Вы находитесь на странице: 1из 7

www.advmat.

de

2D Sandwich-like Sheets of Iron Oxide Grown on Graphene


as High Energy Anode Material for Supercapacitors
Qunting Qu, Shubin Yang, and Xinliang Feng*

Electrochemical capacitors, also called supercapacitors or ultracapacitors, have attracted significant attention as novel energystorage devices attributable to their high power density, long
cycling life, and short charging time. These advantages make
them highly promising for use in electric vehicles, hybrid electric vehicles, and other high power energy sources.[1] However,
supercapacitors deliver an unsatisfactory energy density. Current research work on supercapacitors has been mainly focused
on the enhancement of energy density to make it comparable
to that of batteries.[2] Transition-metal oxides and electronically conducting polymers are the two categories of electrode
materials for supercapacitors based on the redox pseudocapacitive charge storage, whose energy density usually exceed
that of carbon materials involving double layer charge storage.
Figure 1 summarizes the working potential windows of various
pseudocapacitive-type materials in aqueous electrolyte.[3] With
respect to intensive efforts devoted to the cathode materials with
the average working potential of above 0 V vs. SCE,[2,3,g,h,jl,4]
there are only a few reports on the study of anode materials
(working potential below 0 V) due to the unsatisfactory capacitive performance of several types of low-potential electrode
materials.[3bd,f,i]
Graphene, with the maximal surface area of 2630 m2 g1,
is an ideal carbon material for electrochemical double-layer
supercapacitors (EDLCs).[5] However, the exfoliated graphene
tends to restack, thus limiting its capacitance and practical
applications.[6] Recently, graphene-based synergistic composite
materials by combining them with other high-energy metal
oxides or conductive polymers have been reported to possess superior electrochemical performance making use of the
extraordinary surface area, electrical conductivity, thermal stability and mechanical strength of graphene.[7] Especially, supporting other electrochemically active materials uniformly on
the surface of graphene sheets offers an effective protocol to
minimize the aggregation and maximize the electrochemically
accessible area.[7c,8] In this work, we demonstrate the fabrication
of novel 2D sandwich-like graphene-supported Fe3O4 nanocomposites, which show high energy density (85 Wh kg1) with the

Dr. Q. Qu, Dr. X. Feng


College of Chemistry and Chemical Engineering
Shanghai Jiao Tong University
800 Dongchuan Road, Shanghai, 200240, Peoples Republic of China
E-mail: feng@mpip-mainz.mpg.de
Dr. Q. Qu, Dr. S. Yang, Dr. X. Feng
Max Planck Institute for Polymer Research
Ackermannweg 10, 55128 Mainz, Germany

DOI: 10.1002/adma.201103042

5574

wileyonlinelibrary.com

Potential window / V vs. SCE

COMMUNICATION

www.MaterialsViews.com

1.5
1.0

V2O 5

RuO2
M o O3
MnO 2

o r NiO

TiO 2

M oO 2

0.5

WO 3-x

Co3O 4

Ni(O H)2

Polyanili ne
Co (OH)2

0.0
-0.5
In 2O3

-1.0
-1.5

V2O5
B i2O3

VN

Fe3O4

FeOx
Zn O@M oO 3

In aqueous electrolyte
Figure 1. The working potential windows of various pseudocapacitivetype materials in aqueous electrolyte.

average working potential up to 0.8V vs. SCE. The synthesis


procedure for the nanocomposites consists of the direct growth
of FeOOH nanorods on the surface of graphene sheets and the
subsequent electrochemical transformation of FeOOH to Fe3O4
nanoparticles. This strategy provides a novel method for the
preparation of highly active materials directly on the graphene
surface. Apart from the excellent capacitive performance of the
resultant Fe3O4@graphene nanocomposites, the low-cost, abundant resource, and environmentally friendliness of iron oxide
make them highly promising as anode materials for aqueous
supercapacitors.
Graphene oxide (GO) and chemically reduced graphene oxide
(RGO) are used as templates for the growth of FeOOH, the
process of which is schematically shown in Figure 2a. RGO was
prepared by reducing GO in NaOH solution at 80 C.[9] GO and
RGO sheets contain abundant of epoxy, hydroxyl, and carboxyl
groups on their basal planes and edges, making them negatively charged when dispersed in aqueous solution. Fe3+ cations,
formed by the dissolution of FeCl3 in water, can thus favorably
bind with the oxygen-containing groups on graphene sheets
via electrostatic interactions when they are mixed together.[7c,d]
Upon heating the mixture to 80 C, hydrolysis of Fe3+ accompanied by the formation of FeOOH is expected to occur on the surface of GO or RGO. Thereby, 2D sandwich-like nanosheets[10] of
FeOOH and graphene can be obtained. Transmission electron
microscopy (TEM) images (Figure 2bd) present the obtained
GO or RGO-supported FeOOH nanorods. The edges of monolayered graphene sheets can be clearly recognized and FeOOH
nanorods are uniformly distributed on their surface. Despite
that GO or RGO accounts for about 10 wt% of the total mass,
as demonstrated from the thermogravimetric analysis (TGA)

2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2011, 23, 55745580

www.advmat.de
www.MaterialsViews.com

COMMUNICATION
Figure 2. (a) Schematic illustration of the synthesis route towards graphene-based nanocomposites. (b, c, and d) TEM images of the obtained 2D
sandwich-like sheets of FeOOH nanorods grown on GO or RGO.

(Figure S1), FeOOH nanorods only grow on graphene surface


and there is no extra FeOOH remained in the solution. Interestingly, the dimension of the FeOOH nanorods grown on
graphene is much smaller than that obtained without GO or
RGO (see Figure 3a), probably due to the dispersing nucleation
of Fe3+ aided by the oxygen groups of graphene. Remarkably,
the Brunauer-Emmett-Teller (BET) surface area of FeOOH@
RGO nanocomposites is 160 m2 g1 which stands in contrast
to that of pristine FeOOH rods (20 m2 g1) in the absence of
graphene. X-ray diffraction (XRD) pattern of FeOOH@RGO
nanocomposites corresponds well to that of pristine FeOOH
nanorods (Figure S2), signifying the successful growth of
FeOOH on graphene. Importantly, no diffraction peak at about
25o characteristic of restacked graphene appears, indicating
that the monolayer graphene is uniformly dispersed in the 2D
composites.
FeOOH@GO and FeOOH@RGO sheets were subjected
to electrochemical cycles in the potential range of 0.11.3V
(vs. SCE) to complete the transformation of FeOOH to Fe3O4.
The working electrodes for electrochemical cycles were prepared by pressing the mixture of active material (FeOOH,
FeOOH@GO, or FeOOH@RGO), acetylene black, and
poly(tetrafluoroethylene) (PTFE) onto Ni-mesh. The transformation of FeOOH can be validated from the structural and
morphological changes of the electrode materials. Figure 3b

Adv. Mater. 2011, 23, 55745580

shows the cyclic voltammogramms (CV) of the pristine FeOOH


electrode in 1 mol L1 LiOH aqueous solution. During the first
negative scan, an obvious reduction peak appears at 1.2V due
to the electrochemical reduction from Fe3+ to Fe2+. During the
following positive scan, the oxidation peak should be accompanied by the oxidation reaction from Fe2+ to Fe3+. However, to our
surprise, during the second scan, the reduction peak situated at
1.2V disappears and the current increases almost linearly with
a slow slope from 0.8 to 1.1V and a fast slope from 1.1 to
1.2V as the potential goes negatively. It should be noted that
the oxidation peak is rather stable during the extended cycles.
These phenomena suggest that some changes on the crystalline
structure of the electrode material took place during the first
electrochemical cycle.
TEM image of the pristine FeOOH electrode after electrochemical cycles (Figure 3c) indicates that the original largesized FeOOH nanorods collapse into interconnected nanoparticles with the diameter of less than 20 nm. XRD patterns of
the FeOOH electrodes (Figure 3d) after electrochemical cycles
suggest that the original FeOOH has been transformed into an
inverse spinel structure,[11] indicative of the formation of either
Fe3O4 or -Fe2O3 since these two materials exhibit similar XRD
diffraction peaks. Therefore, XPS was further employed to
confirm the composition of the electrode materials. Figure 3e
demonstrates the X-ray photoelectron spectra (XPS) in the Fe2p

2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

5575

www.advmat.de
0.02

(a)

1st

(b)

(c)

0.01

Current / A

COMMUNICATION

www.MaterialsViews.com

0.00

2nd

H evolution -1.1V
2

-0.01
-0.02

-0.03
-1.6

-1.2

-0.8

-0.4

0.0

Potential / V vs. SCE


500

Fe2p3/2

Fe2p1/2

0V vs. SCE

200

-1.3V vs. SCE

Intensity / a. u.

(422)
(511)

(440)

0.1 V

(400)

(220)

300

(111)

400

Intensity / a.u.

(e)

Ni
(311)

(d)

-0.74 V
Charge
-1.15 V
-1.03 V
Discharge

100

FeOOH
0
20

40

60

2-Theta / degree

705

710

715

720

725

730

735

Binding Energy / eV

Figure 3. (a) Scanning electron microscopy (SEM) image of the pristine FeOOH nanorods prepared in the absence of graphene. (b) Cyclic voltammogramms of the FeOOH electrode in 1 mol L1 LiOH aqueous solution at the scan rate of 5 mV s1. (c) TEM image of the electrode material after
electrochemical cycles. (d) XRD pattern of the electrodes before cycling, at the end of discharge (-1.3V vs. SCE), and charge (0.0V vs. SCE). (e) XPS
spectra in the Fe2p region for the electrodes at different charge-discharge states.

region for the electrodes at different charge-discharge states.


The binding energies of Fe2p3/2 and Fe2p1/2 are 712.0 and
725.8 eV, respectively, in good agreement with those of Fe3O4.
Furthermore, the satellite peak at 719.0 eV characteristic of
-Fe2O3 does not appear, further proving that the electrode material is composed of Fe3O4.[11] Based on the above analysis, it can
be concluded that the electrochemical reduction of FeOOH in
LiOH solution offers a feasible protocol to prepare Fe3O4 nanoparticles. Such kind of electrochemical transformation method
has also been revealed to fabricate nanoporous RuO2 and Pt utilizing the electrochemical lithiation reaction of bulk RuO2 and
PtO2.[12]
As we expected, the FeOOH@GO and FeOOH@RGO electrodes exhibit similar cyclic voltammogrammic behavior to the
pristine FeOOH electrode, signifying a similar electrochemical
transformation from FeOOH to Fe3O4 (Figure S3). The electrochemically obtained Fe3O4 is denoted as e-Fe3O4. For comparison,
Fe3O4 powder was prepared through the traditional coprecipitation of Fe2+ and Fe3+ under alkaline condition, and then physically mixed with RGO in aqueous solution. After evaporation of
water, the obtained mixture is denoted as co-Fe3O4/RGO (Figure
S4). The galvanostatic charge curves of co-Fe3O4/RGO, e-Fe3O4,
e-Fe3O4@GO, and e-Fe3O4@RGO electrodes at various current
densities are shown in Figure 4. The distinct plateaus during the
charge process correspond well to the broad anodic peak in CV
curves, indicating the faradic reaction from Fe2+ to Fe3+.

5576

wileyonlinelibrary.com

As for the capacitance mechanism of Fe3O4 in aqueous electrolytes, it is still not clearly understood yet. Some papers proposed that in the electrochemically stable KOH solution, the
adsorption/desorption of OH should be involved in the electrochemical reaction.[3c,13] In order to shed more light on the electrochemical mechanism of Fe3O4 in LiOH solution, we investigated the structure and composition changes of the electrode
materials during the charge/discharge cycles. The electrode was
taken out immediately from the three-electrode cell after being
discharged or charged to a certain potential, and then rinsed
with distilled water to reduce the influence of LiOH electrolyte
and dried at 60 C in vacuum. The Fourier-transform infrared
spectra (FTIR) of the Fe3O4 electrodes at different charge-discharge states are presented in Figure 5b. The absorption peak
at the band of 585 cm1 can be assigned to the Fe-O stretching
vibration mode of Fe3O4. Two peaks at 880 and 796 cm1 correspond to the Fe-O-H bending vibration, probably resulting from
the residual FeOOH in the electrode.[14] The peaks at the band of
1150 and 1200 cm1 are the C-F stretching vibration characteristic of PTFE. Since there are no obvious absorption peaks corresponding to the stretching vibration of free -OH or hydrogenbonded -OH in the region of 3200-3600 cm1,[3i] the adsorption
of OH anions on the surface of Fe3O4 should not be the major
contributor for the charge storage of Fe3O4 in LiOH solution.
Therefore, the existing state and content of Li+ in the Fe3O4 electrode were investigated further by XPS. As indicated in Figure 5c,

2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2011, 23, 55745580

www.advmat.de
www.MaterialsViews.com

-0.2
-0.4
-0.6

-1

0.5A g
1 A g- 1

-0.8

-1

2A g
-1
5A g
-1
7.5A g
-1
10A g

-1.0
-1.2
0

40

e-Fe3O4

0.0

80

120

Potential / V vs. SCE

Potential / V vs. SCE

0.0

(b)

Co-Fe3O4+RGO

-0.2
-0.4
-0.6

-1

0.5A g
-1
2 Ag
-1
3 Ag
5 A g -1
7.5A g -1
10A g -1

-0.8
-1.0
-1.2
0

160

50

-1

e-Fe3O4@GO

0.0

0.5A g
-1
1 Ag
3 A g -1

(d)

-1

Potential / V vs. SCE

Potential / V vs. SCE

0.2

-1

5 Ag
-1
7.5A g
-1
10A g

-0.2
-0.4

100

150

Specific Capacitance / F g

Specific Capacitance / F g

(c)

-0.6
-0.8
-1.0

COMMUNICATION

(a)

0.2

200

-1

-1

0.5A g

e-Fe3O4@RGO

-1

1A g
-1
2A g
-1
3A g
-1
5A g
-1
7.5A g

0.0
-0.2
-0.4

-1

10A g

-0.6
-0.8
-1.0
-1.2

-1.2
0

50

100

150

200

250

300
-1

350

Specific Capacitance / F g

50

100

150

200

250

300

Specific Capacitance / F g-1

350

Figure 4. The galvanostatic charge curves of co-Fe3O4/RGO, e-Fe3O4, e-Fe3O4@GO, and e-Fe3O4@RGO electrodes at various current densities.

the binding energy of Li1s is in proximity to the Li1s state of Li2O


instead of LiOH, which excludes the existence of LiOH electrolyte in the electrode materials. Therefore, the very high binding
energy of Li1s suggests that the adsorbed or intercalated Li+ ions
have a strong interaction with the electrode materials especially
O2 ions in the framework of Fe3O4. The elemental composition of the electrode materials obtained from XPS spectra reveal
that a large amount of Li+ ions reside in the interior of Fe3O4
(Figure 5d). The nanoparticles structure of Fe3O4 and its
inherent ability to adsorb cations[15] probably lead to the extraordinarily high Li/Fe atom ratio for all the electrodes at different
potentials. It is noteworthy that the crystalline structure of Fe3O4
retains stable during the electrochemical cycle, as evidenced by
the XRD and XPS analyses (Figure 3d and e). This is different
from the transformation reaction of Fe3O4 when used as anode
material for lithium ion batteries, where metal Fe is formed at
the end of discharge.[16] All the above results validate that the
fast adsorption or intercalation of Li+ cations plays a very important role in the pseudo-capacitive behavior of Fe3O4.
As displayed in Figure 4, at the current density of 0.5 A g1,
the reversible capacitance of co-Fe3O4/RGO, e-Fe3O4,
e-Fe3O4@GO, and e-Fe3O4@RGO electrodes is 154, 205, 349,
and 326 F g1, respectively. The much higher capacitance of
e-Fe3O4@GO and e-Fe3O4@RGO than that of co-Fe3O4/RGO

Adv. Mater. 2011, 23, 55745580

and e-Fe3O4 suggests that the utilization efficiency of the active


materials can be significantly improved as a result of the uniform distribution of iron oxide on graphene. In addition, the
charge plateaus of e-Fe3O4@GO and e-Fe3O4@RGO electrodes
are more distinct and the potential of these plateaus are more
negative than those of co-Fe3O4/RGO and e-Fe3O4, signifying a
reduced electrode polarization caused by the increased accessible surface area.
The charge behavior of these materials at large current densities reflects their rate capability. When the current density is
increased to 10 A g1, the co-Fe3O4/RGO, e-Fe3O4, e-Fe3O4@
GO, and e-Fe3O4@RGO electrodes maintain 37%, 61%, 68%,
and 90% of their primary capacitance at 0.5 A g1, respectively.
The capacitance of e-Fe3O4@RGO can be up to 304 F g1 even
at the current density of 10 A g1, much higher than the corresponding value of the other three materials. The excellent rate
capability of e-Fe3O4@RGO should originate from the synergistic effect of 2D sandwich-like structure of the nanocomposites with uniform distribution of iron oxide on the graphene
surface and the improved electrical conductivity of RGO.[9a] The
capacitance of e-Fe3O4@RGO is also higher than that of many
metal-oxides-based anode materials[3bf,13b,17] and comparable
to or even higher than that of the cathode materials such as
MnO2, Co3O4, V2O5, WO3-x, and MoO3.[2f,3a,k,18]

2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

5577

www.advmat.de

(a)

0.1V
Charge

Discharge

0.0

-0.74 V

-0.4

-0.74V
-0.8

-1.03V

-1.15V

-1.2
50

100

150

(c)

-1.03 V

C-F
Fe-O-H
Fe-O
1000

-1

(d)
0.1 V
-0.74V
-1.15V
-1.03V

Li2O
LiOH

52

54

56

58

60

3000

62

4000

-1

10.0

7.5

5.0

2.5

0.0
50

2000

Wavenumber / cm

Li1s

Intensity / a.u.

-1.15 V

200

Specific Capacitance / F g

Li/Fe atom ratio

0.1 V

(b)
Transmittance

Potential / V vs. SCE

COMMUNICATION

www.MaterialsViews.com

-1.03 V

-1.15 V

-0.74 V

0.1 V

Binding Energy / eV
Figure 5. (a) The typical galvanostatic discharge-charge curves of the e-Fe3O4 electrode in 1 mol L1 LiOH aqueous solution at the current density of
0.5 A g1. (b) FTIR spectra of the e-Fe3O4 electrodes at different charge-discharge states. (c) XPS spectra of the e-Fe3O4 electrode in the Li1s region.
(d) The relative atom ratio of Li/Fe in the e-Fe3O4 electrode materials obtained from XPS spectra.

Although the capacitance of e-Fe3O4@RGO is still much


lower than that of the cathode material Ni(OH)2 (ranging from
500 to 1800 F g1),[3h,7c,19] we found that its average working
potential could be up to -0.8V vs. SCE, three times as high
as that of Ni(OH)2 (0.27V vs. SCE).[7c] As a result, the energy
density of e-Fe3O4@RGO, which is proportional to the specific
capacitance and square of working potential, is expected to be
much higher than that of Ni(OH)2. For example, the energy density can be up to 85 Wh kg1 at the power density of 2.4 kW kg1,
much higher than 53 Wh kg1 for the Ni(OH)2/graphene
nanocomposites.[7c] The energy density of e-Fe3O4@RGO is
also the highest among the various reported electrode materials
for aqueous supercapacitors,[2,3] enabling it a promising candidate for technological application. Actually, it would be better
to calculate the energy density in the context of a full-sized
and packaged supercapacitor.[20] However, the faradic capacitance of e-Fe3O4@RGO mainly situate in the anodic region,
and assembly of a two-electrode cell shall require another kind
of cathode material such as MnO2, Ni(OH)2, and activated
carbon, etc. Taking Ni(OH)2 for example, a hybrid supercapacitor consisting of e-Fe3O4@graphene anode and Ni(OH)2@
graphene cathode could be assembled by using LiOH solution
as electrolyte. In this way, the two electrodes will adopt Li+ and

5578

wileyonlinelibrary.com

OH as working ions, respectively, which can be favorable for


the full utilization of both the electrode and electrolyte materials. In such two-electrode system, the mass ratio of e-Fe3O4@
graphene anode (specific capacity: 108 mAh g1) to cathode
should be 1.1 if the specific capacity of Ni(OH)2@graphene
cathode is assumed to be 122 mAh g1.[7c] Considering that the
average discharge voltage of this two-electrode cell is supposed
to be 1.07 V, the calculated energy density would be 62 Wh kg1
based on the total mass of active electrode material (including
cathode and anode). Typically, the electrode materials account for
about 50% of the total weight of packaged battery, thus the practical specific energy of this system could be up to 31 Wh kg1,
much higher than that of commercial electrochemical doublelayer capacitors (about 3 Wh kg1) and competitive with that of
a Pb-acid battery and Ni-MH battery (30-50 Wh kg1).[21]
The long-term cycling performance of the co-Fe3O4/RGO,
e-Fe3O4, e-Fe3O4@GO, and e-Fe3O4@RGO electrodes at
the current density of 2 A g1 is compared in Figure 6. The
e-Fe3O4@RGO electrode exhibits good stability maintaining
95% of its initial capacitance after 1000 cycles. On the contrary,
the capacitance of e-Fe3O4 declines rapidly with less than 50%
of its initial value remained. The excellent cycling performance
of e-Fe3O4@RGO can be attributed to the 2D sandwich-like

2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2011, 23, 55745580

www.advmat.de
www.MaterialsViews.com
-1

Specific Capacitance / F g

300

e-Fe3O 4@GO
200

e-Fe3O 4

100

[3]

co-Fe3O 4/RGO
0
0

200

400

600

800

1000

Cycle number
Figure 6. The cycling performance of co-Fe3O4/RGO, e-Fe3O4, e-Fe3O4@
GO, and e-Fe3O4@RGO electrodes at the current density of 2 A g1.

structure of nanocomposites with Fe3O4 uniformly distributed


on graphene sheets, thus preventing the aggregation of Fe3O4
nanoparticles during the long-term cycles. The improved electrical conductivity of RGO with respect to GO could facilitate
the charge-transfer process, which also contributes to the longterm stability of e-Fe3O4@RGO.
In summary, 2D sandwich-like Fe3O4@graphene nanocomposites were fabricated based on the direct growth of FeOOH
nanorods on the surface of graphene sheets and the subsequent
electrochemical transformation of FeOOH to Fe3O4. The synergistic effect of Fe3O4 and graphene leads to high energy density, excellent rate capability, and good cycle life of the Fe3O4@
graphene nanocomposites when they are used as anode materials for aqueous supercapacitors. The low price, abundant
resources, and environmental friendliness of iron oxides may
render their nanocomposites a promising candidate for practical applications.

[4]

[5]

[6]
[7]

Supporting Information
Supporting Information is available from the Wiley Online Library or
from the author.

[8]

Acknowledgements
This work was financially supported by the 2012CB93340, Max Planck
Society through the program ENERCHEM, BASF, DFG Priority Program
SPP 1355, DFG MU 334/32-1, DFG Priority Program SPP 1459, BMBF
LiBZ Project.
Received: August 8, 2011
Revised: September 30, 2011
Published online: November 3, 2011

[9]

[10]
[1] a) P. Simon, Y. Gogotsi, Nat. Mater. 2008, 7, 845; b) J. R. Miller,
P. Simon, Science 2008, 321, 651.
[2] a) W. Chen, Z. L. Fan, L. Gu, X. H. Bao, C. L. Wang, Chem.
Commun. 2010, 46, 3905; b) J. Bae, M. K. Song, Y. J. Park,

Adv. Mater. 2011, 23, 55745580

[11]
[12]

J. M. Kim, M. L. Liu, Z. L. Wang, Angew. Chem. Int. Ed. 2011, 50, 1683;
c) J. A. Yan, E. Khoo, A. Sumboja, P. S. Lee, ACS Nano 2010, 4,
4247; d) W. Sugimoto, H. Iwata, Y. Yasunaga, Y. Murakami,
Y. Takasu, Angew. Chem. Int. Ed. 2003, 42, 4092; e) Z. Chen, Y. C. Qin,
D. Weng, Q. F. Xiao, Y. T. Peng, X. L. Wang, H. X. Li, F. Wei, Y. F. Lu,
Adv. Funct. Mater. 2009, 19, 3420; f) G. Wee, H. Z. Soh, Y. L. Cheah,
S. G. Mhaisalkar, M. Srinivasan, J. Mater. Chem. 2010, 20, 6720;
g) Q. T. Qu, Y. Shi, L. L. Li, W. L. Guo, Y. P. Wu, H. P. Zhang,
S. Y. Guan, R. Holze, Electrochem. Commun. 2009, 11, 1325.
a) S. Yoon, E. Kang, J. K. Kim, C. W. Lee, J. Lee, Chem. Commun.
2011, 47, 1021; b) M. S. Wu, R. H. Lee, J. Electrochem. Soc. 2009,
156, A737; c) N. L. Wu, S. Y. Wang, C. Y. Han, D. S. Wu, L. R. Shiue,
J. Power Sources 2003, 113, 173; d) M. B. Sassin, A. N. Mansour,
K. A. Pettigrew, D. R. Rolison, J. W. Long, ACS Nano 2010, 4, 4505;
e) P. C. Chen, G. Z. Shen, Y. Shi, H. T. Chen, C. W. Zhou, ACS
Nano 2010, 4, 4403; f) F. L. Zheng, G. R. Li, Y. N. Ou, Z. L. Wang,
C. Y. Su, Y. X. Tong, Chem. Commun. 2010, 46, 5021; g) J. K. Chang,
C. M. Wu, I. W. Sun, J. Mater. Chem. 2010, 20, 3729; h) H. Jiang,
T. Zhao, C. Z. Li, J. Ma, J. Mater. Chem. 2011, 21, 3818; i) D. Choi,
G. E. Blomgren, P. N. Kumta, Adv. Mater. 2006, 18, 1178; j) L. Cui,
J. Li, X. G. Zhang, J. Appl. Electrochem. 2009, 39, 1871; k) Y. Y. Liang,
M. G. Schwab, L. J. Zhi, E. Mugnaioli, U. Kolb, X. L. Feng, K. Mllen,
J. Am. Chem. Soc. 2010, 132, 15030; l) J. P. Liu, J. Jiang, C. W. Cheng,
H. X. L, J. X. Zhang, H. Gong, H. J. Fan, Adv. Mater. 2011, 23,
2076.
a) M. Salari, K. Konstantinov, H. K. Liu, J. Mater. Chem. 2011, 21,
5128; b) C. Y. Cao, W. Guo, Z. M. Cui, W. G. Song, W. Cai, J. Mater.
Chem. 2011, 21, 3204.
a) T. H. Han, W. J. Lee, D. H. Lee, J. E. Kim, E. Y. Choi,
S. O. Kim, Adv. Mater. 2010, 22, 2060; b) Y. W. Zhu, S. Murali,
M. D. Stoller, K. J. Ganesh, W. Cai, P. J. Ferreira, A. Pirkle, R. M. Wallace,
K. A. Cychosz, M. Thommes, D. Su, E. A. Stach, R. S. Ruoff, Science
2011, 332, 1537.
D. Li, X. W. Yang, J. W. Zhu, L. Qiu, Adv. Mater. 2011, 23, 2833.
a) M. Pumera, Energy Environ. Sci. 2011, 4, 668; b) S. Stankovich,
D. A. Dikin, R. D. Piner, K. A. Kohlhaas, A. Kleinhammes, Y. Jia,
Y. Wu, S. T. Nguyen, R. S. Ruoff, Carbon 2007, 45, 1558; c) H. L. Wang,
H. S. Casalongue, Y. Y. Liang, H. J. Dai, J. Am. Chem. Soc. 2010,
132, 7472; d) S. Chen, J. W. Zhu, X. D. Wu, Q. F. Han, X. Wang,
ACS Nano 2010, 4, 2822; e) J. J. Xu, K. Wang, S. Z. Zu, B. H. Han,
Z. X. Wei, ACS Nano 2010, 4, 5019; f) X. L. Feng, S. B. Yang,
S. Ivanovici, K. Mllen, Angew. Chem. Int. Ed. 2010, 49, 8408;
g) S. B. Yang, X. L. Feng, K. Mllen, Adv. Mater. 2011, 23, 3575;
h) D. Wu, F. Zhang, P. Liu, X. L. Feng, Chem. Eur. J. 2011, 17,
10804.
a) H. L. Wang, J. T. Robinson, G. Diankov, H. J. Dai, J. Am. Chem.
Soc. 2010, 132, 3270; b) Z. S. Wu, W. C. Ren, D. W. Wang, F. Li,
B. L. Liu, H. M. Cheng, ACS Nano 2010, 4, 5835; c) Z. J. Fan, J. Yan,
L. J. Zhi, Q. Zhang, T. Wei, J. Feng, M. L. Zhang, W. Z. Qian, F. Wei,
Adv. Mater. 2010, 22, 3723; d) D. S. Yu, L. M. Dai, J. Phys. Chem.
Lett. 2010, 1, 467; e) H. L. Wang, Q. L. Hao, X. J. Yang, L. D. Lu,
X. Wang, Nanoscale 2010, 2, 2164; f) S. Liu, X. H. Liu, Z. P. Li,
S. R. Yang, J. Q. Wang, New J. Chem. 2011, 35, 369; g) D. W. Wang,
F. Li, J. P. Zhao, W. C. Ren, Z. G. Chen, J. Tan, Z. S. Wu, I. Gentle,
G. Q. Lu, H. M. Cheng, ACS Nano 2009, 3, 1745; h) Q. Wu, Y. X. Xu,
Z. Y. Yao, A. R. Liu, G. Q. Shi, ACS Nano 2010, 4, 1963.
a) X. B. Fan, W. C. Peng, Y. Li, X. Y. Li, S. L. Wang, G. L. Zhang,
F. B. Zhang, Adv. Mater. 2008, 20, 4490; b) J. P. Rourke, P. A. Pandey,
J. J. Moore, M. Bates, I. A. Kinloch, R. J. Young, N. R. Wilson, Angew.
Chem. Int. Ed. 2011, 50, 3173.
S. B. Yang, X. L. Feng, L. Wang, K. Tang, J. Maier, K. Mllen, Angew.
Chem. Int. Ed. 2010, 49, 4795.
Y. Tian, B. B. Yu, X. Li, K. Li, J. Mater. Chem. 2011, 21, 2476.
Y. S. Hu, Y. G. Guo, W. Sigle, S. Hore, P. Balaya, J. Maier, Nat. Mater.
2006, 5, 713.

2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

COMMUNICATION

e-Fe3 O4@RGO

5579

www.advmat.de

COMMUNICATION

www.MaterialsViews.com

5580

[13] a) S. Y. Wang, K. C. Ho, S. L. Kuo, N. L. Wu, J. Electrochem. Soc.


2006, 153, A75; b) J. Santos-Pena, O. Crosnier, T. Brousse, Electrochim. Acta 2010, 55, 7511.
[14] S. B. Ni, X. H. Wang, G. Zhou, F. Yang, J. M. Wang, Q. Wang, D. He,
Mater. Lett. 2009, 63, 2701.
[15] A. K. Mishra, S. Ramaprabhu, J. Phys. Chem. C 2010, 114, 2583.
[16] J. Z. Wang, C. Zhong, D. Wexler, N. H. Idris, Z. X. Wang, L. Q. Chen,
H. K. Liu, Chem. Eur. J. 2011, 17, 661.
[17] a) J. Chen, K. L. Huang, S. Q. Liu, Electrochim. Acta 2009, 55, 1;
b) Y. H. Kim, S. J. Park, Curr. Appl. Phys. 2011, 11, 462; c) X. Zhao,
C. Johnston, A. Crossley, P. S. Grant, J. Mater. Chem. 2010, 20,
7637; d) M. S. Wu, R. H. Lee, J. J. Jow, W. D. Yang, C. Y. Hsieh,
B. J. Weng, Electrochem. Solid State Lett. 2009, 12, A1; e) L. H. Lu,

wileyonlinelibrary.com

[18]

[19]
[20]
[21]

Y. D. Zhu, F. J. Li, W. Zhuang, K. Y. Chan, X. H. Lu, J. Mater. Chem.


2010, 20, 7645; f) G. R. Li, Z. L. Wang, F. L. Zheng, Y. N. Ou,
Y. X. Tong, J. Mater. Chem. 2011, 21, 4217; g) T. Cottineau, M. Toupin,
T. Delahaye, T. Brousse, D. Belanger, Appl. Phys. A 2006, 82, 599.
a) J. Yan, Z. J. Fan, T. Wei, W. Z. Qian, M. L. Zhang, F. Wei, Carbon
2010, 48, 3825; b) S. Chen, J. W. Zhu, X. Wang, ACS Nano 2010,
4, 6212; c) M. T. Lee, J. K. Chang, Y. T. Hsieh, W. T. Tsai, J. Power
Sources 2008, 185, 1550; d) L. Zheng, Y. Xu, D. Jin, Y. Xie, J. Mater.
Chem. 2010, 20, 7135.
P. Lin, Q. J. She, B. L. Hong, X. A. J. Liu, Y. N. Shi, Z. Shi,
M. S. Zheng, Q. F. Dong, J. Electrochem. Soc. 2010, 157, A818.
M. D. Stoller, R. S. Ruoff, Energy Environ. Sci. 2010, 3, 1294.
F. Beck, P. Ruetschi, Electrochim. Acta 2000, 45, 2467.

2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Adv. Mater. 2011, 23, 55745580

Вам также может понравиться