You are on page 1of 9

DOI: 10.1002/chem.

201101333

Two-Dimensional Nanocomposites Based on Chemically Modified Graphene


Dongqing Wu,[a] Fan Zhang,[a] Ping Liu,[a] and Xinliang Feng*[a, b]

10804

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

Chem. Eur. J. 2011, 17, 10804 10812

CONCEPT

Abstract: The multiple functional groups and unique


two-dimensional (2D) morphology make chemically
modified graphene (CMG) an ideal template for the
construction of 2D nanocomposites with various organic/inorganic components. Additionally, the recovered
electrical conductivity of CMG may provide a fast-electron-transport channel and can thus promote the application of the resultant nanocomposites in optoelectronic
and electrochemical devices. This Concept article summarizes the different strategies for the bottom-up fabrication of CMG-based 2D nanocomposites with small organic molecules, polymers, and inorganic nanoparticles,
which represent the new directions in the development
of graphene-based materials.
Keywords: graphene graphene oxide nanocomposites nanostructures

Introduction
Two-dimensional (2D) nanostructures, which possess a high
degree of anisotropy with nanoscale thickness and infinite
length in other dimensions, have attracted tremendous attention due to their unique morphology associated with
prominent physical properties and potential applications,
such as in electronics, sensing, catalysis, energy storage, and
conversion.[1] Various organic and inorganic 2D nanostructures, such as porphyrin, C60, titanium oxide, and transitionmetal nanosheets, have been reported recently.[2] In general,
there are two strategies to prepare 2D nanosheets. One is
the top-down method, which typically involves the delamination of bulk materials with layered structures;[2a,b] however, the bottleneck of this method is the extremely low yield
of product. The other method is based on the bottom-up
fabrication, in which the nanosheets can be obtained from
the anisotropic assembly of inorganic or organic precursors
in a 2D manner. In the past few years, the latter approach
has shown great promise for the convenient and reproducible fabrication of 2D nanomaterials with a large diversity of
functions.[2ci]
The discovery of graphene has triggered intensive research work in the last few years, since this one-atom-thick
hexagonal carbon sheet possesses unique 2D morphology
and intriguing physical properties.[1] It is undeniable that the
pristine defect-free graphene exhibits unprecedented high
[a] Dr. D. Wu, Dr. F. Zhang, Dr. P. Liu, Dr. X. Feng
College of Chemistry and Chemical Engineering
Shanghai Jiao Tong University, 800 Dongchuan Road
Shanghai, 200240 (P. R. China)
[b] Dr. X. Feng
Max Planck Institute for Polymer Research, Ackermannweg 10
55128, Mainz (Germany)
E-mail: feng@mpip-mainz.mpg.de

Chem. Eur. J. 2011, 17, 10804 10812

charge carrier mobility, mechanical robustness, thermal conductivity, and so on.[1] Nevertheless, the cost-effective synthesis of high-quality graphene on a large scale, both by
physical and chemical means, remains a major challenge.
Graphite oxide, readily accessible by the oxidation of graphite with strong acid and oxidant, has been known for more
than hundred years.[3] The exfoliation of graphite oxide generates single-layer graphene oxide (GO), which can be further subjected to chemical reduction to produce reduced
graphene oxide (RGO) with partial recovery of its electrical
properties.[4] In this regard, GO, RGO, and their functional
derivatives, generally termed as chemically modified graphenes (CMGs),[5] have been attractive to chemists and material
scientists due to their easy accessibility and processability.[6]
Due to the termination of hydroxyl and carboxyl groups
in the plane or at the edge, CMGs are negatively charged
and thus have good dispersibility in aqueous solution, given
that the steric hindrance and electrostatic repulsion caused
by these oxygen-containing substitutions can effectively prevent the re-aggregation of CMGs. In addition, these oxygenbased functional groups can serve as anchor points to
couple with organic and inorganic species through covalent
or non-covalent interactions. Similar to pristine graphene,
CMGs also possess 2D character with a large aspect ratio,
high specific surface area (theoretical value of 2600 m2 g 1
for graphene), and excellent flexibility. Therefore, these
unique characteristics qualify CMGs as a promising template for the anisotropic assembly with organic or inorganic
components in solution. Towards this end, 2D sandwich-like
nanocomposites incorporated with graphene sheets have
been successfully synthesized[7] that have a large aspect
ratio, high surface area, and high monodispersity. The graphene sheet in 2D nanocomposites may thus offer an additional platform for the fast transportation of charge carriers,
which can lead to enhanced performance in various applications, such as catalysis, sensing, supercapacitors, batteries,
and fuel cells. In this Concept article, we will summarize different bottom-up approaches for the fabrication of 2D graphene-based nanosheets. These 2D nanosheets should have
individual dispersibility and can be distinctly visualized by
microscopy technology. Nanocomposites of graphene that
suffer from strong aggregation will not be included. The typical physical properties and potential applications of 2D graphene-based materials will be discussed in context. It is anticipated that this Concept article will arouse more attention
towards graphene-based 2D nanomaterials and encourage
future work to push forward the advancement of this emerging area.

Fabrication of CMG-Based 2D Nanocomposites


with Organic Components
Chemically modified graphene can be viewed as a 2D polymer containing extended aromatic frameworks and multiple

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

www.chemeurj.org

10805

X. Feng et al.

functional groups. Therefore, the introduction of specific organic species at the edge or on the in-plane surface of
CMGs can be achieved through covalent bonds or noncovalent forces, such as aromatic and ionic interactions. In this
section, 2D nanocomposites of CMGs with organic components will be preferentially discussed according to the different driving forces for their formation.
Noncovalent interactions: Owing to the pp interactions between aromatic units, aromatic molecules such as polycyclic
aromatic hydrocarbons (PAHs) and conjugated polymers
have the capability to directly intercalate with CMGs. To
keep the good dispersibility of resultant materials in aqueous solution, aromatic molecules containing anionic groups
are frequently employed to exert the electrostatic repulsion
and thus prevent the re-aggregation of the graphene sheets.
Recently, we have reported the fabrication of 2D nanocomposites of RGO with the sodium salt of pyrene-1-sulfonic
acid (PyS, electron donor, Figure 1) or the diasodium salt of
3,4,9,10-perylenetetracarboxylic diimide bisbenzenesulfonic
acid (PDI, electron acceptor, Figure 1).[8] Both PyS and PDI
have extended aromatic backbones that can strongly immobilize them onto the graphene surface by means of pp interactions. In addition, the negative charges on both molecules can stabilize the graphene dispersion through electrostatic repulsion forces. After the suspension was deposited
on mica, mainly single or double layer RGO sheets were attained. Atomic force microscopy (AFM) investigations indi-

Figure 1. a) Schematic illustration of aqueous dispersions of 2D nanocomposites of RGO-PyS or RGO-PDI. AFM images and cross-section
graphs of b) RGO-Pys and c) RGO-PDI dispersion dip-coated on mica.

10806

www.chemeurj.org

cated that the thickness of a single layer RGO-PyS or


RGO-PDI was  1.7 nm (Figure 1). On this basis, it was supposed that the aromatic molecules mainly arranged face-on
on both sides of RGO sheet in a sandwich-like manner. Due
to this stacking feature, the significant charge-transfer effects between donor/acceptor and RGO led to tunable electronic properties of the nanocomposites. It was also remarkable to note that the thermal annealing of nanocomposite
films at high temperature resulted into thermal reduction of
RGO sheets with dramatically increased conductivities
(>1100 S cm 1 at 1000 8C), about twice as high as that of
RGO (517 S cm 1).
In a similar approach, nanocomposites of conjugated
polymers and CMG sheets can be also prepared through
pp interactions. For example, Shi et al. reported the use of
sulfonated polyaniline (SPANI) as the stabilizer for the reduction of GO sheets and obtained a homogeneous black
dispersion of SPANI/RGO sheets as the product.[9] Later,
Niu et al. obtained highly conductive graphene-based nanocomposites by reducing a mixture of polyACHTUNGRE[2,5-bis(3-sulfonatopropoxy)-1,4-ethynylphenylene-alt-1,4-ethynylphenylene]
and GO with hydrazine monohydrate.[10] For both cases,
sandwich-like structures of CMGs with conjugated polymers
on both sides were attained.
Given the negatively charged nature of CMGs, positively
charged organic molecules can favor assembly with CMGs
through ionic interactions. Shi et al. prepared RGO complexes with positively charged 5,10,15,20-tetrakis(1-methyl4-pyridinio)porphyrin (TMPyP) in aqueous solution.[11] The
main driving forces for the complex formation consist in the
electrostatic and pp stacking interactions between TMPyPs
and RGO. A large bathochromic shift in the absorption
spectrum of TMPyP was observed when RGO suspension
was added. This can be reasonably explained by the flattening effect of TMPyP molecules when they are attached to
the surface of RGO sheets (Figure 2).
Covalent bonds: The functionalization of carbon nanotubes
(CNTs) by covalent anchoring of organic molecules has
been well established in the last decades.[12] This strategy has
been successfully adapted to fabricate functionalized graphene materials. As a typical example, Gao et al. demonstrated the functionalization of GO by means of one-step nitrene chemistry by simply mixing as-prepared GO with
functional azides in N-methyl-2-pyrrolidone (NMP) at
160 8C for 18 h (Scheme 1).[13] The highly reactive nitrene intermediates were generated upon heating the azides at elevated temperature, and some of them were bound to the
graphene framework by cycloaddition to form aziridine
rings. Meanwhile, thermal reduction of GO sheets proceeded during the reaction process.
By applying the above method with different azides reagents, various types of electrically conductive 2D nanocomposites, such as hydroxyl-, carboxyl-, amino-, bromine-,
long-alkyl-chain-, polystryene-, and poly(ethylene glycol)functionalized graphene could be obtained. These 2D nanocomposites show good dispersibility in polar or nonpolar

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

Chem. Eur. J. 2011, 17, 10804 10812

Chemically Modified Graphene

CONCEPT
lecular brushes. The growing process of the brushes was
clearly demonstrated by AFM (Figure 3), showing that the
graphene sheets remain isolated from each other after polymer grafting. It was interesting to note that the molecular
weight, height, stability, and dispersibility of glycidyl methacrylate (GMA)-functionalized GO brushes (GO-g-PGMA)
could be tailored through the reaction time (Figure 3). In
addition, various 2D macromolecular brushes could be obtained, which ranged from polar to apolar, water-soluble to
oil soluble, acidic to basic, and from functional to common
polymers, thus facilitating the design, synthesis, and application in biomimetic coatings and nanocomposites.

Fabrication of CMG-Based 2D Nanocomposites


with Inorganic Components

Figure 2. a) Schematic illustration of TMPyP adsorbed on the RGO


sheet, b) absorption spectra recorded during the titration of aqueous solution of TMPyP (1 mm, 3 mL) with various volumes of RGO dispersion
(0.25 mg mL 1). Copyright  2009, American Chemical Society.

Scheme 1. General strategy for the preparation of functionalized GOs by


nitrene chemistry and the further modifications. Copyright  2010,
American Chemical Society.

solvents, which make them easily processable for fabricating


more complex nanocomposites with organic/inorganic components by the available solution techniques (Scheme 1).
When polymer is covalently grafted on the graphene surface, the resultant CMG polymer sheet can be regarded as
2D macromolecular brushes incorporating a flat graphene
backbone. Recently, Gao at al. reported another strategy for
the synthesis of 2D-CMG-based macromolecular brushes by
using free-radical polymerization (FRP).[14] The widely used
FRP cannot only be simply operated, but it can also be used
for most vinyl monomers. During the reactions, the radicals
generated from the vinyl monomers, including acrylates,
methacrylates, styrenics, acrylamides, and 4-vinylpyridine
(VP), could directly react with the carboncarbon double
bonds in GO/RGO to form the CMG-based 2D macromo-

Chem. Eur. J. 2011, 17, 10804 10812

Inorganic nanoparticles including metal and metal oxide


nanoparticles as well as quantum dots (QDs) have attracted
enormous attention due to their unique catalytic, magnetic,
biologic, and optoelectronic properties.[15] The assembly of
inorganic nanoparticles on the surface of conductive CMGs
not only avoids the agglomeration of nanoparticles, but also
benefit their applications in which conductivity needs to be
a significant concern. To integrate their unique features, fabrication of 2D nanocomposites composed of CMGs and inorganic nanomaterials has been intensively pursued in the
past few years. One of the most common strategies to construct these sandwich-like 2D nanocomposites is to directly
assembly CMGs with pre-prepared inorganic nanoparticles.
On the other hand, the in situ growth of inorganic nanoparticles after the adsorption of their precursor salts on the surface of CMGs offers an alternative approach towards the
CMG based inorganic nanocomposites. Thereby, various
CMGs including GO, RGO, modified GO/RGO and exfoliated graphene have been explored as suitable 2D supports
for such purposes.
GO or RGO as a 2D support: By sonicating the mixture of
GO and the colloidal suspension of the corresponding metal
oxide nanoparticles, Kamat et al. obtained TiO2/GO[16] and
ZnO/GO[17] nanosheets. In their work, GO sheets were applied as the adsorbent of the pre-prepared nanoparticles.
The driving force for the sandwich-like 2D nanocomposite
formation may originate from the physisorption, electrostatic binding, and charge-transfer interactions between GO
and inorganic nanoparticles; however, this was not specified
in this work. The morphology investigations of these nanocomposites indicate that the distribution of metal oxide
nanoparticles on GO is not homogeneous, which can be due
to the fact that the direct binding of inorganic nanoparticles
on the surface of GO lacks selectivity.
As discussed in pervious section, the oxygenated groups
give rise to the negatively charged nature of GO and the
binding with positively charged species. The precursors for
inorganic nanoparticles are usually metal salts. Therefore,
the electrostatic adsorption of inorganic salts on GO, associ-

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

www.chemeurj.org

10807

X. Feng et al.

density and uniformity of inorganic nanoparticles. Alternatively, the modification of GO/


RGO with suitable organic surfactants or stabilizer agents
may provide a complementary
approach for the fabrication of
graphene-based 2D nanocomposites. To a certain extent, one
may expect that not only the
aggregation of the 2D nanosheets can be hampered by
choosing the right combination
system, but also the functionalized graphene surface may additionally promote the conFigure 3. AFM height images of GO and GO-g-PGMA 2D brushes at different reaction time (size for all
images: 1 mm x 1 mm). Copyright  2010, American Chemical Society.
trolled nucleation and growth
of inorganic nanoparticles.
Inspired by the colloidal stabilization of CNTs using
ated with in situ chemical conversion, to give the correanionic surfactants,[20] Aksay et al. prepared 2D TiO2/GO
sponding nanoparticles represents a controllable and repronanocomposites based on thermally expanded GO which
ducible means to produce CMG-based 2D inorganic nanowas decorated with sodium dodecyl sulfate (SDS) by means
composites.[18] For example, Yang et al. transformed the carof hydrophobic interactions.[21] It was believed that the hyboxylic acid groups of GO to carboxylate anions by treat[18a]
ment with NaOH.
drophobic tails of SDS could adsorb onto the graphene surAfter addition of iron salts with
face and, in this way, the hydrophobic graphene resided in
FeIII/FeII to the condensed suspension of GO, 2D nanocomthe hydrophobic domains of the SDS micelles, leading to faposites of Fe3O4 nanoparticles on GO were preferentially
vorable dispersion of graphene sheets in aqueous solution
formed upon the additional treatment with aqueous NaOH
(Figure 5). On the other hand, the hydrophilic head groups
solution. Despite the occurrence of aggregation in some ocof SDS could interact with TiCl3, thus serving as the molecucasions, the size of Fe3O4 nanoparticles on GO was typically
in the range of 2 to 4 nm with a narrow size distribution,
lar template for the controlled nucleation and growth of
which suggested that a large proportion of Fe3O4 nanopartiTiO2 nanoparticles. As the result, nanocrystalline TiO2 with
cles were immobilized on the GO surface with strong bindcontrolled crystalline phase (i.e., rutile and anatase) were
ing capability.
homogeneously deposited on the graphene sheet through
The in-situ growth approach can be also applied to procooperative interactions between the surfactant, graphene,
duce 2D QDs/RGO nanocomposites. Cao and co-workers
and nanocrystalline TiO2.
first mixed GO and CdACHTUNGRE(CH3COO)22 H2O in dimethylsulfoxAs one of the most frequently used polymer stabilizers
for CNTs, poly(N-vinyl-2-pyrrolidone) (PVP) can be also
ide (DMSO) and then heated the suspension in an autoclave
adopted to functionalize CMGs by means of hydrophobic
at 180 8C for 12 h.[19] During the thermal treatment, DMSO
interactions. The resulting PVP/CMG can form highly stable
could serve as both a solvent and source of sulfur. As the
colloidal suspensions in water, ethanol, and dimethylformresult, the reduction of GO and the deposition of CdS on
amide.[22] Wang et al. used this advantage to fabricate highRGO occurred simultaneously, which led to a one-pot preparation of CdS/RGO. Transmission electron microscopy
quality Pt-on-Pd bimetallic nanodendrites supported on
(TEM) images (Figure 4) revealed that the nanocomposites
RGO sheets (TP-BNGN).[22] In their work, PVP-functionalconsist of 2D nanosheets with decoration of individually isoized RGO was firstly prepared through the reduction of GO
lated CdS QDs on RGO. In the X-ray diffraction pattern of
CdS/RGO, there are three main peaks at scattering angles
of 26.506, 43.960, and 52.1328, corresponding to the (111),
(220), and (311) crystal planes of CdS, respectively. This
result shows that the CdS QDs on the graphene sheet are of
a blende structure (JCPDS 10-0454).
Modified GO/RGO as a 2D support: In spite of its ready accessibility for the fabrication of 2D nanocomposites,
GO/RGO suffers from a disadvantage, as the distribution of
inorganic nanoparticles on GO/RGO is strongly dependent
on the number and density of oxygen-containing functional
groups, which in turn unavoidably influences the deposition

10808

www.chemeurj.org

Figure 4. a) TEM image of a CdS/RGO nanocomposite with densely


coated CdS QDs. b) TEM image of a CdS/RGO nanocomposite sparsely
coated with CdS QDs, showing natural wrinkles of a single graphene
sheet. c) High-resolution TEM image of CdS crystals on a graphene
sheet.

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

Chem. Eur. J. 2011, 17, 10804 10812

Chemically Modified Graphene

CONCEPT

Figure 5. a) Illustration of preparing a TiO2/GO nanocomposite from


SDS-modified graphene sheets. (b) TEM and (c) SEM image of a rutile
TiO2/GO nanocomposite. (d) TEM image of an anatase TiO2/GO nanocomposite. Copyright  2010, American Chemical Society.

with hydrazine in the presence of PVP. Single-crystal Pd


nanoparticles with a size of about 3 nm were then anchored
on the PVP-functionalized RGO through the in situ reduction of H2PdCl4. In the third step, Pd-supported RGO was
used as seeds to direct the dendritic growth of Pt upon the
reduction of K2PtCl4 by ascorbic acid in an aqueous solution. Remarkably, both AFM and TEM images demonstrate
that the monodispersed 2D nanosheets comprise Pt-on-Pt
bimetallic nanodendrites with an average size of 15 nm that
are uniformly distributed on the functionalized graphene
surface (Figure 6). Moreover, the number of Pt branches on
nanodendrites could be easily adjusted by modifying the reaction parameters, such as the concentration of precursors,
thus providing a facile means to tune their electrocatalytic
activity in methanol oxidation.
In another strategy, positively charged small molecules
and polymers can be used as linkers to couple CMGs with
inorganic nanoparticles, both of which are negatively
charged, which may benefit the fabrication of 2D nanocomposites. For instance, cationic polyelectrolyte-functionalized
CMG sheets were obtained by modifying RGO with polyACHTUNGRE(diallyldimethyl ammonium chloride) (PDDA) associated
with PVP as a stabilizer.[23] It turned out that citrate-capped
gold and Au@Pd hybrid nanoparticles with high loading
could be uniformly deposited on PDDA-functionalized graphene sheets.
Very recently, we reported the fabrication of 2D sandwich-like graphene-based mesoporous silica (GM-silica)
sheets with cationic surfactant cetyl trimethylammonium
bromide (CTAB)-functionalized GO as the template
(Figure 7).[24] In this work, CTAB was adapted to modify
GO in alkaline solution through strong electrostatic interactions. Thereby, CTAB could not only provide the template
for the mesoporous silica formation through the hydrolysis
of tetraethylorthosilicate (TEOS), but it also allowed the
tight coupling between GO and the inorganic species. Thus,
GO-based mesoporous silica (GOM-silica) nanosheets with
could be attained in high yield; these sheets underwent fur-

Chem. Eur. J. 2011, 17, 10804 10812

Figure 6. AFM images of a) RGO/Pd nanosheets and b) RGO/bimetallic


nanodendrite nanosheets. c) TEM and d) HRTEM images of the RGO/
bimetallic nanodendrite nanosheets. The circled parts in panel d) denote
Pd nanoparticles. Copyright  2009, American Chemical Society.

ther transformation into graphene-based mesoporous silica


nanosheets by thermal treatment. The resultant nanosheets
possess a large aspect ratio, mesoporous structure, high surface area, and high monodispersity. Microscopy images unambiguously reveal that the resulting nanosheets preserve
the exact same morphology as GO. Figure 7 f and 7g further
demonstrate a uniform thickness of (28  1) nm with low
roughness for GOM-silica nanosheets. Notably, the thickness
of the sheets could be tuned by simply adjusting the ratio of

Figure 7. Graphene oxide based mesoporous silica (GOM-silica) sheets.


a) Fabrication process for GOM-silica sheets. b),c) SEM and d),e) TEM
images reveal the flat GOM-silica sheets. f) AFM image and g) thickness
analysis taken around the white line in f) reveal a thickness of 28 nm for
GOM-silica sheets.

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

www.chemeurj.org

10809

X. Feng et al.

GO to TEOS during the fabrication process. More importantly, the graphene-based mesoporous silica nanosheets
could further serve as template to prepare graphene-based
mesoporous carbon and metal oxide nanosheets by means
of nanocasting technology. We believe that this method will
certainly broaden the accessibility and application of different graphene-based 2D nanocomposites.
Biomacromolecules with multiple functional groups also
show potential in the fabrication of CMG-based 2D nanocomposites. In a recent report, GO and RGO were decorated with thiolated d(GT)29SH DNA oligomers by means of
pp interactions, similar to the case of DNA wrapping on
carbon nanotubes.[25] The thiol groups tagged on DNA
strands could then act as the anchor points for the 2D selfassembly of pre-prepared gold nanoparticles on GO or
RGO.[26] Later, the authors used the Tyr residues of bovine
serum albumin (BSA) as both reductant and stabilizer for
GO.[27] It turned out that the resultant BSA-RGO nanocomposites could be further used as template for the uniform
2D assembly of pre-synthesized Au, Pt, Pd, Ag, and latex
(polystyrene sphere) nanoparticles due to the strong interactions between multiple thiol, amine, and imidazole groups
on BSA potine and the nanoparticles.
Exfoliated graphene as a 2D support: Different from previous approaches that use GO, RGO, or modified GO/RGO
as the template for the synthesis of 2D nanocomposites, Dai
et al. reported a step-wise strategy to cover high-quality graphene sheets with Ni(OH)20.75 H2O nanoparticles with
crystalline structure.[28] In the first step, small nanoparticles
of Ni(OH)20.75 H2O were grown on exfoliated graphene
sheets after the adsorption of corresponding metal salt precursors in N,N-dimethylformamide (DMF)/water (10:1). Afterwards, the nanocomposites were hydrothermally treated
at 180 8C in water. It was interesting to find that the small
particles of Ni(OH)20.75 H2O transformed into hexagonal
nanoplates during the hydrothermal process. The morphology, size, and crystallinity of resulting nanocrystals were dependent on the oxidation degree of graphene. For the graphene with lower oxidation degree, large, single-crystalline
hexagonal Ni(OH)2 nanoplates could be obtained that
showed a typical size of several hundred nanometers and
thickness of less than 10 nm (Figure 8). XRD revealed that
the nanoplates were crystalline b-Ni(OH)2 (Figure 8 d). In
contrast, for highly oxidized GO, small Ni(OH)2 nanoparticles remained at their original positions after the hydrothermal treatment. It was assumed that the large amount of
oxygen functional groups, such as carboxylic, hydroxyl, and
epoxy groups, interacted strongly with the anchored inorganic nanoparticles, which therefore hampered the diffusion
and recrystallization of nanoparticles.[29]

Figure 8. a) SEM image of Ni(OH)20.75 H2O particles uniformly coated


on GS after the first step of growth at 80 8C. b) SEM image of Ni(OH)2/
GS after the second step of simple hydrothermal treatment of the product depicted in a) at 180 8C. c) TEM image of hexagonal Ni(OH)2 nanoplates formed on top of GS. d) XRD spectrum of a packed thick film of
hexagonal Ni(OH)2 nanoplates on GS. Copyright  2010, American
Chemical Society.

nanomaterials may open up enormous opportunities for applications across the fields of biology, medicine, chemistry,
and physics. Regarding the CMG-based 2D nanocomposites,
the recovered electrical conductivity of graphene may provide additional electron-transport pathways and can thus
promote their applications in optoelectronic and electrochemical devices. In the following section, we will specifically discuss a few examples of CMG-based 2D nanocomposites for the applications in electrochemical energy storage
and conversion, as well as catalysis and sensing.
Supercapacitors: The electric double-layer capacitor
(EDLC), also known as supercapacitor, is an electrochemical capacitor with much higher energy density than the traditional electrolytic capacitors. Dai et al. investigated their
Ni(OH)2/GS nanocomposite as the electrode material for
electrochemical pseudocapacitors. It turned out that the
lightly oxidized, highly conductive graphene sheets with
single-crystalline Ni(OH)2 hexagonal nanoplates exhibited a
high specific capacitance of  1335 F g 1 at a current density
of 2.8 A g 1 and  953 F g 1 at 45.7 A g 1 with excellent cyclability, outperforming that of Ni(OH)2 nanoparticles
grown on highly oxidized GO and the simple mixture of
Ni(OH)2 nanoplates with graphene sheets.[29a] This work suggests that the quality of graphene as well as the morphology
and crystallinity of the 2D nanomaterials are all the critical
issues required to achieve high electrochemical performance
materials for supercapcaitors.

Properties and Applications


Due to their unique structural features such as large aspect
ratio, ultrathin thickness, and high specific surface area, 2D

10810

www.chemeurj.org

Lithium-ion batteries: Lithium-ion batteries are currently


the best portable energy storage device for the consumer
electronic market. In the last decades, carbonaceous materi-

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

Chem. Eur. J. 2011, 17, 10804 10812

Chemically Modified Graphene

CONCEPT

als have been widely used as the anode materials for lithium-ion batteries. The electrochemical performances of lithium storage including the charge/discharge capacity, columbic efficiency, and cycle stability, depend strongly upon the
microstructure of the carbon. CMG-based 2D nanocomposites are very attractive anode materials for lithium storage
due to their unique flat nanostructures and electronic properties. For instance, a high first reversible capacity of
915 mAh g 1 at a rate of C/5 (one lithium per six formula
units (LiC6) was achieved in the case of graphene-based
mesoporous carbon (GM-C) nanosheets.[24] It was further
found that the reversible capacities of GM-C sheets were
stabilized at about 770 mAh g 1, delivering 84 % capacity retention after 30 cycles (Figure 9). GM-C sheets also exhibited excellent rate performance in stark contrast to the traditional non-graphitic carbon, which typically show continuous
and progressive capacity decay along with cycling processes
at various rates. Such good electrochemical performances of
GM-C nanosheets for lithium storage was attributed to their
high surface area, thin thickness, and numerous mesopores,
which were favorable for the accessibility of the electrolyte,
rapid diffusion of lithium ions, and host uptake. Furthermore, the graphene layer within each nanosheet could act as
mini-current collectors homogeneously dispersed in the electrode, which facilitated the fast transport of electrons during
the chargedischarge processes owing to the high electrical
conductivity of graphene.
In another case, Mn3O4/RGO 2D nanocomposites were
prepared by two-step solution-phase reactions and exhibited
a high specific capacity up to  900 mAh g 1 [29b] (close to the
theoretical capacity of Mn3O4), good rate capability, and
cycle stability. The improved electrochemical performance
of Mn3O4/RGO nanocomposites was ascribed to the intimate interaction between graphene and Mn3O4 nanoparti-

cles,[29b] enabling the charge carriers to be effectively and


rapidly transported from Mn3O4 to the current collector
through the highly conducting graphene network. On the
other hand, the aggregation of Mn3O4 nanoparticles was
prevented by RGO sheets, which led to the increased cycle
stability. Therefore, these results clearly suggest that graphene-based 2D nanocomposites can be promising candidates for high-capacity, low-cost, and environmentally
friendly anodes for lithium ion batteries.
Sensor and other applications: CMG-based 2D nanocomposites are also good candidates as sensing materials for gas,
ions, and biomolecules, since the target molecules can be
easily accessed due to their large specific surface areas. For
a typical example, Shi and co-workers found that the coordination reaction between 2D TMPyP/RGO nanocomposites
and CdII ions proceeded completely in 8 min under ambient
conditions, while pure TMPyP required about 20 h to reach
its equilibrium under the same conditions.[11] Compared with
pure TMPyP, the selectivity of the 2D TMPyP/RGO probe
towards CdII ions was also improved. The enhanced performance was attributed to the efficient anchoring of TMPyP
on RGO sheets, which led to flattening feature of TMPyP
on graphene.
CMG-based 2D nanocomposites have also been used in
many other fields such as catalysis[22] and solar cells.[8] In
these cases, not only the outstanding electronic properties
but also the unique 2D morphology play pivotal role on the
performance of the resulting nanocomposites. It is no doubt
that by taking advantages of these 2D nanocomposite materials, new applications will be further developed in the near
future.

Summary and Outlook

Figure 9. Electrochemical performance of GM-C sheets for lithium ion


storage. a) Lithium insertion and extraction in GM-C sheets, where graphene acts as mini-current collectors during discharge and charge processes, facilitating the rapid diffusion of electrons during cycling processes. b) First two dischargecharge curves (black: first discharge, red: first
charge, green: second discharge, blue: second charge) and c) cycle performance (black: discharge, red: charge) of GM-C sheets at a rate of C/5.
The inset is the rate capability of GM-C sheets at various rates of C/5,
1C, 5C, 10C, and 20C.

Chem. Eur. J. 2011, 17, 10804 10812

In this Concept article, the recent developments in the


bottom-up fabrication of various CMG-based 2D nanocomposite materials are described. The above results undoubtedly indicate that the 2D nanocomposites based on CMG
sheets are very attractive to material scientists due to their
striking features and potential applications. Either the preprepared or in situ grown organic and inorganic nanostructures can be loaded on the surface of CMG sheets through
covalent or non-covalent forces. The diversified choices of
the available molecular building blocks and fabrication
methods will certainly guarantee the synthesis of new CMGbased 2D nanocomposites with unique properties. Nevertheless, it is undeniable that there remain some important
issues need to be overcome during the fabrication of 2D
nanomaterials for which the nature of CMG in solution can
be the major reason. For example, the shape and size of the
CMG-based 2D nanocomposites are strongly governed by
the morphology of CMG substrates, while the production of
CMGs with desired size and shape constitutes a great challenge.[30] In addition, the attachment of organic or inorganic
components on the surface of CMGs is more like a random

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

www.chemeurj.org

10811

X. Feng et al.

process. The uniform assembly of various nanostructures on


CMGs with designed patterns will call for the rational selection of fabrication methods and careful control over the
processing conditions. Since the morphology of 2D nanomaterials is the key factor that determines the performance in
applications,[7] it is predictable that more effort will be devoted to the morphology control of CMG-based 2D nanocomposites in the near future. The work on CMG-based 2D
nanocomposites is currently an exciting field that interfaces
disciplines such as chemistry, physics, materials science, and
engineering. It is indeed exciting to stand at the point where
we see one of the fastest developing research fields of graphene chemistry, which will surely benefit the human society
in near future owing to the excellent performance of these
functional materials.

[7]
[8]

[9]
[10]
[11]
[12]

[13]
[14]
[15]

Acknowledgements
The authors acknowledge the support by the 985 project of Shanghai
Jiao Tong University and BASF.

[1] a) K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang,


S. V. Dubonos, I. V. Grigorieva, A. A. Firsov, Science 2004, 306,
666 669; b) K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang,
M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos, A. A. Firsov,
Nature 2005, 438, 197 200; c) A. H. Castro Neto, F. Guinea,
N. M. R. Peres, K. S. Novoselov, A. K. Geim, Rev. Mod. Phys. 2009,
81, 109.
[2] a) Y. Omomo, T. Sasaki, Wang, M. Watanabe, J. Am. Chem. Soc.
2003, 125, 3568 3575; b) M. Osada, T. Sasaki, J. Mater. Chem. 2009,
19, 2503 2511; c) Z. C. Wang, Z. Y. Li, C. J. Medforth, J. A. Shelnutt, J. Am. Chem. Soc. 2007, 129, 2440 2441; d) Y. Chen, K. Li, W.
Lu, S. S.-Y. Chui, C.-W. Ma, C.-M. Che, Angew. Chem. 2009, 121,
10 093 10 097; Angew. Chem. Int. Ed. 2009, 48, 9909 9913; e) T.
Wakahara, M. Sathish, K. Miyazawa, C. P. Hu, Y. Tateyama, Y.
Nemoto, T. Sasaki, O. Ito, J. Am. Chem. Soc. 2009, 131, 9940 9944;
f) H. G. Yang, G. Liu, S. Z. Qiao, C. H. Sun, Y. G. Jin, S. C. Smith, J.
Zou, H. M. Cheng, G. Q. Lu, J. Am. Chem. Soc. 2009, 131, 4078
4083; g) J. S. Son, X.-D. Wen, J. Joo, J. Chae, S.-i. Baek, K. Park,
J. H. Kim, K. An, J. H. Yu, S. G. Kwon, S.-H. Choi, Z. Wang, Y.-W.
Kim, Y. Kuk, R. Hoffmann, T. Hyeon, Angew. Chem. 2009, 121,
6993 6996; Angew. Chem. Int. Ed. 2009, 48, 6861 6864; h) G.
Fernndez, F. Garcia, F. Aparicio, E. Matesanz, L. Sanchez, Chem.
Commun. 2009, 7155 7157; i) D. Q. Wu, R. L. Liu, W. Pisula, X. L.
Feng, K. Mllen, Angew. Chem. 2011, 123, 2843 2846; Angew.
Chem. Int. Ed. 2011, 50, 2791 2794.
[3] a) B. C. Brodie, Philos. Trans. R. Soc. London 1859, 149, 249 259;
b) L. Staudenmaier, Ber. Dtsch. Chem. Ges. 1898, 31, 1481 1487;
c) G. Charpy, C. R. Hebd. Seances Acad. Sci. 1909, 148, 920 923;
d) W. S. Hummers, R. E. Offeman, J. Am. Chem. Soc. 1958, 80,
1339; e) H. P. Boehm, A. Clauss, G. O. Fischer, U. Z. Hofmann, Z.
Anorg. Allg. Chem. 1962, 316, 119 127.
[4] a) N. A. Kotov, I. Dekany, and J. H. Fendler, Adv. Mater. 1996, 8,
637 641; b) S. Stankovich, D. A. Dikin, G. H. B. Dommett, K. M.
Kohlhaas, E. J. Zimney, E. A. Stach, R. D. Piner, S. T. Nguyen, R. S.
Ruoff, Nature 2006, 442, 282 286.
[5] D. R. Dreyer, S. Park, C. W. Bielawski, R. S. Ruoff, Chem. Soc. Rev.
2010, 39, 228 240.
[6] a) Y. W. Zhu, S. Murali, W. W. Cai, X. S. Li, J. W. Suk, J. R. Potts,
R. S. Ruoff, Adv. Mater. 2010, 22, 3906 3924; b) T. Kuilla, S.
Bhadra, D. H. Yao, N. H. Kim, S. Bose, J. H. Lee, Prog. Polym. Sci.

10812

www.chemeurj.org

[16]
[17]
[18]

[19]
[20]

[21]

[22]
[23]
[24]

[25]

[26]
[27]
[28]
[29]

[30]

2010, 35, 1350 1375; c) D. Cai, M. Song, J. Mater. Chem. 2010, 20,
7906 7915.
a) Y. X. Xu, G. Q. Shi, J. Mater. Chem. 2011, 21, 3311 3323; b) H.
Bai, C. Li, G. Shi, Adv. Mater. 2011, 23, 1089 1115.
a) Q. Su, S. Pang, V. Alijani, C. Li, X. Feng, K. Mllen, Adv. Mater.
2009, 21, 3191 3195; b) J. Liu, W. Yang, L. Tao, D. Li, C. Boyer,
T. P. Davis, J. Polym. Sci. Polym. Chem. 2010, 48, 425 433; c) W.
Hong, H. Bai, Y. Xu, Z. Yao, Z. Gu, G. Shi, J. Phys. Chem. C. 2010,
114, 1822 1826.
H. Bai, Y. Xu, L. Zhao, C. Li, G. Shi, Chem. Commun. 2009, 1667
1669.
H. Yang, Q. Zhang, C. Shan, F. Li, D. Han, L. Niu, Langmuir 2010,
26, 6708 6712.
Y. Xu, L. Zhao, H. Bai, W. Hong, C. Li, G. Shi, J. Am. Chem. Soc.
2009, 131, 13490 13497.
a) D. Tasis, N. Tagmatarchis, A. Bianco, M. Prato, Chem. Rev. 2006,
106, 1105 1136; b) N. Karousis, N. Tagmatarchis, D. Tasis, Chem.
Rev. 2010, 110, 5366 5397.
H. He, C. Gao, Chem. Mater. 2010, 22, 5054 5064.
L. Kan, Z. Xu, C. Gao, Macromolecules 2011, 44, 444 452.
a) C. N. R. Rao, G. U. Kulkarni, P. J. Thomas, P. P. Edwards, Chem.
Soc. Rev. 2000, 29, 27 35; b) J. Dupont, J. D. Scholten, Chem. Soc.
Rev. 2010, 39, 1780 1804; c) T. K. Sau, A. L. Rogach, Adv. Mater.
2010, 22, 1781 1804; d) M. E. Franke, T. J. Koplin, U. Simon, Small
2006, 2, 36 50.
G. Williams, B. Seger, P. V. Kamat, ACS Nano 2008, 2, 1487 1491.
G. Williams, P. V. Kamat, Langmuir 2009, 25, 13869 13873.
a) X. Yang, X. Zhang, Y. Ma, Y. Huang, Y. Wang, Y. Chen, J. Mater.
Chem. 2009, 19, 2710 2714; b) L.-S. Zhang, L.-Y. Jiang, H.-J. Yan,
W. D. Wang, W. Wang, W.-G. Song, Y.-G. Guo, L.-J. Wan, J. Mater.
Chem. 2010, 20, 5462 5467; c) M. Zhang, D. Lei, Z. Du, X. Yin, L.
Chen, Q. Li, Y. Wang, T. Wang, J. Mater. Chem. 2011, 21, 1673
1676.
A. Cao, Z. Liu, S. Chu, M. Wu, Z. Ye, Z. Cai, Y. Chang, S. Wang, Q.
Gong, Y. Liu, Adv. Mater. 2010, 22, 103 106.
a) J.-M. Bonard, T. Stora, J.-P. Salvetat, F. Maier, T. Stckli, C.
Duschl, L. Forr, W. A. de Heer, A. Chtelain, Adv. Mater. 1997, 9,
827 831; b) C. Richard, F. Balavoine, P. Schultz, T. W. Ebbesen, C.
Mioskowski, Science 2003, 300, 775 778.
D. Wang, D. Choi, J. Li, Z. Yang, Z. Nie, R. Kou, D. Hu, C. Wang,
L. V. Saraf, J. Zhang, I. A. Aksay, J. Liu, ACS Nano 2009, 3, 907
914.
S. Guo, S. Dong, E. Wang, ACS Nano 2010, 4, 547 555.
Y. Fang, S. Guo, C. Zhu, Y. Zhai, E. Wang, Langmuir 2010, 26,
11277 11282.
S. B. Yang, X. L. Feng, L. Wang, K. Tang, J. Maier, K. Mllen,
Angew. Chem. 2010, 122, 4905 4909; Angew. Chem. Int. Ed. 2010,
49, 4795 4799.
a) J. Sharma, R. Chhabra, A. Cheng, J. Brownell, Y. Liu, H. Yan,
Science 2009, 323, 112 116; b) M. M. Maye, M. T. Kumara, D. Nykypanchuk, W. B. Sherman, O. Gang, Nat. Nanotechnol. 2010, 5,
116 120.
J. Liu, Y. Li, Y. Li, J. Li, Z. Deng, J. Mater. Chem. 2010, 20, 900
906.
J. Liu, S. Fu, B. Yuan, Y. Li, Z. Deng, J. Am. Chem. Soc. 2010, 132,
7279 7281.
H. Wang, J. T. Robinson, G. Diankov, H. Dai, J. Am. Chem. Soc.
2010, 132, 3270 3271.
a) H. Wang, H. S. Casalongue, Y. Liang, H. Dai, J. Am. Chem. Soc.
2010, 132, 7472 7477; b) H. Wang, L.-F. Cui, Y. Yang, H. Sanchez
Casalongue, J. T. Robinson, Y. Liang, Y. Cui, H. Dai, J. Am. Chem.
Soc. 2010, 132, 13978 13980.
J. Cai, P. Ruffieux, R. Jaafar, M. Bieri, T. Braun, S. Blankenburg, M.
Muoth, A. P. Seitsonen, M. Saleh, X. Feng, K. Mllen, R. Fasel,
Nature 2010, 466, 470 473.

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

Published online: August 18, 2011

Chem. Eur. J. 2011, 17, 10804 10812

You might also like