You are on page 1of 9

Ultrasonics Sonochemistry 18 (2011) 1082–1090

Contents lists available at ScienceDirect

Ultrasonics Sonochemistry
journal homepage: www.elsevier.com/locate/ultsonch

Sonochemical synthesis of TiO2 nanoparticles on graphene for use as photocatalyst


Jingjing Guo a, Shenmin Zhu a,⇑, Zhixin Chen b, Yao Li a, Ziyong Yu a, Qinglei Liu a, Jingbo Li a,
Chuanliang Feng a, Di Zhang a,⇑
a
State Key Laboratory of Metal Matrix Composites, Shanghai Jiao Tong University, 800 Dongchuan Road, Shanghai 200240, PR China
b
Faculty of Engineering, University of Wollongong, Wollongong, NSW 2522, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Using ultrasonication we succeed in a controlled incorporation of TiO2 nanoparticles on the graphene lay-
Received 29 November 2010 ers homogeneously in a few hours. The average size of the TiO2 nanoparticles was controlled at around
Received in revised form 14 March 2011 4–5 nm on the sheets without using any surfactant, which is attributed to the pyrolysis and condensation
Accepted 17 March 2011
of the dissolved TiCl4 into TiO2 by ultrasonic waves. The photocatalytic activity of the resultant graphene–
Available online 8 April 2011
TiO2 composites containing 25 wt.% TiO2 is better than that of commercial pure TiO2. This is partly due to
the extremely small size of the TiO2 nanoparticles and partly due to the graphene–TiO2 composite struc-
Keywords:
ture consisting of homogeneous dispersion of crystalline TiO2 nanoparticles on the graphene sheets. As
Graphene
Ultrasonication
the graphene in the composites has a very good contact with the TiO2 nanoparticles it enhances the
Photocatalysis photo-electron conversion of TiO2 by reducing the recombination of photo-generated electron–hole pairs.
Nanocomposite Ó 2011 Elsevier B.V. All rights reserved.

1. Introduction trochemical methods, chemical vapor deposition methods [8], ion-


exchange methods, intercalation [20], hydrothermal reduction
Graphene (GR), a flat monolayer of carbon atoms tightly packed methods [24], sol–gel methods [25], and inner modification fol-
into a two-dimensional honeycomb lattice. It has recently at- lowed by in situ reduction methods, have been developed to incor-
tracted a great deal of attention for potential applications in many porate nanoparticles inside graphene sheets [21]. To be effective as
fields, such as nanoelectronics, fuel-cell technology, supercapaci- spacers, the nanoparticles have to adhere to the surface of the
tors and catalysts [1–5]. Several approaches, including microme- graphene as uniformly as possible. Therefore, the control of the for-
chanical exfoliation of graphite [6,7], chemical vapor deposition mation and distribution of nanoparticles on the graphene layers is
[8,9], and solution-based chemical reduction [10–12] have been critical.
developed to produce single-layered graphene. Unfortunately, the Ultrasonication has proved to be an effective technique for gen-
yield of single-layer graphene sheets from various production erating nanoparticles with attractive properties in a short period of
methods is quite low [10,12–16]. An even more serious problem reaction time [26–28]. The enhanced chemical effect of ultrasound
is that single-layer sheets of graphene are not stable in solution is due to acoustic cavitation phenomena: the rapid formation,
and tend to aggregate back to graphite gradually because of the growth, and the collapse of bubbles in liquid. The extremely high
strong Johannes Diderik van der Waals interactions [17]. local temperature (>5000 K), pressure (>20 MPa) and very high
One possible technique of preventing aggregation and harness- cooling rates (>1010 Ks 1) confer sonicated solutions unique prop-
ing the unique properties of single-layer graphene would be to erties, and are able to reduce metal ions to metal or metal oxide
incorporate graphene sheets into composite materials. Many metal nanoparticles [29,30]. The major advantage of this method, apart
or metal oxide nanoparticles, such as Au, Ag, Pt, Pd, TiO2, SnO2, and from its fast quenching rate and operation at ambient conditions,
MnO2 have been deposited on graphene sheets [18–22]. These me- is that it is a simple and energy efficient process. In our previous
tal or metal oxide nanoparticles not only prevent the aggregation work [31], sonochemical method has been employed successfully
of graphene sheets into graphite but also combine with the special for the synthesis of TiO2 butterfly wings from butterfly templates.
two-dimensional (2D) graphene, giving rise to some unique elec- TiO2 has attracted much attention on account of its high photo-
tronic, optical and catalytical properties which may be used in conversion efficiency [32]. TiO2 can produce photo-induced elec-
applications, such as biologic sensing, photocatalysis, optoelec- tron–hole pairs under the irradiation of ultraviolet (UV) light. From
tronic and electrochemical devices [23]. Approaches, such as elec- the point view of photo-conversion efficiency, the photocatalytic
properties of TiO2 can be further enhanced if the recombination
of the photo-induced electron–hole pairs can be effectively
⇑ Corresponding authors. Tel.: +86 21 34202584; fax: +86 21 34202749. suppressed. Graphene–TiO2 composites are thus promising photo-
E-mail addresses: smzhu@sjtu.edu.cn (S. Zhu), zhangdi@sjtu.edu.cn (D. Zhang). catalytic materials because graphene can act as an electron transfer

1350-4177/$ - see front matter Ó 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.ultsonch.2011.03.021
J. Guo et al. / Ultrasonics Sonochemistry 18 (2011) 1082–1090 1083

channel thus reduces the recombination of the photo-generated stirred for 24 h at 50 °C. GR–TiO2 composite was obtained by filtra-
electron holes, and leads to improved photo-conversion efficiency tion and drying under vacuum at 60 °C. After calcination at 450 °C
[33,34]. It was reported that a mixture of commercially pure TiO2 for 3 h under nitrogen, crystallized GR–TiO2 composite (GR–TiO2–
(P25) powder and graphene showed a relatively high photocata- T) was obtained.
lytic property [24]. However, the particle size of P25 was relative
large and in micrometers and thus the photocatalytic property 2.4. Preparation of graphene
can be further improved if the size of TiO2 is reduced to nanome-
ters. In fact, TiO2 particle size needs to be reduced down to Graphene was obtained by chemical conversion of GO according
nanometers in order to improve the photocatalytic activity for to the literature [36]. In a typical reduction experiment, 0.5 g of GO
liquid-phase oxidation. powder was dispersed in 150 ml of deionized water and the mix-
In this paper, we report a sonochemical method that embeds ture was sonicated for 1 h. Next, 18 ml of hydrazine (85%) was
TiO2 nanoparticles into graphene oxide (GO) nanosheets homoge- added under magnetic stirring, and the mixture was continuously
neously without functionalizing the surface with a surfactant. stirred at 50 °C for 24 h. Finally, black graphene powder was
Graphene–TiO2 composite (GR–TiO2) was obtained by chemical obtained by filtration and drying under vacuum at 60 °C.
reduction of the graphene oxide–TiO2 composite (GO–TiO2). The
photocatalytic properties of the obtained GR–TiO2 composites
2.5. Characterization
were investigated by measuring the photo-degradation of methy-
lene blue under UV-light, which demonstrated a high photocata-
The synthesized samples were characterized by X-ray diffrac-
lytic performance for the obtained composites.
tion (XRD) using a RigakuD/max2550VL/PC system operated at
40 kV and 40 mA with Cu Ka radiation (k = 1.5406 Å), at a scan rate
2. Experimental of 5° min 1 and a step size of 0.050° in 2h. Nitrogen adsorption
measurements at 77 K were performed using an ASAP2020 volu-
2.1. Preparation of GO metric adsorption analyzer, after the samples had been outgassed
for 8 h in the degas port of the adsorption apparatus. Field-emis-
GO was prepared from natural graphite (crystalline, 300 mesh, sion scanning electron microscopy (FE-SEM) was performed on a
Alfa Aesar) by a modified Hummers method [35]. The detailed pro- JEOL JSM-6360LV field emission microscope at an accelerating
cessing is described as below: 2 g of graphite and 1 g of NaNO3 voltage of 15 kV. Transmission electron microscopy (TEM) and en-
were mixed with 46 ml of H2SO4 (98%) in a 250 ml wide necked ergy-dispersive X-ray measurements (EDX) were carried out on a
bottle placed in an ice bath. Then, the mixture was stirred for JEOL 2010 microscope at 200 kV. TEM specimens were prepared
30 min. While maintaining magnetic vigorous stirring, a certain by grinding the synthesized samples into powder with a mortar
amount of KMnO4 (6 g) was added to the suspension carefully. and pestle and the powder was dispersed in pure ethanol and
After that, the ice bath was replaced by an oil bath, and the mixture picked up with holey carbon supporting films on copper grids. Fou-
was stirred at 15 °C for 2 h. As the reaction progressed, the mixture rier transform-infrared measurements (FT-IR) were recorded on
gradually became pasty, and the color turned into light brownish. KBr pellets with a PE Paragon 1000 spectrophotometer. Thermal
The next step was to increase the reaction temperature to 40 °C gravimetric analysis (TGA) was conducted on a PE TGA-7 instru-
and keep for another 1 h. Then, 92 ml of H2O was slowly added ment with a heating rate of 20 °C/min. X-ray photoelectron spectra
to the pasty mixture with vigorous agitation. The reaction temper- (XPS) were collected on a physical electronics PHI5400 using Mg
ature was rapidly increased to 98 °C with effervescence, and the Ka radiation as X-ray source. All the spectra were corrected with
color changed to yellow. Finally, 30 ml of 5% H2O2 was added to the C1s (285.0 eV) band. Diffuse reflectance electronic spectra were
the mixture and allowing the high-temperature reaction to go on measured with a Perkin–Elmer 330 spectrophotometer equipped
for 1 h. For purification, the mixture was rinsed and centrifugated with a 60 mm Hitachi integrating sphere accessory.
with 5% HCl and deionized water for several times. After filtration
and drying under vacuum at 60 °C, GO was obtained as a gray 2.6. Photocatalytic experiments
powder.
30 mg of catalyst (GR, TiO2, and GR–TiO2–T) was dispersed in a
2.2. Preparation of GO–TiO2 40 ml 0.02 g/L aqueous solution of methylene blue dye in a quartz
test cell. The reaction system was kept saturated with oxygen by
GO–TiO2 composite was synthesized by sonochemical reaction purging a slow stream of oxygen at ambient temperature. Before
of TiCl4 in the presence of the GO. The detailed process is described irradiation, the reactor was stirred constantly in the presence of
as follows: firstly, TiO2 precursors were prepared with the molar oxygen in dark for 30 min, and then irradiated by four UV lamps
ratios of ethanol: H2O: TiCl4 = 35:11:1, then, 0.25 g of the GO was (6 W each) at room temperature. The photocatalytic activity was
added and stirred for 0.5 h at ambient temperature, and finally evaluated on the basis of decrease of the absorbance band of the
the suspension was sonicated at room temperature for 3 h using methylene blue at 665 nm recorded at a regular time interval fol-
a high-intensity ultrasonic probe (Ti horn, 20 kHz, 100 W/cm2). lowing the UV illumination. The absorbance measurement of the
The resulting composite was recovered by centrifugation and rins- reaction solution was taken after separating the TiO2 from the
ing with ethanol solvent several times, then dried under vacuum at reaction system by centrifugation.
60 °C. The obtained sample is called GO–TiO2.
3. Results and discussion
2.3. Preparation of GR–TiO2
3.1. Characterizations
Chemical conversion of GO–TiO2 to GR–TiO2 was carried out by
using a reduction method [36]. In a typical reduction experiment, Wide-angle XRD patterns of the pure TiO2, GR, GR–TiO2 and
0.7 g of GO–TiO2 powder was dispersed in 180 ml of deionized GR–TiO2–T are given in Fig. 1. The very broad peak at 2h = 25.5°
water and the mixture was sonicated for 1 h. Next, 24 ml of hydra- of GR–TiO2 indicates that TiO2 in the sample is largely of amor-
zine (85%) was added into the mixture under magnetic stirring and phous in nature. After the calcination at 450 °C, this peak at
1084 J. Guo et al. / Ultrasonics Sonochemistry 18 (2011) 1082–1090

the graphene sheets in the GR–TiO2–T composite exhibited nano-


scale textures, indicative of a much rougher surface (Fig. 3b). At a
high magnification (Fig. 3d), we can clearly see the uniform disper-
sion of TiO2 nanoparticles on both the graphene surface and the
interlayers. There is no agglomeration in the composite as shown
in the SEM image of the GR–TiO2 sheet (Fig. 3f). Fig. 3f also shows
that the TiO2 particle size was in nanometers which will be further
characterized by using TEM.
Fig. 4 shows TEM images of TiO2 loaded on the graphene layers
before (GR–TiO2) and after the calcination at 450 °C for 3 h (GR–
TiO2–T). For the as-synthesized GR–TiO2 (Fig. 4a), there are plenty
of wrinkles (in circles and at the edge) on the clean sheet owing to
the 2D nature of the sheet. The edges of the sheets are folded and
give the appearance of ribbons. High-magnification TEM micro-
graphs (Fig. 4b and c) reveal no discrete and well defined nanocrys-
talline TiO2 particles on the graphene sheets. However EDX (see
Supporting Information, Fig. s1) shows the presence of Ti, O and
C in the sample. The corresponding selected area electron diffrac-
tion pattern (SAED) of the composite shows that diffraction spots
of the graphite superimposed with the diffraction rings of TiO2
Fig. 1. XRD of TiO2, GR, GR–TiO2 before and after calcinations under nitrogen.
nanoparticles. After the calcination at 450 °C for 3 h, discrete and
well defined TiO2 fully crystallized particles were uniformly scat-
tered on the sheets as shown in Fig. 4d and the particle size is
2h = 25.5° becomes relatively narrower and resolved, indicating around 4–5 nm from the high resolution image (Fig. 4f). This obser-
that nanocrystalline TiO2 formed on the graphene. However, the vation is consistent with the wide angle XRD results presented ear-
peak is still too broad to be used to estimate the size of crystallites lier (Fig. 1). The inset in Fig. 4f is the SAED of GR–TiO2–T, and it can
with Scherrer equation [37]. Nevertheless the broad peak suggests be seen that the sample has a clear sixfold pattern of the graphene
that growth of the nanocrystalline TiO2 on the graphene was very and the diffraction rings of the TiO2 nanoparticles. The existence of
limited during the calcinations. elements Ti, O and C was detected by EDX analysis (see Supporting
The loading of the TiO2 nanoparticles on the graphene was Information Fig. s2). As demonstrated in Fig. 4e, the measured lat-
determined by performing TGA in air by heating up from 40 to tice-fringe spacing of 0.34 nm in the ribbons corresponds to the
900 °C (Fig. 2). A significant mass loss was observed at tempera- (0 0 2) of the graphene sheets, where the crystal lattice of TiO2
tures from 510 to 700 °C for both GR and GR–TiO2–T, due to the nanoparticles is also resolved. From Fig. 4f the measured lattice
destruction of the carbon skeleton (carbonyl/double bond) of fringe spacing of 0.355 nm in GR–TiO2–T composites corresponds
graphene. The weight loss of GR–TiO2–T was stabilized at 75% at to the (1 0 1) of anatase TiO2 (JCPDS 21-1272). No agglomeration
temperatures between 700 and 900 °C, which indicates that the of the TiO2 particles was observed in this sample, which is usually
amount of TiO2 loaded on the graphene sheets was about 25 wt.%. observed in samples prepared by other methods [20,24].
Fig. 3 shows secondary electron images of GR before and after The TEM images shown in Fig. 4 confirm that a uniform disper-
the functionalization with TiO2 (GR–TiO2–T). The pure graphene sion of TiO2 nanoparticles on the graphene sheets have been
has a layered structure with a very smooth surface (Fig. 3a, c and achieved under sonochemical reactions without the use of any sur-
e). The GR–TiO2–T composite shows a mass of wrinkles both on factant. Which means ultrasound is very effective in dispersing
the edge of graphene sheets (Fig. 3b, d and f) and on the interlayer TiO2 nanoparticles on graphene layers.
sheets (see the circles in Fig. 3b, d and f). The layered structure of In order to understand the texture of the composites, nitrogen
isothermal adsorption technique was employed to investigate the
ultrasonic irradiation effect on the pore structure as well as to
determine the location of the TiO2 in graphene (Fig. 5). The shape
of the isotherm for GR–TiO2–T was similar to that of the parent
GR, indicating the presence of open pores. It has a specific surface
area of 49 m2g 1 according to the BET (Brunauer, Emmett and Tell-
er) analysis. The P/P0 position of the inflection points is related to
the pore size in mesopore range, and the sharpness of these steps
indicates the uniformity [38]. As the relative pressure increases,
both the isotherms of GR and GR–TiO2–T exhibit sharp inflections
in the P/P0 in the range 0.5–0.6, a characteristic of capillary con-
densation within a uniform of interlayer spacing. Interestingly,
the space layer distribution of the starting GR centered at 2.6 nm
but disappeared when the TiO2 was loaded, suggesting the possi-
bility that titanium oxide formed first in this interlayer spacing.
As a result, the space layer distribution of GR–TiO2–T is more uni-
form than that of the pure GR from pore size distributions shown
in Fig. 5 (inset). The metal oxide nanoparticles increased the dis-
tance between the GR sheets to several nanometers, thereby mak-
ing the both faces of the graphene accessible.
The light-absorbance property of the samples was studied with
Fig. 2. TGA curves of GR and GR–TiO2–T (heating rate = 20 °C/min under air a UV–vis spectrophotometer and the measured UV–visible diffuse
atmosphere). reflectance electronic spectra are shown in Fig. 6. The absorption
J. Guo et al. / Ultrasonics Sonochemistry 18 (2011) 1082–1090 1085

Fig. 3. FE-SEM images of GR (a, c, e) and GR–TiO2–T (b, d, f) in different magnifications.

bands of GR–TiO2–T appeared to be fairly different from that of chemically bonded on the graphene sheets, which was facilitated
P25. Because the appearance of TiO2 is white whereas the GR– by the sonochemical reaction. Further it was found that the peak
TiO2–T is black, the Kubelka–Munk theory could not apply to this of the stretching vibration C-OH at 3419.7 cm 1 of GR shifted to
sample with such a strong adsorption. Therefore, the quantitative a higher wave number of 3443.8 cm 1 for GR–TiO2–T, which could
consideration is void between TiO2 and GR–TiO2–T. However, a be explained by the influence of the formation of Ti–O–C bond.
broad and strong absorption in the visible light region was ob- Similar to that of GR, the skeletal vibration of the GR sheets at
served for GR–TiO2–T due to the presence of 75 wt.% GR in the 1569.6 cm 1 was also clearly observed for GR–TiO2–T, illustrating
GR–TiO2–T composite which acts as a ‘‘dyade’’ structure [39]. This the formation of GR–TiO2–T composites. Due to the Ti–O–C bond
enhanced light-harvesting intensity of GR–TiO2–T could be possi- the TiO2 nanoparticles in GR–TiO2–T did not grow much upon cal-
bly explained by the formation of the chemical bonding between cinations, i.e. the host graphene sheet prevented the nanocrystals
TiO2 and GR, i.e. a Ti–O–C bond which favors charge transfer upon from sintering.
light excitation [40]. Complicated titanium coordination states in GR–TiO2–T, were
The formation of the Ti–O–C bond was confirmed by FT-IR spec- revealed from the XPS results. Fig. 8a shows the Ti 2p XPS spectra
troscopy as shown in Fig. 7. GO depicts a strong OH peak at of GR–TiO2–T, GR and GO. As expected, there are two peaks, Ti 2p3/2
3389.4 cm 1 and other C–O functionalities such as COOH centered at 458.5 eV and Ti 2p1/2 at 464.2 eV in the XPS spectrum
(1726.6 cm 1) and COC/COH (1383.3–1055.3 cm 1) are clearly vis- of GR–TiO2–T, which were not detected in the spectra of GR and
ible. The spectrum also shows a C@C peak at 1617.5 cm 1 corre- GO. Fig. 8b shows the XPS spectra of GR–TiO2–T, GO and GR in
sponding to the remaining sp2 character. After the hydrazine the binding energy range between 280 and 294 eV (C 1s). The
reduction, the characteristic absorption peaks (at 1726.6, 1383.3– two peaks at 287 and 285 eV of GO indicate a considerable degree
1055.3, and 591.2 cm 1) of GO disappeared due to the reduction of oxidation of the graphite. The binding energy of 285 eV is due to
of oxygenous groups from the GO. As for GR–TiO2–T, the broad non-oxygenated ring C (284.3 eV), and 287 eV is due to the C in C–
absorptions at low frequencies (below 1000 cm 1) were ascribed O bonds (285.2 eV), the carbonyl C (C@O, 287.4 eV) and carboxyl C
to the vibration of Ti–O–Ti bond (671.3, 683.9 cm 1) and Ti–O–C (COOH, 289.1 eV) respectively [41]. Both the C 1s in the XPS spec-
bond (780.2 cm 1) which was not observed in the spectrum of tra of GR and GR–TiO2–T exhibit only a single strong peak at
GR. This demonstrates that the TiO2 nanoparticles were strong 284.3 eV which is attributed to elemental carbon, with a weak
1086 J. Guo et al. / Ultrasonics Sonochemistry 18 (2011) 1082–1090

Fig. 4. TEM images of GR–TiO2 before (a–c) and after calcinations (GR–TiO2–T) (d–f) in different resolutions.

Fig. 5. N2 adsorption/desorption isotherms of samples GR and GR–TiO2–T and the


corresponding pore size distributions (inset). Fig. 6. Diffuse reflectance electronic spectra of pure TiO2 and GR–TiO2–T.
J. Guo et al. / Ultrasonics Sonochemistry 18 (2011) 1082–1090 1087

shoulder at 285.4 eV due to the oxygen functionalities (C–O


bonds), indicating considerable de-oxygenation by the reduction
process. There is an additional component at 286.4 eV correspond-
ing to the C in the C–N bonds of hydrazine, attributed to partial
reduction of carbonyl functionalities to hydrazone groups for GR
and GR–TiO2–T. The XPS survey scans of GR and GR–TiO2–T
(Fig. 8c) show that the characteristic peak at 400.98 eV (N 1p, cor-
responding to nitrate), which for GR–TiO2–T is considerably smal-
ler because some of N element was lost during the calcination. The
peak at around 281 eV, ascribed to Ti–C bonds was not detected in
GR–TiO2–T, suggesting that carbon was not doped into the lattice
of the TiO2.

3.2. Proposed mechanism

A uniform distribution of fine TiO2 particles on graphene sheets


has been achieved by a sonochemical method as described in Fig. 9.
Firstly, the process is involved in a simple physisorption of Ti
ion on graphene sheets. GO is known to consist of oxidized graph-
ene sheets that have their basal planes mostly decorated with
epoxide and hydroxyl groups, in addition to a few carboxyl groups.
Fig. 7. FT-IR spectra of GO, GR and GR–TiO2–T. Carboxylic moieties can interact with the metal cations through

Fig. 8. (a) The Ti 2p XPS spectra of GO, GR and GR–TiO2–T, (b) the C 1s XPS spectra of GO, GR and GR–TiO2–T, (c) the survey XPS spectra of GO, GR and GR–TiO2–T.
1088 J. Guo et al. / Ultrasonics Sonochemistry 18 (2011) 1082–1090

Fig. 9. The proposed TiO2 nanoparticles formation mechanism on the graphene sheets by sonochemical method.

physisorption, electrostatic binding, or through charge-transfer


interaction. The Ti ions might become attached to the carboxyl
groups, and the reaction to TiO2 takes place on the surface of the
exfoliated sheets. However, the number of ionizable carboxyl
groups at the edge is not enough to complex the metal cations
and to stabilize the metal oxide particles formed, resulting in less
homogeneous and large particles. Generally, surface functionaliza-
tion is needed in order to be effective for homogeneous dispersion
[22]. A surfactant such as SB12 is generally used to control the size
of nanoparticles and to prevent aggregation [42]. Here in our pro-
cess, fine TiO2 nanoparticles were formed on the surface of graph-
ene without surface modification or additional agent. The unique
properties of ultrasonication may have effects on the formation
of nanoparticles.
Ultrasound agitation disperses solutes homogeneously in the
mixture and enhances the diffusion rate of TiCl4 to the surface of
GO, which can interact with the metal cations forming Ti–O–C
bonds on the GO surface as evidenced by FT-IR (Fig. 7). The chem-
ical effect of the ultrasound arises from acoustic cavitation, the for- Fig. 10. Photo-degradation of methylene blue solution under UV-light over GR, TiO2
mation, growth, and implosive collapse of bubbles in a liquid. The and GR–TiO2–T, respectively.

aggregation of the TiCl4 on the surface of the GO sheets can be sig-


nificantly retarded by the collapse of small bubbles, resulting in a the pure GR, illustrating that the photocatalytic reaction is related
layer coating, a few nanometers thick, on the GO surface, which to the existence of active sites (here TiO2). The pseudo-first-order
is consistent with the disappearance of layer spacing at 2.6 nm reaction was observed for the both TiO2 and GR–TiO2–T (Fig. 11).
from the BET analysis. As a result, uniform dispersion of fine TiO2 For GR–TiO2–T, the k value at 0.0139 is higher than that of P25
forms in situ without any surface functionalization. The one-step at 0.0054; a high k value is well known to have a high photocata-
formation of the TiO2-graphene composites under ultrasonic irra- lytic activity. Photocatalytic efficiency can be affected by three
diation illustrates the simplicity and efficiency of the sonochemical things: crystalline phase, surface area and hierarchical structures.
approach as compared with the commonly used sol–gel’s proce- It is also well known that the activity is proportional to the surface
dure which is generally complex and very time-consuming. area accessible to the liquid. A large surface area could adsorb sig-
nificant amounts of water thus hydroxyl groups which can react
3.3. Photocatalytic measurements with photo excited holes and produce hydroxyl radicals which
are powerful oxidants in degrading organics. In this investigation,
The photocatalytic activities of GR–TiO2–T were measured by the surface area of GR–TiO2–T was almost the same as that of
the photo-degradation of methylene blue as model reaction under TiO2 but the former showed much higher photocatalytic activity
UV-light (6400 nm), and the results are shown in Fig. 10. For a than the latter. This may be explained by the crystalline TiO2 nano-
comparison, the photocatalytic activities of commercial TiO2 and particles prepared by sonochemical and calcinations approach,
pure GR were also measured under the same condition. In order which enhance the activity by facilitating the access to the reactive
to exclude the influence of adsorption process, a time of 60 min sites of TiO2. Further, the graphene in the composite can act as an
was allowed to achieve adsorption equilibrium before the photo- electron transfer channel to reduce the recombination of the
catalytic reaction. Fig. 10 shows that the concentration of methy- photo-generated electron holes, and lead to improved photo-con-
lene blue decreases fast with irradiation time both for TiO2 and version efficiency [33,34]. It should be noted that the weight ratio
GR–TiO2–T, but almost no concentration change was observed for of the TiO2 in GR–TiO2–T composite is only 25 wt.%, so the real
J. Guo et al. / Ultrasonics Sonochemistry 18 (2011) 1082–1090 1089

References

[1] A.K. Geim, K.S. Novoselov, The rise of graphene, Nat. Mater. 6 (2007) 183–191.
[2] S. Watcharotone, D.A. Dikin, S. Stankovich, R. Piner, I. Jung, G.H.B. Dommett, G.
Evmenenko, S.E. Wu, S.F. Chen, C.P. Liu, S.T. Nguyen, R.S. Ruoff, Graphene–silica
composite thin films as transparent conductors, Nano Lett. 7 (2007) 1888–
1892.
[3] X. Wang, L.J. Zhi, K. Mullen, Transparent, conductive graphene electrodes for
dye-sensitized solar cells, Nano Lett. 8 (2008) 323–327.
[4] J.C. Meyer, A.K. Geim, M.I. Katsnelson, K.S. Novoselov, T.J. Booth, S. Roth, The
structure of suspended graphene sheets, Nature 446 (2007) 60–63.
[5] M. Ishigami, J.H. Chen, W.G. Cullen, M.S. Fuhrer, E.D. Williams, Atomic
structure of graphene on SiO2, Nano Lett. 7 (2007) 1643–1648.
[6] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, Y. Zhang, S.V. Dubonos, I.V.
Grigorieva, A.A. Firsov, Electric field effect in atomically thin carbon films,
Science 306 (2004) 666–669.
[7] A. Dato, V. Radmilovic, Z.H. Lee, J. Phillips, M. Frenklach, Substrate-free gas-
phase synthesis of graphene sheets, Nano Lett. 8 (2008) 2012–2016.
[8] K.S. Kim, Y. Zhao, H. Jang, S.Y. Lee, J.M. Kim, K.S. Kim, J.H. Ahn, P. Kim, J.Y. Choi,
B.H. Hong, Large-scale pattern growth of graphene films for stretchable
transparent electrodes, Nature 457 (2009) 706–710.
[9] A. Reina, X.T. Jia, J. Ho, D. Nezich, H.B. Son, V. Bulovic, M.S. Dresselhaus, J. Kong,
Large area, few-layer graphene films on arbitrary substrates by chemical vapor
deposition, Nano Lett. 9 (2009) 30–35.
[10] D. Li, M.B. Muller, S. Gilje, R.B. Kaner, G.G. Wallace, Processable
aqueous dispersions of graphene nanosheets, Nat. Nanotechnol. 3 (2008)
101–105.
[11] S. Park, R.S. Ruoff, Chemical methods for the production of graphenes, Nat.
Fig. 11. Kinetic plots based on the data of Fig. 10.
Nanotechnol. 4 (2009) 217–224.
[12] S. Stankovich, D.A. Dikin, R.D. Piner, K.A. Kohlhaas, A. Kleinhammes, Y. Jia, Y.
Wu, S.T. Nguyen, R.S. Ruoff, Synthesis of graphene-based nanosheets via
chemical reduction of exfoliated graphite oxide, Carbon 45 (2007) 1558–
photocatalytic activity of the TiO2 nanoparticles should be even
1565.
higher than commercial pure TiO2. [13] Y. Hernandez, V. Nicolosi, M. Lotya, F.M. Blighe, Z.Y. Sun, S. De, I.T. McGovern,
B. Holland, M. Byrne, Y.K. Gun’ko, J.J. Boland, P. Niraj, G. Duesberg, S.
Krishnamurthy, R. Goodhue, J. Hutchison, V. Scardaci, A.C. Ferrari, J.N.
4. Conclusions Coleman, High-yield production of graphene by liquid-phase exfoliation of
graphite, Nat. Nanotechnol. 3 (2008) 563–568.
[14] C. Nethravathi, M. Rajamathi, Chemically modified graphene sheets produced
Using ultrasonication we succeed in controlled incorporation of
by the solvothermal reduction of colloidal dispersions of graphite oxide,
TiO2 on the graphene layers homogeneously in a few hours. The Carbon 46 (2008) 1994–1998.
average particle size of the TiO2 was controlled at around 4– [15] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, M.I. Katsnelson, I.V.
5 nm on the sheets without using any surfactant, which is attrib- Grigorieva, S.V. Dubonos, A.A. Firsov, Two-dimensional gas of massless Dirac
fermions in graphene, Nature 438 (2005) 197–200.
uted to the pyrolysis and condensation of the dissolved TiCl4 into [16] P.W. Sutter, J.I. Flege, E.A. Sutter, Epitaxial graphene on ruthenium, Nat. Mater.
TiO2 by ultrasonic waves. The synthesis process is simple and effi- 7 (2008) 406–411.
cient. The photocatalytic performance of GR–TiO2–T containing [17] D.F. Leigh, C. Norenberg, D. Cattaneo, J.H.G. Owen, K. Porfyrakis, A.L. Bassi, A.
Ardavan, G.A.D. Briggs, Self-assembly of trimetallic nitride template fullerenes
25 wt.% TiO2 is better than commercial pure TiO2. The much im- on surfaces studied by STM, Surf. Sci. 601 (2007) 2750–2755.
proved photocatalytic activity of GR–TiO2–T, prepared by sono- [18] R. Muszynski, B. Seger, P.V. Kamat, Decorating graphene sheets with gold
chemical method, is attributed to the graphene–TiO2 composite nanoparticles, J. Phys. Chem. C 112 (2008) 5263–5266.
[19] C. Xu, X. Wang, J.W. Zhu, Graphene-metal particle nanocomposites, J. Phys.
structure consisting of very small and homogeneous dispersion Chem. C 112 (2008) 19841–19845.
of crystal TiO2 on the graphene sheets. In this case the photocata- [20] G. Williams, B. Seger, P.V. Kamat, TiO2–graphene nanocomposites UV-
lytic activity of the TiO2 is enhanced not only by the fine size assisted photocatalytic reduction of graphene oxide, ACS Nano 2 (2008)
1487–1491.
but also by the graphene which reduces the recombination of [21] T. Cassagneau, J.H. Fendler, Preparation and layer-by-layer self-assembly of
photo-generated electron–hole pairs. Thus, such integration of silver nanoparticles capped by graphite oxide nanosheets, J. Phys. Chem. B 103
2D supports with large surface areas, and the highly dispersed (1999) 1789–1793.
[22] Y.C. Si, E.T. Samulski, Exfoliated graphene separated by platinum
nanoparticles, can be an exciting material for use in future nano-
nanoparticles, Chem. Mater. 20 (2008) 6792–6797.
technology. These results and other related data demonstrate that [23] C. Xu, X. Wang, Fabrication of flexible metal-nanoparticte film using graphene
this sonochemical method may also be extended to synthesize oxide sheets as substrates, Small 5 (2009) 2212–2217.
other metal oxides on graphene sheets, with high performance in [24] H. Zhang, X.J. Lv, Y.M. Li, Y. Wang, J.H. Li, P25–graphene composite as a high
performance photocatalyst, ACS Nano 4 (2010) 380–386.
many potential applications, such as optical, electrical, catalysis, [25] X.Y. Zhang, H.P. Li, X.L. Cui, Y.H. Lin, Graphene/TiO2 nanocomposites:
sensors, and energy conversion devices and so on. synthesis, characterization and application in hydrogen evolution from
water photocatalytic splitting, J. Mater. Chem. 20 (2010) 2801–2806.
[26] J.C. Yu, J.G. Yu, W.K. Ho, L.Z. Zhang, Preparation of highly photocatalytic active
Acknowledgments nano-sized TiO2 particles via ultrasonic irradiation, Chem. Commun. (2001)
1942–1943.
[27] G.A. Tai, W.L. Guo, Sonochemistry-assisted microwave synthesis and optical
The authors gratefully acknowledge the financial support of this
study of single-crystalline US nanoflowers, Ultrason. Sonochem. 15 (2008)
research by the National Science Foundation of China (Nos. 350–356.
51072117, 50772067), Shanghai Science and Technology Commit- [28] A. Gedanken, Using sonochemistry for the fabrication of nanomaterials,
Ultrason. Sonochem. 11 (2004) 47–55.
tee (Nos.06PJ14063, 07DJ14001), and Sino-French Project of MOST
[29] A. Gedanken, X.H. Tang, Y.Q. Wang, N. Perkas, Y. Koltypin, M.V. Landau, L.
of China (No. 2009DFA52410). We also thank SJTU Instrument Vradman, M. Herskowitz, Using sonochemical methods for the preparation of
Analysis Center for the measurements. mesoporous materials and for the deposition of catalysts into the mesopores,
Chem.-Eur. J. 7 (2001) 4546–4552.
[30] W. Chen, W.P. Cai, C.H. Liang, L.D. Zhang, Synthesis of gold nanoparticles
Appendix A. Supplementary data dispersed within pores of mesoporous silica induced by ultrasonic irradiation
and its characterization, Mater. Res. Bull. 36 (2001) 335–342.
[31] S.M. Zhu, D. Zhang, J.J. Gu, J.Q. Xu, J.P. Dong, J.L. Li, Biotemplate fabrication of
Supplementary data associated with this article can be found, in SnO2 nanotubular materials by a sonochemical method for gas sensors, J.
the online version, at doi:10.1016/j.ultsonch.2011.03.021. Nanopart. Res. 12 (2010) 1389–1400.
1090 J. Guo et al. / Ultrasonics Sonochemistry 18 (2011) 1082–1090

[32] U.I. Gaya, A.H. Abdullah, Heterogeneous photocatalytic degradation of organic [38] L. Zhao, X.F. Chen, X.C. Wang, Y.J. Zhang, W. Wei, Y.H. Sun, M. Antonietti, M.M.
contaminants over titanium dioxide: A review of fundamentals, progress and Titirici, One-step solvothermal synthesis of a carbon@TiO2 dyade structure
problems, J. Photoch. Photobio. C 9 (2008) 1–12. effectively promoting visible-light photocatalysis, Adv. Mater. 22 (2010)
[33] X.Y. Zhang, H.P. Li, X.L. Cui, Preparation and photocatalytic activity for 3317.
hydrogen evolution of TiO2/graphene sheets composite, Chin. J. Inorg. Chem. [39] M. Niederberger, G. Garnweitner, F. Krumeich, R. Nesper, H. Colfen, M.
25 (2009) 1903–1907. Antonietti, Tailoring the surface and solubility properties of nanocrystalline
[34] D.H. Wang, D.W. Choi, J. Li, Z.G. Yang, Z.M. Nie, R. Kou, D.H. Hu, C.M. Wang, L.V. titania by a nonaqueous in situ functionalization process, Chem. Mater. 16
Saraf, J.G. Zhang, I.A. Aksay, J. Liu, Self-assembled TiO2-graphene hybrid (2004) 1202–1208.
nanostructures for enhanced Li-Ion insertion, ACS Nano 3 (2009) 907–914. [40] S. Stankovich, R.D. Piner, X.Q. Chen, N.Q. Wu, S.T. Nguyen, R.S. Ruoff, Stable
[35] M. Hirata, T. Gotou, S. Horiuchi, M. Fujiwara, M. Ohba, Thin-film particles of aqueous dispersions of graphitic nanoplatelets via the reduction of exfoliated
graphite oxide 1: high-yield synthesis and flexibility of the particles, Carbon graphite oxide in the presence of poly(sodium 4-styrenesulfonate), J. Mater.
42 (2004) 2929–2937. Chem. 16 (2006) 155–158.
[36] Y. Wang, Y.M. Li, L.H. Tang, J. Lu, J.H. Li, Application of graphene-modified [41] R. Bissessur, P.K.Y. Liu, S.F. Scully, Intercalation of polypyrrole into graphite
electrode for selective detection of dopamine, Electrochem. Commun. 11 oxide, Synthetic. Met. 156 (2006) 1023–1027.
(2009) 889–892. [42] Y.M. Cui, D.J. Shan, Y.R. Zhu, Studies on photocatalytic oxidation of I-over TiO2
[37] A.L. Patterson, The Scherrer formula for X-ray particle size determination, thin film, Chin. J. Inorg. Chem. 17 (2001) 401–406.
Phys. Rev. 56 (1939) 978–982.

You might also like