Hydrothermal synthesis of a uniformly dispersed hybrid graphene–TiO2 nanostructure for optical and enhanced electrochemical applications

Rajesh Kumar*ab, Rajesh Kumar Singh*c, Pawan Kumar Dubeyd, Dinesh Pratap Singhe, Ram Manohar Yadavf and Radhey Shyam Tiwarib
aDepartment of Materials Science and Engineering, Yonsei University, Seoul-120749, South Korea. E-mail: rajeshbhu1@gmail.com
bDepartment of Physics, Banaras Hindu University, Varanasi-221005, India
cDepartment of Physics, Indian Institute of Technology (Banaras Hindu University), Varanasi-221005, India. E-mail: rksbhu@gmail.com
dNanotechnology Application Centre, University of Allahabad, Allahabad-211002, India. E-mail: dubey.pawan@yahoo.com
eDepartamento de Física, Universidad de Santiago de Chile, Avenida Ecuador 3493, Estación Central, Santiago-9170124, Chile. E-mail: dineshpsingh@gmail.com
fDepartment of Materials Science and Nano Engineering, Rice University, Houston, Texas-77005, USA. E-mail: rmanohar28@gmail.com

Received 9th July 2014 , Accepted 17th December 2014

First published on 18th December 2014


Abstract

Highly dispersed TiO2 nanoparticles on graphene nanosheets were achieved by hydrothermal treatment of graphene nanosheets obtained by modified Hummer's method followed by thermal exfoliation. The hybrid graphene TiO2 nanostructure composite (H-GTN) showed enhanced optical and electrochemical properties for future application as a supercapacitor. The structural, optical and electrochemical properties of the composite are systematically investigated. The as-prepared H-GTN showed a quenching phenomenon of its photoluminescence properties, which was attributed to the specific properties of graphene. Remarkably, the CV test obtained for H-GTN showed a very high specific capacitance value up to 530 F g−1 at a scan rate of 3 mV s−1, and nearly stable capacitance of 400 F g−1 above 20 mV s−1. The cyclic stability test shows stable behavior after some initial cycles and the stability was then retained without obvious aging or performance degradation, showing long cyclic stability. This is attributed to the excellent electrochemical performance of the H-GTN electrode material for practical application in energy storage devices.


Introduction

Graphene, a monolayer of sp2-hybridized carbon atoms arranged in a two-dimensional honeycomb lattice, has attracted tremendous attention and research in recent years, owing to its exceptional properties such as high thermal conductivity, ultrahigh charge carrier mobility, extremely large surface area, high mechanical strength and flexibility.1–4 Though free standing graphene has excellent physical, chemical and electrical properties, its tendency to restack is a major hurdle in realizing its full potential especially in applications such as supercapacitors and lithium ion batteries. These unique properties along with atomic scale dimension offer various potential applications in nanoelectronics, energy conversion/storage, catalysis, and nanocomposites.5–9 The two dimensional sheet structure of graphene provides an excellent conductive platform for accommodating nanosized electrochemically active materials and allows surface coating while at the same time providing conduction channels for the transport of electrons. In addition, the high surface area and high charge carrier mobility of graphene offer new opportunities to develop graphene-based hybrid nanocomposite materials with other common metal oxides/hydroxides and conducting polymers for supercapacitor applications.6,10

Transition metal oxides have been explored as potential electrode materials for use in supercapacitors with their pseudocapacitance based charge storage mechanisms.11–14 RuO2 has been found to give high capacitance due to redox transitions that penetrates into the bulk of the material. However, the cost of Ru is one of the major concerns for its commercial acceptance. The use of graphene has been widely explored as the nanoscale substrates for the formation of nanocomposites with metal oxides to obtain a hybrid, which may have combined features of both graphene and nanosized metal oxide particles.15,16 In recent years, metal oxides such as TiO2, MnO, RuO, IrO, etc. have been used in graphene-based supercapacitors with the function of contributing pseudo-capacitance to the total capacitance.17 TiO2 is considered to be one of the most attractive compounds for supercapacitors not only due to its pseudocapacitance properties but due to its abundance in nature, commercial viability, excellent chemical stability, nontoxicity to the atmosphere and high surface area. Therefore, it is necessary to develop a simple and effective method to prepare highly dispersed hybrid graphene–TiO2 nanocomposites (H-GTN) for supercapacitor application.

A variety of methods has been demonstrated for the preparation of graphene–TiO2 composite, including chemical, hydrothermal, photo-catalytic methods, and so on.9,16,18,19 In this work, we report the synthesis of H-GTN using facile hydrothermal method under mild condition and explored their optical and electrochemical properties to elucidate its potential applications as supercapacitor. The H-GTN is synthesized via in situ hydrothermal process employing titanium tetraisopropoxide (TiC12H28O4) precursor as titania source. In H-GTN, graphene nanosheets provide a large surface area for the decoration of TiO2 nanostructure and serve as a highly conductive supportive base. Our experimental results show that the hybrid nanostructure exhibits overall specific capacitance of 400 F g−1 with high cyclic stability in aqueous electrolyte.

Result and discussions

The stepwise schematic in situ hydrothermal synthesis procedure of H-GTN is shown in Fig. 1. As obtained graphene nanosheets and H-GTN were characterized by SEM for microstructural characterizations. Fig. 2a shows the SEM micrographs of as synthesized GNSs, and H-GTN. Fig. 2a reveals that the GNSs consisting of randomly aggregated, thin crumpled 3-dimensional structure of the sheets closely associated with each other and having disordered structure. The folded regions of the sheets were found to have an average width of ∼1 μm by high resolution SEM. Fig. 2b shows the attached TiO2 nanoparticles on the wrinkled surface of GNSs. In H-GTN, the TiO2 nanoparticles are attached to the surface of GNSs with the help of oxygen functionalities formerly attached on the graphene surface. At high TiO2 content, it is uniformly decorated and firmly anchored on the wrinkled GNSs layers with uniform density. Energy-dispersive X-ray spectroscopy (EDS) analysis of H-GTN hybrids is shown in the inset of Fig. 2c. This shows the analysis of H-GTN hybrids. The EDS data of H-GTN shows that the elemental content of C, O and Ti only elements detected, indicating that H-GTN has a higher O content. Based on the obtained results, the atomic weight percentages of C, O and Ti were 63.32, 25.16 and 11.51%, respectively.
image file: c4ra06852a-f1.tif
Fig. 1 Schematic representation for the formation of H-GTN using hydrothermal process.

image file: c4ra06852a-f2.tif
Fig. 2 SEM image of (a) GNSs and (b and c) H-GTN.

The EDS analysis of the H-GTN hybrid shows the presence of C, O, and Ti elements. The existence of Ti and O with an approximate ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2 implies its stoichiometry. The TEM and HRTEM images of as-synthesized GNSs and H-GTN are shown in Fig. 3. It is clearly seen from Fig. 3a that GNSs are transparent with wrinkles on the surface. The HRTEM images of GNSs (Fig. 3b) show the few layer graphene and the number of layers varies between 4–7 layers. Fig. 3c shows anchored and nicely distributed TiO2 nanoparticles on the GNSs. Due to the homogeneous distribution, the aggregation of TiO2 nanoparticles or the GNSs is prohibited and the specific surface area of TiO2 is highly increased.


image file: c4ra06852a-f3.tif
Fig. 3 TEM micrographs of as synthesized (a) GNSs, (b) HRTEM of GNSs, (c and d) H-GTN and (e and f) HRTEM of H-GTN with different lattice plane orientations.

This is beneficial for fast diffusion of the redox phase, which makes the hybrid to attain high electrochemical capacitance. The adjacent lattice spacing as shown in Fig. 3d corresponds to the distance between the two nearest crystal planes of GNSs and TiO2 as clearly observable in the higher magnification image. The HRTEM images (Fig. 3e and f) of H-GTN show the lattice fringes of spacing 0.37, 0.35 and 0.25 nm, that correspond to (002) crystal plane of GNSs, (101) plane of anatase TiO2 and (101) plane of rutile TiO2, respectively. The clear lattice plane of TiO2 indicates that the TiO2 is of high crystallinity.

The XRD pattern of the as-synthesized GNSs and H-GTN are shown in Fig. 4. The peaks at 2θ values of 25.3, 37.8, 48.0, 53.9, 55.1, 62.7, 68.8, 70.3, and 75.0° can be indexed to (101), (004), (200), (105), (211), (204), (116), (220), and (215) crystal planes of anatase TiO2, resp. In addition, characteristic diffraction peaks of rutile phase of TiO2 at 27.4, 36.1, and 41.2° are also observed, that are attributed to the (110), (101), and (111) faces.


image file: c4ra06852a-f4.tif
Fig. 4 XRD pattern of (a) GNSs and (b) H-GTN.

Notably, no typical diffraction peaks of separate GNSs are observed in the H-GTN. There are some small intensity peaks present in the range of 13–20 deg, which are also noticeable in GNSs XRD pattern. These peaks show the presence of stacked GNSs in the sample. The most intense diffraction peak of GNSs may not be distinguishable as the same peak at ∼25° (002) is overlapped or suppressed by the (101) diffraction peak of anatase TiO2.20,21

Bonding nature of GNSs and H-GTN was further investigated by Raman spectroscopy. Fig. 5 shows the Raman spectra of GNSs and H-GTN. Both samples exhibit two peaks namely D-band and G-band. The D-band is associated with the presence of defects in the hexagonal graphitic layers and the G-bands is associated with Raman-active E2g mode which is usually assigned to the breathing mode of k point phonons of A1g symmetry and the E2g phonon of C sp2 atoms, respectively.22 It can be noted that the Raman D-band peak of GNSs at 1359 cm−1 as been shifted to 1336 cm−1 in H-GTN i.e. shifted towards lower wave number. Raman G-bands peak of GNSs at 1583 cm−1 has been shifted to 1587 cm−1 in H-GTN i.e. shifted towards higher wave number. These observations show that the D-band is slightly red-shifted by 23 cm−1 in the H-GTN, whereas the G band has shown a blue shift of 4 cm−1. Similar shifts in the case of H-GTN have also been reported elsewhere.23,24 These shifts in the Raman peak have been attributed to the chemical interaction between GNSs and TiO2 nanoparticles. The ID/IG ratio has been found to increase from 0.48 to 0.92 for GNSs to H-GTN. This indicates the existence of reduction to graphene and interaction between graphene and TiO2 nanoparticles.25


image file: c4ra06852a-f5.tif
Fig. 5 Raman spectra of (a) GNSs and (b) H-GTN.

The second order Raman feature, the 2D-band at 2710 cm−1 and D + G band at 2960 cm−1 are very sensitive to the stacking order of the graphene sheets along the c-axis (the number of layers) and show more broadened shape (often a doublet) with an increasing number of graphene layers.26,27 The 2D peak position of the H-GTN (Fig. 5b) is shifted to 2671 cm−1 from that of GNSs at 2710 cm−1, which indicates formation of few layer graphene in TiO2 in H-GTN.28 The intensity ratio of D band to G band is usually used to represents the relative degree of disorder.29 These disorderness are associated with vacancies, grain boundary and amorphous carbon presents in the as synthesized samples.30 The intensity ratio of D band to G band, for H-GTN (0.92) shows an enhanced value compared to that for GNSs (0.48), suggesting the presence of more localized sp3 defects within the sp2 domains carbon network on H-GTN.28 This is also an evidence of the reduction that take place during the formation of H-GTN. It suggests that the TiO2 lattices are possibly entrapped inside the GNSs by concurring chemical reaction. In addition, peaks around 149 cm−1 (the main Eg anatase vibration mode), 324 cm−1 (B1g), 507 cm−1 (A1g), and 632 cm−1 (Eg) cm−1 suggest that the anatase crystallites are the major species in H-GTN. These signatures of anatase TiO2 already has been reported in the Raman spectra of H-GTN.31

The Fourier transform infrared (FTIR) study of as-synthesized GNSs and H-GTN in the range of 4000–400 cm−1 is shown in Fig. 6. The peaks at 3173 and 3389 cm−1 are due to the presence of O–H stretching vibrations of the C–OH groups and water in GNSs and H-GTN. The other peaks in GNSs, correspond to carboxyl C[double bond, length as m-dash]O (1728 cm−1), aromatic stretching C[double bond, length as m-dash]C (1627 cm−1), tertiary C–OH group stretching (1376 cm−1), and alkoxy C–OH group stretching vibrations (1070 cm−1).32


image file: c4ra06852a-f6.tif
Fig. 6 FTIR spectra of (a) GNSs and (b) H-GTN.

In H-GTN, the peak at 1638 cm−1 correspond to C[double bond, length as m-dash]C, which is shifted towards higher wave number side as comparison to GNSs. In Fig. 6b, compared with GNSs, the intensities of the peaks corresponding C[double bond, length as m-dash]O and C–OH groups disappeared in the FTIR spectrum of H-GTN, indicating that the oxygen-containing functional groups in GNSs were removed from H-GTN. The broad band between 600 and 1000 cm−1 is the characteristic of Ti–O–Ti stretching in H-GTN based materials.33,34 This peak could be ascribed to a combination of Ti–O–Ti and Ti–O–C vibration.9,35 The presence of Ti–O–C (580 cm−1) bonds indicates the firmly bonding between the TiO2 nanoparticles and GNSs during the hydrothermal process.

X-ray photoelectron spectroscopy (XPS) is carried out to evaluate the chemical composition of H-GTN sample. The typical XPS spectra of as prepared H-GTN sample are displayed in Fig. 7. As shown from Fig. 7a, the full scale XPS spectra show the C, O, and Ti photoelectron lines as detected in the XPS survey spectra of H-GTN. The Ti 2p spectrum (Fig. 7b) showing the spin–orbit split lines of Ti 2p3/2 and Ti 2p1/2 located at 458.7 and 464.7 eV, respectively, that are characteristic of the Ti4+ oxidation state.36 The ratios (at%) of C, Ti, and O are 65.54[thin space (1/6-em)]:[thin space (1/6-em)]10.04[thin space (1/6-em)]:[thin space (1/6-em)]24.42, respectively. The O/Ti ratio is 2.43, (>2 of the stoichiometry of Titania) slightly higher than that of pristine TiO2 nanoparticles, resulting from the additional oxygen containing functional group of the reduced graphene. In order to investigate the states of carbon in the sample, spectrum of the C 1s core levels were measured and shown in Fig. 7d. Deconvolution of the C 1s peak in the XPS spectrum performed by four types of carbon bonds, namely, C–C (284.51 eV), C[double bond, length as m-dash]C (284.84 eV), COOH (289.50) and C–O (285.95 eV).37 The low intensity small peaks related to oxygenate C–O groups and very small amount of carboxyl group COOH indicate the presence of residual oxygenate groups on the H-GTN. The smaller peak corresponds to the oxygenate groups suggest a considerable de-oxygenation and the formation of H-GTN.38


image file: c4ra06852a-f7.tif
Fig. 7 XPS survey spectra of (a) H-GTN. Core level XPS spectra of (b) Ti 2p, (c) O 1s and (d) C 1s of H-GTN.

A small amount of residual oxygenate groups on H-GTN are believed to be favorable for maintaining a good dispersion of the nanoparticles.

In the process of simultaneous hydrothermal conversions of GO to graphene, the C–O functional groups on the surface of GO may react with [TiO] framework to form the Ti–O–C bonds, thus acting as the anchoring spots to initiate and support the growth of titania nanoparticles. In the O 1s spectrum (Fig. 7c), the peak located at 531.1 eV is attributed to O–H or the Ti–O–C groups.

Electron energy loss spectroscopy (EELS) was again carried out to confirm the presence of Ti–O–C group. The EEL spectra for GNSs and H-GTN are shown in Fig. 8. This shows the presence of Ti and O together with C peaks. Each C–K edge EELS consists of a peak at around 295 eV due to excitations from the 1s level to empty π* states of the sp2-bonded atoms. The H-GTN show that Ti and O elements result from anatase TiO2 and C element results from the GNSs. O–K, Ti–L2,3 and Ti–L1 peaks occur at ∼332, ∼460 and ∼560 eV, respectively.39


image file: c4ra06852a-f8.tif
Fig. 8 EEL spectra of (a) GNSs and (b) H-GTN.

Fig. 9a shows the photoluminescence (PL) of TiO2 and H-GTN. PL spectra of H-GTN were measured by using an ultraviolet light with a 265 nm wavelength as the excitation source and the results are shown in Fig. 9a. A strong emission peak around 500 nm was observed, implying that most of charges quickly recombine in TiO2 to produce PL emission. The H-GTN results show that nearly disappearance of PL intensities, which indicates that, the electron–hole recombination. When the pure TiO2 nanoparticles were coupled with GNSs, electrons would flow from conduction band (CB) of TiO2 into GNSs, leading to the formation of Schottky barrier at the H-GTN interface.40


image file: c4ra06852a-f9.tif
Fig. 9 Optical properties of (a) PL spectra and (b) UV-visible of TiO2 and H-GTN.

UV-visible spectrum for TiO2 and H-GTN are shown in Fig. 9b. Fig. 9b shows that in the whole visible region of the spectrum, as observed for other carbonaceous materials, GNSs combined and attached with TiO2 nanoparticles.41–43 After the formation of H-GTN, a bit shift in the absorption edge into the visible light region was observed. An induced dramatically improved light absorption in the visible light region was also observed, along with noticeable red shift of ca. 171 nm in the absorption edge of H-GTN, compared to bare TiO2. The red shift in the absorption band of HGTN may be attributed to the Ti–O–C bond. Ti–O–C bonds have the similar effect on TiO2 as impurities or carbon doping. The carbon doping or impurities in TiO2 composite introduce defect states into the TiO2 band gap, allowing photogeneration from lower-energy photons. The Ti–O–C bond in the HGTN is formed due to interaction of π electrons of the graphene nanosheets and free electrons in TiO2 which results into the replacement of some Ti–O–Ti bond to Ti–O–C covalent bonds due to higher electronegativity of the C than Ti.21,45 In fact, the graphene nanosheets might act as photosensitizer chemically bonded with TiO2 through the interactions between Ti ions and oxygen groups and thus contributes to the visible light absorption. Compared to pure TiO2 and H-GTN, the H-GTN exhibits increased spectra in UV region, suggesting that the electronic conjugation within the GNSs was restored upon chemical reduction.44 The narrowing of the band gap of TiO2 occurred with the introducing of GNSs. This narrowing should be attributed to the chemical bonding between TiO2 and GNSs, that is the formation of Ti–O–C bond, similar to the case of others reports for graphene–TiO2 composites.9,46 As shown in the inset of Fig. 9b, The optical band gap is for a semi-conductor near the absorption band edge can be estimated from the following equation known as the Tauc plot:47

(αhν) ∝ ( − Eg)n
where α is the optical absorption coefficient, is the energy of the incident photon, Eg the optical band and n = 0.5 and 2 correspond to direct allowed transition semiconductor and indirect allowed transition semiconductor respectively. Tauc plot has been drawn between (αhν)2 and energy of the photon for the calculation of the optical bandgap of TiO2 and H-GTN.48 Due to mixed phase Anatase (3.2 eV) and Rutile (3.0 eV) and some impurities phases present in TiO2 causes the measured optical bandgap lower than the standard value of the anatase phase 3.2 eV. The band gap of TiO2 was decreased from 3.03 eV to 2.78 eV by uniformly decoration of TiO2 nanoparticles. Fig. 9b shows that after hybridization with GNSs, the adsorption edges of TiO2 was shifted to visible light region.

To further evaluate the potential application of the H-GTN hybrids as electrode materials for electrochemical supercapacitors, cyclic voltammetry (CV) measurements were carried out between 0 and 0.6 V (vs. Ag/AgCl) at various scan rates ranging from 3 to 70 mV s−1, and corresponding specific capacitance has been shown in Fig. 10a. The specific capacitances of electrodes were calculated using CV curve using following equations as given below

image file: c4ra06852a-t1.tif
where C is the specific capacitance based on the mass of electroactive materials (F g−1), ∫Idv is the integrated area of CV curve, I is the response current density (A g−1), v is the scan rate (mV s−1), m is the mass of the electroactive materials in the electrode (g), ΔV is the sweep potential window (V). The measurements were carried out in a 0.5 M BMIM-BF4/CH3CN solution as electrolyte at room temperature. The specific capacitances evaluated from the CV curves were 531, 470, 431, 404, 403, 404, 402, 401, and 401 F g−1 at scan rates of 3, 5, 10, 20, 30, 40, 50, 70 and 100 mV s−1 respectively as shown in Fig. 10b. The specific capacitance shows nearly stable capacitance of 400 F g−1 above 20 mV s−1 which is more stable and constant as reported by other groups.49


image file: c4ra06852a-f10.tif
Fig. 10 Electrochemical measurements of H-GTN. (a) CV curve of H-GTN (b) specific capacitance values of H-GTN with scan rates and (c) long-term cycling stability of the supercapacitor at a constant current density of 5 A g−1 over 1000 cycles; the inset shows charge/discharge curve of the H-GTN in the potential range from 0 to 0.6 V at 5A g−1.

We observed that the specific capacitance for both electrodes decreased with an increase in the scan rate from 3 to 100 mV s−1. This is a common phenomenon and is caused by the insufficient time available for ion diffusion and adsorption inside the smallest pores within a particle at high scan rates.50 For high scan rates, the diffusion rates of electrolyte ions are limited by electrode structural properties, and only the external sites can take part in ion transfer reaction. But for low scan rates, all the active areas, including external and internal surfaces, can be utilized for charge/discharge and electrochemical utilization of TiO2 nanoparticles. The improvement can probably be attributed to the unique structure of H-GTN. The TiO2 nanoparticles, are well dispersed on the surface of GNSs not only effectively inhibit the stacking/agglomerating of GNSs, but also improve the electrochemical utilization of H-GTN. GNSs also provide a highly conductive network for electron transport during the charge/discharge processes. The excellent interfacial contact and increased contact area between TiO2 and GNSs can significantly improve the accessibility of H-GTN to the electrolyte ions and shorten the ion diffusion and migration pathways. Furthermore, GNSs can also serve as reliable conductive channels between individual active TiO2 nanoparticles. The cycle life time of H-TGN electrode was examined. As shown Fig. 10c, the cyclic stability test displays insignificant decrease in the specific capacitance over 100 cycles, approximately 9.1% from of the initial value and after 100 cycle its shows the nearly stable behavior.

This demonstrates the excellent electrochemical performance of the H-GTN electrode material for application in practical energy storage devices. It is concluded that the synergistic effect between conducting GNSs and TiO2 nanoparticles is responsible for the excellent electrochemical performance of the process.

Experimental methods

Synthesis of graphene oxide

Graphene oxide was prepared from graphite by modified Hummer's method.51 The graphite powder (1 g) was dispersed in concentrated sulphuric acid (H2SO4) (25 mL, 98 wt%) containing sodium nitrate (1 g, NaNO3), and then, potassium permanganate (KMnO4, 5 g) were slowly added with continuous vigorous stirring and cooling in ice bath to prevent the exceeding temperature. The ice bath was removed and replaced by a water bath, and the mixture was heated to 40 °C for 1 h for releasing gas under continuous stirring, followed by slow addition of deionized (DI) water (50 mL), which produced a rapid temperature increase in solution. The reaction was maintained for 24 h in order to increase the oxidation degree of the graphene oxide. The resultant brown-yellow suspension was terminated by addition of more DI water (300 mL), followed by a hydrogen peroxide solution (H2O2, 30%, 10 mL). The solid product was separated by centrifugation and washed with 500 mL of 1[thin space (1/6-em)]:[thin space (1/6-em)]5 HCl solution, and water until pH = 7 was achieved. Finally, the powder was dried at room temperature. This dried powder was thermally exfoliated in argon (Ar) atmosphere at 1000 °C for the conversion into graphene nanosheets (GNSs).

Synthesis of hybrid graphene–TiO2 nanostructure (H-GTN)

In a typical synthesis route for H-GTN, GNSs (50 mg) and titanium tetraisopropoxide (TiC12H28O4) (7 mg which was optimised after several experiments) were added to 50 mL of DI water and then the solution was sonicated for 45 min and stirred for 1 h using magnetic stirrer, respectively. The electrostatic and molecular grafting between GNSs and TiC12H28O4 was accomplished by stirring the mixture for 1 h at room temperature.52 After complete stirring, the solution was transferred into a 100 mL Teflon-lined tightly sealed stainless steel autoclave. The autoclave was then heated to 240 °C and kept there for 36 h. After the complete hydrothermal treatment, the hybrid nanostructure, denoted as H-GTN, was centrifuged, washed, and finally dried in air at 55 °C overnight.

Materials characterization

The as grown materials were characterized by X-ray diffractmeter (XRD, Philips 1710). The sample was scanned for 2θ from 10° to 80°. Scanning electron microscope (SEM) observations were performed using a Philips XL 20 at 2 kV in gentle-beam mode without any metal coating with the fully dried sample loaded on a carbon tape. The X-ray photoelectron spectroscopy (XPS) spectrum was recorded on a MultiLab2000 photoelectron spectrometer (Thermo-VG Scientific, USA) with Al Kα (1486.6 eV) as the X-ray source. All XPS spectra were corrected using the C 1s line at 284.6 eV. The EDS analysis was performed at several points in the region and averaged to obtain the representative results. The transmission electron microscopy (TEM) analysis was carried out using a Tecnai G2 20 microscope operated at 200 kV using a holey carbon-coated copper grid. The sample was prepared by dispersing a small amount of dry powder in ethanol or water. Then, one droplet of the suspension was dropped on a holey-carbon coated, 300 mesh copper TEM grids and allowed to dry in air at room temperature. The Raman spectrum was acquired on a LabRAM HR UV/vis/NIR (Horiba JobinYvon, France) using a CW Ar-ion laser (514.5 nm) as an excitation source focused through a confocal microscope (BXFM, Olympus, Japan) equipped with an objective lens (50×, numerical aperture = 0.50) at room temperature. The changes in the surface chemical bonding and surface composition were characterized by Fourier transform infrared spectroscopy (FT-IR, Jasco FT-IR 4100). The test samples were pressed into tablets with KBr. The UV-visible absorption spectra were recorded using a (Jasco, V-570) spectrometer. The photoluminescence (PL) emission spectra of the samples were recorded on a fluorescence spectrophotometer (LabRAM HR UV/Vis/NIR PL) using a 325 nm He–Cd laser as an excitation light source. Electron energy loss spectroscopy (EELS) spectra were recorded on a JEOL (TEM: ARM 200F) microscope operated at 300 using holey carbon-coated copper grid.

Electrochemical measurements

All electrochemical measurements were done in a three-electrode system. The H-GTN electrode behave as working electrode, Pt (platinum) wire as a counter electrode, and Ag/AgCl electrode as a reference electrode. Electrochemical measurements, such as cyclic voltammetry (CV) were carried out by a VersaSTAT 3 (Princeton Applied research). The glassy carbon electrodes (GCEs) (5 mm diameter) were polished first with 1.0 and 0.05 μm alumina slurry. Then, 1.0 mg of the H-GTN hybrid material was dispersed in 1.0 mL of dimethylformamide (DMF) with the aid of ultrasonicator to give a 1.0 mg mL−1 black suspension. Then, 10 μL of the black suspension was dropped on a cleaned GCE electrode with the help of a micro-syringe and the solvent was evaporated in air. Thus, a uniform film of H-GTN hybrids coating was formed on the surface of GCE. Finally, 5 μL of Nafion (5 wt%) was cast and used as a net to stably hold the H-GTN hybrids on the electrode surface. The solvent was allowed to evaporate before use. The final electrode was taken as the H-GTN hybrid electrode. All electrochemical measurements were done with a three-electrode system, which set the above working electrode, Pt (platinum) wire as a counter electrode, and Ag/AgCl electrode as a reference electrode.

Conclusions

Uniformly dispersed TiO2 nanoparticles on the graphene surfaces were achieved by a new, straightforward and facile one pot route by in situ hydrothermal method. Mixed rutile and anatase phases TiO2 sphere like nanoparticles were homogeneously dispersed on the few layers GNSs. Nanoparticles had a narrow size particle distribution of ∼10–20 nm. The graphene evidently influenced the formation of TiO2 and made the nanoparticles more uniform in morphology and size due to the selective nucleation and growth of TiO2 on GNSs. We observed the optical properties of TiO2 nanoparticles to be directly affected by GNSs and the photoluminescence spectra of H-GTN attributed the significant improvement in quenching with respect to TiO2 nanoparticles in the wavelength of 400–700 nm. UV-visible spectra showed the slight decrease in band gap after the formation of H-GTN. The electrochemical measurements results showed that GNSs and TiO2 nanoparticles enhanced the electrode conductivity, stability and improved the supercapacitive behavior of H-GTN. Significantly, this result also showed that the H-GTN had a high specific capacitance with constant stability at high scan rate and long cycling stability. The present as synthesized materials suggested that the H-GTN can be used as a promising material for supercapacitor for high performance.

Acknowledgements

We would like to gratefully acknowledge anonymous referees for useful comments and constructive suggestions. Two of the authors, RKS and PKD are grateful to UGC-New Delhi for providing Dr D. S. Kothari Postdoctoral fellowship.

References

  1. A. A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. Teweldebrhan, F. Miao and C. N. Lau, Nano Lett., 2008, 8, 902 CrossRef CAS PubMed.
  2. K. I. Bolotin, K. J. Sikes, Z. Jiang, M. Klima, G. Fudenberg, J. Hone, P. Kim and H. L. Stormer, Solid State Commun., 2008, 146, 351 CrossRef CAS PubMed.
  3. M. D. Stoller, S. Park, Y. Zhu, J. An and R. S. Ruoff, Nano Lett., 2008, 8, 3498 CrossRef CAS PubMed.
  4. C. Lee, X. Wei, J. W. Kysar and J. Hone, Science, 2008, 321, 385 CrossRef CAS PubMed.
  5. Y. Xuan, Y. Q. Wu, T. Shen, M. Qi, M. A. Capano, J. A. Cooper and P. D. Ye, Appl. Phys. Lett., 2008, 92, 013101 CrossRef PubMed.
  6. T. Kuila, A. K. Mishra, P. Khanra, N. H. Kim and J. H. Lee, Nanoscale, 2013, 5, 52 RSC.
  7. B. F. Machado and P. Serp, Catal. Sci. Technol., 2012, 2, 54 CAS.
  8. H. Kim, A. A. Abdala and C. W. Macosko, Macromolecules, 2010, 43, 6515 CrossRef CAS.
  9. H. Zhang, X. Lv, Y. Li, Y. Wang and J. Li, ACS Nano, 2009, 4, 380 CrossRef PubMed.
  10. A. K. Mishra and S. Ramaprabhu, J. Phys. Chem. C, 2011, 115, 14006 CAS.
  11. M. Toupin, T. Brousse and D. Bélanger, Chem. Mater., 2004, 16, 3184 CrossRef CAS.
  12. H. Kim and B. N. Popov, J. Power Sources, 2002, 104, 52 CrossRef CAS.
  13. C.-C. Hu, K.-H. Chang, M.-C. Lin and Y.-T. Wu, Nano Lett., 2006, 6, 2690 CrossRef CAS PubMed.
  14. Y.-T. Wang, A.-H. Lu, H.-L. Zhang and W.-C. Li, J. Phys. Chem. C, 2011, 115, 5413 CAS.
  15. C. Xu, X. Wang and J. Zhu, J. Phys. Chem. C, 2008, 112, 19841 CAS.
  16. G. Williams, B. Seger and P. V. Kamat, ACS Nano, 2008, 2, 1487 CrossRef CAS PubMed.
  17. P.-C. Chen, G. Shen, Y. Shi, H. Chen and C. Zhou, ACS Nano, 2010, 4, 4403 CrossRef CAS PubMed.
  18. D. Wang, D. Choi, J. Li, Z. Yang, Z. Nie, R. Kou, D. Hu, C. Wang, L. V. Saraf, J. Zhang, I. A. Aksay and J. Liu, ACS Nano, 2009, 3, 907 CrossRef CAS PubMed.
  19. C. Zhu, S. Guo, P. Wang, L. Xing, Y. Fang, Y. Zhai and S. Dong, Chem. Commun., 2010, 46, 7148 RSC.
  20. J. S. Chen, Z. Wang, X. C. Dong, P. Chen and X. W. Lou, Nanoscale, 2011, 3, 2158 RSC.
  21. Y.-J. Xu, Y. Zhuang and X. Fu, J. Phys. Chem. C, 2010, 114, 2669 CAS.
  22. W. Zhang, J. Cui, C.-a. Tao, Y. Wu, Z. Li, L. Ma, Y. Wen and G. Li, Angew. Chem., Int. Ed., 2009, 48, 5864 CrossRef CAS PubMed.
  23. Y. Min, K. Zhang, W. Zhao, F. Zheng, Y. Chen and Y. Zhang, Chem. Eng. J., 2012, 193–194, 203 CrossRef CAS PubMed.
  24. O. Akhavan, M. Abdolahad, A. Esfandiar and M. Mohatashamifar, J. Phys. Chem. C, 2010, 114, 12955 CAS.
  25. H. Gao, W. Chen, J. Yuan, Z. Jiang, G. Hu, W. Shangguan, Y. Sun and J. Su, Int. J. Hydrogen Energy, 2013, 38, 13110–13116 CrossRef CAS PubMed.
  26. M. A. Pimenta, G. Dresselhaus, M. S. Dresselhaus, L. G. Cancado, A. Jorio and R. Saito, Phys. Chem. Chem. Phys., 2007, 9, 1276 RSC.
  27. A. C. Ferrari, J. C. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri, F. Mauri, S. Piscanec, D. Jiang, K. S. Novoselov, S. Roth and A. K. Geim, Phys. Rev. Lett., 2006, 97, 187401 CrossRef CAS.
  28. D. Graf, F. Molitor, K. Ensslin, C. Stampfer, A. Jungen, C. Hierold and L. Wirtz, Nano Lett., 2007, 7, 238 CrossRef CAS PubMed.
  29. A. C. Ferrari and J. Robertson, Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 61, 14095 CrossRef CAS.
  30. T. A. Pham, B. C. Choi, K. T. Lim and Y. T. Jeong, Appl. Surf. Sci., 2011, 257, 3350 CrossRef CAS PubMed.
  31. Y. Lei, L. D. Zhang and J. C. Fan, Chem. Phys. Lett., 2001, 338, 231 CrossRef CAS.
  32. 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 CrossRef CAS PubMed.
  33. K. Zhou, Y. Zhu, X. Yang, X. Jiang and C. Li, New J. Chem., 2011, 35, 353 RSC.
  34. J. Shen, B. Yan, M. Shi, H. Ma, N. Li and M. Ye, J. Mater. Chem., 2011, 21, 3415 RSC.
  35. S. Sakthivel and H. Kisch, Angew. Chem., Int. Ed., 2003, 42, 4908 CrossRef CAS PubMed.
  36. S.-K. Joung, T. Amemiya, M. Murabayashi and K. Itoh, Chem.–Eur. J., 2006, 12, 5526 CrossRef CAS PubMed.
  37. N. Yang, J. Zhai, D. Wang, Y. Chen and L. Jiang, ACS Nano, 2010, 4, 887 CrossRef CAS PubMed.
  38. S. Stankovich, R. D. Piner, X. Chen, N. Wu, S. T. Nguyen and R. S. Ruoff, J. Mater. Chem., 2006, 16, 155 RSC.
  39. F. Zou, Y. Yu, N. Cao, L. Wu and J. Zhi, Scr. Mater., 2011, 64, 621 CrossRef CAS PubMed.
  40. T. Kamegawa, D. Yamahana and H. Yamashita, J. Phys. Chem. C, 2010, 114, 15049 CAS.
  41. C. G. Silva and J. L. Faria, Appl. Catal., B, 2010, 101, 81 CrossRef CAS PubMed.
  42. S. Shanmugam, A. Gabashvili, D. S. Jacob, J. C. Yu and A. Gedanken, Chem. Mater., 2006, 18, 2275 CrossRef CAS.
  43. L. Zhang, P. Liu and Z. Su, J. Mol. Catal. A: Chem., 2006, 248, 189 CrossRef CAS PubMed.
  44. P. Cheng, Z. Yang, H. Wang, W. Cheng, M. Chen, W. Shangguan and G. Ding, Int. J. Hydrogen Energy, 2012, 37, 2224 CrossRef CAS PubMed.
  45. M. Xing, F. Shen, B. Qiu and J. Zhang, Sci. Rep., 2014, 4, 6341 CrossRef PubMed.
  46. Y. Zhang, Z.-R. Tang, X. Fu and Y.-J. Xu, ACS Nano, 2010, 4, 7303 CrossRef CAS PubMed.
  47. H. Jia, H. Xu, Y. Hu, Y. Tang and L. Zhang, Electrochem. Commun., 2007, 9, 354 CrossRef CAS PubMed.
  48. J. Tauc, Mater. Res. Bull., 1970, 5, 721 CrossRef CAS.
  49. H. Su, T. Wang, S. Zhang, J. Song, C. Mao, H. Niu, B. Jin, J. Wu and Y. Tian, Solid State Sci., 2012, 14, 677 CrossRef CAS PubMed.
  50. J. Yan, T. Wei, W. Qiao, B. Shao, Q. Zhao, L. Zhang and Z. Fan, Electrochim. Acta, 2010, 55, 6973 CrossRef CAS PubMed.
  51. W. S. Hummers and R. E. Offeman, J. Am. Chem. Soc., 1958, 80, 1339 CrossRef CAS.
  52. Y.-B. Tang, C.-S. Lee, J. Xu, Z.-T. Liu, Z.-H. Chen, Z. He, Y.-L. Cao, G. Yuan, H. Song, L. Chen, L. Luo, H.-M. Cheng, W.-J. Zhang, I. Bello and S.-T. Lee, ACS Nano, 2010, 4, 3482 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.