You are on page 1of 5

Electrochemistry Communications 13 (2011) 133137

Contents lists available at ScienceDirect

Electrochemistry Communications
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / e l e c o m

Direct electrodeposition of reduced graphene oxide on glassy carbon electrode and


its electrochemical application
Liuyun Chen a, Yanhong Tang a, Ke Wang b, Chengbin Liu b,, Shenglian Luo b
a
b

College of Materials Science and Engineering, Hunan University, Changsha 410082, PR China
State Key Laboratory of Chemo/Biosensing and Chemometrics, Hunan University, Changsha 410082, PR China

a r t i c l e

i n f o

Article history:
Received 29 October 2010
Received in revised form 25 November 2010
Accepted 25 November 2010
Available online 3 December 2010

a b s t r a c t
Graphene nanosheets were directly deposited onto a glassy carbon electrode through cyclic voltammetric
reduction of a graphene oxide colloidal solution. The resulting electrodes were characterized by
electrochemical methods and scanning electron microscopy. The application of the graphene modied
electrodes in simultaneous determination of hydroquinone and catechol was investigated.
2010 Elsevier B.V. All rights reserved.

Keywords:
Electrodeposition
Graphene
Sensor
Hydroquinone
Catechol

1. Introduction
Direct simultaneous determination of isomers is a challenge because
of their similar structures and physicochemical properties. Hydroquinone (1,4-dihydroxybenzene, HQ), catechol (1,2-dihydroxybenzene,
CC), and resorcinol (1,3-dihydroxybenzene, RC) are three isomers of
phenolic compounds, and due to their high toxicity, their determination
is of interest in environmental analysis. At present, a number of methods
for dihydroxybenzene analysis have been proposed [15], among which
electrochemical technique has the advantages such as simplicity and
convenience. Unfortunately, however, due to the same electro-active
groups, the isomers, especially HQ and CC, usually present overlapped
redox peaks so that they can't be resolved at conventional electrodes
such as gold and glassy carbon electrode (GCE).
Graphene has recently attracted tremendous interest because of its
unique thermal, mechanical, and electrical properties [6]. One of the
promising applications of graphene is electrochemical sensing [7,8].
Since every atom in a graphene sheet is a surface atom, molecular
interaction and thus electron transport through graphene can be
highly sensitive to adsorbed molecules [9]. In this regard, it is believed
that graphene, when used for the modication of the conventional
electrodes, has the great potential for distinguishing a diverse range of
aromatic isomers. Here, we show that the graphene-based electrode
readily allows the simultaneous detection of HQ and CC.

Corresponding author. Tel.: +86 731 88823805.


E-mail address: chem_cbliu@hnu.cn (C. Liu).
1388-2481/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.elecom.2010.11.033

For sensing applications, the methods for depositing active


materials on electrodes are most crucial. So far, graphene lms on
electrodes usually have been obtained by drop-casting solution-based
graphene obtained from chemical reduction of graphene oxide (GO)
sheets [10]. Such a preparation methodology has intrinsic limitations
such as lack control of the lm thickness, and moreover, toxic
chemicals are involved. More recently, electrochemical reduction of
GO to graphene has drawn great attention due to its fast and green
nature [1113]. Typically, the electrochemical synthesis of graphene
was carried out via two steps, namely, GO being rst assembled on the
electrodes by solution deposition methods, and then being subjected
to electrochemical reduction. Herein, we show that graphene lms
can be prepared on electrodes directly from GO dispersions by onestep electrodeposition technique.
2. Experimental
Oxidized graphite was synthesized from natural ake graphite by
the Hummers method [14]. All other chemicals were of analytic grade
and used as received. Double-distilled water was used throughout.
The prepared graphite oxide powder was exfoliated in a 0.067 M
pH 9.18 phosphate buffer solution (PBS) by ultrasonication to form a
1.0 mg mL1 GO colloidal dispersion. Prior to electrodeposition, the
GCE was in turn polished with 0.3 and 0.05 M alumina powders, and
then sequentially sonicated in double-distilled water and anhydrous
ethanol. The cyclic voltammetric reduction was performed in the GO
dispersion with a magnetic stirring and N2 bubbling on a CHI 660 C
electrochemical station (CHI Instruments Inc., USA) with a threeelectrode system where bare GCE was used as the working electrode,

134

L. Chen et al. / Electrochemistry Communications 13 (2011) 133137

and the Pt foil and saturated calomel electrode (SCE) acted as the
counter and reference electrodes, respectively.
The scanning electron microscopic (SEM) and transmission
electron microscopic (TEM) images were carried out on a JEOL
JSM-6700F eld emission scanning electron microscope and a JEOL
JEM-3010 high-resolution transmission electron microscope,
respectively.

electrochemically active oxygen-containing groups on graphene


planes that are too stable to be reduced by the cyclic voltammetry
method [14]. The formation of EG lm on the GCE was directly
conrmed by the SEM image shown in Fig. 1D. The graphene
electrodeposition can occur on any conducting surfaces, and moreover, the graphene coating is very stable as a result of its poor
insolubility in common solvents.

3. Results and discussion


3.2. Enhanced electrochemical sensing with EG/GCE
3.1. Electrodeposition of graphene
Fig. 1A reveals that the graphite oxide was entirely exfoliated as
individual GO sheets in pH 9.18 PBS. Despite GO can form welldispersed aqueous colloids, it is well known that chemical reduction
of GO sheets in aqueous solutions results in their irreversible
agglomerate [15]. Therefore, it is reasonable to speculate that when
the GO sheets in direct contact with an electrode accept electrons to
suffer from electrochemical reduction, the resulted graphene sheets
will also be insoluble, and thus directly attach to the electrode surface.
Fig. 1B shows the typical cyclic voltammograms (CVs) of GO
electrolysis on GCE, where one anodic peaks (I) and two cathodic
peaks (II and III) are observed during the process. The persistent
increase of the peak currents with successive potential scans indicates
that the deposition of conducting graphene on the GCE has indeed
been achieved. Fig. 1C shows that the electrodeposited graphene (EG)
modied GCE (EG/GCE) in 0.067 M PBS (pH 9.18) only exhibits peaks I
and II. Therefore, the cathodic current peak III is attributed to the
irreversible electrochemical reduction of GO [1113], and the anodic
peak I and cathodic peak II are ascribed to the redox pair of some

Compared to the bare GCE, the excellent conductivity, the large


surface area, and the remaining oxygen-related detects of the EG lm
make it a sensitive promoter for electrochemical sensing processes
[16,17]. The redox probe Fe(CN)3/4
is sensitive to surface chemistry
6
of carbon-based electrodes [18], and therefore was used to directly
evaluate the charge transfer property of the graphene electrode. The
CVs shown in Fig. 2A prove the enhanced current response of the EG/
GCE toward Fe(CN)3/4
, indicating increased electrochemical active
6
sites by EG surface modication. No saturation of the peak current is
observed when the electrodeposition increased from 1 to 18 cycles,
while the anodic-to-cathodic peak separation (Ep) for EG/GCE
initially became smaller and then larger than that for bare GCE
(Fig. 2B). The narrowed Ep of Fe(CN)3/4
suggests the improved
6
electron transfer kinetics by EG, which is ascribed to the reactive edge
defects on graphene [19,20]. Here we show that the electron transfer
kinetics also correlates with thickness of graphene lm that can be
easily controlled by using the electrodeposition technique, a distinct
advantage over previously developed methods. In this work, 10 cycles
were used for the electrode modication in view of current sensitivity.

Fig. 1. (A) TEM image of GO dispersed in 0.067 M, pH 9.18 PBS, (B) CVs depicting electrochemical reduction of 1.0 mg mL1 GO in PBS (0.067 M, pH 9.18) on a GCE at 10 mV s1,
(C) CV of an EG/GCE in PBS (0.067 M, pH 9.18) at 10 mV s1, and (D) SEM image of EG lm (10 electrodeposition cycles) modied GCE.

L. Chen et al. / Electrochemistry Communications 13 (2011) 133137

135

130

(A)

120

I / A

(B)

110

75
0
-75

GCE
GCE+EG (3 Cycles)
GCE+EG (9 Cycles)
GCE+EG (12 Cycles)
GCE+EG (18 Cycles)

-150
-0.2

0.0

0.2

0.4

P/ mV

150

100
90
80
70
60
0

0.6

12

15

18

electrodeposition cycle

E / V (vs. SCE)
30
30

(C)

(D)

HQ
20

CC

I / A

I / A

15
0
-15

0.1

0.2

0.3

GCE
EG/GCE

-20

-30
0.0

0
-10

GCE
EG/GCE

-0.1

10

0.4

0.0

0.1

30

(E)

HQ

0.2

0.3

0.4

0.5

E / V (vs. SCE)

E / V (vs. SCE)
50

CC

40

(F)

CC

HQ

I / A

I / A

15
0
-15

GCE
EG/GCE

20
10

GCE
EG/GCE

-30
-0.1

30

0.0

0.1

0.2

0.3

0.4

0
0.5

E / V (vs. SCE)

0.0

0.1

0.2

0.3

0.4

E / V (vs. SCE)

Fig. 2. (A) CVs of 5.0 mM K3Fe(CN)6 in 0.1 M KCl on bare GCE and EG/GCE with the EG lms prepared under various electrodeposition cycles, (B) dependence of Ep on
electrodeposition cycle, (C) and (D) CVs of HQ (0.2 mM) and CC (0.2 mM) in 0.2 M acetate buffer solution (ABS, pH 5.8) at bare GCE and EG/GCE at 50 mV s1, respectively, (E) CVs of
HQ (0.2 mM) + CC (0.2 mM) in 0.2 M, pH 5.8 ABS at bare GCE and EG/GCE at 50 mV s1, (F) Differential pulse voltammograms (DPVs) of HQ (0.2 mM) + CC (0.2 mM) in 0.2 M, pH
5.8 ABS at bare GCE and EG/GCE, the DPV parameters were at a scan rate of 4 mV s1, 50 mV pulse amplitude, and 20 ms pulse width.

Electrocatalysis of HQ and CC on GCE and EG/GCE is shown in


Fig. 2C and D, respectively. At bare GCE, both HQ and CC showed an
irreversible electrochemical behavior with a very big Ep (311 and
230 mV for HQ and CC, respectively) and relatively weak redox
current. Whereas, at the EG/GCE, the Ep signicantly narrowed (38
and 35 mV for HQ and CC, respectively), accompanied with greatly
increased but nearly equal cathodic and anodic peak currents for both
species, indicative of a quasi-reversible electrochemical process.
These suggest efcient electrocatalytic oxidation reactions of HQ
and CC at the EG/GCE. Compared to the bare GCE, the background
current of the EG/GCE increased greatly, implying a larger surface area
of the EG lm on the GCE. Good linear relationships between the peak
current and the square root of scan rates from 10 to 300 mV s1 at EG/
GCE were obtained for both HQ and CC (data not shown), indicating
diffusion-control redox processes of HQ and CC.
Fig. 2E shows the CVs of mixed HQ and CC at GCE and EG/GCE,
respectively. Two well-separated redox pairs were resolved on the
EG/GCE, and the separation of oxidation peaks between HQ and CC
was about 110 mV, in contrast to a single broad peak at about 0.39 V
on bare GCE owing to the overlap of the oxidation peaks of HQ and CC.

Moreover, peak currents substantially increased at EG/GCE due to the


accelerated electron transfer kinetics by the graphene lm. The results
suggest that simultaneously electrochemical determination of HQ and
CC at the graphene modied electrode is possible. Differential pulse
voltammetry (Fig. 2F) exhibited a much higher current sensitivity
than cyclic voltammetry (Fig. 2E) and therefore the differential pulse
voltammetry was used for the detection.
3.3. Simultaneous determination of HQ and CC
The simultaneous determination of HQ and CC was carried out by
keeping one species constant in concentration (0.1 mM) and changing
the concentration of the other one in a 0.2 M, pH 5.8 acetate buffer
solution (ABS). Fig. 3A and B display the DPVs for the determination,
where the linear concentration ranges were obtained to be 6.0 M
0.2 mM (r = 0.9996) for HQ and 1.0 M0.2 mM (r = 0.9977) for CC
(Fig. 3C). The detection limits of HQ and CC were 0.2 and 0.1 M
(S/N = 3), respectively.
Eleven successive determination of 0.1 mM HQ and 0.1 mM CC at
one EG/GCE resulted in a relative standard deviation (RSD) of 1.74%

136

L. Chen et al. / Electrochemistry Communications 13 (2011) 133137

(A)

HQ

30
200M
150
100
80
40
6
0 M

I / A

40
30
20

(A)

15

I / A

50

0
pH 2.6
pH 4.0
pH 4.6
pH 5.8

-15

10
-30
0
0.1

0.2

0.3

0.0

0.4

45

(B)

200M
150
100
80
40
10
6
1
0 M

CC

I / A

60
40
20

30

0.4

0.6

0.8

(B)

15

I / A

80

0.2

E / V (vs. SCE)

E / V (vs. SCE)

0
pH 2.6
pH 4.0
pH 4.6
pH 5.8

-15
-30

0
0.0

0.1

0.2

0.3

0.0

0.4

60

0.4

0.6

0.8

Fig. 4. CVs of (A) 0.2 mM HQ and (B) 0.2 mM CC in 0.2 M ABS with different pHs.

(C)

CC

I(A)=1.29+0.298C(M)
r=0.9977

protons and electrons. Therefore, the probable reactions of HQ and


CC on EG/GCE are described as follows [4,5]:

I / A

40

0.2

E / V (vs. SCE)

E / V (vs. SCE)

HQ

OH

20

40

80

120

160

+ 2 H+ + 2e-

HQ:

I(A)=0.373+0.129C(M)
r=0.9996

200

OH

C / M

OH

O
OH

Fig. 3. DPVs showing the oxidation peak current of (A) HQ vs. HQ concentration in the
presence of 0.1 mM CC and (B) CC vs. CC concentration in the presence of 0.1 mM HQ.
(C) Calibration plots. The electrolyte: ABS (0.2 M, pH 5.8). The DPV parameters were at
a scan rate of 4 mV s1, 50 mV pulse amplitude, and 20 ms pulse width.

CC:

O
+ 2 H+ + 2e-

4. Conclusion
and 0.37%, respectively, suggesting a high reproducibility. The
stability of the electrode was checked by performing 30 consecutive
potential scans in 0.2 M, pH 5.8 ABS containing 0.1 mM HQ. On
comparison, the graphene electrode exhibited a 5.5% decrease in the
signal, whereas the bare GCE exhibited about a 24% change in the
signal.
3.4. Mechanism for reactions
Fig. 4A and B shows CVs of EG/GCE in 0.2 M ABS containing
0.2 mM HQ and 0.2 mM CC, respectively, with different pH values.
In both cases, negative shifts in anodic and cathodic peak
potentials as well as enhanced current responses are observed
with increasing pH values from 2.6 to 5.8, which suggest the
involvement of proton in the electrode reactions. The anodic peak
potential (Epa) has a linear relationship with pH from 2.6 to 5.8,
where the linear regression equations are Epa (V) = 0.495 0.055
pH (r = 0.982) and Epa (V) = 0.601 0.057 pH (r = 0.980) for HQ
and CC, respectively. The slopes of Epa vs. pH for both HQ and CC
are very close to the theoretical value of 59 mV/pH (25C) for a
reversible electrochemical process involved with equal number of

We have demonstrated a straightforward one-step electrodeposition method to prepare graphene lm for simultaneous determination
of HQ and CC. The graphene modied electrode showed enhanced
electron transfer properties and high resolution capacity to the HQ
and CC isomers. Wide linear concentration ranges, low detection
limits, and excellent reproducibility and stability were achieved on
the EG/GCE, indicating graphene a promising sensing platform for
isomers determination.
Acknowledgements
This work was supported by the National Natural Science
Foundation of China (50878079, 51078129, 21047004) and the
National Basic Research Program of China (2009CB421601).
References
[1] G.N. Chen, J.S. Liu, J.P. Duan, H.Q. Chen, Talanta 53 (2000) 651.
[2] P.J. Kerzic, W.S. Liu, M.T. Pan, H. Fu, Y. Zhou, A.R. Schnatter, R.D. Irons, Chem. Bio.
Interact. 184 (2010) 182.
[3] M.A. Ghanem, Electrochem. Commun. 9 (2007) 2501.

L. Chen et al. / Electrochemistry Communications 13 (2011) 133137


[4] D. Zhao, X. Zhang, L. Feng, L. Jia, S. Wang, Colloids Surf., B 74 (2009) 317.
[5] H.L. Qi, C.X. Zhang, Electroanalysis 17 (2005) 832.
[6] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, Y. Zhang, S.V. Dubonos, I.V.
Grigorieva, A.A. Firsov, Science 306 (2004) 666.
[7] M. Zhou, Y.M. Zhai, S.J. Dong, Anal. Chem. 81 (2009) 5603.
[8] C.S. Shan, H.F. Yang, J.F. Song, D.X. Han, A. Ivaska, L. Niu, Anal. Chem. 81 (2009)
2378.
[9] A. Rochefort, J.D. Wuest, Langmuir 25 (2009) 210.
[10] M. Pumera, A. Ambrosi, E.L.K. Chng, H.L. Poh, Trends Anal. Chem. 29 (2010) 954.
[11] H.L. Guo, X.F. Wang, Q.Y. Qian, F.B. Wang, X.H. Xia, ACS Nano 3 (2009) 2653.
[12] Y.Y. Shao, J. Wang, M. Engelhard, C.M. Wang, Y.M. Lin, J. Mater. Chem. 20 (2010) 743.

137

[13] M. Zhou, Y.L. Wang, Y.M. Zhai, J.F. Zhai, W. Ren, F. Wang, S.J. Dong, Chem. Eur. J. 15
(2009) 6116.
[14] S. William, J.R. Hummers, E.O. Richard, J. Am. Chem. Soc. 80 (1958) 1339.
[15] S. Stankovich, D.A. Dikin, R.D. Piner, K.A. Kohlhaas, A. Kleinhammes, Y. Jia, Y. Wu, S.
T. Nguyen, R.S. Ruoff, Carbon 45 (2007) 1558.
[16] M. Pumera, Chem. Soc. Rev. 39 (2010) 4146.
[17] M. Pumera, Chem. Rec. 9 (2009) 211.
[18] P.H. Chen, M.A. Fryling, R.L. McCreery, Anal. Chem. 67 (1995) 3115.
[19] D.K. Kampouris, C.E. Banks, Chem. Commun. 46 (2010) 8986.
[20] D.A.C. Brownson, C.E. Banks, Analyst 135 (2010) 2768.

You might also like