Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Graphene-Based Materials As Supercapacitor Electrodes: Li Li Zhang, Rui Zhou and X. S. Zhao

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

View Online / Journal Homepage / Table of Contents for this issue

FEATURE ARTICLE

www.rsc.org/materials | Journal of Materials Chemistry

Graphene-based materials as supercapacitor electrodes


Li Li Zhang, Rui Zhou and X. S. Zhao*

Downloaded by University of Edinburgh on 02 July 2012


Published on 26 April 2010 on http://pubs.rsc.org | doi:10.1039/C000417K

Received 7th January 2010, Accepted 8th March 2010


DOI: 10.1039/c000417k
Graphene is an emerging carbon material that may soon find practical applications. With its unusual
properties, graphene is a potential electrode material for electrochemical energy storage. This article
highlights recent research progress in graphene-based materials as supercapacitor electrodes. With
a brief description of the working principle of supercapacitors, research progress towards the synthesis
and modification of graphene-based materials, including graphene oxide, fullerenes, and carbon
nanotubes, is presented. Applications of such materials with desirable properties to meet the specific
requirements for the design and configuration of advanced supercapacitor devices are summarized and
discussed. Future research trends towards new approaches to the design and synthesis of graphenebased nanostructures and architectures for electrochemical energy storage are proposed.

1. Introduction
With the rapidly growing market in portable electronic devices,
electric vehicles and hybrid electric vehicles (HEVs), there has
been an ever increasing demand for environmentally friendly,
high-performance energy-storage systems. Supercapacitors, also
known as electrochemical capacitors (ECs) or ultracapacitors,
are such devices, with high power capability, long cyclic life
(> 100 000 cycles), low maintenance, and fast dynamics of charge
propagation.14 With many thousands of times higher power
density than lithium ion batteries and much larger energy density
compared to conventional capacitors, supercapacitors are an
ideal energy storage device offering advantages over other energy
storage systems for applications requesting short load cycle and
high reliability, such as energy capture sources, including load

Department of Chemical and Biomolecular Engineering, National


University of Singapore, 4 Engineering Drive 4, Singapore, 117576.
E-mail: chezxs@nus.edu.sg; Fax: +65-67791936; Tel: +65-651164727

Li Li Zhang

Li Li Zhang received her B. Eng.


in Chemical and Biomolecular
Engineering from the National
University of Singapore in 2004.
Since 2006, she has been a PhD
candidate under the supervision
of Professor X. S. Zhao in the
same department at the
National University of Singapore. Her research interest
focuses on the development of
innovative carbon materials for
application in electrochemical
energy storage.

This journal is The Royal Society of Chemistry 2010

crane, forklifts, and electric vehicles.5 Nowadays, supercapacitors are used in consumer electronics, memory back-up
systems, and industrial power and energy management.3 A more
recent application of supercapacitors in Airbus A380 planes has
shown their safe and reliable performance.
While the energy density of supercapacitors is much higher
than conventional dielectric capacitors, it is still lower than
batteries and fuel cells. Most of the commercially available
supercapacitor products have a specific energy density less
than 10 Wh kg1, which is 3 to 15 times lower than batteries
(150 Wh kg1 is possible for lithium ion batteries).6 Thus, there
has been a great deal of research effort on increasing the energy
performance of supercapacitors to be close to or even beyond
that of batteries.
A supercapacitor stores energy using either ion adsorption
(electrical double layer capacitors, EDLCs) or fast and reversible
Faradic reactions (pseudocapacitors). These two mechanisms
can function simultaneously, depending on the nature of the
electrode material. Progress towards supercapacitor technologies

Rui Zhou received his B. Eng.


(Honors) in Polymer Materials
and Engineering from Zhejiang
University (China) in 2007,
minor
in
entrepreneurship
management from Chu Kochen
Honors College. After two years
masters education in the
Department of Chemical and
Biomolecular Engineering at
North Carolina State University
(USA), he took a PhD scholarship from Singapore Ministry
Rui Zhou
of Education project grant under
the supervision of Professor
X. S. Zhao in 2009. His
doctorate research interests focus on tailoring the nanostructure of
graphene-based materials for energy storage applications.
J. Mater. Chem., 2010, 20, 59835992 | 5983

Downloaded by University of Edinburgh on 02 July 2012


Published on 26 April 2010 on http://pubs.rsc.org | doi:10.1039/C000417K

View Online

can benefit from the continuous development of nanostructured


electrode materials. A number of recent reviews and books have
discussed the scientific and technological aspects of supercapacitor devices and electrode materials.1,2,710 In the development of EDLCs, a proper control over the pore size and specific
surface area of the electrode for an appropriate electrolyte
solution is crucial to ensure a good performance of the supercapacitor in terms of both power delivery rate and energy storage
capacity.11 To further enhance the specific capacitance of the
electrode, the pseudo-capacitance that is due to the presence of
foreign electro-active species on the electrode can be coupled
with the electrical double layer capacitance. The capacitive
performance of various carbon-based electrodes and the most
commonly studied pseudo-capacitive materials in the literature
are shown in Fig. 1.10
Activated carbons (ACs) are the most widely used electrode
materials because of their large surface area, low cost and easy
processability.1214 However, as is seen from Fig. 1, the limited
energy storage capacity (typically below 200 F g1) and rate
capability restrict their applications only to certain niche
markets. Obviously, there are other important factors in
dictating the selection of a good electrode material for supercapacitor applications, such as electrical conductivity, chemical
and mechanical stability, and nanostructure. Table 1 summarizes
some properties and characteristics of different carbon electrodes
materials based on the literature. Graphene-based materials,
including zero-dimensional (0D) fullerenes, one-dimensional
(1D) carbon nanotubes (CNTs), two-dimensional (2D) graphene, and three-dimensional (3D) graphite, are of particular
interest because of their exceptional electrical and mechanical
properties and unique structures.15 Graphene, a 2D flat monolayer of sp2 hybridized carbon bonded in a hexagonal lattice, is
the parent of all the graphitic carbons (Fig. 2).1517 The 0D
fullerene, 1D CNT, and 3D graphite or diamond can be formed
by wrapping, rolling, and stacking of a graphene sheet, respectively.13

X. S. Zhao received his PhD in


Chemical Engineering from the
University of Queensland in 1999
under the supervision of Prof.
Max Lu. After two-years of
postdoctoral research training at
the same university, he joined
the Department of Chemical
and Biomolecular Engineering,
National University of Singapore, as an Assistant Professor,
and then became an Associate
Professor. His research interests
X: S: Zhao
included templated synthesis of
porous materials, applied catalysis, including photocatalysis,
colloidal self assembly, and electrochemical energy storage and
conversions. His recent research activities focus on porous photocatalysts for water purifications, carbon and graphene electrode
materials for supercapacitors, and hierarchically structured zeolite
catalysts for converting dimethyl ether/methanol into light olefins.
5984 | J. Mater. Chem., 2010, 20, 59835992

While theoretical studies on graphene started some sixty


years ago,25,26 it was only when free-standing graphene was
first experimentally turned into a reality27 did it arouse
a worldwide resurgence. With the confirmation of its unusual
physical properties,28,29 applications of graphene as nanotransistors, radio-frequency devices, non-volatile memories,
transparent electrodes for solar cells, liquid crystal displays,
and supercapacitor electrodes have been widely exploited. In
this feature article, we focus on recent research progress on
graphene-based electrode materials for supercapacitor
applications.

2. Supercapacitors: principle and performance


The mechanism by which a supercapacitor stores energy is, in
principal, based on two types of capacitive behaviors: the electrical double layer (EDL) capacitance from pure electrostatic
charge accumulation at the electrode-electrolyte interface and the
pseudo-capacitance due to fast and reversible surface redox
processes at characteristic potentials. The structure of the
supercapacitor is similar to that of a battery. It consists of two
electrodes in contact with an electrolyte solution separated by
a separator (Fig. 3a). The components that make up the supercapacitor, including the electrodes, the separator, the current
collector, as well as the electrolyte, all are important factors
affecting the overall performance of the device that must be
considered in designing a supercapacitor device. Porous carbon
materials are often the choice of the electrode because of their
good electrical conductivity coupled with large interface area.
Fig. 3b illustrates an EDL structure (the Stern model) formed on
a positively charged porous electrode surface without considering the curvature of the pore. The negative charge in both
the Stern layer and diffuse layer all contribute to the EDL
capacitance.
For the EDL type of supercapcitor, the specific capacitance C
(F g1) of each electrode is generally assumed to follow that of
a parallel-plate capacitor:14
C

3r 30
A
d

(1)

Fig. 1 The capacitive performance for carbon and pseudocapacitor


electrodes.10

This journal is The Royal Society of Chemistry 2010

View Online

Downloaded by University of Edinburgh on 02 July 2012


Published on 26 April 2010 on http://pubs.rsc.org | doi:10.1039/C000417K

Table 1 A comparison of various carbon electrode materials for supercapacitors

Carbon

Specific
Electrical
surface area/m2 g1 Density/g cm3 conductivity/S cm1 Cost

Fullerene
CNTs
Graphene
Graphite
ACs
Templated porous carbon
Functionalized porous carbon
Activated carbon fibers
Carbon aerogels

1100140017
120500
263020
1020
10003500
5003000
3002200
10003000
4001000

1.72
0.6
>1
2.2619
0.40.7
0.51
0.50.9
0.30.8
0.50.7

108101418
10410519
106
10423
0.11
0.31024
> 300
510
110

where 3r (a dimensionless constant) is the relative permittivity,


30(F m1) is the permittivity of a vacuum, A (m2 g1) is the specific
surface area of the electrode accessible to the electrolyte ions, and
d (m) is the effective thickness of the EDL (also known as the
Debye length). Different from the EDL capacitance, pseudocapacitance arises for thermodynamic reasons between the extent
of charge acceptance (Dq) and the change of potential (DV).7 The
derivative C d(Dq)/(d DV) corresponds to a capacitance, which
is referred to as the pseudo-capacitance. The main difference
between the pseudo-capacitance and the EDL capacitance lies in
that pseudo-capacitance is Faradic in origin, involving fast and
reversible redox reactions between the electrolyte and some
electro-active species on the electrode surface. The most
commonly known active species are ruthenium oxide,30

Medium
High
High
Low
Low
High
Medium
Medium
Low

Aqueous electrolyte

Organic electrolyte

F g1

F g1

F cm3

F cm3

50100
< 60
< 60
< 30
10020521,22 >100205 8011021 >80110
< 200
120350
150300
120370
100125

< 80
< 200
< 180
< 150
< 80

< 100
60140
100150
80200
< 80

<
<
<
<
<

50
100
90
120
40

manganese oxide,31,32 vanadium nitride,33 electrically conducting


polymers,34,35 and surface functional groups on carbon.36 While
the pseudo-capacitance can be higher than EDL capacitance, it
surfers from the drawbacks of a low power density and lack of
stability during cycling.
The performance of a supercapacitor is mainly evaluated on
the basis of the following criteria: (1) power density substantially
greater than batteries with acceptably high energy densities
(> 10 Wh kg1), (2) an excellent cycle ability (more than 100 times
of batteries), (3) fast charge-discharge processes (within seconds),
(4) low self-discharging, (5) safe operation, and (6) low cost. It
must be pointed out that the time constant expressed as resistance (R) times capacitance (C) is another important parameter
in evaluating the performance of a supercapacitor. Therefore, in

Fig. 2 Graphene: the mother of all graphitic carbon materials.15

This journal is The Royal Society of Chemistry 2010

J. Mater. Chem., 2010, 20, 59835992 | 5985

View Online

3. Graphene-based supercapacitor electrode


materials
The honeycomb network of graphene is the basic building block
of other important carbon allotropes with different dimensionalities, as shown in Fig. 2. Each of the allotropes exists as
a different structure with different electrochemical properties.
The performance of these graphene-based materials as supercapacitor electrodes is discussed separately in the following
sections.

Downloaded by University of Edinburgh on 02 July 2012


Published on 26 April 2010 on http://pubs.rsc.org | doi:10.1039/C000417K

3.1 Graphite, carbon onions and fullerenes

Fig. 3 Schematic diagrams of (a) a two-cell supercapacitor device made


of nonporous electrode and (b) the EDL structure based at a positively
charged electrode surface.

order to have smaller leakage or low self-discharging rate, a large


time constant value is desirable.
The maximum energy stored and power delivered for a single
cell supercapacitor are given in eqns (3) and (4), respectively:
1
E CT V 2
2

V2
4Rs

(3)

(4)

where V in volts is the cell voltage, CT in farads is the total


capacitance of the cell, and Rs in ohms is the equivalent series
resistance (ESR). Each of the elements is crucial to the final
performance of the supercapacitor. The capacitance of the cell
depends extensively upon the electrode material. The cell
voltage is limited by the thermodynamic stability of the electrolyte solution. The ESR comes from various types of resistance associated with the intrinsic electronic properties of the
electrode matrix and electrolyte solution, mass transfer resistance of the ions in the matrix, contact resistance between the
current collector and the electrode. Hence, a high-performance
supercapacitor must simultaneously satisfy the requirements of
large capacitance value, high operating cell voltage, and
minimum ESR. It is thus obvious that the development of both
electrode material and the electrolyte solution are essential in
order to optimize the overall performance of the supercapacitor.
With regard to the supercapacitor electrode, it must possess
high surface area and proper pore size. In addition, the electrode
should be electrically conductive with a good stability. The
presence of pseudo-active species for enhancing the overall
capacitance is sometimes necessary. Furthermore, the density of
the electrode should be sufficiently high to give a high volumetric
energy density, which is desirable for designing highly compact
energy and power sources. On the other hand, the electrolyte
must be carefully chosen and evaluated as well. Non-aqueous
electrolytes with a low resistivity are a good choice for designing
high-power and high-energy density supercapacitors because the
non-aqueous electrolyte can be operated at high voltages (up to
3.5 to 4 V).
5986 | J. Mater. Chem., 2010, 20, 59835992

Graphite has been known as a natural mineral for about


500 years and is widely used in our daily life, such as pencils and
dry lubricants. Its high in-plane thermal conductivity and low
electrical resistance make it a good candidate as an electrode
material. Graphite is an anisotropic material with out-of-plane
electrical and thermal conductivities 1000 times lower than the
in-plane situation. In addition, graphite can be intercalated with
other elements to produce graphite intercalation compounds
(GICs).37 Furthermore, substitution of other elements, such as
boron for carbon produces p-type graphite.38
Liu and Osaka39 described the EDLC behavior of isotropic,
high-density graphite electrodes and polyethylene oxide/LiClO4
polymer electrolytes. The concentration and temperature of
LiClO4 significantly affected the capacitance. Later, Mitra and
Sampath40 made an advancement towards solid-state electrochemical capacitors by using exfoliated graphite (EG) as the
electrode material. Specific capacitances ranging from 0.74 to
0.98 mF cm2 with a long cycle life and a short response time
were achieved. Expanding the interlayer spacing of graphite
(d002 0.404 nm) was described by Ka and Oh.41 Ion-accessible
sites were generated through ion intercalation into the interlayer
space during the electrode polarization process. It was found that
the volumetric specific capacitance could reach 30 F mL1 and
the cell voltage could be as high as 3.7 V with a charge-discharge
columbic efficiency of over 99%.
The low electrolyte decomposition voltage limits the cell
voltage of an EDLC device. On the other hand, the high operating voltages for lithium-ion battery devices are attributed to the
formation of a protective layer on the electrode surface (the socalled solid electrolyte interphase, SEI) during the initial charge
cycle, which can prevent further solvent reduction during
subsequent cycling. Therefore, one possible route towards a high
EDLC voltage may be the formation of a SEI-like protection
layer on the EDLC electrode so the sustained solvent reduction
can be prevented as well. Campana and co-workers42 investigated the intercalation of (C2H5)4N+ and passive film formation
on a negative electrode made of highly oriented pyrolytic
graphite (HOPG). Their results showed that the formation of
a SEI-like protective layer can indeed allow the EDLC to be
operated beyond 3 V, which is advantageous for improving
energy storage capacity.
Recent development in asymmetric supercapacitors promises
a wider operation window and higher energy density. Asymmetric supercapacitors using anodes and cathodes of different
materials can overcome the limitation of symmetric supercapacitors by leveraging the different chemistries of the
This journal is The Royal Society of Chemistry 2010

Downloaded by University of Edinburgh on 02 July 2012


Published on 26 April 2010 on http://pubs.rsc.org | doi:10.1039/C000417K

View Online

electrodes, thus achieving a wider cell voltage.43 Khomenko


et al.44 described a hybrid supercapacitor with an organic electrolyte and graphite and activated carbon as the negative and
positive electrodes, respectively. A voltage of 4.5 V was found to
be the optimal voltage to achieve the highest energy density
without capacitance fading during cycling. Park and coworkers45 developed a novel hybrid supercapacitor using
graphite as the cathode and niobium(V) oxide as the anode to
achieve high current load requirements. Recently, Tian and
Lian43 discussed the high voltage window and excess capacitance
observed on an asymmetric supercapacitor with graphite as the
negative electrode and RuO2 as the positive electrode in
H4SiW12O40 (SiWA) electrolyte.
Carbon onions, consisting of concentric graphene spherical
shells, are high-surface-area (from 350 to 520 m2 g1) graphitic
carbon materials with few sub-nanometre pores. The completely
accessible surface of carbon onions to electrolyte ions manifests
them as good candidates for high-power supercapacitor applications. A study46 on their high current charge-discharge
performance revealed a decrease of specific capacitance due to
diamond particle graphitization and defect formation on particle
surfaces. The carbon onion could deliver the stored energy under
a high current density, since its capacitance was less dependent
on current density. Capacitances of onion-like carbon materials
ranging within 2040 F g1 in an acid electrolytic solution (1 M
H2SO4) and 70100 F g1 in an alkaline electrolytic solution (6 M
KOH) have been reported.47,48
Fullerenes and their derivatives are another promising family
of carbon materials for supercapacitor applications due to their
unique nanostructures, which combine reversible redox charge
storage properties with high surface areas.17,49 The discovery of
C60 in 198550 led to the 1996 Nobel prize award in Chemistry to
Smalley, Curl and Kroto.51 Theoretical predictions indicated that
the lowest unoccupied molecular orbital (LUMO) of C60 is
capable of accepting at least six electrons upon reduction. In
addition to having high reversible redox charge capability, being
chemically stable, uniformity in size and shapes, fullerenes and
their derivatives have a large specific surface area (Table 2). The
electrochemical properties of pure fullerenes and endohedral
metallofullerenes are strongly dependent upon solutesolvent
interactions.49,51 Tran and co-workers17 examined the dependence of size, shape and metal species (La) on the redox properties of C60, C70, and La@C82 in an ammonia solution
containing 0.1 M potassium triflate (CF3SO3K). The relative
magnitude of transfer coefficients that represent a measure of the
symmetry of the energy barrier for the oxidation and reduction
of the monoanionic species of the fullerenes suggested that the
availability of the surface area permitting delocalization of p
electrons is a determining factor of their reduction potentials.

Table 2 The physical properties of fullerenes.17

Fullerene geometry

3
Volume/A

Surface
2
area/A

Specific surface
area/m2 g1

C60 (sphere)
C70 (prolate spheroid)
C82 (C2v, prolate
spheroid)

186.5
205.2
239.8

157.8
166.5
186.8

1364
1234
1182

This journal is The Royal Society of Chemistry 2010

The lower reduction potential of La@C82 (about 0.72 V) than


pure fullerenes C60 and C70 (about 0.96 V and 0.94 V,
respectively) indicated that the former is a stronger electron
acceptor. The rehybridization of C82 by the three electrons
donated from the La atom makes it easier for La@C82 to accept
an electron than other fullerenes.
Despite the superior electronic properties and charge carrier
capabilities of graphite and fullerenes, these carbon materials are
hardly used as supercapacitor electrodes, mainly due to their low
specific capacitance. For graphite and exfoliated graphite, it is
still unknown whether electrolyte ions can penetrate into the
graphene layers, and how different the contributions of basal
planes and edges to the total capacitance will be.48 Therefore,
both theoretical and fundamental studies on the mechanism of
charge storage in such materials should be carried out to advance
supercapacitor development.
3.2 Carbon nanotubes and associated composites
The discovery of CNTs has significantly advanced the science of
carbon materials. With their unique physical properties, such as
high electrical conductivity and good chemical and mechanical
stability, CNTs have proven their advantages in electrochemical
energy conversion and storage.6,32,5257 Both single-walled carbon
nanotubes (SWNTs) and multi-walled carbon nanotubes
(MWNTs) have been explored as energy storage electrode
materials.
Niu et al.58 observed that a MWNT-based supercapacitor
electrodes with a specific surface area of 430 m2 g1 displayed
a specific capacitance of 102 F g1 and a power density of 8 kW
kg1 in an acidic electrolyte. According to An and co-workers,59
SWCNT-based supercapacitor electrodes displayed a maximum
specific capacitance of 180 F g1 and a measured power density of
20 kW kg1 at an energy density of 7 W kg1 in KOH electrolyte.
A recent study60 showed that the electrode fabricated with
entangled CNTs was less efficient in facilitating ion transport
than the electrode fabricated with aligned CNTs due to the
irregular pore structures and high entanglement of the structure
of the former. Hence, the use of aligned CNTs is advantageous in
terms of high-power supercapacitors. Hata and co-workers54
described a method to fabricate a densely packed, aligned SWNT
electrode by utilizing the zipping effect of liquids, which allowed
the bulk materials to retain their intrinsic properties. The energy
density of the aligned SWNT electrode was about 35 Wh kg1
(normalized to a cell consisting of two identical electrodes) in an
organic electrolyte. In addition, the rate capability was observed
to be better than that of ACs.
Improvement of the energy density of CNT electrodes by
increasing the specific surface area via KOH activation has been
described.61 The porosity and electrical conductivity must be
properly balanced in order to achieve both high capacitance and
good rate performance. An interesting CNT-aerogel composite
material was recently synthesized by uniformly dispersing carbon
aerogel throughout the CNT host matrix without destroying the
integrity or reducing the aspect ratio of the CNT.55The important
features of the nanocomposite include large surface area, high
mechanical resilience, lightweight matrix with pore spaces on
micrometre scale, and binderless electrode. The specific capacitance was found to be 524 F g1.
J. Mater. Chem., 2010, 20, 59835992 | 5987

Downloaded by University of Edinburgh on 02 July 2012


Published on 26 April 2010 on http://pubs.rsc.org | doi:10.1039/C000417K

View Online

In spite their superior electrical properties, CNT-based


supercapacitors do not meet the expected performance.59,62 One
of the major issues is the high contact resistance between the
electrode and the current collector, greatly limiting the power
performance. Compared with binding electrode material to the
current collectors indirectly, growing CNTs directly on to
a conductive substrate is a promising approach to reducing the
contact resistance. A template method for fabricating CNT/Au
nanowire hybrid electrodes has been reported (Fig. 4).63 Both the
electrode (CNT) and current collector (Au nanowire) are integrated into a single nanostructure, thus leading to a largely
improved electrochemical performance with a power density as
high as 48 kW kg1.
Another approach to enhancing the specific capacitance is to
modify CNTs with active materials to realize pseudo-capacitance. Zhang and co-workers32 described the use of carbon
nanotube arrays (CNTAs), which were directly connected with
the current collector (a Ta foil) to make composite electrodes
with a hierarchical porous structure. Manganese oxide,32
PANI,64,65 and polypyrrole (Ppy)66 were employed as the pseudoactive materials to prepare manganese oxide/CNTA, PANI/
CNTA and Ppy/MWNT composites, respectively. The electrochemical results showed a very high specific capacitance of about
1000 F g1 for the PANI/CNTA composite with an excellent rate
capability (a 95% capacity retention at 118 A g1) and a good
stability (up to 5000 cycles test). The manganese oxide/CNTA
composite possessed a moderate specific capacitance of about
100 F g1 at a current density of 77 A g1 with good cycle ability.
The observed electrochemical performances of the composite
electrode materials were ascribed to the direct growth of CNTA
on the current collector and the pseudo-capacitive behavior of
the active materials. In addition, the high density of the
manganese oxide/CNTA composite material led to a high volumetric capacitance, suitable for compact supercapacitor design.
It is worthy to note that the pseudo-capacitance strongly
depends on the surface utilization of the active material. Simon
and Gogotsi9 suggested that one possible strategy to improve
both the energy and power densities for supercapacitors is to
achieve a conformal deposit of pseudo-capacitive materials onto
highly ordered, high-surface-area CNT and AC electrodes.
3.3 Graphene, graphene oxide and associated composites
Since Geim and co-workers27 experimentally demonstrated the
preparation of a single layer of graphite with atomic thickness, it
has been gaining increasing interest in a wide range of fields.67

Fig. 4 A scheme showing a supercapacitor device with CNT/Au as the


electrodes.63

5988 | J. Mater. Chem., 2010, 20, 59835992

With inherent properties, such as tunable band gap, extraordinary electronic transport properties, excellent thermal conductivity, great mechanical strength, and large surface area,
graphene has been explored for applications ranging from electronic devices to electrode materials.68,69
3.3.1 Synthesis of graphene and graphene oxide. Many
methods have been reported to synthesize graphene sheets over
the past few years, including mechanical cleavage of graphite,27
unzipping carbon nanotubes,70 chemical exfoliation of
graphite,71 solvothermal synthesis,72 epitaxial growth on SiC
surfaces73 and metal surfaces,74 chemical vapor deposition,75
bottom-up organic synthesis,76 etc. Taghioskoui77 recently
summarized various graphene preparation methods. One of the
greatest challenges is the stabilization of single graphene sheets
with controllable size and morphology. Fundamentals of
chemistry and physics are especially important and preparation
methods with or without polymeric and surfactant dispersants
have been developed to prepare exfoliated graphene.7880
Herein, we mainly focus on two methods, namely mechanical
exfoliation that produces few layer graphene of high quality and
a chemical method that has been demonstrated to give high
throughput.
Mechanically exfoliated single-layer graphene was first developed by Geim and co-workers27 using a technique called
micromechanical cleavage. Typically, a celluphene tape is used to
peel off graphene layers from a graphite flake, followed by
pressing the tape against a substrate. Upon removing the tape,
a single sheet graphene is obtained. Ritter and Lyding81 utilized
the mechanical exfoliation method to deposit graphene monolayers and bilayers with minimum lateral dimensions of 210 nm
onto a H-passivated Si(100) surface. These earlier works
provided opportunities to experimentally investigate the electronic structure of nanosized graphene and formed the foundation to develop graphene-based nanoelectronics. However, the
low throughput of the mechanical exfoliation method largely
limits its applications for mass production. Thereafter, alternative approaches affording a high yield are highly desirable.
On the other hand, the chemical method that is considered
a scalable approach to synthesizing graphene, has been used
widely to synthesize chemically derived graphene. As is schematically illustrated in Fig. 5, graphite is first oxidized to graphene oxide (GO) using either the Hummers method82 or the
modified Hummers method.83 Chemically derived graphene can
then be obtained after reduction of the GO using hydrazine
solution or other reducing agents. GO is an excellent precursor to
synthesize graphene nanosheets.80 The larger interlayer distance
makes it readily dispersed in solution with a relatively high
stability. Ruoffs group84 demonstrated a solution-based process
for the production of chemically derived single-layer graphene
with an excellent stability. A two-step method was recently
reported for nearly complete reduction of surface functionalities
of GO by deoxygenation with NaBH4 and dehydration with
concentrated sulfuric acid.85 However, the harsh oxidation and
reduction reactions may deteriorate the graphene structure and
decrease the performance of graphene-based devices.
Ang et al.86 proposed a straightforward one-step intercalation
and exfoliation method to produce large-sized, conductive
graphene sheets without the use of surfactants. The method is
This journal is The Royal Society of Chemistry 2010

View Online

Downloaded by University of Edinburgh on 02 July 2012


Published on 26 April 2010 on http://pubs.rsc.org | doi:10.1039/C000417K

Fig. 5 An illustration of the chemical route to the synthesis of chemically derived graphene.71

based on the rich intercalation chemistry of GO aggregates. The


large-sized GO aggregates consist of multilayer graphene flakes,
which are highly oxidized on the outer layer. The inner layer, on
the other hand, consists of mildly oxidized graphene sheets.
Intercalation of GO aggregates by tetrabutylammonium cations
(TBA) via electrostatic attraction and cation-p interactions
followed by exfoliation in DMF yields large-sized conductive
graphene sheets with a high yield (> 90%).
Thin films are a useful morphology in many technological
applications, particularly as electrodes in supercapacitors and
batteries, fuel cells, and electronic devices. A flow-directed
assembly method was recently described to produce GO paperlike materials.87 The chemical functional groups on the surface of
the GO provide opportunities for tailoring the physical and
chemical properties of the material. The preparation of macroscopic, free-standing GO membranes with controlled thickness
and area at a liquid/air interface was demonstrated.88 Very
recently, Yang and co-workers76 reported a chemical growth
method for the synthesis of 2D graphene nanoribbons with
lengths of up to 12 nm and the synthesis scheme is shown in
Fig. 6.
3.3.2 Graphene as supercapacitor electrodes. Graphene is
considered to be an excellent electrode material for supercapacitors because of its high electrical conductivity, high surface

area, great flexibility, excellent mechanical properties, and rich


chemistry. Graphene films have been used as stretchable electrodes.75 Chemically modified graphene (CMG) sheets can
physically adjust themselves to be accessible to different types of
electrolyte ions, free from the use of conductive fillers and
binders. Besides, its flexibility facilitates an easy fabrication of
supercapacitor devices. Ruoff and coworkers from UT-Austin
pioneered the fundamentals of CMG materials.89
Studies have shown that the specific capacitances of graphene
can reach 135 F g1, 99 F g1 and 75 F g1 in aqueous, organic,
and ionic liquid electrolytes, respectively.90,91 Despite the intense
interest and continuous report on experimental observations,
real applications of graphene have yet to be realized. This is
mainly due to the difficulty in reliable production of high-quality
graphene from a scalable approach.16 While mechanical exfoliation produces graphene with the highest quality, the method is
neither high throughput nor high yield. Zhao et al.92 employed
the chemical vapor deposition method to synthesize carbon
nanosheets composed of graphene layers on conventional carbon
fibers and carbon papers. It was found that such carbon nanosheets possess a capacitance value of 0.076 F cm2 (based on the
geometric testing area) in a H2SO4 solution. Hence, the total
capacitance was estimated to be 1.49  104 F based on a virtual
supercapacitor device rolled in a sandwich pad with a given
dimension.

Fig. 6 The synthesis of linear two-dimensional graphene nanoribbons.76

This journal is The Royal Society of Chemistry 2010

J. Mater. Chem., 2010, 20, 59835992 | 5989

View Online

Downloaded by University of Edinburgh on 02 July 2012


Published on 26 April 2010 on http://pubs.rsc.org | doi:10.1039/C000417K

Fig. 7 A scheme illustrating the synthesis of GNS/PANI composite materials.97

The experimentally observed capacitances are mainly limited


by the agglomeration of graphene sheets and do not reflect the
intrinsic capacitance of an individual graphene sheet. Very
recently, an experimental determination of EDL capacitance
( 21 mF cm2) and quantum capacitance of single layer
and double-layer graphene were reported.93 In order to
harness the unique properties of graphene, graphene- and
GO-based composite materials have been explored.84,94 Ruoff
and co-workers84 reported a general approach to preparing
graphene-polymer composites via exfoliation of graphite and
molecular-level dispersion of CMG sheets within polymers.
A polystyrene-graphene composite was found to exhibit
a percolation threshold of around 0.1 vol.% for room temperature conductivity, the lowest reported value for any carbonbased composite and almost the same as a SWNT-polymer
composite. At only 1 vol.%, the composite displayed a conductivity of 0.1 S m1, which is sufficient for many applications.
Graphene-conducting-polymer composites have received the
greatest interest.9598 Cheng and co-workers95 prepared a graphene/polyaniline composite paper (GPCP) by in situ anodic
electropolymerization of aniline monomer as a PANI film on
graphene paper. The obtained composite paper combines flexibility, conductivity and electrochemical activity and exhibited
a gravimetric capacitance of 233 F g1 and a volumetric capacitance of 135 F cm3. A graphene nanosheet/polyaniline (GNS/
PANI) composite was also synthesized using the polymerization
method (see Fig. 7).97 Graphene (about 15 wt%) was homogeneously coated on the surface of PANI nanoparticles. A specific
capacitance of 1046 F g1 was observed at a low scan rate. The
energy density of the composite could reach 39 Wh kg1 at
a power density of 70 kW kg1.
Although graphene nanosheets are excellent electrode materials, the use of highly toxic reducing agents, such as hydrazine
and dimethylhydrazine, remains a serious issue for large-scale
production. Murugan and co-workers96 demonstrated a microwave-assisted solvothermal process to produce graphene nanosheets without the need for highly toxic chemicals. The authors
investigated the energy storage properties of these thus-prepared
graphene nanosheets and associated PANI composites. The
graphene/PANI composite with 50 wt% graphene displayed both
EDL capacitance and pseudocapacitance with an overall specific
capacitance of 408 F g1.
Very recently, we prepared a series of CMG and PANI
nanofiber composites using an in situ polymerization method.99
5990 | J. Mater. Chem., 2010, 20, 59835992

PANI fibers were observed to adsorb on the graphene surface


and/or filled between the graphene sheets. The composite displayed a specific capacitance as high as 480 F g1 at a current
density of 0.1 A g1. The results showed that good capacitive
performance can be obtained by doping either graphene with
a small amount of PANI or bulky PANI with a small amount of
graphene.

4. Summary
Graphene-based materials with various microtextures have
proven to be promising electrode materials for supercapacitor
applications. Despite the rapidly growing number of publications
on the synthesis of graphene and GO, a cost-effective and environmentally benign method for the production of high quality,
free-standing graphene still remains a great challenge. Control
over both the quality and quantity of the final product in
synthesizing graphene requires complete understanding of the
physics and chemistry of graphene under different processing
conditions. A number of critical issues, such as complete exfoliation of graphite, stabilization of single or few-layer graphene
sheets in various solvents, and retaining the intrinsic properties
of 2D graphene must be addressed before the chemical exfoliation method can be commercially used to produce graphene.
With the theoretical demonstration of pillared graphene for
hydrogen storage100 and experimental results on porous graphene for gas separation,101 research breakthroughs are likely to
be made in the near future with regard to nanoporous graphene
for applications beyond gas storage and separation. With the
well-established materials processing and growth techniques,
such as the Langmuir-Blodgett method for producing thin films,
the layer-by-layer technique for fabricating core-shell nanostructures, and self assembly under controlled colloidal chemistry, graphene-based architectures with designed physical,
chemical and morphological properties as electrode materials for
electrochemical energy storage and conversion will probably be
one of the research trends.
Graphene-based composite materials with the intrinsic properties of graphene and pseudocapacitive materials, such as graphene-conducting-polymer composites and graphene-transition
metal oxide composites, are promising supercapacitor electrodes.
Future research efforts should be placed on enhancing the
interactions between graphene and the pseudocapacitive material
to improve the Faradic processes across the interface so as to
This journal is The Royal Society of Chemistry 2010

View Online

achieve efficient pseudo-capacitance in addition to the EDL


capacitance. Using conducting polymers and transition metal
oxides to pillar graphene sheets is perhaps a good research
direction towards utilizing the unique properties of graphene for
supercapacitor applications.
In any case, colloidal science will play a vital role in realizing
such envisioned structures and applications.

Acknowledgements

Downloaded by University of Edinburgh on 02 July 2012


Published on 26 April 2010 on http://pubs.rsc.org | doi:10.1039/C000417K

Financial support from Ministry of Education under grant


number MOE2008-T2-1-004 is appreciated.

Reference
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33

M. Winter and R. J. Brodd, Chem. Rev., 2004, 104, 42454269.


A. Burke, J. Power Sources, 2000, 91, 3750.
J. R. Miller and P. Simon, Science, 2008, 321, 651652.
L. L. Zhang and X. S. Zhao, Chem. Soc. Rev., 2009, 38, 25202531.
J. R. Miller and A. F. Burke, Electrochem. Soc. Interf. Spring, 2008,
17, 5357.
V. V. N. Obreja, Phys. E., 2008, 40, 25962605.
B. E. Conway, Electrochemical Supercapacitors: Scientific
Fundamentals and Technological Applications, Kluwer Academic/
Plenum Publisher, New York, 1999.
A. S. Arico, P. Bruce, B. Scrosati, J.-M. Tarascon and W. van
Schalkwijk, Nat. Mater., 2005, 4, 366377.
P. Simon and Y. Gogotsi, Nat. Mater., 2008, 7, 845854.
A special issue on Electrochemical Capacitors, Electrochem. Soc.
Interf., 2008, Spring.
C. Largeot, C. Portet, J. Chmiola, P. L. Taberna, Y. Gogotsi and
P. Simon, J. Am. Chem. Soc., 2008, 130, 27302731.
A. G. Pandolfo and A. F. Hollenkamp, J. Power Sources, 2006, 157,
1127.
O. Barbieri, M. Hahn, A. Herzog and R. Kotz, Carbon, 2005, 43,
13031310.
E. Frackowiak, Phys. Chem. Chem. Phys., 2007, 9, 17741785.
A. K. Geim and K. S. Novoselov, Nat. Mater., 2007, 6, 183191.
C. N. R. Rao, A. K. Sood, K. S. Subrahmanyam and A. Govindaraj,
Angew. Chem., Int. Ed., 2009, 48, 77527777.
N. E. Tran, S. G. Lambrakos and J. J. Lagowski, J. Mater. Eng.
Perform., 2009, 18, 95101.
T. L. Makarova, B. Sundqvist, P. Scharff, M. E. Gaevski, E. Olsson,
V. A. Davydov and A. V. Rakhmanina, Carbon, 2001, 39, 2203
2209.
D. Eder, Chem. Rev., 2010, 110, 1348.
M. Pumera, Chem. Rec., 2009, 9, 211223.
D. Li and R. B. Kaner, Science, 2008, 320, 11701171.
Y. Wang, Z. Q. Shi, Y. Huang, Y. F. Ma, C. Y. Wang, M. M. Chen
and Y. S. Chen, J. Phys. Chem. C, 2009, 113, 1310313107.
M. J. Allen, V. C. Tung and R. B. Kaner, Chem. Rev., 2010, 110,
132145.
A. B. Fuertes and S. Alvarez, Carbon, 2004, 42, 30493055.
P. R. Wallace, Phys. Rev., 1947, 71, 622634.
J. W. McClure, Phys. Rev., 1956, 104, 666671.
K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang,
S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004,
306, 666669.
K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang,
M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos and
A. A. Firsov, Nature, 2005, 438, 197200.
Y. Zhang, Y. W. Tan, H. L. Stormer and P. Kim, Nature, 2005, 438,
201204.
C. C. Hu, K. H. Chang, M. C. Lin and Y. T. Wu, Nano Lett., 2006, 6,
26902695.
L. L. Zhang, T. Wei, W. Wang and X. S. Zhao, Microporous
Mesoporous Mater., 2009, 123, 260267.
H. Zhang, G. Cao, Z. Wang, Y. Yang, Z. Shi and Z. Gu, Nano Lett.,
2008, 8, 26642668.
D. Choi, G. E. Blomgren and P. N. Kumta, Adv. Mater., 2006, 18,
11781182.

This journal is The Royal Society of Chemistry 2010

34 L. Z. Fan, Y. S. Hu, J. Maier, P. Adelhelm, B. Smarsly and


M. Antonietti, Adv. Funct. Mater., 2007, 17, 30833087.
35 L. L. Zhang, S. Li, J. Zhang, P. Guo, J. Zheng and X. S. Zhao, Chem.
Mater., 2010, 22, 11951202.
36 M. Seredych, D. Hulicova-Jurcakova, G. Q. Lu and T. J. Bandosz,
Carbon, 2008, 46, 14751488.
37 O. Tanaike and M. Inagaki, Carbon, 1997, 35, 831836.
38 J. Kouvetakis, R. B. Kaner, M. L. Sattler and N. Bartlett, J. Chem.
Soc., Chem. Commun., 1986, 17581759.
39 X. J. Liu and T. Osaka, J. Electrochem. Soc., 1996, 143, 39823986.
40 S. Mitra and S. Sampath, Electrochem. Solid-State Lett., 2004, 7,
A264A268.
41 B. H. Ka and S. M. Oh, J. Electrochem. Soc., 2008, 155, A685A692.
42 F. P. Campana, M. Hahn, A. Foelske, P. Ruch, R. Kotz and
H. Siegenthaler, Electrochem. Commun., 2006, 8, 13631368.
43 Q. Tian and K. Lian, Electrochem. Solid St., 13, A4A6.
44 V. Khomenko, E. Raymundo-Pinero and F. Beguin, J. Power
Sources, 2008, 177, 643651.
45 G. J. Park, D. Kalpana, A. K. Thapa, H. Nakamura, Y. S. Lee and
M. Yoshio, Bull. Korean Chem. Soc., 2009, 30, 817820.
46 C. Portet, J. Chmiola, Y. Gogotsi, S. Park and K. Lian, Electrochim.
Acta, 2008, 53, 76757680.
47 E. G. Bushueva, P. S. Galkin, A. V. Okotrub, L. G. Bulusheva,
N. N. Gavrilov, V. L. Kuznetsov and S. I. Moiseekov, Phys.
Status Solidi B, 2008, 245, 22962299.
48 C. Portet, G. Yushin and Y. Gogotsi, Carbon, 2007, 45, 25112518.
49 F. Zhou, C. Jehoulet and A. J. Bard, J. Am. Chem. Soc., 1992, 114,
1100411006.
50 H. W. Kroto, J. R. Heath, S. C. OBrien, R. F. Curl and
R. E. Smalley, Nature, 1985, 318, 162163.
51 L. Echegoyen and L. E. Echegoyen, Acc. Chem. Res., 1998, 31, 593
601.
52 M. Pumera, Chem.Eur. J., 2009, 15, 49704978.
53 R. H. Baughman, A. A. Zakhidov and W. A. de Heer, Science, 2002,
297, 787792.
54 D. N. Futaba, K. Hata, T. Yamada, T. Hiraoka, Y. Hayamizu,
Y. Kakudate, O. Tanaike, H. Hatori, M. Yumura and S. Iijima,
Nat. Mater., 2006, 5, 987994.
55 T. Bordjiba and L. H. D. M. Mohamedi, Adv. Mater., 2008, 20, 815
819.
56 J. Chen, A. I. Minett, Y. Liu, C. Lynam, P. Sherrell, C. Wang and
G. G. Wallace, Adv. Mater., 2008, 20, 566570.
57 J. Yan, H. Zhou, P. Yu, L. Su and L. Mao, Adv. Mater., 2008, 20,
28992906.
58 C. Niu, E. K. Sichel, R. Hoch, D. Moy and H. Tennent, Appl. Phys.
Lett., 1997, 70, 14801482.
59 K. H. An, W. S. Kim, Y. S. Park, J.-M. Moon, D. J. Bae, S. C. Lim,
Y. S. Lee and Y. H. Lee, Adv. Funct. Mater., 2001, 11, 387392.
60 H. Zhang, G. Cao, Y. Yang and Z. Gu, J. Electrochem. Soc., 2008,
155, K19.
61 E. Frackowiak, S. Delpeux, K. Jurewicz, K. Szostak, D. CazorlaAmoros and F. Beguin, Chem. Phys. Lett., 2002, 361, 3541.
62 C. Portet, P. L. Taberna, P. Simon and E. Flahaut, J. Electrochem.
Soc., 2006, 153, A649A653.
63 M. M. Shaijumon, F. S. Ou, L. Ci and P. M. Ajayan, Chem.
Commun., 2008, 23732375.
64 H. Zhang, G. Cao, Z. Wang, Y. Yang, Z. Shi and Z. Gu,
Electrochem. Commun., 2008, 10, 10561059.
65 C. Meng, C. Liu and S. Fan, Electrochem. Commun., 2009, 11, 186
189.
66 M. S. P. S. M. Hughes, A. C. Renouf, C. Singh, G. Z. Chen,
D. J. Fray and A. H. Windle, Adv. Mater., 2002, 14, 382385.
67 A. K. Geim, Science, 2009, 324, 15301534.
68 X. Wang, L. J. Zhi and K. Mullen, Nano Lett., 2008, 8, 323327.
69 T. Yumura, K. Kimura, H. Kobayashi, R. Tanaka, N. Okumura
and T. Yamabe, Phys. Chem. Chem. Phys., 2009, 11, 82758284.
70 D. V. Kosynkin, A. L. Higginbotham, A. Sinitskii, J. R. Lomeda,
A. Dimiev, B. K. Price and J. M. Tour, Nature, 2009, 458, 872
875.
71 V. C. Tung, M. J. Allen, Y. Yang and R. B. Kaner, Nat.
Nanotechnol., 2009, 4, 2529.
72 M. Choucair, P. Thordarson and J. A. Stride, Nat. Nanotechnol.,
2009, 4, 3033.
73 K. V. Emtsev, A. Bostwick, K. Horn, J. Jobst, G. L. Kellogg, L. Ley,
J. L. McChesney, T. Ohta, S. A. Reshanov, J. Rohrl, E. Rotenberg,

J. Mater. Chem., 2010, 20, 59835992 | 5991

View Online

74
75
76
77
78

79
80

Downloaded by University of Edinburgh on 02 July 2012


Published on 26 April 2010 on http://pubs.rsc.org | doi:10.1039/C000417K

81
82
83
84
85
86

A. K. Schmid, D. Waldmann, H. B. Weber and T. Seyller, Nat.


Mater., 2009, 8, 203207.
P. W. Sutter, J.-I. Flege and E. A. Sutter, Nat. Mater., 2008, 7, 406
411.
K. S. Kim, Y. Zhao, H. Jang, S. Y. Lee, J. M. Kim, J. H. Ahn,
P. Kim, J. Y. Choi and B. H. Hong, Nature, 2009, 457, 706710.
X. Yang, X. Dou, A. Rouhanipour, L. Zhi, H. J. Rader and
K. Mullen, J. Am. Chem. Soc., 2008, 130, 42164217.
M. Taghioskoui, Mater. Today, 2009, 12, 3437.
M. Lotya, Y. Hernandez, P. J. King, R. J. Smith, V. Nicolosi,
L. S. Karlsson, F. M. Blighe, S. De, Z. M. Wang, I. T. McGovern,
G. S. Duesberg and J. N. Coleman, J. Am. Chem. Soc., 2009, 131,
36113620.
J. N. Coleman, Adv. Funct. Mater., 2009, 19, 368016.
D. Li, M. B. Muller, S. Gilje, R. B. Kaner and G. G. Wallace, Nat.
Nanotechnol., 2008, 3, 101105.
K. A. Ritter and J. W. Lyding, Nanotechnology, 2008, 19, 015704
015709.
W. S. Hummers and R. E. Offeman, J. Am. Chem. Soc., 1958, 80,
13391339.
N. I. Kovtyukhova, P. J. Ollivier, B. R. Martin, T. E. Mallouk,
S. A. Chizhik, E. V. Buzaneva and A. D. Gorchinskiy, Chem.
Mater., 1999, 11, 771778.
S. Stankovich, D. A. Dikin, G. H. B. Dommett, K. M. Kohlhaas,
E. J. Zimney, E. A. Stach, R. D. Piner, S. T. Nguyen and
R. S. Ruoff, Nature, 2006, 442, 282286.
W. Gao, L. B. Alemany, L. J. Ci and P. M. Ajayan, Nat. Chem.,
2009, 1, 403408.
P. K. Ang, S. A. Wang, Q. L. Bao, J. T. L. Thong and K. P. Loh,
ACS Nano, 2009, 3, 35873594.

5992 | J. Mater. Chem., 2010, 20, 59835992

87 D. A. Dikin, S. Stankovich, E. J. Zimney, R. D. Piner,


G. H. B. Dommett, G. Evmenenko, S. T. Nguyen and
R. S. Ruoff, Nature, 2007, 448, 457460.
88 C. M. Chen, Q. H. Yang, Y. G. Yang, W. Lv, Y. F. Wen, P. X. Hou,
M. Z. Wang and H. M. Cheng, Adv. Mater., 2009, 21, 30073011.
89 R. Ruoff, Nat. Nanotechnol., 2008, 3, 1011.
90 M. D. Stoller, S. Park, Y. Zhu, J. An and R. S. Ruoff, Nano Lett.,
2008, 8, 34983502.
91 S. R. C. Vivekchand, C. S. Rout, K. S. Subrahmanyam,
A. Govindaraj and C. N. R. Rao, J. Chem. Sci., 2008, 120, 913.
92 X. Zhao, H. Tian, M. Zhu, K. Tian, J. J. Wang, F. Kang and
R. A. Outlaw, J. Power Sources, 2009, 194, 12081212.
93 J. Xia, F. Chen, J. Li and N. Tao, Nat. Nanotechnol., 2009, 4, 505
509.
94 L. J. Cote, R. Cruz-Silva and J. Huang, J. Am. Chem. Soc., 2009,
131, 1102711032.
95 D. W. Wang, F. Li, J. Zhao, W. Ren, Z. G. Chen, J. Tan, Z. S. Wu,
I. Gentle, G. Q. Lu and H. M. Cheng, ACS Nano, 2009, 3, 17451752.
96 A. V. Murugan, T. Muraliganth and A. Manthiram, Chem. Mater.,
2009, 21, 50045006.
97 J. Yan, T. Wei, B. Shao, Z. Fan, W. Qian, M. Zhang and F. Wei,
Carbon, 2010, 48, 487493.
98 Y. Zhang, H. Li, L. Pan, T. Lu and Z. Sun, J. Electroanal. Chem.,
2009, 634, 6871.
99 K. Zhang, L. L. Zhang, J. Wu and X. S. Zhao, Chem. Mater., 2010,
22, 13921401.
100 G. K. Dimitrakakis, E. Tylianakis and G. E. Froudakis, Nano Lett.,
2008, 8, 31663170.
101 D.-e. Jiang, V. R. Cooper and S. Dai, Nano Lett., 2009, 9, 4019
4024.

This journal is The Royal Society of Chemistry 2010

You might also like