D. H. Adamson, R. K. Prudhomme, R. Car, D. A. Saville and I. A. Aksay, J. Phys. Chem. B, 2006, 110, 85358539. 35 J. S. Wu, W. Pisula and K. Mullen, Chem. Rev., 2007, 107, 718747. 36 N. Liu, F. Luo, H. X. Wu, Y. H. Liu, C. Zhang and J. Chen, Adv. Funct.Mater., 2008, 18, 15181525. 37 G. K. Ramesha and S. Sampath, J. Phys. Chem. C, 2009, 113, 7985 7989. 38 M. Zhou, Y. Wang, Y. Zhai, J. Zhai, W. Ren, F. Wang and S. Dong, Chem.Eur. J., 2009, 15, 61166120. 39 F. R. F. Fan, S. Park, Y. W. Zhu, R. S. Ruoff and A. J. Bard, J. Am. Chem. Soc., 2009, 131, 937. 40 H. L. Guo, X. F. Wang, Q. Y. Qian, F. B. Wang and X. H. Xia, ACS Nano, 2009, 3, 2653. 41 P. Simon and Y. Gogotsi, Nat. Mater., 2008, 7, 845854. ~a, M. V. Buchanan, Steven Visco, 42 J. B. Goodenough, H. D. Abrun M. Stanley Whittingham, B. Dunn, Y. Gogotsi, A. Gewirth, D. Nocera, R. D. Kelley and J. S. Vetrano, Report of the Basic Energy Sciences Workshop for Electrical Energy Storage, April 24, 2007, http://www.sc.doe.gov/bes/reports/les/EES_rpt.pdf. 43 Y. Shao, J. Liu, Y. Wang and Y. Lin, J. Mater. Chem., 2009, 19, 46 59. 44 Y. Y. Shao, J. H. Sui, G. P. Yin and Y. Z. Gao, Appl. Catal., B, 2008, 79, 8999. 45 J. Wang, Electroanalysis, 2005, 17, 714. 46 G. Eda, G. Fanchini and M. Chhowalla, Nat. Nanotechnol., 2008, 3, 270274. 47 T. Szabo, O. Berkesi, P. Forgo, K. Josepovits, Y. Sanakis, D. Petridis and I. Dekany, Chem. Mater., 2006, 18, 27402749. 48 Y. C. Xing, L. Li, C. C. Chusuei and R. V. Hull, Langmuir, 2005, 21, 41854190. 49 D. A. Dikin, S. Stankovich, E. J. Zimney, R. D. Piner, G. H. B. Dommett, G. Evmenenko, S. T. Nguyen and R. S. Ruoff, Nature, 2007, 448, 457460. 50 M. J. McAllister, J. L. LiO, D. H. Adamson, H. C. Schniepp, A. A. Abdala, J. Liu, M. Herrera-Alonso, D. L. Milius, R. CarO, R. K. Prudhomme and I. A. Aksay, Chem. Mater., 2007, 19, 4396 4404. 51 N. A. Kotov, I. Dekany and J. H. Fendler, Adv. Mater., 1996, 8, 637 641. 52 M. M. Shaijumon, F. S. Ou, L. J. Ci and P. M. Ajayan, Chem. Commun., 2008, 23732375. 53 E. Frackowiak and F. Beguin, Carbon, 2001, 39, 937950. 54 A. S. Arico, P. Bruce, B. Scrosati, J. M. Tarascon and W. Van Schalkwijk, Nat. Mater., 2005, 4, 366377. 55 M. Zhou, Y. M. Zhai and S. J. Dong, Anal. Chem., 2009, 81, 5603 5613. 56 Y. Y. Shao, G. P. Yin, J. Zhang and Y. Z. Gao, Electrochim. Acta, 2006, 51, 58535857. 57 G. X. Wang, J. Yang, J. Park, X. L. Gou, B. Wang, H. Liu and J. Yao, J. Phys. Chem. C, 2008, 112, 81928195. 58 S. Niyogi, E. Bekyarova, M. E. Itkis, J. L. McWilliams, M. A. Hamon and R. C. Haddon, J. Am. Chem. Soc., 2006, 128, 77207721. 59 R. Alcantara, J. M. Jimenez-Mateos, P. Lavela and J. L. Tirado, Electrochem. Commun., 2001, 3, 639642. 60 A. G. Pandolfo and A. F. Hollenkamp, J. Power Sources, 2006, 157, 1127. 61 N. Alexeyeva and K. Tammeveski, Electrochem. Solid-State Lett., 2007, 10, F18F21. 62 M. N. Zhang, Y. M. Yan, K. P. Gong, L. Q. Mao, Z. X. Guo and Y. Chen, Langmuir, 2004, 20, 87818785. 63 K. P. Gong, F. Du, Z. H. Xia, M. Durstock and L. M. Dai, Science, 2009, 323, 760764. 64 A. Agui, P. Yanez-Sedeno and J. M. Pingarron, Anal. Chim. Acta, 2008, 622, 1147. 65 C. E. Banks, A. Crossley, C. Salter, S. J. Wilkins and R. G. Compton, Angew. Chem., Int. Ed., 2006, 45, 25332537. 66 M. Pumera, Langmuir, 2007, 23, 64536458. 67 K. H. Kangasniemi, D. A. Condit and T. D. Jarvi, J. Electrochem. Soc., 2004, 151, E125E132. 68 E. Yeager, J. Mol. Catal., 1986, 38, 525.

Downloaded by University Institute of Chemical Technology (UICT) on 23 November 2011 Published on 20 November 2009 on http://pubs.rsc.org | doi:10.1039/B917975E

748 | J. Mater. Chem., 2010, 20, 743748

This journal is The Royal Society of Chemistry 2010

You might also like