Vous êtes sur la page 1sur 9

CARBON

7 0 ( 2 0 1 4 ) 2 1 5 2 2 3

Available at www.sciencedirect.com

ScienceDirect
journal homepage: www.elsevier.com/locate/carbon

Unusual functionalization of reduced graphene


oxide for excellent chemical surface-enhanced
Raman scattering by coupling with ZnO
Ruey-Chi Wang
a
b

a,*
,

Ya-Chun Chen a, Shu-Jen Chen

b,*
,

Yu-Ming Chang

Department of Chemical and Materials Engineering, National University of Kaohsiung, Kaohsiung 81148, Taiwan
Department of Chemical and Materials Engineering, National Kaohsiung University of Applied Sciences, Kaohsiung 80778, Taiwan

A R T I C L E I N F O

A B S T R A C T

Article history:

A low-cost noble metal-free substrate comprised of annealed graphene oxide (GO)/ZnO

Received 4 September 2013

composites is prepared to demonstrate an efficient chemical surface-enhanced Raman

Accepted 31 December 2013

scattering effect. A high enhancement factor of about 104, better than those reported for

Available online 10 January 2014

reduced GO (rGO)/Au and GO/Ag composites, is mainly attributed to the unusually


abundant oxygen-containing groups generated on surface of rGO by coupling with ZnO
nanoparticles at moderate temperature. High-resolution transmission electron microscopy,
X-ray diffraction, and Fourier transform infrared spectroscopy are employed to examine
the evolution of ZnO as well as reduction and functionalization of GO after different heat
treatments.
 2014 Elsevier Ltd. All rights reserved.

1.

Introduction

Graphene is a novel carbon nanomaterial of a two-dimensional


conjugated structure, offering unique electronic properties,
and has been intensively studied for many applications, such
as electronic devices [1], solar cells [2], photocatalysts [3], and
sensors [4]. There has been growing interest in graphene-based
surface-enhanced Raman scattering (SERS) substrates for biochemical sensing due to the high surface area and great electrical conductivity of graphene. For example, Au nanoparticle
(NP)graphene oxide (GO) [5] and Aggraphene [6] composites
were recently synthesized for chemical sensing and cell imaging by taking advantage of the electromagnetic enhancement
of plasmonic metal NPs. In addition, an ultraviolet/ozone treatment [7] has been used to oxidize graphene and demonstrated
an efficient improvement in SERS compared with pristine
graphene, due to charge transfer-induced chemical enhancement. However, these high-performance SERS substrates usu-

ally need expensive noble metals to enhance plasmonic


resonance or toxic O3 gas to activate graphene. It thus remains
a challenge to fabricate low-cost graphene-based SERS substrates with excellent enhancement performance.
In this work, an excellent SERS substrate comprised of annealed low-cost reduced GO (rGO)/ZnO composites is firstly
demonstrated. The morphologies, structures, and functional
groups of the rGO/ZnO composites after different heat treatments show significant variations, indicative of strong reactions between rGO and ZnO. The high SERS enhancement
factor (EF) of about 104, even better than those reported for
Ag/GO [8], Au/GO [5], Au/rGO [9], and Au/graphene [10]
composites, is mainly attributed to the abundant oxygencontaining groups unusually generated on the surface of
rGO by coupling with ZnO NPs at moderate temperature, as
seen in the Fourier transform infrared spectroscopy (FTIR)
spectra. This work demonstrates the preparation of low-cost
SERS substrates with excellent performance by reactions of

* Corresponding authors: Fax: +886 7 5919277.


E-mail addresses: rcwang@nuk.edu.tw (R.-C. Wang), biochen@kuas.edu.tw (S.-J. Chen).
0008-6223/$ - see front matter  2014 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.carbon.2013.12.110

216

CARBON

7 0 ( 2 0 1 4 ) 2 1 5 2 2 3

rGO with ZnO, which have good potential for bio-chemical


sensing applications.

2.

Experiment

2.1.

Preparation of rGO/ZnO composites

GO flakes were prepared following Hummers method [11].


The GO dispersion was prepared by dispersing 0.5 g GO flakes
in 500 ml distilled water using centrifugal separation at a
speed of 8000 rpm for 40 min. GO dispersion of 0.5 ml was
then dropped onto a ZnO nanorod (NR)/Si substrate, synthesized by a hydrothermal route [12], and dried at 90 C for
5 min. The dropping/drying processes were repeated for 10
cycles to obtain a pristine GO/ZnO composite. The prepared
GO/ZnO composites were then annealed at temperatures of
100400 C for 1 h in air and cooled down to room temperature (RT) to transform them into various rGO/ZnO composites.

2.2.

Characterization and SERS measurements

The as-prepared nanostructures were examined by fieldemission scanning electron microscopy (FE-SEM) with a
Hitachi 4800, X-ray diffraction (XRD) with a D/MAX-2500,
field-emission transmission electron microscopy (FE-TEM)
and energy-dispersive spectrometry (EDS) with a Philips Tecnai G2 F20 FEG-TEM for morphology, crystallography, microstructure, and composition characterization, respectively.
The functional groups of the GO and rGO/ZnO samples were
characterized by FTIR with a Perkin Elmer Spectrum 100.
The thermogravimetric analysis (TGA) was performed with
a TA, SDT-A600 thermogravimetric analyzer at a heating rate
of 15 C/min under nitrogen. About 15 mg GO and 0.24 mg
rGO/ZnO powders were taken for measurement. The electrical properties were measured by depositing two Au electrodes
on the rGO/ZnO composites and the current was measured at
a fixed bias of 1 V. The gap between the electrodes was about
2 mm.
SERS measurements were performed with a Jobin Yvon
LabRAM HR800 Raman spectrometer equipped with a
532 nm HeCd laser as the excitation sources. The laser power

at the sample position was about 70 lW. Each of the samples


for the SERS measurements was prepared by casting 10 lL of
Rhodamine 6G (R6G) in ethanol, dropping this on a 5 5 mm2
substrate, and then allowing the solvent to evaporate. The
size of the area of molecules participating in the scattering
process is essentially determined by the diameter of the laser
spot (1 lm). To evaluate the EF values of SERS, the Raman
intensity of R6G was also acquired on a highly resistive flat
Si wafer as the reference for comparison, which exhibits a
negligible SERS effect.

3.

Results and discussion

3.1.
Morphologies, microstructures, and crystal structures
of the rGO/ZnO composites
Based on the SEM results, the morphologies of the GO/ZnO
composites show significant variations after annealing at different temperatures. Fig. 1(a) and (b) show the SEM images of
the ZnO NRs and GO/ZnO composites before annealing,
respectively. The GO flakes are uniformly attached to the surface of the NRs. After annealing at 150200 C for 1 h, the GO
flakes were transformed into film-like structures, and small
dots of lighter contrast were generated on the films, as shown
in Fig. 1(c) and (d). As the annealing temperature was increased, the dots gradually enlarged and aggregated, as
shown in Fig. 1(e) and (g).
Fig. 2(a) and (d) show the TEM characterizations of the
GO/ZnO composites without annealing. As seen in the
bright-field (BF) image of Fig. 2(a), the GO flakes were
connected to a ZnO NR. Notably, the surfaces of the NR were
relatively rougher compared with the original ZnO NR (not
shown). Fig. 2(b) and (c) are the electron diffraction patterns
(DPs) of the ZnO NR and GO flakes marked in Fig. 2(a). The
ZnO NR remains a single-crystalline wurtzite structure growing along the c-axis. The DP of the GO flakes shows multi-ring
characteristics, indicative of poly-crystalline structures.
Fig. 2(d) is the dark-field (DF) image of the GO/ZnO composites. It can be seen that there are a few NPs of light contrast
embedded in the GO flakes, which were subsequently characterized as ZnO NPs.

Fig. 1 SEM images of as-prepared (a) ZnO NRs and (b) GO/ZnO composites; GO/ZnO composites annealed at (c) 150 C, (d)
200 C, (e) 250 C, (f) 300 C, and (g) 400 C for 1 h.

CARBON

7 0 (2 0 1 4) 2 1 522 3

217

Fig. 2 TEM characterization of pristine GO/ZnO composites. (a) BF image; (b, c) electron DPs from area b and c in (a) for the
ZnO NR and GO flakes, respectively; (d) DF image of the composites. (A color version of this figure can be viewed online.)

Fig. 3 TEM characterization of rGO/ZnO composites after annealing at 200 C. (a) BF image with inset showing electron DP of
a ZnO NR; (b) electron DP of rGO flakes; (c) DF image and (d) electron DP of rGO flakes and the generated ZnO NPs. (A color
version of this figure can be viewed online.)

To realize the effects of annealing, the GO/ZnO composites


after annealing at 200400 C were characterized by TEM, with
the results shown in Figs. 3 and 4, respectively. Fig. 3(a) shows
the BF image of the composites with the inset showing the
electron DP of the ZnO NR. It can be seen that the single-crystalline structure of the ZnO NR did not change. Fig. 3(b) shows
the DP of the GO flakes marked in Fig. 3(a). The d-spacing of
, indicative of the
the (0 0 2) plane was calculated to be 3.98 A
reduction in GO after annealing. Fig. 3(c) shows a DF image
of the rGO flakes, in which even more NPs with lighter contrast were generated. Fig. 3(d) shows the corresponding electron DP, which has the diffraction rings of ZnO in addition to
rGO. The results shown in Figs. 2 and 3 indicate that the ZnO
NPs were generated after the drying process, and then grew
gradually after annealing.
Fig. 4(a) and (b) show the TEM BF and high-resolution (HR)
images of the ZnO NR/ZnO NP/rGO composites after annealing
at 400 C. The surface of the ZnO NR has a relatively rough morphology and the single-crystalline characteristic of the ZnO NP
can be clearly seen. Fig. 4(c) shows the HR image of the rGO
flake marked in Fig. 4(a) with the d-spacing of (0 0 2) plane
. Apparently, the d-spacing of (0 0 2) plane
being about 3.6 A
gradually decreases with the increase of annealing temperature. Fig. 4(d) and (e) are the electron DP of the rGO flakes
and EDS characterization of the ZnO NPs/rGO composites,
respectively. The ZnO NPs were dispersed on the rGO flakes,
resulting in a high percentage of Zn and O in rGO.
Fig. 5(a) and (b) show the XRD patterns of the GO and
GO/ZnO composites, respectively, before and after annealing
at different temperatures. Fig. 5(a) shows that the d-spacing

of the (0 0 2) plane of GO decreases significantly from 8.3 to


, indicative of significant reduction, after annealing at
3.6 A
200 C. This chemical reduction then gradually increases with
the annealing temperature above 200 C. From the XRD patterns of GO/ZnO composites, all the samples show the
(0 0 0 2) preferred orientation of wurtzite ZnO, with other
ZnO weak peaks appearing after the addition of GO. Apparently, there are some ZnO crystals generated with orientations different from that of the ZnO NR, which is consistent
with the formation of ZnO NPs characterized by TEM. In
addition, due to the small amount of GO in the composites
compared with ZnO, the peaks of GO and rGO were not clearly
seen in the XRD patterns.

3.2.

Growth mechanisms of the ZnO NP/rGO composites

The results presented above indicate that ZnO NPs are generated on the rGO flakes, and their growth mechanisms can be
explained as follows. When the GO aqueous solutions were
dropped onto the ZnO NRs, the ZnO were dissolved in the
solution by reacting with the carboxyl groups (ACOOH) on
the surface of GO, as in the following formula [13]:
ZnO 2R-COOH ! R-COO 2 Zn2 H2 O

The surfaces of the ZnO NRs were thus etched, and so


exhibited a relatively rough morphology in the TEM characterization. After the drying process, ZnO were precipitated
on the surface of GO in the form of NPs. Consequently, there
are other diffraction peaks in addition to the (0 0 0 2) of ZnO
generated after the drying process. As the annealing

218

CARBON

7 0 ( 2 0 1 4 ) 2 1 5 2 2 3

Fig. 4 TEM characterization of rGO/ZnO composites after annealing at 400 C. (a) BF image; HR images of the (b) ZnO NP and
(c) rGO flake marked in (a); (d) electron DP of rGO flakes; (e) BF image and EDS characterization of ZnO NPs/rGO composites. (A
color version of this figure can be viewed online.)

Fig. 5 XRD patterns of (a) GO flakes, and (b) GOZnO composites before and after annealing at different temperatures. (A color
version of this figure can be viewed online.)

temperature is increased, the growth and aggregation of ZnO


NPs enhances even more, which can be observed from the
SEM and TEM images. In addition, the oxygen-containing
groups between the graphene layers were dissociated from
the GO as the temperature is increased, which leads to a gradual decrease in the d-spacing of the (0 0 0 2) plane with a rising annealing temperature, consistent with the electron DP
results obtained by TEM. Fig. 6 shows the proposed growth
mechanisms.

3.3.
Variation of functional groups, thermogravimetric
analysis, and proposed mechanisms
Fig. 7(a) shows the FTIR spectra of the GO/ZnO composites
with different heat treatments. The sample without annealing
exhibits signals corresponding to C@O, OH, COOH, C@C, and
ZnO. The oxygen-containing groups, especially C@O, significantly increase as the annealing temperature rises from
RT to 200 C, and abruptly decrease as it increases to
250400 C. The increase in the functional groups at
150200 C is abnormal compared with common GO samples,

where the amount of carbonyl and carboxyl groups slightly


decreases as temperature is increased to 200 C (see Fig. S1
in the Supplementary materials). According to the electron
DP results obtained by TEM, the d-spacing of GO in the
GO/ZnO composites shows a significant decrease after annealing at 200 C, so the newly generated abundant functional
groups should be on the surface rather than the interlayers
of GO, and ZnO may have an important role in the generation
of these. To investigate the reactions that occur during heating, TGA was performed with the results shown in Fig. 7(b).
As expected, the weight of the GO samples gradually decreases as temperature rises, due to the removal of adsorbed
water and hydroxyl groups. However, when the temperature
is higher than 200 C, a significant decrease in weight happens, which is attributed not only to the detachment of H2O
and OH, but also to the carbonyl and carboxyl groups in the
GO flakes. On the other hand, the weight of GO/ZnO composites shows a slight increase from about 100 to 200 C.
Considering the simple GO sample without ZnO, the
weight loss of GO and the disappearance of functional groups
(Fig. S1) during heat treatment might be attributed to the

CARBON

7 0 (2 0 1 4) 2 1 522 3

219

Fig. 6 Proposed growth mechanisms of the various rGOZnO composites by annealing. (A color version of this figure can be
viewed online.)

Fig. 7 (a) FTIR spectra of GO/ZnO composites before and after annealing at different temperatures; (b) TGA thermograms of
GO flakes and GO/ZnO composites. (A color version of this figure can be viewed online.)

attack of oxygen-containing free radicals. It is well known


that there are intrinsic oxygen-containing functional groups
in GO flakes prepared by Hummers method, such as epoxides,
ethers, hydroxyls, carboxyls, and ketones. According to the
thermal reduction mechanism proposed by Acik et al. [14],
the trapped water on GO could initiate the formation of hydroxyl free radicals under heat treatment. The resulting free

radicals then undergo radical propagation reactions and attack the hydroxyls, carboxyls, and epoxides. As shown in
Fig. 8, these radical reactions are as follows: (a) attacking
hydroxyls and resulting in the formation of carbonyl radicals,
(b) attacking carboxyls and resulting in the formation of carboxylate radicals and the subsequent decomposition of carboxylates, (c) attacking epoxides in the presence of water,

220

CARBON

7 0 ( 2 0 1 4 ) 2 1 5 2 2 3

Decomposition of GO
Formation of free radicals
H2O Heat
Heat
2 H2O
O2

HO

+ H

4 HO

Attack on (a) hydroxyls, (b) decarboxylation, and (c) ring-opening of epoxides by free radicals
(a) Attack on hydroxyls
ketone
O
epoxide

O
O

HO

HO

hydroxyl
OH
COOH

ether

HO

HO
HOOC

O
OH
carboxyl radical
epoxide
(c) Ring-opening of epoxides

H2O

C
O

O
O

OH

+ CO/CO2

H2O
heat

HO

HO

H2O
heat

HO

COOH

OO
O

+ CO/CO2

(b) Decarboxylation
carboxyl

HO

C
carbonyl radical

HOOC

H2O
heat

C
H2O
heat

HO
C

In the presence of ZnO


O2

e-

+ CO/CO2

O
OH

COOH

e-

O
HO

ZnO

heat

HO

HOOC

COOH

RT to 200oC

200 to 400oC
O

carboxyl radical

C
+ CO/CO2

O
HO
HOOC

heat

OH
HO

+ CO/CO2

carbonyl radical

OO

VB

O
carbonyl radical

O2-

heat
CB

+ CO/CO2

O
carbonyl radical

H2OOHe

accumulation of carbonyl and carboxyl radicals

GO

Fig. 8 Schematic of the decomposition of GO in the presence of free radicals and proposed functionalizing mechanisms of
rGO by coupling with ZnO at moderate temperature. (A color version of this figure can be viewed online.)

and resulting in the ring-opening of epoxides. The radical


reactions further lead to the formation of other intermediates
and free radicals, such as peroxides, hydroperoxyl and hydronium radicals [14]. The radical reactions are finally terminated with the depletion of the generated free radicals
(hydroxyl, carbonyl and carboxylate radicals), which can then
react and lead to the formation of CO or CO2. The disappearance of functional groups in GO may be related to the binding
energies of oxygen-containing groups, which are in the
order of ketone species (C@O, 8.0 eV/O) > carboxyl (COOH,
5.8 eV/COOH) > epoxide (CAOAC, 3.1 eV/O) > hydroxyl (CAOH,
1.5 eV/OH) [15]. Therefore, it can be postulated that the
decomposition of carbonyls and carboxyls will proceed at a
higher temperature compared with those for epoxides and
hydroxyls. In addition, the FTIR spectrum demonstrated by
Acik et al. [14] indicates that most of the functional groups began to decompose from 100 C; however, the carbonyl groups
(C@O) increase under mild thermal treatment (below 200 C),
and are removed at temperature higher than 200 C. This

might explain why there is an abrupt loss of weight in the


ZnO-free GO sample at 200 C in our TGA data.
However, the changes in the functional groups of ZnO/GO
during heating may follow a different mechanism. When
ZnO/GO was heated from RT to 150200 C, the thermal energy might excite the electron transition in ZnO from defect
levels. The excited electrons then reacted with H2O and O2,
and led to the generation of OH and O
2 , and the resulting free
radicals attacked the functional groups on the GO. With the
aid of ZnO, more initial free radicals were generated in radical
propagation reactions during heating from RT to 200 C,
resulting in the acceleration of the reactions between the free
radicals and GO. The carbonyl and carboxyl intermediates
originated from the free radical reactions accumulated rapidly, resulting in an increase in the C@O and COOH groups
in the FTIR spectra from RT to 200 C. At temperatures higher
than 200 C, the carbonyl and carboxyl intermediates began to
decompose, and thus a decrease in the C@O and COOH
groups was observed. In addition, the slight increase in

CARBON

7 0 (2 0 1 4) 2 1 522 3

221

Fig. 9 Raman spectra of R6G molecules on (a) various substrates, and (b) GO/ZnO composites with different annealing
treatments; (c) enhancement factors of various SERS substrates; (d) electrical measurements of rGO/ZnO composites annealed
at different temperatures. (A color version of this figure can be viewed online.)

ZnO/GO weight from 100 to 200 C might be attributed to the


condensation of water in the interlayers of GO following the
reactions (OH OHH2 O H) [14] according to the strong
peak at around 3400 cm1 in the FTIR spectrum.

3.4.
Improved
applications

SERS

effects

for

chemical

sensing

Raman detection of the functionalized rGO/ZnO composites


reveals unusually strong SERS effects. Fig. 9(a) shows the Raman spectra of R6G molecules absorbed on the ZnO NRs, GO,
and GO/ZnO composites with annealing treatments. To evaluate the EF values of SERS, a highly resistive flat Si wafer,
showing typical of a negligible SERS effect, was also used as
a control. All the peaks present in the spectra correspond to
the vibration modes of R6G [16]. Due to the pretty low SERS effect of the Si substrate and ZnO NRs, no R6G signals were detected with a 106 M solution. The R6G signals can only be

detected by increasing the concentration to 103 M. As shown


in Fig. 9 (a), the Raman peaks of R6G on Si and ZnO NRs are
pretty weak even with a higher R6G concentration of
103 M. Notably, the annealed rGO/ZnO sample exhibits a
much higher intensity even for a lower R6G concentration of
106 M, only one thousandth of that on the Si and ZnO samples. Fig. 9(b) shows the Raman spectra of the R6G molecules
absorbed on rGO/ZnO composites with different annealing
treatments. The rGO/ZnO sample treated at 200 C has the
highest intensity. Fig. 9(c) shows the EF values with standard
deviations for the vibration mode at 1190 cm1 of various
samples calculated following the standard procedures (see
Supplementary materials) [17]. The rGO/ZnO composites
show significantly higher EF values than the ZnO NR and
rGO samples. Particularly, the EF value increases with the
annealing temperature from RT to 200 C, resulting in a maximum value of about 104, but this then decreases as the
annealing temperature continues to rise. On the other hand,

222

CARBON

7 0 ( 2 0 1 4 ) 2 1 5 2 2 3

the EF values of the rGO samples annealed at different temperatures do not show obvious difference. In addition to the
peak at 1190 cm1, other main vibrational modes of R6G were
also compared (see Fig. S2 and Table S1 of the Supplementary
materials), and the EF values for other peaks are also around
104. Since the electrical conductivities of substrates are critical for physical SERS effects, the electrical properties of the
rGO/ZnO samples with different annealing temperatures
were measured and shown in Fig. 9(d). At a fixed bias of 1 V,
the current increases with the annealing temperature from
RT to 400 C. It appears that the physical effects are not
responsible for the strong SERS effects of the rGO/ZnO sample
annealed at 200 C. On carefully checking the peak positions
of R6G on different substrates, including Si, ZnO NRs, and
GO/ZnO composites annealed at different temperatures, we
found that there is only negligible shift (see Table S2 of the
Supplementary materials). Therefore, the enhancement
mechanism should not be the effect of charge-transfer or
chemical bonding between R6G and functional groups on
rGO [7,18] because the above mechanisms will affect the geometrical or electronic structure of the adsorbed molecules,
leading to significant spectral shifts [19]. Therefore, the
enhancement may be attributed to the local electric field on
molecules generated by the highly electronegative oxygencontaining functional groups under the laser excitation from
their strong local dipole moments [20]. In the work, we
adopted the dropping [21,22] rather than soaking [20,23]
method to prepare SERS samples, so the number of molecules
per unit area should be identical irrespective of the morphologies of the nanostructures changed with different annealing
temperatures because the samples were all fully covered by
the drops. Although the signal of R6G on the annealed rGO/
ZnO composites is not pretty high, the EF value is estimated
by comparing the respective signal intensity per molecule under the same parameters of SERS measurements. Because the
number of R6G molecules on the rGO/ZnO samples is only
one thousandth of that on the Si substrate by considering
the difference in concentration, the high EF value of 104 is derived. By testing some R6G solutions with different concentrations, we can know the lowest R6G concentration on the
annealed ZnO/rGO sample that all the main Raman peaks still
can be detected is about 108 M (see Fig. S3 in the Supplementary materials). This work demonstrates that rGO could be
effectively functionalized by coupling with ZnO at moderate
temperature, which leads to excellent chemical SERS effects
for bio-chemical sensing applications.

4.

Conclusion

This work presents a novel and efficient SERS substrate based


on low-cost rGO/ZnO composites. The high EF value, about
104 for R6G, is mainly attributed to the unusually abundant
oxygen-containing groups generated on rGO by coupling with
ZnO at a moderate temperature. The SEM, TEM, and XRD results show numerous ZnO NPs were generated on the surface
of rGO, and the d-spacing of GO gradually decreased after the
annealing treatments. Unusually, the functional groups of
rGO/ZnO increased significantly, in contrast to the behavior
of GO powders, after annealing at 150200 C, indicative of

the functionalizing of the rGO surface, even if the d-spacing


decreased. This evolution could be attributed to the coupling
of rGO with ZnO, and the generated abundant functional
groups enhanced the chemical SERS effects due to the local
electric field created under the laser excitation by their strong
local dipole moments. This work thus demonstrates an efficient method to functionalize rGO for excellent SERS effects,
which is promising for low-cost bio-chemical sensing
applications.

Acknowledgement
This work was supported by the National Science Council of
Taiwan under the Grant (NSC100-2221-E-390-009-MY3).

Appendix A. Supplementary data


Supplementary data associated with this article can be found,
in the online version, at http://dx.doi.org/10.1016/j.carbon.
2013.12.110.

R E F E R E N C E S

[1] Chen YL, Hu ZA, Chang YQ, Wang HW, Zhang ZY, Yang YY,
et al. Zinc oxide/reduced graphene oxide composites and
electrochemical capacitance enhanced by homogeneous
incorporation of reduced graphene oxide sheets in zinc oxide
matrix. J Phys Chem C 2011;115:256371.
[2] Zhang DW, Li XD, Li HB, Chen S, Sun Z, Yin XJ, et al.
Graphene-based counter electrode for dye-sensitized solar
cells. Carbon 2011;49:53828.
[3] Zhang H, Lv XJ, Li YM, Wang Y, Li JH. P25-graphene composite
as a high performance photocatalyst. ACS Nano 2010;4:3806.
[4] Kaniyoor A, Jafri RI, Arockiadoss T, Ramaprabhu S.
Nanostructured Pt decorated graphene and multi walled
carbon nanotube based room temperature hydrogen gas
sensor. Nanoscale 2009;1:3826.
[5] Hu CF, Rong JH, Cui JH, Yang YH, Yang LF, Wang YL, et al.
Fabrication of a graphene oxidegold nanorod hybrid
material by electrostatic self-assembly for surface-enhanced
Raman scattering. Carbon 2013;51:25564.
[6] He SJ, Liu KK, Su S, Yan J, Mao XH, Wang DF, et al. Graphenebased high-efficiency surface-enhanced Raman scatteringactive platform for sensitive and multiplex DNA detection.
Anal Chem 2012;84:46227.
[7] Huh S, Park J, Kim YS, Kim KS, Hong BH, Nam JM. UV/ozoneoxidized large-scale graphene platform with large chemical
enhancement in surface-enhanced Raman scattering. ACS
Nano 2011;5:9799806.
[8] Sun ST, Wu PY. Competitive surface-enhanced Raman
scattering effects in noble metal nanoparticle-decorated
graphene sheets. Phys Chem Chem Phys 2011;13:2111620.
[9] Liu AP, Xu T, Ren QH, Yuan M, Dong WJ, Tang WH. Graphene
modulated 2D assembly of plasmonic gold nanostructure on
diamond-like carbon substrate for surface-enhanced Raman
scattering. Electrochem Commun 2012;25:748.
[10] Hao QZ, Wang B, Bossard JA, Kiraly B, Zeng Y, Chiang IK, et al.
Surface-enhanced Raman scattering study on graphenecoated metallic nanostructure substrates. J Phys Chem C
2012;116:724954.
[11] Hummers WS, Offeman RE. Preparation of graphitic oxide. J
Am Chem Soc 1958;80:1339.

CARBON

7 0 (2 0 1 4) 2 1 522 3

[12] Wang RC, Lin HY, Chen SJ, Lai YF, Huang MRS. Boundary
layer-assisted chemical bath deposition of well-aligned ZnO
rods on Si by a one-step method. Appl Phys A 2009;96:77581.
[13] Mahaling RN, Jana GK, Das CK, Jeong H, Ha CS. Carboxylated
nitrile elastomer/filler nanocomposite: effect of silica
nanofiller in thermal, dynamic mechanical behavior, and
interfacial adhesion. Macromol Res 2005;13:30613.
[14] Acik M, Lee G, Mattevi C, Pirkle A, Wallace RM, Chhowalla M,
et al. The role of oxygen during thermal reduction of
graphene oxide studied by infrared absorption spectroscopy.
J Phys Chem C 2011;115:1976181.
[15] Acik M, Lee G, Mattevi C, Chhowalla M, Cho K, Chabal YJ.
Unusual infrared-adsorption mechanism in thermal reduced
graphene oxide. Nat Mater 2010;9:8405.
[16] Shin KS, Lee HS, Joo SW, Kim K. Surface-induced
photoreduction of 4-nitrobenzenethiol on Cu revealed by
surface-enhanced Raman scattering spectroscopy. J Phys
Chem C 2007;111:152237.
[17] Yan B, Thubagere A, Premasiri WR, Ziegler LD, Negro LD,
Reinhard BM. Engineered SERS substrates with multiscale
signal enhancement: nanoparticle cluster arrays. ACS Nano
2009;3:1190202.

223

[18] Otto A. The chemical (electronic) contribution to surfaceenhanced Raman scattering. J Raman Spectrosc
2005;36:497507.
[19] Jensen L, Aikens CM, Schatz GC. Electronic structure
methods for studying surface-enhanced Raman scattering.
Chem Soc Rev 2008;37:106173.
[20] Yu XX, Cai HB, Zhang WH, Li XJ, Pan N, Luo Y, et al. Tuning
chemical enhancement of SERS by controlling the chemical
reduction of graphene oxide nanosheets. ACS Nano
2011;5:9528.
[21] Kim NH, Kim K. Surface-enhanced resonance Raman
scattering of rhodamine 6G on Pt nanoaggregates. J Raman
Spectrosc 2005;36:6238.
[22] Wang RC, Lin HY. Efficient surface enhanced Raman
scattering from Cu2O porous nanowires transformed from
CuO nanowires by plasma treatments. Mater Chem Phys
2012;136:6615.
[23] Xie LM, Ling X, Fang Y, Zhang J, Liu ZF. Graphene as a
substrate to suppress fluorescence in resonance Raman
spectroscopy. J Am Chem Soc 2009;131:98901.

Vous aimerez peut-être aussi