To What Extent Do Graphene Scaffolds Improve

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

pubs.acs.

org/JPCL

To What Extent Do Graphene Scaffolds Improve


the Photovoltaic and Photocatalytic Response of
TiO2 Nanostructured Films?
Yun Hau Ng,†,‡,§ Ian V. Lightcap,† Kevin Goodwin,† Michio Matsumura,‡ and
Prashant V. Kamat*,†

Radiation Laboratory and Department of Chemistry & Biochemistry, University of Notre Dame, Notre Dame, Indiana 46556,
and ‡Research Center for Solar Energy Chemistry, Osaka University, 1-3 Machikaneyama, Toyonaka, 560-8531, Japan

ABSTRACT Graphene-TiO2 nanocomposites synthesized via a solution-based


method involving photocatalytic reduction of graphene oxide have been employed
as photoanodes. Nearly 90% enhancement in the photocurrent is seen as reduced
graphene oxide serves as electron collector and transporter. Additionally, the graphene-
TiO2 nanocomposite electrodes exhibit significant activity for the complete photo-
catalytic decomposition of 2,4-dichlorophenoxyacetic acid (2,4-D). Combined
with safe, solution-based synthetic practices, the promising photocurrent and
photocatalytic degradation rates provide the framework and motivation for the
implementation of graphene-TiO2 nanocomposites on larger scales.
SECTION Energy Conversion and Storage

Scheme 1. Electron Transport in Pristine TiO2 (left) and RGO-

C
arbon nanostructures such as fullerenes, carbon nano-
tubes, and graphene possess unique electronic and TiO2 Nanocomposite (right) Films Cast on an Optically Transpar-
ent Electrode (OTE)
catalytic properties.1-7 The recent emergence of stable
two-dimensional (2-D) carbon, graphene, has initiated a rapid
exploration of its potential in a broad range of fields such as
nanoelectronics, sensors, catalysis and nanocomposites.8-10
Owing to the existence of an extended sp2-bonded carbon
network, individual sheets of graphene possess excellent con-
ductivity. However, strong van der Waals interactions bet-
ween graphitic sheets make it difficult to isolate individual
graphene layers. Chemical functionalization or deposition of
metal nanoparticles keep sheets separate and further intro-
separated. Graphene sheet thicknesses obtained through
duce new electronic and optical properties for these carbon
this method typically correspond to mono- or bilayer gra-
nanostructures.9,11-14 Two common preparations for exfolia-
phene. Additionally, the process allows for control over the
ted graphene sheets are mechanical cleavage of highly pyro-
degree of reduction by adjusting exposure times. Solution-
lytic graphite15,16 and chemical treatment.11,17-19 Although
based graphene-TiO2 nanocomposites can thus be readily
mechanical cleavage of graphite has enabled the examination
formed in situ and are easily transferred or film-cast to a wide
of pristine graphene, its low yield significantly hinders it from
range of substrates.23 In recent years, several groups have
large-scale application. Chemical oxidation is a convenient,
employed such studies to explore photoelectrochemical and
scalable method that can be used to induce the exfoliation
photocatalytic properties of GO-TiO2 composites.24-28
of graphite into an aqueous dispersion of individual sheets of
The electron-accepting ability of carbon nanostructures
graphene oxide (GO). After oxidation, GO can be reduced
can be used to enhance electron transport properties of nano-
chemically to reduced graphene oxide (RGO) and partially
structured semiconductor electrodes (Scheme 1). By employing
restore the conductivity of pristine graphene while still main-
carbon nanotubes as scaffolding, we have shown the increa-
taining exfoliation. Although efficient, the chemical reduction
sed photocurrent generation efficiency in TiO2, CdS, and CdSe
by hydrazine can leave unwanted chemical residues and
nanostructured films.26,29-31 Employed as a scaffold, RGO
induce restacking of graphene sheets.18,20
can create a 2-D conductive support path for charge trans-
We have recently reported alternative reduction of GO
port and collection at the electrode surface. We present here
involving UV-assisted photocatalytic21 and sonolytic22 methods.
One advantage of photocatalytic reduction is the direct inter-
action between TiO2 nanoparticles and individual GO sheets. Received Date: May 28, 2010
Such an interaction facilitates delamination of RGO sheets Accepted Date: June 23, 2010
after reduction as the attached TiO2 particles keep them Published on Web Date: July 07, 2010

r 2010 American Chemical Society 2222 DOI: 10.1021/jz100728z |J. Phys. Chem. Lett. 2010, 1, 2222–2227
pubs.acs.org/JPCL

the role of RGO in improving the photoelectrochemical and in Figure 2B along with elemental analysis verifying the pre-
photocatalytic performance of TiO2 nanostructured films. sence of titanium.
Characterization of RGO-TiO2 Nanocomposites with Elec- Photoelectrochemical Behavior of Graphene-TiO2 Films.
tron Microscopy. The morphology of RGO-TiO2 nanocompo- Suspensions of known amounts of TiO2 and graphene were
sites was examined using a JEOL 2010 transmission electron deposited onto areas of ca. 2 cm2 of the optically transparent
microscope (TEM) and a Hitachi S-4500 field emission scan- electrode (OTE) by the drop cast method, as described pre-
ning electron microscope (SEM). Diluted samples of the viously.32,33 Careful deposition of the suspension results in the
RGO-TiO2 nanocomposite were placed on Toray carbon formation of homogeneous RGO-TiO2 thin films as shown in
paper for SEM characterization. Images of the nanocomposite Figure 3. While a pure TiO2 thin film displays uniform white
reveal the entwined nature of graphene sheets in TiO2 color, the increasing addition of small amounts of graphene
nanoparticle networks. gradually darkens the film color. The concentration of TiO2
Figure 1A shows an SEM image of the RGO-TiO2 nanocom- was held constant at 2 mg/cm2, while the amount of graphene
posite with some graphene edges marked by dotted lines to was varied. We probed the photocurrent generation of the
assist the reader. The SEM image in Figure 1B depicts a larger RGO-TiO2 films by employing them as photoanodes in
graphene sheet emerging from the TiO2 assembly. photoelectrochemical cells. The photocurrent response of
Transmission electron microscopy (TEM) was performed RGO-TiO2 films at different graphene loadings is shown in
in order to determine the nature of TiO2 adsorption on gra- Figure 4A.
phene sheets. Diluted samples of RGO-TiO2 nanocomposites Upon irradiation with UV light, TiO2 particles undergo
were placed on holey carbon grids under N2 flow. The rela- charge separation to yield electrons (e) and holes (h). As the
tively large holes in the TEM grid allow for unbound TiO2 holes are scavenged by OH- ions, the electrons are collected
nanoparticles to pass through, while large graphene sheets by the OTE electrode to generate photocurrent. A photocur-
with adsorbed TiO2 can be isolated. A TEM micrograph of a rent of ∼20 μA/cm2 was obtained using pristine TiO2 films.
graphene sheet obtained from the RGO-TiO2 nanocomposite The photocurrent response was prompt and reproducible
can be seen in Figure 2A. The selected area diffraction pattern during repeated on/off cycles of illumination. At concentra-
(SADP) exhibits reciprocal lattice spacings characteristic of tions below 0.05 mg/cm2 we see an increase in photocurrent
anatase TiO2 and graphene (inset). A high-resolution micro- with increasing RGO content. A significant increase (∼90%)
graph of individual TiO2 nanoparticles on graphene is shown in photocurrent generation up to ∼38 μA/cm2 was observed
with 0.05 mg/cm2 RGO RGO-TiO2 nanocomposite films. The
observed enhancement in the photocurrent represents impro-
ved charge transportation from the TiO2 through the RGO-
TiO2 nanocomposites to the collecting electrode surface. At
higher graphene loadings (>0.05 mg/cm2), light absorption by
RGO adversely affects the excitation of TiO2 nanoparticles. This
effect is also observed in the I-V curves of RGO-TiO2 nano-
composite films in Figure 4B, where cell performance begins to
decrease beyond an optimum graphene loading.
To further examine the photoelectrochemical properties of
the nanocomposites, the incident photon-to-photocurrent
efficiency (IPCE) of the films was examined. The RGO-TiO2
Figure 1. SEM images of RGO-TiO2 nanocomposites. (A) Typical films reduced by photocatalytic and chemical reduction
image of the nanocomposite to which dotted lines are added to
assist reader in distinguishing graphene sheet edges from TiO2 (hydrazine) were compared with pristine TiO2 films. The IPCE
clusters. (B) Graphene sheet emerging from TiO2 clusters. spectra were obtained using a monochromatic excitation

Figure 2. (A)TEM micrograph showing TiO2 nanoparticles adsorbed onto a graphene sheet. Inset shows the diffraction pattern taken from
sheet with spacings corresponding to graphene and anatase TiO2. (B) High-resolution image of TiO2 nanoparticles on graphene. Insets show
energy dispersive spectra (upper left) confirming the presence of TiO2 and magnification of the TiO2 lattice image (lower left).

r 2010 American Chemical Society 2223 DOI: 10.1021/jz100728z |J. Phys. Chem. Lett. 2010, 1, 2222–2227
pubs.acs.org/JPCL

source and calculated by normalizing the photocurrent to the Photocatalytic Degradation of 2,4-Dichlorophenoxyacetic
incident light energy and intensity using eq 1, Acid. One of the applications of such carbon-TiO2 composites
IPCEð%Þ ¼ 100 1240  I sc =ðP  λÞ ð1Þ is in the remediation of contaminated water. Runoff from
fields is often a source of pesticide and herbicide contamina-
Where Isc is the short-circuit photocurrent (A/cm ), P is the
2 tion in drinking water. Advanced oxidation processes are
incident light intensity (W/cm2), and λ is the wavelength often useful in degrading herbicides such as 2,4 dichlorophe-
(nm). All films (TiO2, RGO-TiO2 (in situ GO reduction), and noxyacetic acid (2,4-D).34 Early studies have highlighted the
RGO-TiO2 (hydrazine GO reduction) exhibit similar photore- ability of RGO-TiO2 nanocomposites to act as efficient
sponse patterns in the UV region, obtaining a maximum IPCE photocatalysts27 and electrocatalysts.35-37
value at a wavelength of ∼350 nm, as shown in Figure 5. Films The merits of TiO2-assisted photocatalytic degradation of
composed of TiO2 without graphene yield a maximum IPCE 2,4-D compared to other oxidation processes such as radioly-
value of 7.4%. The RGO-TiO2 nanocomposite films yield sis and sonolysis has been discussed in detail previously.34,38,39
maximum IPCE values of 13.9% for the in situ RGO (photo- Upon excitation by UV light, electron-hole pairs are genera-
catalytic reduction) and 11.4% for the hydrazine RGO (chemical ted within TiO2. As the electrons are scavenged by dissolved
reduction using hydrazine). As shown earlier,32 the majority oxygen, the holes participate in the generation of OH 3
(>90%) of the charge carriers generated following UV- radicals that oxidize 2,4-D. The holes can either directly
excitation of TiO2 undergo recombination. Quick capture oxidize the organic substrate or can transfer to the adsorbed
followed by transport of electrons is an important aspect in water to form hydroxyl radicals.
enhancing the overall photoconversion efficiency. The high- TiO2 þ hν f TiO2 ðe þ hÞ ð2Þ
er IPCE values for the RGO-TiO2 films confirm graphene's
role as an electron collector, suppressing charge recombina- TiO2 ðeÞ þ O2 þ graphene f TiO2 þ O2 - þ grapheneðeÞ
tion in photoexcited TiO2.
ð3Þ

TiO2 ðhÞ þ OH - f TiO2 þ • OH ð4Þ

TiO2 ðhÞ þ • OH þ 2; 4-D f degradation products ð5Þ


34,38,39
As confirmed in our previous mechanistic studies,
the herbicide is initially oxidized by hydroxyl radical to form
the intermediate 2,4-dichlorophenol (2,4-DCP). The oxidation
process further transforms the intermediates into quinones
and ultimately to carboxylic acids. Figure 6A shows UV-vis
absorption spectra recorded during the photocatalytic decom-
position of 2,4-D solution (aqueous, O2 saturated) at UV-
irradiated RGO-TiO2 film. The TiO2 and RGO-TiO2 films
were inserted in a photolysis cell and irradiated with UV-light
Figure 3. Photographs of TiO2 thin films with different graphene filtered through CuSO4 filter. Since the herbicide absorbs only
loadings. at wavelengths less than 300 nm (an intense and sharp peak

Figure 4. (A) Photocurrent response versus time profiles and (B) I-V characteristics of OTE/graphene-TiO2 with different amounts of
graphene. Light is illuminated from the backside of the OTE.

r 2010 American Chemical Society 2224 DOI: 10.1021/jz100728z |J. Phys. Chem. Lett. 2010, 1, 2222–2227
pubs.acs.org/JPCL

at 230 nm and a broad peak at 285 nm), we can carry out selec- RGO-TiO2 films, respectively. The 4-fold increase in the rate of
tive excitation of TiO2 in the 300-350 nm region. Further- photocatalytic degradation parallels the improvement in the
more, the absorbance at 230 nm can be monitored to obtain photoelectrochemical behavior of RGO-TiO2 films.
the concentration of remaining 2,4-D in solution. As we con- When we compare the improved photoelectrochemi-
tinue to irradiate the RGO-TiO2 film with UV light, we see a cal and photocatalytic performance of the RGO-TiO2 films,
decrease in the absorption bands of 2,4-D, indicating the a 2-fold enhancement is observed in the photocurrent res-
photocatalytic degradation of the herbicide in about 80 min. ponse, while the increase in the photocatalytic degradation
Figure 6B shows the decrease in 2,4-D concentration with rate is 4-fold. In addition to the improved charge separation,
irradiation time in the absence and presence of TiO2 and RGO- the graphene sheets can overcome the mass transfer limita-
TiO2 nanocomposite films. The blank experiment carried tion by increasing the availability of 2,4-D near the photo-
out in the absence of the catalyst film confirms that photo- catalyst surface. Graphitic-based materials have already been
lysis of 2,4-D under our experimental conditions has no noti- employed in the removal of organic pollutants.40 Earlier stu-
ceable effect. On the other hand, both TiO2 and RGO-TiO2 dies have shown that TiO2 deposited on adsorbent supports
are quite effective in degrading 2,4-D, but with a different such as activated carbon and silica can have significant influ-
rate. The degradation half-lives observed for these two sets ence in concentrating pollutant from low solution-phase con-
of experiments were 147 and 38 min for pristine TiO2 and centrations.41,42 We expect a similar 2,4-D-rich environment
RGO-TiO2 films, respectively. The decay of 2,4-D during the at the graphene/TiO2 interface to be an important contribut-
initial period of UV excitation was fitted to pseudo-first-order ing factor for achieving higher degradation rate.
kinetics (inset Figure 6B). The rate constants for 2,4-D de- The experimental results discussed here highlight the dual
gradation were 0.002 min-1 and 0.008 min-1 for TiO2 and role of graphene-based carbon nanostructures in improving
the photoelectrochemical and photocatalytic performance of
semiconductor nanoparticles. The improved photocurrent
generation is attributed to the ability of graphene sheets to
capture and transport charge to the collecting electrodes. The
faster rate of 2,4-D degradation observed with RGO-TiO2
points out the additional benefit of concentrating the organics
near the photocatalyst surface. Such beneficial aspects of
2-D nanostructures need to be further exploited to design
next-generation photocatalyst and solar cell devices. Indeed,
one such strategy of anchoring semiconductor and metal
nanoparticles on a single graphene sheet for designing a
2-D catalyst mat was demonstrated recently.28 Efforts are
underway to develop graphene-based catalysts with multi-
functional activities.

EXPERIMENTAL METHODS
Figure 5. Photocurrent action spectra of TiO2-only film compa-
red with RGO-TiO2 nanocomposite films prepared by photocata- Materials. Graphite powder (conducting grade, -325 mesh)
lytic reduction and hydrazine reduction methods. was purchased from Alfa Aesar; hydrazine hydrate and 2,4-D

Figure 6. (A) Absorption spectra of aqueous solution (O2 saturated) of 0.15 mM 2,4-D recorded following UV-excitation of RGO-TiO2 film.
(B). The decrease in 2,4-D concentration following the UV photolysis with TiO2 and RGO-TiO2 films. The blank experiments recorded in the
absence of photocatalyst are also shown. Inset shows the pseudo-first-order fit of 2,4-D decay during initial 30 min of UV excitation.

r 2010 American Chemical Society 2225 DOI: 10.1021/jz100728z |J. Phys. Chem. Lett. 2010, 1, 2222–2227
pubs.acs.org/JPCL

(98%) were purchased from Aldrich; P25 TiO2 nanoparti- Present Addresses:
§
cles were obtained from Degussa. Other reagent-grade Current address: ARC Centre of Excellence for Functional Nano-
chemicals were purchased from Fisher Scientific and used materials, School of Chemical Engineering, The University of
as supplied. New South Wales, Sydney NSW 2052, Australia.
GO was synthesized using a modified Hummers
method.43 The GO powder obtained was filtered and washed ACKNOWLEDGMENT Y.H.N. thanks the Global Center of Excel-
with 1 M HCl and deionized water. A 5.0 mg/mL TiO2 lence (GCOE) of Osaka University for the summer fellowship to
suspension was prepared by the addition of Degussa P25 conduct research at Notre Dame. K.G. thanks the NDNano Center
TiO2 powder to methanol under stirring. Dispersions of GO for awarding a NURF summer fellowship. The research described
in methanol were added to TiO2 suspensions and sonicated herein was supported by the Department of Energy, Office of Basic
Energy Sciences. This is contribution number NDRL 4856 from the
for 30 min to produce GO-TiO2 dispersions. Dispersions
Notre Dame Radiation Laboratory.
were then purged with N2 and exposed to UV-irradiation
through an aqueous CuSO4 filter (>320 nm) for 30 min using
an Oriel 450 W xenon arc lamp to obtain RGO-TiO2 REFERENCES
composites. Typical concentrations for TiO2 and GO were
(1) Iijima, S. Helical Microtubules of Graphitic Carbon. Nature
1 mg/mL and 0.5 mg/mL, respectively. The ratio of GO to 1991, 354, 56–58.
TiO2 was varied to investigate the photoelectrochemical (2) Eckert, J.-F.; Nierengarten, J.-F.; Liu, S. G.; Echegoyen, L.;
properties of the RGO-TiO2 nanocomposite. The obtained Barigelletti, F.; Armaroli, N.; Ouali, L.; Krasnikov, V.; Hadziioannou,
nanocomposites were deposited onto OTEs by dropwise G. Fullerene-Oligophenylenevinylene Hybrids: Synthesis, Elec-
addition under flowing air. Films were calcined at 673 K tronic Properties, and Incorporation in Photovoltaic Devices.
in air for 1 h and are referred to as RGO-TiO2 films. For J. Am. Chem. Soc. 2000, 122, 7467–7479.
comparison, RGO-TiO2 films were made using hydrazi- (3) Girishkumar, G.; Vinodgopal, K.; Meisel, D.; Kamat, P. V.
ne-RGO. This involved the addition of 1 M hydrazine Carbon Nanostructures in Portable Fuel Cells: Single-Walled
hydrate solution to the GO dispersion followed by a treat- Carbon Nanotube Electrodes for Methanol Oxidation and Oxy-
ment at 80 °C for 6 h. gen Reduction. J. Phys. Chem. B 2004, 108, 19960–19966.
Photoelectrochemical Measurements. Thin films of RGO- (4) Thomas, K. G.; George, M. V.; Kamat, P. V. Photoinduced Elec-
tron Transfer Processes in Fullerene-Based Donor-Acceptor
TiO2 nanocomposite material were cast and employed as
Systems. Helv. Chim. Acta 2005, 88, 1291–1308.
electrodes in photoelectrochemical and photocatalytic cells to
(5) Brown, P. R.; Takechi, K.; Kamat, P. V. Single-Walled Carbon
determine the benefit of composite material versus TiO2 Nanotube Scaffolds for Dye-Sensitized Solar Cells. J. Phys.
alone. Half cell reactions were conducted using Pt foil as the Chem. C 2008, 112, 4776–4782.
counter electrode and 1 M KOH as the electrolyte. An Oriel (6) Imahori, H.; Umeyama, T. Donor-Acceptor Nanoarchitec-
450 W Xe lamp with a CuSO4 filter was employed as the light ture on Semiconducting Electrodes for Solar Energy Conver-
source. The intensity of the incident light near the electrode sion. J. Phys. Chem. C 2009, 113, 9029–9039.
surface was 100 mW/cm2. The photoelectrochemical proper- (7) Rao, C. N. R.; Sood, A. K.; Voggu, R.; Subrahmanyam, K. S.
ties were measured with a Keithley model 617 electrometer. Some Novel Attributes of Graphene. J. Phys. Chem. Lett. 2010,
The incident photon to charge carrier generation efficiency 1, 572–580.
(IPCE) was recorded using a Princeton Applied Research (8) Geim, A. K.; Novoselov, K. S. The Rise of Graphene. Nat. Mater.
model PARSTAT 2263 potentiostat. 2007, 6, 183–191.
Photocatalytic Degradation of 2,4-D. Photocatalytic decom- (9) Jasuja, K.; Berry, V. Implantation and Growth of Dendritic Gold
Nanostructures on Graphene Derivatives: Electrical Property Tailo-
position of 2,4-D involved the addition of 3 mL of 0.15 mM
ring and Raman Enhancement. ACS Nano 2009, 3, 2358–2366.
2,4-D into a cuvette containing an RGO-TiO2 film. An Oriel
(10) Robinson, J. T.; Perkins, F. K.; Snow, E. S.; Wei, Z. Q.; Sheehan,
Apex Illuminator with a glass filter that removed light with P. E. Reduced Graphene Oxide Molecular Sensors. Nano Lett.
wavelengths lower than 320 nm was used as the light source. 2008, 8, 3137–3140.
The system was purged with O2 gas throughout the experi- (11) Niyogi, S.; Bekyarova, E.; Itkis, M. E.; McWilliams, J. L.;
ment. The thin film sample was immersed in the 2,4-D Hamon, M. A.; Haddon, R. C. Solution Properties of Graphite
solution in dark conditions for 30 min prior to the irradiation and Graphene. J. Am. Chem. Soc. 2006, 128, 7720–7721.
to allow the equilibration between analyte and catalyst parti- (12) Bekyarova, E.; Itkis, M. E.; Ramesh, P.; Haddon, R. C. Chemical
cles. Degradation of 2,4-D was monitored using a Cary 50 Bio Approach to the Realization of Electronic Devices in Epitaxial
UV-vis spectrophotometer. Graphene. Phys. Status Solidi RRL 2009, 3, 184–186.
(13) Bekyarova, E.; Itkis, M. E.; Ramesh, P.; Berger, C.; Sprinkle,
M.; de Heer, W. A.; Haddon, R. C. Chemical Modification of
SUPPORTING INFORMATION AVAILABLE Absorption spec- Epitaxial Graphene: Spontaneous Grafting of Aryl Groups.
tra of RGO. This material is available free of charge via the Internet at J. Am. Chem. Soc. 2009, 131, 1336–1337.
http://pubs.acs.org. (14) Muszynski, R.; Seger, B.; Kamat, P. Decorating Graphene
Sheets with Gold Nanoparticles. J. Phys. Chem. C 2008, 112,
AUTHOR INFORMATION 5263–5266.
(15) Novoselov, K. S.; Geim, A. K.; Morozov, S. V.; Jiang, D.; Zhang,
Corresponding Author: Y.; Dubonos, S. V.; Grigorieva, I. V.; Firsov, A. A. Electric Field
*To whom correspondence should be addressed. E-mail: pkamat@ Effect in Atomically Thin Carbon Films. Science 2004, 306,
nd.edu. 666–669.

r 2010 American Chemical Society 2226 DOI: 10.1021/jz100728z |J. Phys. Chem. Lett. 2010, 1, 2222–2227
pubs.acs.org/JPCL

(16) Hersam, M.; Green, A. Emerging Methods for Producing (32) Vinodgopal, K.; Hotchandani, S.; Kamat, P. V. Electrochemi-
Monodisperse Graphene Dispersions. J. Phys. Chem. Lett. cally Assisted Photocatalysis. TiO2 Particulate Film Electro-
2010, 1, 544–549. des for Photocatalytic Degradation of 4-Chlorophenol.
(17) McAllister, M. J.; Li, J. L.; Adamson, D. H.; Schniepp, H. C.; J. Phys. Chem. 1993, 97, 9040–4.
Abdala, A. A.; Liu, J.; Herrera-Alonso, M.; Milius, D. L.; Car, R.; (33) Drew, K.; Girishkumar, G.; Vinodgopal, K.; Kamat, P. V.
Prud'homme, R. K.; Aksay, I. A. Single Sheet Functionalized Boosting the Fuel Cell Performance with a Semiconductor
Graphene by Oxidation and Thermal Expansion of Graphite. Photocatalyst. TiO2/Pt-Ru Hybrid Catalyst for Methanol Oxi-
Chem. Mater. 2007, 19, 4396–4404. dation. J. Phys. Chem. B 2005, 109, 11851–11857.
(18) Stankovich, S.; Dikin, D. A.; Piner, R. D.; Kohlhaas, K. A.; (34) Peller, J.; Wiest, O.; Kamat, P. V. Hydroxyl Radical's Role in the
Kleinhammes, A.; Jia, Y.; Wu, Y.; Nguyen, S. T.; Ruoff, R. S. Remediation of a Common Herbicide, 2,4-Dichlorophenoxy-
Synthesis of Graphene-Based Nanosheets via Chemical Re- acetic Acid (2,4-D) (Feature Article). J. Phys. Chem. A 2004,
duction of Exfoliated Graphite Oxide. Carbon 2007, 45, 108, 10925–10933.
1558–1565. (35) Tsuji, M.; Kubokawa, M.; Yano, R.; Miyamae, N.; Tsuji, T.; Jun,
(19) Li, D.; Muller, M. B.; Gilje, S.; Kaner, R. B.; Wallace, G. G. M.-S.; Hong, S.; Lim, S.; Yoon, S.-H.; Mochida, I. Fast Prepa-
Processable Aqueous Dispersions of Graphene Nanosheets. ration of PtRu Catalysts Supported on Carbon Nanofibers by
Nature Nanotechnol. 2008, 3, 101–105. the Microwave-Polyol Method and Their Application to Fuel
(20) Stankovich, S.; Dikin, D. A.; Dommett, G. H. B.; Kohlhaas, Cells. Langmuir 2006, 23, 387–390.
K. M.; Zimney, E. J.; Stach, E. A.; Piner, R. D.; Nguyen, S. T.; (36) Seger, B.; Kamat, P. V. Electrocatalytically Active Graphene-
Ruoff, R. S. Graphene-Based Composite Materials. Nature Platinum Nanocomposites. Role of 2-D Carbon Support in
2006, 442, 282–286. PEM Fuel Cells. J. Phys. Chem. C 2009, 113, 7990–7995.
(21) Williams, G.; Seger, B.; Kamat, P. V. TiO2-Graphene Nano- (37) Yoo, E.; Okata, T.; Akita, T.; Kohyama, M.; Nakamura, J.;
composites. UV-Assisted Photocatalytic Reduction of Graphene Honma, I. Enhanced Electrocatalytic Activity of Pt Subnano-
Oxide. ACS Nano 2008, 2, 1487–1491. clusters on Graphene Nanosheet Surface. Nano Lett. 2009, 9,
(22) Vinodgopal, K.; Neppolian, B.; Lightcap, I. V.; Grieser, F.; 2255–2259.
Ashokkumar, M.; Kamat, P. V. Sonolytic Design of Graphene (38) Peller, J.; Wiest, O.; Kamat, P. V. Mechanism of Hydroxyl
Au Nanocomposites. Simultaneous and Sequential Reduction Radical-Induced Breakdown of the Herbicide 2,4-Dichloro-
of Graphene Oxide and Au(III). J. Phys. Chem. Lett. 2010, phenoxyacetic Acid. Chem.;Eur. J. 2003, 9, 5379–5387.
1987–1993. (39) Peller, J.; Wiest, O.; Kamat, P. V. Synergy Of Combining
(23) Kamat, P. V. Graphene Based Nanoarchitectures. Anchoring Sonolysis and Photocatalysis in the Degradation And Minera-
Semiconductor and Metal Nanoparticles on a Two-Dimensional lization of Chlorinated Aromatic Compounds. Environ. Sci.
Carbon Support. J. Phys. Chem. Lett. 2010, 1, 520–527. Technol. 2003, 37, 1926–1932.
(24) Lambert, T. N.; Chavez, C. A.; Hernandez-Sanchez, B.; Lu, P.; (40) Park, C.; Engel, E. S.; Crowe, A.; Gilbert, T. R.; Rodriguez, N. M.
Bell, N. S.; Ambrosini, A.; Friedman, T.; Boyle, T. J.; Wheeler, Use of Carbon Nanofibers in the Removal of Organic Solvents
D. R.; Huber, D. L. Synthesis and Characterization of Titania- from Water. Langmuir 2000, 16, 8050–8056.
Graphene Nanocomposites. J. Phys. Chem. C 2009, 113, (41) Torimoto, T.; Ito, S.; Kuwabata, S.; Yoneyama, H. Effects of
19812–19823. Adsorbents Used as Supports for Titanium Dioxide Loading
on Photocatalytic Degradation of Propyzamide. Environ. Sci.
(25) Kim, S. R.; Parvez, M. K.; Chhowalla, M. UV-Reduction of
Technol. 1996, 30, 1275–1281.
Graphene Oxide and Its Application as an Interfacial Layer to
(42) Anderson, C.; Bard, A. J. An improved Photocatalyst of TiO2/
Reduce the Back-Transport Reactions in Dye-Sensitized Solar
SiO2 Prepared by Sol-Gel Synthesis. J. Phys. Chem. 1995, 99,
Cells. Chem. Phys. Lett. 2009, 483, 124–127.
9882–9885.
(26) Yang, N.; Zhai, J.; Wang, D.; Chen, Y.; Jiang, L. Two-Dimen-
(43) Hummers, W. S.; Offeman, R. E. Preparation of Graphitic
sional Graphene Bridges Enhanced Photoinduced Charge
Oxide. J. Am. Chem. Soc. 1958, 80, 1339–1339.
Transport in Dye-Sensitized Solar Cells. ACS Nano 2010, 4,
887–894.
(27) Zhang, H.; Lv, X. J.; Li, Y. M.; Wang, Y.; Li, J. H. P25-Graphene
Composite as a High Performance Photocatalyst. ACS Nano
2010, 4, 380–386.
(28) Lightcap, I. V.; Kosel, T. H.; Kamat, P. V. Anchoring Semicon-
ductor and Metal Nanoparticles on a Two-Dimensional Cata-
lyst Mat. Storing and Shuttling Electrons with Reduced
Graphene Oxide. Nano Lett. 2010, 10, 577–583.
(29) Kongkanand, A.; Domínguez, R. M.; Kamat, P. V. Single Wall
Carbon Nanotube Scaffolds for Photoelectrochemical Solar
Cells. Capture and Transport of Photogenerated Electrons.
Nano Lett. 2007, 7, 676–680.
(30) Vietmeyer, F.; Seger, B.; Kamat, P. V. Anchoring ZnO Particles
on Functionalized Single Wall Carbon Nanotubes. Excited
State Interactions and Charge Collection. Adv. Mater. 2007,
19, 2935–2940.
(31) Wang, M. F.; Kumar, S.; Lee, A.; Felorzabihi, N.; Shen, L.; Zhao,
F.; Froimowicz, P.; Scholes, G. D.; Winnik, M. A. Nanoscale
Co-organization of Quantum Dots and Conjugated Polymers
Using Polymeric Micelles as Templates. J. Am. Chem. Soc.
2008, 130, 9481–9491.

r 2010 American Chemical Society 2227 DOI: 10.1021/jz100728z |J. Phys. Chem. Lett. 2010, 1, 2222–2227

You might also like