Você está na página 1de 13

Graphene

Graphene Oxide, Highly Reduced Graphene Oxide, and


Graphene: Versatile Building Blocks for Carbon-Based
Materials**
Owen C. Compton and SonBinh T. Nguyen*
From the Contents
1. Introduction . . . . . . . . . . . . . . . . . . . 712
2. Graphene Oxide-Based Materials . . . . 713
3. Graphene-Based Materials. . . . . . . . . 716
4. Summary and Outlook . . . . . . . . . . . 721
Isolated graphene, a nanometer-thick two-dimensional analog of
fullerenes and carbon nanotubes, has recently sparked great
excitement in the scientic community given its excellent mechanical
and electronic properties. Particularly attractive is the availability of
bulk quantities of graphene as both colloidal dispersions and
powders, which enables the facile fabrication of many carbon-based
materials. The fact that such large amounts of graphene are most
easily produced via the reduction of graphene oxideoxygenated
graphene sheets covered with epoxy, hydroxyl, and carboxyl
groupsoffers tremendous opportunities for access to
functionalized graphene-based materials. Both graphene oxide and
graphene can be processed into a wide variety of novel materials with
distinctly different morphological features, where the carbonaceous
nanosheets can serve as either the sole component, as in papers and
thin lms, or as llers in polymer and/or inorganic nanocomposites.
This Review summarizes techniques for preparing such advanced
materials via stable graphene oxide, highly reduced graphene oxide,
and graphene dispersions in aqueous and organic media. The
excellent mechanical and electronic properties of the resulting
materials are highlighted with a forward outlook on their
applications.
Graphene Oxide- and Graphene-Based Materials
small 2010, 6, No. 6, 711723 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim 711
1. Introduction
The exceptional electronic and mechanical properties and
the solution dispersibility of fullerenes and carbon nanotubes
(CNTs) have drawn considerable attention to these carbon
allotropes for nearly three decades.
[1,2]
Zero-dimensional (0D)
buckminsterfullerene, discovered by Kroto, Curl, and Smalley
in1985,
[3]
ignitedsignicant interest innanoscalecarbonaceous
materials. 1DCNTs, highlighted shortly thereafter by Iijima,
[4]
further increased enthusiasm in carbon-based nanoparticle
research. While bothfullerenes andCNTs are proposedtohave
been derived from the 2D graphene sheet,
[1]
only since its
recent isolationin2004has this parent carbonmonolayer found
its place in the limelight.
[58]
This delay in discovery can be
partially attributed to the single-atom-thick nature of the
graphene sheet, which was initially thought to be thermo-
dynamically unstable.
[9]
However, graphene is not only stable
but also exhibits impressive electronic and mechanical proper-
ties (charge-carrier mobility 250 000 cm
2
V
1
s
1
at room
temperature,
[10]
thermal conductivity 5000 W m
1
K
1
,
[11]
and mechanical stiffness 1 TPa
[12]
). Importantly, graphene
can be made in bulk quantities and in a wide range of
processable forms using simple mechanical and chemical
processes. As such, the development of new materials
exploiting the impressive properties of graphene sheets, as
indicated by the number of published articles in this area
(Figure 1), has surged in exponential fashion.
Graphene sheets comprise a 2D layer of sp
2
-hybridized
carbon atoms, arranged in a hexagonal lattice. In graphite,
adjacent graphene layers are arranged with overlapping p
z
orbitals,
[13]
whose vast number of interactions inhibit the
complete delamination of bulk graphite into individual
graphene sheets under typical mechanical actions. Attempts
tomechanically exfoliate graphite typically only result instacks
of sheets, or a fewisolated sheets in lowyields (see Section 3.1).
Chemical exfoliation strategies such as sequential oxidation
reduction of graphite often result in a class of graphenelike
materials best described as highly reduced graphene oxide
(HRG),
[14,15]
with graphene domains, defects, and residual
oxygen-containing groups on the surface of the sheets (see the
introduction to Section 3). Indeed, none of the currently
available methods for graphene production (see below) yields
bulk quantities of defect-free sheets.
In general, methods for producing graphene and HRG can
beclassiedintovemainclasses: 1) mechanical exfoliationof a
single sheet of graphene from bulk graphite using Scotch
tape,
[6,7]
2) epitaxial growth of graphene lms,
[8,16]
3) chemical
vapor deposition (CVD) of graphene monolayers,
[17,18]
4) longitudinal unzipping of CNTs,
[19,20]
and 5) reduction
of graphene derivatives, such as graphene oxide and graphene
uoride,
[5,21,22]
whichinturncanbeobtainedfromthe chemical
exfoliation of graphite. For producing gram-scale quantities of
graphene powder or large (2 cm
2
) graphene monolayers for
thefabricationof devices, methods (1) to(4) currentlyfall short.
Mechanical exfoliation using the Scotch-tape method is a
laborious procedure where the probability of nding individual
graphene sheets of good quality is often low. Epitaxial growth
can produce good-quality graphene more consistently but
requires high-vacuum conditions and specialized, expensive
fabrication systems to generate only small-area lms. While
recent advances inCVDtechniques have pavedthe way toward
the generation of graphene monolayers with large surface
areas, such methods have yet to be fully developed.
[23]
Similarly, although longitudinal unzipping of CNTs can
potentially afford bulk quantities of graphene nanoribbons
whose width is dependent on CNT diameter, this method has
only recently been discovered and its scalability has yet to be
demonstrated. At the moment, the reduction of graphene
derivatives (method (5)) stands out as the primary strategy that
can yield bulk amounts of graphenelike sheets that, albeit not
defect-free, are highly processable and can be fabricated into a
variety of materials.
Graphene oxide, HRG, and graphene occupy the interface
of macromolecules and nanoscale objects. As such, they can be
modied readily using a plethora of chemical reactions and
subsequently employed as nanoscale building blocks in the
assemblyof papers andthinlms, or as llers innanocomposites
with polymeric and/or inorganic materials. This Review
highlights the impressive range of functionalized materials
that can be prepared in a bottom-up fashion from bulk
dispersions of graphene oxide, HRG, and graphene in various
media. Readers interested in the electronic properties of
defect-free graphene sheets, chemical methods for the
preparation of HRG and graphene, and the background and
unique properties of defect-free graphene can consult the
excellent reviews by Neto et al.,
[24]
Park and Ruoff,
[25]
and
Geimand Novoselov,
[1]
respectively. To simplify the terminol-
ogyintheremainder of this Review, wewill usethegeneral term
graphene to reference the macroscopic papers, thin lms,
reviews O. C. Compton and S. T. Nguyen
[

] Dr. O. C. Compton, Prof. S. T. Nguyen


Department of Chemistry, Northwestern University
2145 Sheridan Road, Evanston, IL 60208 (USA)
E-mail: stn@northwestern.edu
[

] Support for this work was provided by the United States NSF (Grant
# DMR-0520513 and CHE-0936924) and the ARO (Grant #
W911NF-09-1-0541). O.C.C. is an ACC-NSF fellow.
DOI: 10.1002/smll.200901934
Figure 1. Annual number of publishedarticlescontaininggraphene as
a keyword. The starred value for 2009 has been projected assuming
continuation of the average monthly publication rate through August.
712 www.small-journal.com 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim small 2010, 6, No. 6, 711723
powders, andcomposites madefrombothHRGanddefect-free
graphene sheets, indicating the differences between these two
basic materials only when necessary.
2. Graphene Oxide-Based Materials
The most common route toward bulk quantities of HRG
begins with the oxidation of graphite to graphite oxide (GO).
(In referring to bulk graphite oxide, Boehm and co-workers in
the late 1950s
[26]
coined the abbreviation GO, which has
sometimes been incorrectly appropriated to denote nanoscale
graphene oxide sheets. In this Review, we will use GO in its
traditional sense.) The rst procedures for the synthesis of GO
were developed several decades ago by Brodie,
[27]
Staudenmeier,
[28]
and Hummers et al.,
[29]
and still remain in
use today with only minor modications (Table 1). The extent
of graphite oxidation, as quantied by the C:O atomic ratio, is
dependent upon the synthetic technique as well as the length of
reaction.
[30,31]
TheStaudenmaier method
[28]
typically produces
the most oxidized GO, though the entire synthesis may take
several days. Unfortunately, both it and the Brodie method
[27]
generate ClO
2
gas, which must be handled with caution due to
its high toxicity and tendency to decompose in air to produce
explosions ranging fromthe toot of a small whistle to the noise
of a recracker.
[32]
As such, the Hummers method,
[29]
with its
relativelyshorter reactiontimeandabsenceof hazardous ClO
2
,
has seen more use in current research. One drawback of the
Hummers method is potential contamination by excess
permanganate ions, which should be removed by treatment
with H
2
O
2
,
[33]
followed by washing and thorough dialysis.
The oxidation of graphite to GO breaks up the sp
2
-
hybridizedstructure of the stackedgraphene sheets, generating
defects that manifest as clear wrinkles in the stack
[34]
and
increase the distance between adjacent sheets from 3.35 A

in
graphite powder to 6.8 A

for GO powder.
[35]
This increased
spacing varies signicantly depending on the amount of water
intercalated within the stacked-sheet structure
[36]
and reduces
interactionbetweensheets, thus facilitatingthedelaminationof
GO into individual graphene oxide sheets upon exposure to
low-power sonication in water. At slightly basic pH, negatively
charged, hydrophilic oxygen-containing functional groups (see
Section 2.1) on the graphene oxide surface can stabilize
dispersions of these sheets in aqueous media.
[37]
Such
dispersions are the precursors from which many graphene
oxide- and graphene-based materials are prepared.
Thestructureof grapheneoxidesheets has beendebatedfor
decades with uncertainty pertaining to both the type and
distribution of oxygen-containing functional groups.
[38,39]
A
recent NMR study of
13
C-labelled graphene oxide
[40]
lends
strong support to the model proposed by Lerf et al.,
[41,42]
wherein the basal plane of the sheet is decorated with hydroxyl
and epoxy (1,2-ether) functional groups (Figure 2A). Carbonyl
groups arealsopresent, most likelyas carboxylic acids alongthe
sheet edge but also as organic carbonyl defects within the sheet.
These functional groups provide reactive handles for a variety
of surface-modicationreactions, which canbe used todevelop
functionalized graphene oxide- and graphene-based materials,
as will be discussed throughout this Review.
Graphene Oxide- and Graphene-Based Materials
Owen C. Compton completed his
undergraduate studies in chemistry at the West
Lafayette, IN campus of Purdue University in
2003. In 2008, he received his Ph.D. in
inorganic chemistry from the University of
California, Davis under the direction of Frank
Osterloh. He is currently an NSF-American
Competitiveness in Chemistry Postdoctoral
Fellow in the Nguyen laboratory at
Northwestern University, Evanston, IL. His
scientic interests lie in the synthesis and
application of layered solid materials such as
DionJacobsen-type perovskites and graphite,
including their exfoliation into two-dimensional
nanosheets.
During his undergraduate years,
SonBinh T. Nguyen carried out research with
Gregory Geoffroy at Penn State and in the
groups of Henry Bryndza and Steve Ittel at
DuPont Central Research. He received a
doctoral degree in chemistry under Robert
Grubbs and Nathan Lewis at Caltech, where he
held NSF and NDSEG predoctoral fellowships.
After an NSF postdoctoral fellowship with
Barry Sharpless at Scripps, SonBinh began his
independent career at Northwestern, where he
is now a Professor of Chemistry. He holds the
Directorship of the Integrated Science Program,
the Dow Chemical Company Research
Professorship, and the McCormick Professorship of Teaching Excellence.
Table 1. Methods for the oxidation of graphite to graphite oxide.
Brodie Staudenmaier Hummers Modied Hummers
Year 1859 1898 1958 1999 2004
Oxidants KClO
3
, HNO
3
KClO
3
(or NaClO
3
),
HNO
3
, H
2
SO
4
NaNO
3
, KMnO
4
, H
2
SO
4
pre-ox: K
2
S
2
O
8
,
P
2
O
5
, H
2
SO
4
NaNO
3
, KMnO
4
, H
2
SO
4
ox: KMnO
4
, H
2
SO
4
C:O ratio 2.16
[27]
N/A
[28]
2.25
[29]
1.3
[60]
1.8
[119]
2.28
[36]
1.85
[36]
2.17
[36]
Reaction time 34 days
[27]
12 days
[28]
2h
[29]
6h pre-ox 2h ox
[60]
5 days
[119]
10h
[36]
10 days
[36]
910h
[36]
Intersheet spacing [A

] 5.95
[36]
6.23
[36]
6.67
[36]
6.9
[60]
8.3
[119]
small 2010, 6, No. 6, 711723 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com 713
2.1. Dispersion of Graphene Oxide in Organic Solvents
The majority of materials discussed here are prepared
from colloidal dispersions of graphene oxide or HRG. Given
that graphene oxide is intrinsically hydrophilic, its dispersion
in water is readily achieved by sonication (see the introduc-
tion to Section 2); however, suspending graphene oxide in
organic media is not soeasily accomplished. The rst reported
strategy for preparing colloidal dispersions of graphene oxide
in organic solvents employed modication with organic
isocyanates,
[43]
where the surface- and edge-bound hydroxyl
and carboxyl groups of graphene oxide were converted into
amide and carbamate groups, respectively. The isocyanate-
modied sheets became dispersible in polar organic solvents
such as dimethyl sulfoxide (DMSO), N,N-dimethylforma-
mide (DMF), and N-methylpyrrolidone (NMP) but not in
water.
Similar to the aforementioned example, diisocyanate-
functionalized graphene oxide can be coupled to an amphi-
philic oligoester to produce amphiphilic graphene oxide that is
dispersibleinbothwater andDMF.
[44]
Suspensions of graphene
oxide have also been modied by sonication in the presence of
TiO
2
nanoparticles, which coat and stabilize the surface of the
graphene oxide sheets, preventing agglomeration in the polar
solvent ethanol.
[45]
While surface modica-
tion provides an easy means for preparing
organic dispersions, the presence of addi-
tives can complicate the subsequent pro-
cessing of materials and affect both
mechanical and electronic properties (see
Sections 2.2 and 3.1).
The need to employ surface-stabilizing
additives in the production of non-aqueous
graphene oxide suspensions has recently
been circumvented. The groups of Song
and Paredes initially reported that unmo-
died graphene oxide can be dispersed in
organic mediabysonicatinga nely ground
GO powder for approximately an hour.
Song et al. only generated dispersions in
DMF,
[46]
while the Paredes group also
prepared dispersions in ethylene glycol,
NMP, and tetrahydrofuran (THF) that
remained stable for three weeks.
[47]
However, these results could not be
reproduced by Ruoff et al., even after
24 h of sonication in DMF.
[15]
Instead, the
Ruoff group found that stable dispersions
of unmodied graphene oxide in primarily
organic media (9:1 v/v organic:water)
could only be prepared via dilution of an
aqueous dispersion of aqueous graphene
oxide with the appropriate organic solvent.
While highly polar solvents (DMF, etha-
nol, DMSO, NMP, and acetonitrile) gave
stable dispersions, dilution with less polar
organics (acetone, THF, and toluene) led
to either occulation or aggregation.
2.2. Free-Standing Graphene Oxide Papers
The ow-directed ltration of aqueous dispersions of
lamellar clay, such as vermiculite and mica, into free-standing,
paperlike lms is a well-known process that has been
successfully commercialized.
[48]
This technique was duplicated
withgrapheneoxidedispersions byDikinet al. togivegraphene
oxide paper (Figure 2B),
[49]
a brownblack paperlike material
featuring a layered structure with intersheet spacing of 8.3 A

,
near that of un-exfoliated GO (6.8 A

),
[35]
a variance that is
likely an effect of intercalated water.
[49]
As seen by scanning
electron microscopy (SEM; Figure 2C),
[49]
the structure of
graphene oxide paper consists of tightly packed interlocking
sheets that subtly undulate along the paper surface. By
modifying the amount of graphene oxide in the ltered
dispersion the thickness of the paper can be readily tuned
from 1 to 30 mm, with thinner papers (<5 mm) being
semitransparent and thicker papers fully opaque.
[49]
Nearly
transparent papers (transmission >95%), which may more
appropriately be referred to as thin lms (thicknesses of
2 nm), have also been prepared via lter assembly.
[50]
While graphene oxide paper is non-conductive,
[51]
it
exhibits impressive mechanical properties (Table 2), with a
Youngs modulus value (E, ranging from 2342 GPa,
[49,52,53]
reviews O. C. Compton and S. T. Nguyen
Figure 2. A) Schematic model of a graphene oxide sheet. B) Photograph of a graphene oxide
paper ribbon. Reproduced with permission from Reference [49]. Copyright 2007, Nature
Publishing Group. C) SEM image of the edge of graphene oxide paper showing tightly packed,
undulatingstructure. ReproducedwithpermissionfromReference [49]. Copyright 2007, Nature
PublishingGroup. D) AFMtopographof grapheneoxidesheetsattachedtoanamine-terminated
template on a gold surface. Reproduced with permission fromReference [63]. Copyright 2008,
AmericanChemical Society. E) SEMimages of a LangmuirBlodgett-assembledgraphene oxide
thin lm collected on a silica substrate with surface pressure increasing from left to right.
Reproducedwith permission fromReference [22]. Copyright 2009, American Chemical Society.
714 www.small-journal.com 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim small 2010, 6, No. 6, 711723
comparable to that of concrete
[54]
) that is signicantly higher
than that of inorganic vermiculite lms (14.1 GPa)
[48]
or bucky
paper (2.7 GPa),
[55]
a lter-assembled paper composed of
overlapping CNTs. The maximumtensile strength of graphene
oxide paper (s, ranging from 15193 MPa,
[49,52,53]
comparable
to that of cast iron
[56]
) is also signicantly higher than bucky
paper (33.2 MPa) and is similar to that of vermiculite lms
(160 MPa). These mechanical properties can be further
improvedviamodicationof thefunctional groups ongraphene
oxide sheets. For example, in situ treatment of fabricated
graphene oxide paper with only 1 wt% of divalent cations
(Mg
2
or Ca
2
) crosslinks edge-bound carboxylic acid groups
of adjacent sheets,
[53]
improving Youngs modulus by 1040%
and tensile strength by 1080%. Exposure of graphene oxide
paper to hexylamine affords an amine-modied paper via ring-
opening of the epoxide groups on the basal plane.
[52]
The hexyl
groups ll the intersheet spacing and stabilize the structure of
the modied paper during thermal annealing at 300 8C, with
68%retentionof tensilestrengthand35%retentionof Youngs
modulus; this is in stark contrast to the complete loss of all
mechanical integrity in unmodied graphene oxide paper after
annealing.
Paperlike graphene oxide membranes can also be prepared
by evaporation of water from aqueous graphene oxide
dispersions, which induces sheet self-assembly.
[57]
Such
membranes featureanintersheet spacingof 7.5 A

, intermediate
between ow-directed assembled graphene oxide papers
(8.3 A

)
[49]
and GO powder (6.8 A

).
[35]
As with graphene oxide
paper, the thickness of this evaporation-assembled membrane
(0.520 mm) can be tuned by increasing or decreasing
dispersion concentration. Despite such close sheet packing,
the mechanical properties of the membrane were signicantly
lower (Youngs modulus12.7 GPa, tensile strength70 MPa)
than observed for ow-directed assembled papers (see above).
2.3. Thin Films of Graphene Oxide
Complementary to the mechanically strong, micrometer-
thick, self-supporting graphene oxide papers just discussed,
which contain thousands of stacked sheets, nanometer-thick
thin lms consisting of only a few graphene oxide sheets (i.e.,
mono-, bi-, and trilayers of graphene oxide) can also be
prepared. Such lms have been investigated as components of
eld-effect transistors
[58]
or as efcient electrical conductors
after reduction to graphene (see Section 3.3). The rst efforts
toward a uniform, single graphene oxide monolayer can be
traced back to work on ionically conductive composite lms
featuring alternating graphene oxide and polyelectrolyte
layers.
[5961]
Thin lms with spotty coverage (3090%,
depending on the type of substrate) were prepared by drop-
casting graphene oxide suspensions on a silica substrate or
briey immersing the substrate withinthe suspension.
[60]
While
lms with near-monolayer thickness of graphene oxide (11
14 A

,
[60]
1420 A

,
[61]
1828 A

[59]
) were claimed, coverage
>60% was only attained in multilayered structures. Such lms
were foundto be effective hole conductors; however, they were
unstable and readily reduced to graphene during electroche-
mical analysis.
Building upon the drop-casting and immersion methods,
spin-casting and templating have recently been applied to
produce high-quality and patterned graphene oxide lms,
respectively. Spin-castingprovides aneasymeans for producing
continuous lms composed of graphene oxide sheets,
[62]
while templating allows for intricate patterning of graphene
oxide sheets on the surface of a substrate.
[63]
In spin-casting,
the thickness of the deposited layer can be tuned by varying the
concentration of the graphene oxide dispersion, with dilute
dispersions (2 mg mL
1
) giving lms 3 nm thick and higher-
concentration dispersions (1215 mg mL
1
) yielding lms
20 nm thick. It is likely that the thinnest of these spin-cast
lms are not true monolayers but bi- or trilayers, as they are
thicker than individual graphene oxide sheets (1 nm
[5]
).
As an example of templating, a mica-peeled gold
surface patterned with 11-amino-1-undecanethiol can act as
a positively charged substrate for the negatively charged
graphene oxide surface. Exposing this patterned substrate to a
dispersion of graphene oxide at relatively low pH (6.58.5)
immobilizes the free sheets onto the positively charged
ammonium-thiol pattern (Figure 2D). This mode of electro-
static interaction was supported by the observation that a
graphene oxide dispersion held at pH 911, above the
predicted pK
a
value for alkylammonium salts, prevented
patterning, as the non-protonated amine cannot interact with
graphene oxide.
Graphene Oxide- and Graphene-Based Materials
Table 2. Mechanical properties of self-supporting carbonaceous papers.
Reference Type of paper Average Youngs
modulus [GPa]
Average ultimate tensile
strength [MPa]
[49] Graphene oxide paper 32 130
[51] Graphene oxide paper, annealed (3008C) not determinable not determinable
[52] Graphene oxide paper, hexylamine-modied 17.2 63.4
[52] Graphene oxide paper, hexylamine-modied, annealed (3008C) 5.9 42.8
[51] Graphene paper, ow-directed assembled 20.1 152
[52] Graphene paper, reduced in situ 3.0 12.7
[51] Graphene paper, ow-directed assembled, annealed (2208C) 41.8 293.3
[52] Graphene paper, reduced in situ annealed (3008C) 3.4 25.0
[53] Graphene oxide paper 25.6 81.9
[53] Graphene oxide paper, Mg
2
-modied 27.9 80.6
[53] Graphene oxide paper, Ca
2
-modied 21.8 125.8
[57] Graphene oxide membrane, self-assembled 12.7 70.0
[55] Bucky paper 2.7 33.2
small 2010, 6, No. 6, 711723 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com 715
More recently, a monolayer-thin lm has been produced
from basic dispersions of graphene oxide using Langmuir
Blodgett (LB) assembly techniques (Figure 2E).
[22]
Formation
of multilayer structures are discouraged by repulsive negative
charges, resulting from deprotonated carboxylic acid groups,
along the edges of adjacent graphene oxide sheets. This
repulsion discourages overlapping of the sheets upon LB
compression, forcing them to fold or wrinkle at the edge while
leaving the interior at. At high compression, the folded edge
layers partiallyoverlapandinterlocktocreateacontinuous lm
with a relatively uniform monolayer thickness (1 nm).
2.4. Nanocomposites of Graphene Oxide
While graphene oxide has servedas the primary constituent
in all materials discussed thus far, nanocomposites of graphene
oxideutilizethesesheets as aminor ller component embedded
withineither a polymer or aninorganic matrix. Giventheir high
oxygen content, graphene oxide sheets are the carbon analogs
of 2D lamellar clay sheets, such as montmorillonite (MMT),
whose nanocomposites with polymers have been thoroughly
investigatedfor a wide range of applications.
[64]
The processing
of polymerclay nanocomposites is often carried out via
extrusion and forced intercalation of polymer into neat clay,
whichproduces micrometer-thick, multisheet aggregates.
[65]
In
contrast, the high surface-to-volume ratio of graphene oxide, in
conjunction with its high dispersibility in both water and
organic solvents and its wide range of reactive surface-bound
functional groups, discourages aggregation, facilitating proces-
sing in solution and promoting signicant interaction between
llers and polymers.
A variety of graphene oxide-based nanocomposites have
beenpreparedas thinlms, thoughthey are commonly reduced
tographene for conductivity studies (see Section3.5). Spin-cast
lms of silica containing up to 11 wt% of graphene oxide
have been prepared and reduced to give transparent
conductive layers.
[66]
Paperlike thin lms with up to 1.4 wt%
of polystyrene have been made by co-ltration of an aqueous
solution containing both graphene oxide and polystyrene
nanoparticles.
[67]
Upon photothermal reduction these lms
increase by nearly an order of magnitude in thickness but can
still conduct electrons (1 S m
1
, assuming a lm thickness of
100 mm). In a third example, 3D macroporous composite
monoliths have beenfabricatedbyice-segregationself-induced
assembly (ISISA),
[68]
where vessels containing poly(vinyl
alcohol) (PVA)/graphene oxide aqueous dispersions up to
13 wt% graphene oxide were slowly immersed into liquid
nitrogen, establishing macroporous structures that remained
intact after lyophilization. However, the reduced analogs of
these PVA/GO materials did not exhibit conductivity.
More recently, drop-cast polyurethane lms containing
small amounts of graphene oxide (4.4 wt%) were prepared and
foundtoincreaseYoungs modulus by900%incomparisonto
lms of the Pristine polymer, with only minimal decrease in
tensile strength.
[69]
This level of increase in the modulus is
especially impressive in comparison to that observed for
polyurethaneMMTcomposites, where similar loadings of clay
(45 wt%) only increased Youngs modulus by 150%, with a
concurrent 200% improvement in tensile strength.
[70,71]
3. Graphene-Based Materials
With its fully conjugated, rigid 2D structure, graphene
possesses both electronic (electrical
[10]
and thermal
[11]
con-
ductivity) and mechanical properties
[12]
that are superior to
those of CNTs. As such, graphene has attracted signicant
interest in the scientic community as the most promising
carbon allotrope for the development of next-generation
carbon-based materials. Indeed, the number of peer-reviewed
articles that discuss graphene and its associated materials
[72,73]
has increased exponentially (Figure 1) since the 2004 isolation
of single graphene sheets via mechanical exfoliation.
[7]
As
mentioned in Section 1, while graphene can be prepared froma
variety of graphite derivatives, such as graphite uoride or
expandable graphite,
[74,75]
bulk quantities of fully exfoliated
graphenelike sheets are commonly obtained by reducing
graphene oxide dispersions or powders. More accurately
described as highly reduced graphene oxide (HRG),
[14,15]
these sheets comprise graphene domains interspersed with
residual oxygen-containing functionalities (Figure 3A) and
have C:O ratios ranging from 10:1
[76]
to 5:1.
[52]
Although
HRG sheets are pock-marked with defects and have lower
conductivity and mechanical strength in comparison to the
pristine graphene sheets obtained via Scotch-tape mechanical
exfoliation,
[7,77]
the excellent scalability of their synthesis has
enabled the preparation of many novel graphene-based
materials with excellent physical properties.
3.1. Dispersions of Graphene in Aqueous and Organic
Solvents
As with graphene oxide, solution dispersions of graphene
serve as the key precursors for many graphene-basedmaterials.
While dispersions of defect-free graphene have been prepared
via direct exfoliation of graphite or expandable gra-
phite,
[75,78,79]
the most popular graphene dispersions are those
containing HRG sheets, commonly produced by reduction of
graphene oxide dispersions. For the conversion from hydro-
philic graphene oxide to hydrophobic HRG/graphene to be
successful, bothspecies must bestabilizedina single mediumso
as to prevent aggregation during the reduction process.
[25]
The rst stable graphene dispersion was produced by
reducing an aqueous graphene oxide dispersion with hydrazine
hydrate in the presence of the amphiphilic surfactant poly-
(sodium 4-styrenesulfonate) (PSS).
[80]
In this strategy, the
newly reduced graphene was stabilized via association with
the hydrophobic backbone of PSS while the hydrophilic
sulfonate side groups sustained the whole graphenePSS
complex in water. Reduction in the absence of PSS surfactant
gave graphene aggregates that quickly precipitated from the
aqueous medium(Figure3B). Inasubsequent report, graphene
dispersions in THF, CCl
4
, and 1,2-dichloroethane (EDC)
were obtained by converting the edge carboxylic acid groups
into octadecylamides, which sterically stabilized the graphene
sheets.
[81]
A variety of modiers and surfactantsincluding isocya-
nates,
[43]
quaternary amines,
[82]
diazonium salts,
[83,84]
ionic
liquids,
[85]
andsingle-strandedDNA
[86]
have beenemployed
to stabilize either aqueous or organic dispersions of graphene.
reviews O. C. Compton and S. T. Nguyen
716 www.small-journal.com 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim small 2010, 6, No. 6, 711723
However, amphiphilic graphene has only beenreportedvery
recently, via covalent grafting of an amphiphilic polystyrene-
polyacrylamide copolymer to the graphene surface during
graphene oxide reduction by hydrazine.
[87]
In contrast to PSS-
stabilized graphene, where only the hydrophilic sulfonate side
groups are exposed to water, in this case both the hydrophobic
styrene and hydrophilic acrylamide moieties can interact with
the solvent, allowing for good dispersion in either water or
xylene.
Electrostatic repulsion can be used to counter the
irreversible aggregation induced by intersheet van der Waals
attractions in the preparation of stable aqueous suspensions of
graphene. Noting that carboxylic acids are not affected by
hydrazine, Li et al. surmised that these groups would remain
intact on the edge of graphene sheets after hydrazine reduction
and, if deprotonated, would provide sufcient electrostatic
repulsion (zeta potential 30 mV)
[88]
to overcome the
aforementioned van der Waals attraction.
[21]
By adding
ammonia (to raise the pHof the suspension
to 10) and mineral oil (to avoid sheet
agglomeration at the air/water interface)
during the hydrazine reduction of graphene
oxide, a stable aqueous dispersion of
graphene was obtained. Unfortunately, this
suspension was fairly sensitive to both pH
and the presence of hard electrolytes
such as sodium chloride.
Charge stabilizers have been utilized to
improve the robustness of electrostatically
stabilized graphene dispersions, making
them less sensitive to pH. Reduction of
graphene oxide suspensions in 98% anhy-
drous hydrazine is reported to produce
hydraziniumions that stabilize the resulting
graphene dispersions.
[89]
Exposing aqueous
dispersions of graphene oxidetoKOHprior
to hydrazine reduction has also been found
to result in the stabilization of edge-bound
carboxylic acids on the graphene sheets by
K

,
[14]
similar to the interactions proposed
in alkali earth-modied graphene oxide
paper.
[53]
While the use of surfactants or covalent
surface modiers provides a convenient
strategy for stabilizing graphene disper-
sions, they often negatively affect conduc-
tivity and mechanical properties of the
subsequently produced materials (see
Section 3.2). To this end, stabilizer-free
dilute dispersions of graphene have been
prepared in primarily organic solvents.
[15]
Hydrazine reduction of a graphene oxide
suspension in DMF/water (9:1 v/v) pro-
duced a stable dispersion that could be
further diluted with DMF, DMSO, THF, or
NMP (up to ninefold in volume) to give
suspensions with as little as 1 vol% water.
As guidance, it was suggested that only a
small solvent polarity window, as deter-
mined by the sum of the polarity cohesion (d
p
) and hydrogen-
bonding cohesion (d
h
) factors of the Hansen solubility
parameters,
[90]
is available for dispersion of graphene in
organic media. Onlysolvents with(d
p
d
h
) intherangeof 1329
were found to yield stable graphene suspensions.
Graphene dispersions in highly polar organic solvents have
also been prepared from sources other than graphene oxide,
such as expandable graphite and graphite powder. This
approach avoids the use of defect-laden graphene oxide
dispersions and affords defect-free graphene sheets that may
result in materials with higher conductivity. Dai et al. rst
reported a direct, precursor-free route to graphene by
subjecting expandable graphite to rapid thermal treatment
(1000 8C) followed by sonication in 1,2-dichloroethane in the
presence of a polymer stabilizer.
[91]
The resulting graphene
nanoribbons (10 nmwide and 1 mmlong) were only obtained in
low yields with many comprising graphene bi- and trilayers.
Later, the Dai group was able to produce individual graphene
Graphene Oxide- and Graphene-Based Materials
Figure 3. A) Schematic model of ahighly reducedgrapheneoxidesheet obtainedviareduction
of graphene oxide. Possible residual oxygen-containing defects are shown to illustrate the
imperfect structure. B) Photograph of aqueous solutions of graphene sheets reduced in the
absence(left) andpresence(right) of poly(sodium4-styrenesulfonate) (PSS), demonstratingthe
critical role of the amphiphilic PSS surfactant in maintaining a dispersion after graphene oxide
reduction. Modeledafter Reference [80]. C) Photographof graphenepaper producedbyltration
of an aqueous graphene dispersion. Reproduced with permission from Reference [72].
Copyright 2008, American Association for the Advancement of Science. D) SEM image of the
layeredstructureof agraphenepaper, viewedfromafractureedge. ReproducedfromReference
[51]. E) SEMimageof aggregatedgraphenesheets intheformof graphenepowder. Reproduced
with permission fromReference [5]. Copyright 2007, Elsevier. F) Photograph of a transparent,
conductivegraphene thin lmfabricatedvia LangmuirBlodgett assembly withplot correlating
decreased resistance and transparency with increased lm thickness. Reproduced with
permission from Reference [92]. Copyright 2008, Nature Publishing Group.
small 2010, 6, No. 6, 711723 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com 717
sheets by intercalating acidied expandable graphite with a
sterically large tetrabutylammonium cation and sonicating the
resulting material in DMF in the presence of a polymer
stabilizer.
[92]
A simpler method, inspired by the sonication-induced
exfoliation of CNTs,
[93]
has recently come to light where
graphite powder can be successfully exfoliated in DMF
[79]
or
NMP
[75]
by sonication, which previously had only given
multilayer aggregates.
[94]
The key discovery here is the use of
solvents whosesurfaceenergy(37for DMF
[95]
and40 mJm
2
for NMP
[75]
) matches well with that of graphite (5590mJ
m
2
), allowing for strong interaction of the solvent with the
graphite gallery and minimizing the energy required for
exfoliation. After sonication, the exfoliated materials are
predominantly single sheets and bi- and trilayer stacks, with
some multilayer structures that can be separated by centrifuga-
tion. More recently, peruorinated solvents (C
6
F
6
, C
5
F
5
N,
etc.),
[78]
pyridine,
[78]
and o-dichlorobenzene
[96]
have also been
identied as suitable media for graphite exfoliation by sonica-
tion. The development of preparative methods that exclusively
generate defect-free graphene sheets, in the absence of either
oxidativepreparationsteps or stabilizinggroups, shouldpavethe
way for highly conductive graphene-based materials.
3.2. Electrically Conductive, Free-Standing Graphene
Papers
Similar to the ow-directed preparation of graphene oxide
paper (see Section 2.2), ltration of graphene dispersions also
leads to a free-standing paperlike material (Figure 3C
and D).
[21,51]
As prepared, this graphene paper is not
well ordered and exhibits a broad, weak diffraction peak near
3.9 A

with relatively low mechanical strength (Youngs


modulus 20.5 GPa and tensile strength150 MPa) and in-
plane electrical conductivity (7200 S m
1
, as measured by a
four-point probe). Upon annealing at 500 8C, the paper
becomes more ordered, exhibiting a much sharper diffraction
peak near 3.4 A

(pristine graphite 3.35 A

). The conductivity
of this annealed paper (35 000 S m
1
) is the highest value
reportedtodatefor aself-supporting, graphene-basedmaterial.
Its mechanical strength also improves upon annealing with
maximum values for Youngs modulus (41.8 GPa) and tensile
strength (293.3 MPa) reached at 220 8C; however, these values
decreaseuponannealingat higher temperatures. This is instark
contrast to graphene oxide paper, whose mechanical strength
decreases upon annealing, becoming easily crumpled after
heating to 300 8C. Presumably, the pyrolysis of oxygen-
containing functional groups above 150 8C
[97]
generates a
signicant amount of gases, whichcoulddisturbthe structure of
the annealed graphene oxide paper and reduce its mechanical
integrity.
Graphene papers have also been prepared by reducing
fabricated graphene oxide papers. Two distinctly different
techniques have been reported thus far for this process, in situ
hydrazine reduction
[52]
and ash reduction.
[67]
In situ
hydrazine reduction is carried out in a manner analogous to
the hydrazine reduction of graphene oxide dispersions (see
Section 3.1), where an aqueous solution of hydrazine is ltered
through a membrane-supported piece of graphene oxide paper
under heating.
[52]
The resulting paper is littered with local
heterogeneities, which lower both its in-plane electrical
conductivity (200 S m
1
) and mechanical strength (Youngs
modulus 3.0 GPa and tensile strength13.2 MPa). As with
graphene paper prepared via ow-directed ltration,
[51]
annealing at 300 8C increases all of these properties
(conductivity 1700 S m
1
, Youngs modulus 3.4 GPa, and
tensile strength24.7 MPa). Given its simplicity, the in situ
reduction method has been extended to produce modied
graphene paper from chemically modied graphene-oxide
paper.
[52]
The ash technique involves exposing graphene oxide
paper to a xenon lamp, in the form of a digital camera ash
(typical applied energy ranging from 0.1 to 2 J cm
2
). This
photothermal reductionincreases the thickness of a 1-mm-thick
paper by nearly two orders of magnitude, presumably as a
consequence of the pyrolysis-evolved gas, producing a uffy
material with relatively low conductivity (1000 S m
1
). Given
the ash-inducednatureof this reductionmethod, masks canbe
used to cover the surface before reduction, producing a
patterned surface with conductive stripes that can be used as
interdigitated electrodes.
[67]
As discussed in Section 3.1, the presence of functional
groups on the surface of graphene sheets is expected to
negatively affect the conductivity of macroscopic materials
fabricated from these functionalized sheets. This is evident
when comparing the conductivities of graphene papers
prepared from aqueous
[51]
and organic
[15]
solution (7200
and 1690 S m
1
, respectively) with those of graphene-based
papers fabricated in the presence of stabilizers such as K

ions
[14]
or 1-pyrenebutyrate
[98]
(687 and 200 S m
1
, respec-
tively). However, conductivity values can be improved by
reducing the intersheet spacing in graphene paper, most easily
accomplished by annealing. The smaller spacing likely facil-
itates the transfer of electrons between adjacent sheets, with
increases by nearly an order of magnitude having been
observed after annealing at temperatures as low as 150 8C
(spacing decreased from3.86 to 3.70 A

).
[15]
Interestingly, while
conductivity generally improves as graphene paper samples are
annealed at higher temperatures, no direct correlation exists
between intersheet spacing and conductivity (Figure 4).
3.3. Graphene Thin Films
One of the most promising applications of graphene-based
materials is their use as transparent, conductive thin lms
(Figure 3F). Such lms can be used as electrodes in solar cells,
transistors in next-generation displays, and highly sensitive
detectors for explosives. As discussed in Section 1, mechanical
exfoliation,
[7]
CVD,
[99,100]
and epitaxial growth
[10]
all produce
highly oriented graphene monolayers. Unfortunately, the
production of large-area thin lms, which are required for
many applications, using these methods is still in its infancy.
Fabrication from HRG and graphene dispersions or reduction
of graphene oxide thin lms affords an easy means to prepare
large lms but these often have many conductivity-limiting
defects that are intrinsic to the oxidation-reduction procedures
(see Figure 3A and Section 3.2). As such, intense efforts are
currently being devoted to produce large-area lms using this
reviews O. C. Compton and S. T. Nguyen
718 www.small-journal.com 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim small 2010, 6, No. 6, 711723
route that can compete with pristine graphene in
conductivity.
[5,76]
The most convenient method for producing large-area thin
lms of graphene involves the reduction of pre-fabricated
graphene oxide thin lms. Since the production of graphene
oxide lms has already been presented in Section 2.3, we will
limit discussion here to their reduction, which is typically
achieved by thermal annealing,
[101]
exposure to hydrazine
gas,
[22]
or a combination of both procedures (Table 3).
[62,102]
Becerril et al. observed that thermal treatment of graphene
oxide thin lms at 1100 8C produces highly reduced lms with
sheet-resistance values nearly four orders of magnitude lower
than those only exposed to hydrazine gas.
[62]
Such high
temperatures were necessary to produce lms with low sheet
resistance, as 10-nm-thick lms annealed at 9001100 8C had
sheet resistances (2 kV &
1
) that were nearly an order
of magnitude lower than those annealed at 550700 8C
(15 kV &
1
).
[101]
Higher annealing temperatures were also
found to increase lm transparency in the spectral region
>300 nm in wavelength, attributable to increases in the degree
of graphitization.
Dai et al. have compared graphene thin lms prepared
from thermally annealed LB-assembled graphene oxide
thin lms against LB-assembled graphene thin lms
prepared directly from graphene dispersions.
[92]
The former
had a sheet resistance nearly three orders of magnitude
higher than that of the latter, presumably due to residual
oxygen-containingfunctional groups onthereducedgraphene
oxide surface. Increasing lm thickness from 1 to 3 nm
resulted in more pathways for electron transport (see
Section 3.2) and a dramatic drop in sheet resistance (from
150 to 8 kV &
1
) but at the expense of transparency (from
93 to 83% at 1000 nm).
3.4. Graphene Powder
While graphene powder can be isolated from reduced
dispersions of graphene oxide, thermal exfoliation provides an
alternative method for its preparation directly from GO and
circumvents the use of solvents. Developed by Aksay,
Prudhomme, et al.,
[76]
this method involves the rapid heating
(>2000 8Cmin
1
) of GOto1050 8C, whichundergoes complete
delamination via the evolution of CO
2
as surface hydroxyl and
epoxy groups decompose.
[34]
The violent means by which these
functionalized single graphene sheets are produced leaves
the structure littered with defects such as 5- and 7-membered
rings and carbon vacancies.
[77]
Despite such defects, conduc-
tivity values for compacted samples of the material remain in
the range of 10002300 S m
1
,
[76]
less than an order of
magnitude lower than compacted samples of graphite pow-
der.
[5]
For comparison, compacted samples of graphene
powder, obtained by reduction of a graphene oxide dispersion
Graphene Oxide- and Graphene-Based Materials
Figure 4. Values of the in-plane electrical conductivity for self-
supporting graphene papers as a function of intersheet spacing.
Table 3. Sheet resistance values for graphene-based thin lms.
Reference Preparation method Sheet resistance
[V &
1
]
Transmission
[%]
Film thickness
[nm]
[62] Spin-cast GO, hydrazine-reduced 2.210
8
94@550nm 3
[62] Spin-cast GO, hydrazine-reduced and annealed (4008C) 1.710
5
92@550nm 3
[62] Spin-cast GO, annealed (11008C) 3.510
4
95@550nm 4
[66] Spin-cast GO-silica composite, hydrazine-reduced 2.710
8
95@650nm 33
[66] Spin-cast GO-silica composite, hydrazine-reduced and annealed (4008C) 7.910
5
98@650nm 28
[22] LB-assembled GO, hydrazine-reduced 1.910
7
95@550nm 1
[92] LB-assembled graphene 1.510
5
93@1000nm 1
[92] LB-assembled GO, annealed (8008C) 1.110
7
N/A 1
[101] Dip-coated GO, annealed (5508C) 2.010
4
70@1000nm 10
[101] Dip-coated GO, annealed (7008C) 1.110
4
74@1000nm 10
[101] Dip-coated GO, annealed (11008C) 1.810
3
81@1000nm 10
[120] Spray-deposited GO 4.010
10
N/A 1
[120] Spray-deposited GO, hydrazine-reduced 4.010
6
N/A 1
[50] Filter-assembled GO, hydrazine-reduced 1.910
9
93@550nm 3
[50] Filter-assembled GO, hydrazine-reduced and annealed (2008C) 1.610
5
90@550nm 3
[75] Filter-assembled graphene, overnight vacuum 7.210
6
61@632nm 30
[75] Filter-assembled graphene, annealed (3008C) in air 7.110
3
42@632nm 30
[75] Filter-assembled graphene, annealed (2508C) in H
2
/N
2
5.110
3
42@632nm 30
small 2010, 6, No. 6, 711723 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com 719
in the absence of any stabilizers (Figure 3E), also exhibited
conductivity values (2000 S m
1
)
[5]
in this range.
3.5. Nanocomposites of Graphene
Due to the late arrival of defect-free graphene sheets, the
majority of work on graphene nanocomposites has been based
on HRG. As a ller material in polymer nanocomposites, the
inclusion of HRG leads to signicant enhancements of
electronic and mechanical properties at much lower volume
loadings (0.15 vol%) when compared to lamellar clays.
[64,103]
Given that they can be easily functionalized,
[43,83,85]
HRG
sheets are ideal as nanoller materials, beingreadilydispersible
inawidevarietyof polymers, evenbeingutilizedtopromotethe
dispersion of other nanoparticles in polymer matrices.
[104]
The rst graphene-based nanocomposite was prepared by
Stankovich et al., where an isocyanate-treated graphene oxide
dispersion in DMF was reduced in the presence of polystyrene
to afford a composite powder that could be pressed or cast into
thin lms.
[105]
While the graphene sheets in these lms, as
imaged by SEM, resembled crumpled paper (Figure 5A), they
had enough long-range order to give rise to electron diffraction
patterns. Electrical-conductivity measurements of composites
with increased loading determined a percolation threshold of
0.1 vol% (conductivity 10
5
S m
1
) and an ultimate
conductivity of 1 S m
1
at 2.5 vol% (Figure 5B). Similar
ultimate conductivity values have been observed for
graphene-loaded composites of poly(vinylidene uoride) (up
to 4 wt%),
[106]
polyurethane (1 wt%),
[107]
and silica (up to
11 wt%; see Section 2.4).
[66]
Similar to the aforementioned low percolation threshold
observed for electrical conductivity in graphene polymer
composites, excellent improvements in mechanical properties
can be observed with very little addition of graphene to the
parent matrices. An unprecedented 30 8C increase in glass-
transition temperature (T
g
) of poly(methyl methacrylate)
(PMMA) was observed with the addition of only 0.05 wt% of
graphene.
[108]
At 0.01 wt% graphene loading, the Youngs
modulus of PMMA increased by 33%, nearly seven times
greater than theoretical predictions (Figure 5C). Increased
graphene loading to 1 wt% further improved the Youngs
modulus to 80% greater than that of PMMA alone.
Ultimate tensile strength also increased, albeit more modestly,
by 20% at 1 wt% graphene loading. For comparison, PMMA
composites containing graphite and heat-treated expandable
graphite (i.e., expanded graphite) did not exhibit similar
signicant increases, suggesting that the aforementioned
remarkable mechanical property enhancement only occurs
with fully exfoliated graphene.
[109]
Similar increases in both
Youngs modulus (57%) and ultimate tensile strength (70%)
have been observed for polystyrene-grafted graphene
nanocomposites.
[110]
Recently, monoliths and foams incor-
porating graphene nanoller have been
demonstrated to have impressive increases
in compressional moduli compared to the
parent materials. Silica monoliths contain-
ing 0.1wt% of graphene sheets functiona-
lized with 3-aminopropyltriethoxysilane
(APTES) exhibited 19 and 92% increases
in compressive failure strength and tough-
ness, respectively, over silica monoliths
containing only APTES.
[111]
The compres-
sional modulus of siliconefoamwas foundto
increaseover 200%withtheadditionof only
0.25 wt% of graphene.
[112]
This nanocom-
posite foam also exhibited a 6% increase in
thermal conductivity and a near 40 8C
increaseinthermal decompositiontempera-
ture (from 16 to 55 8C).
Graphene-polymer composites with
distinct 3D macroporous structures have
also been prepared via two separate
methods: 1) the ISISAtechnique described
in Section 2.4 and 2) electrostatic templat-
ing of polystyrene beads (Figure 5D).
[68]
In
the rst example, a PSS-stabilized gra-
phene/PVA dispersion was immersed into
liquid nitrogen to give a macroporous
monolith after lyophilization. While non-
conductive even at 1.5 wt% loading, this
monolith featured a compressional mod-
ulus 300%higher than that of the pristine
polymer. In the second example, poly-
(allylamine hydrochloride)-functionalized
reviews O. C. Compton and S. T. Nguyen
Figure 5. A) SEM images of a graphene-polystyrene nanocomposite at low and high
magnication. Reproduced with permission from Reference [105]. Copyright 2006, Nature
PublishingGroup. B) Plotsof conductivities(circle) andYoungsmoduli (diamond) for graphene-
based nanocomposites with polystyrene and PMMA, respectively, at increasing weight
loadings. Adapted from References [105] and [109]. C) Comparison of the mechanical
properties of graphene and CNTPMMA nanocompsites. Adapted from Reference [108].
D) SEMimagesof graphenenanocomposites: 1) Longitudinal sliceof ISISA-preparedcomposite.
2) Cross section of ISISA-prepared composite. 3) PSSgraphene composite coated on
polystyrene beads. 4) Remaining templated PSSgraphene composite after dissolving the
polystyrene bead template in toluene. Reproduced from Reference [68].
720 www.small-journal.com 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim small 2010, 6, No. 6, 711723
polystyrene beads were combined with negatively charged
PSS-stabilized graphene sheets
[80]
to give spherical compo-
sites that could be converted into macroporous structures by
dissolving the styrene core with toluene. These examples
signal the next step in the development of macroscopic
graphene-based materials, where three-dimensional objects
with specic shapes and sizes, beyond papers, thin lms, and
powders, can now be prepared.
4. Summary and Outlook
Within a short time of being available in bulk quantities,
graphene oxide, HRG, and graphene have become highly
versatile, inexpensive building blocks for the development of
several advancedcarbonaceous materials, withapplications to
be limited only by the imagination. For example, it is not
difcult to imagine the incorporation of a variety of additives
into paperlike composites of graphene and graphene oxide to
create novel material compositions. Indeed, polymers have
already been added to graphene papers and lms, either in
nanoparticle form
[67]
or via in situ electropolymerization,
[113]
for mechanical stabilization and to facilitate processing. Bio-
nanocomposites have also been prepared recently, via
intercalation of DNA and cytochrome c into graphene oxide
paper.
[86]
Graphene paper itself is biologically compatible,
being able to function as a support for mouse broblast
cells,
[51]
which suggests its use in encapsulating enzymes,
[114]
serving as a template for DNA-based therapy,
[86]
or culturing
cells.
[51]
In light of the heated pace with which materials based on
graphene oxide andgraphene are currently being studied, rapid
developments in the chemistries of graphene oxide and
graphene and their processing methods should be anticipated
in the near future. The development of methods to fully
exfoliate graphite powder, in a wide variety of solvents,
[75,78]
and in the absence of reducing or stabilizing agents, will be
integral to the fabrication of high-quality graphene materials.
Withimprovements inthesynthesis andprocessingof graphene
oxide and graphene, these highly versatile materials can be
accessed by more scientists and engineers, thus ensuring the
rapid development of a myriad of new materials with
impressive properties. For example, monodisperse colloidal
solutions of single graphene sheets and few-layer graphene
stacks have recently been separated from bulk graphene
nanosheet dispersions via density-gradient ultracentrifuga-
tion
[121]
and used to prepare thin lms that have lower sheet
resistance thanthose preparedfrombulkgraphene dispersions.
A third frontier for future development lies in the
production of large-area graphene thin lms for use in
electronic and energy applications. However, to produce the
highly conductive and transparent monolayers needed for
such purposes, new techniques for large-area thin-
lm preparation and patterning must be combined with
improvements in the synthesis of large-scale, defect-free
graphene sheets, as discussed in the preceding paragraph. As
graphene can also serve as a template for catalytic nanopar-
ticles,
[45,115117]
achieving high activities and turnover frequen-
cies, the fabrication of catalytic thin lms would afford
high-surface-area heterogeneous catalysts for use in fuel cells.
In another application, graphene drop-cast from a dispersion
has recently been utilized as an ultrathin support lm for TEM
imaging to give unmatched atomic resolution.
[118]
With such
innovative applications in mind, the development of new
graphene oxide- and graphene-based materials will certainly
lead to many future advances in science and technology.
Keywords:
composites
.
graphene
.
graphene oxide
.
thin films
[1] A. K. Geim, K. S. Novoselov, Nat. Mater. 2007, 6, 183191.
[2] P. Avouris, Z. H. Chen, V. Perebeinos, Nat. Nanotechnol. 2007, 2,
605615.
[3] H. W. Kroto, J. R. Heath, S. C. Obrien, R. F. Curl, R. E. Smalley,
Nature 1985, 318, 162163.
[4] S. Iijima, Nature 1991, 354, 5658.
[5] S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas,
A. Kleinhammes, Y. Jia, Y. Wu, S. T. Nguyen, R. S. Ruoff, Carbon
2007, 45, 15581565.
[6] K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich,
S. V. Morozov, A. K. Geim, Proc. Natl. Acad. Sci. USA 2005, 102,
1045110453.
[7] 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,
666669.
[8] C. Berger, Z. Song, X. Li, X. Wu, N. Brown, C. Naud, D. Mayou, T. Li,
J. Hass, A. N. Marchenkov, E. H. Conrad, P. N. First, W. A. de Heer,
Science 2006, 312, 11911196.
[9] N. D. Mermin, Phys. Rev. 1968, 176, 250.
[10] M. Orlita, C. Faugeras, P. Plochocka, P. Neugebauer, G. Martinez,
D. K. Maude, A. L. Barra, M. Sprinkle, C. Berger, W. A. de Heer,
M. Potemski, Phys. Rev. Lett. 2008, 101, 267601267604.
[11] A. A. Balandin, S. Ghosh, W. Bao, I. Calizo, D. Teweldebrhan,
F. Miao, C. N. Lau, Nano Lett. 2008, 8, 902907.
[12] C. Lee, X. Wei, J. W. Kysar, J. Hone, Science 2008, 321, 385388.
[13] C. A. Coulson, R. Taylor, Proc. Phys. Soc. Sect. A 1952, 65, 815
825.
[14] S. Park, J. An, R. D. Piner, I. Jung, D. Yang, A. Velamakanni,
S. T. Nguyen, R. S. Ruoff, Chem. Mater. 2008, 20, 65926594.
[15] S. Park, J. H. An, I. W. Jung, R. D. Piner, S. J. An, X. S. Li,
A. Velamakanni, R. S. Ruoff, Nano Lett. 2009, 9, 15931597.
[16] T. A. Land, T. Michely, R. J. Behm, J. C. Hemminger, G. Comsa, Surf.
Sci. 1992, 264, 261270.
[17] M. Eizenberg, J. M. Blakely, Surf. Sci. 1979, 82, 228236.
[18] K. S. Kim, Y. Zhao, H. Jang, S. Y. Lee, J. M. Kim, K. S. Kim, J.-H. Ahn,
P. Kim, J.-Y. Choi, B. H. Hong, Nature 2009, 457, 706710.
[19] D. V. Kosynkin, A. L. Higginbotham, A. Sinitskii, J. R. Lomeda,
A. Dimiev, B. K. Price, J. M. Tour, Nature 2009, 458, 872876.
[20] L. Jiao, L. Zhang, X. Wang, G. Diankov, H. Dai, Nature 2009, 458,
877880.
[21] D. Li, M. B. Muller, S. Gilje, R. B. Kaner, G. G. Wallace, Nat.
Nanotechnol. 2008, 3, 101105.
[22] L. J. Cote, F. Kim, J. X. Huang, J. Am. Chem. Soc. 2009, 131, 1043
1049.
[23] A. N. Obraztsov, Nat. Nanotechnol. 2009, 4, 212213.
[24] A. H. C. Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov,
A. K. Geim, Rev. Mod. Phys. 2009, 81, 109162.
[25] S. Park, R. S. Ruoff, Nat. Nanotechnol. 2009, 4, 217224.
[26] A. Clause, R. Plass, H.-P. Boehm, U. Hofmann, Z. Anorg. Allg.
Chem. 1957, 291, 205220.
[27] B. C. Brodie, Philos. Trans. R. Soc. London 1859 149, 249259.
[28] L. Staudenmaier, Ber. Dtsch. Chem. Ges. 1898 31, 14811487.
Graphene Oxide- and Graphene-Based Materials
small 2010, 6, No. 6, 711723 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com 721
[29] W. S. Hummers, R. E. Offeman, J. Am. Chem. Soc. 1958, 80, 1339
1339.
[30] C. Hontoria-Lucas, A. J. Lopez-Peinado, J. D. D. Lopez-Gonzalez,
M. L. Rojas-Cervantes, R. M. Martin-Aranda, Carbon 1995, 33,
15851592.
[31] T. Szabo, O. Berkesi, P. Forgo, K. Josepovits, Y. Sanakis,
D. Petridis, I. Dekany, Chem. Mater. 2006, 18, 27402749.
[32] F. S. Hyde, J. Soc. Chem. Ind. 1904, 23, 3003002.
[33] J. A. Johnson, C. J. Benmore, S. Stankovich, R. S. Ruoff, Carbon
2009, 47, 22392243.
[34] M. J. McAllister, J. L. Li, D. H. Adamson, H. C. Schniepp,
A. A. Abdala, J. Liu, M. Herrera-Alonso, D. L. Milius, R. Car,
R. K. Prudhomme, I. A. Aksay, Chem. Mater. 2007, 19, 4396
4404.
[35] A. B. Bourlinos, D. Gournis, D. Petridis, T. Szabo, A. Szeri,
I. Dekany, Langmuir 2003, 19, 60506055.
[36] W. Scholz, H. P. Boehm, Z. Anorg. Allg. Chem. 1969, 369, 327
340.
[37] R. Bissessur, S. F. Scully, Solid State Ionics 2007, 178, 877882.
[38] A. Clauss, R. Plass, H.-P. Boehm, U. Hofmann, Z. Anorg. Allg.
Chem. 1957, 291, 205220.
[39] T. Nakajima, A. Mabuchi, R. Hagiwara, Carbon 1988, 26, 357
361.
[40] W. W. Cai, R. D. Piner, F. J. Stadermann, S. Park, M. A. Shaibat,
Y. Ishii, D. X. Yang, A. Velamakanni, S. J. An, M. Stoller, J. H. An,
D. M. Chen, R. S. Ruoff, Science 2008, 321, 18151817.
[41] A. Lerf, H. He, M. Forster, J. Klinowski, J. Phys. Chem. B 1998, 102,
44774482.
[42] H. Y. He, J. Klinowski, M. Forster, A. Lerf, Chem. Phys. Lett. 1998,
287, 5356.
[43] S. Stankovich, R. D. Piner, S. T. Nguyen, R. S. Ruoff, Carbon 2006,
44, 33423347.
[44] C. Xu, X. Wu, J. Zhu, X. Wang, Carbon 2008, 46, 386389.
[45] G. Williams, B. Seger, P. V. Kamat, ACS Nano 2008, 2, 1487
1491.
[46] D. Y. Cai, M. Song, J. Mater. Chem. 2007, 17, 36783680.
[47] J. I. Paredes, S. Villar-Rodil, A. Martinez-Alonso, J. M. D. Tascon,
Langmuir 2008, 24, 1056010564.
[48] D. G. H. Ballard, G. R. Rideal, J. Mater. Sci. 1983, 18, 545561.
[49] D. A. Dikin, S. Stankovich, E. J. Zimney, R. D. Piner,
G. H. B. Dommett, G. Evmenenko, S. T. Nguyen, R. S. Ruoff, Nature
2007, 448, 457460.
[50] G. Eda, G. Fanchini, M. Chhowalla, Nat. Nanotechnol. 2008, 3,
270274.
[51] H. Chen, M. B. Muller, K. J. Gilmore, G. G. Wallace, D. Li, Adv.
Mater. 2008, 20, 35573561.
[52] O. C. Compton, D. A. Dikin, K. W. Putz, L. C. Brinson, S. T. Nguyen,
Adv. Mater. 2010, 22, 892896.
[53] S. Park, K. Lee, G. Bozoklu, W. Cai, S. T. Nguyen, R. S. Ruoff, ACS
Nano 2008, 2, 572578.
[54] X. Brunetaud, L. Divet, D. Damidot, Cem. Concr. Res. 2008, 38,
13431348.
[55] L. Berhan, Y. B. Yi, A. M. Sastry, E. Munoz, M. Selvidge,
R. Baughman, J. Appl. Phys. 2004, 95, 43354345.
[56] K. H. W. Seah, J. Hemanth, S. C. Sharma, Mater. Des. 1995, 16,
175179.
[57] C. Chen, Q.-H. Yang, Y. Yang, W. Lv, Yuefang en, P.-X. Hou,
M. Wang, H.-M. Cheng, Adv. Mater. 2009, 21, 30073011.
[58] M. Jin, H. K. Jeong, W. J. Yu, D. J. Bae, B. R. Kang, Y. H. Lee, J. Phys.
D-Appl. Phys. 2009, 42, 135109.
[59] N. A. Kotov, I. Dekany, J. H. Fendler, Adv. Mater. 1996, 8, 637
641.
[60] N. I. Kovtyukhova, P. J. Ollivier, B. R. Martin, T. E. Mallouk,
S. A. Chizhik, E. V. Buzaneva, A. D. Gorchinskiy, Chem. Mater.
1999, 11, 771778.
[61] T. Cassagneau, J. H. Fendler, Adv. Mater. 1998, 10, 877
881.
[62] H. A. Becerril, J. Mao, Z. Liu, R. M. Stoltenberg, Z. Bao, Y. Chen,
ACS Nano 2008, 2, 463470.
[63] Z. Q. Wei, D. E. Barlow, P. E. Sheehan, Nano Lett. 2008, 8, 3141
3145.
[64] D. R. Paul, L. M. Robeson, Polymer 2008, 49, 31873204.
[65] G. Choudalakis, A. D. Gotsis, Eur. Polym. J. 2009, 45, 967984.
[66] S. Watcharotone, D. A. Dikin, S. Stankovich, R. Piner, I. Jung,
G. H. B. Dommett, G. Evmenenko, S.-E. Wu, S.-F. Chen, C.-P. Liu,
S. T. Nguyen, R. S. Ruoff, Nano Lett. 2007, 7, 18881892.
[67] L. J. Cote, R. Cruz-Silva, J. Huang, J. Am. Chem. Soc. 2009, 131,
1102711032.
[68] J. L. Vickery, A. J. Patil, S. Mann, Adv. Mater. 2009, 21,
21802184.
[69] D. Cai, K. Yusoh, M. Song, Nanotechnology 2009, 20, 085712/1
5.
[70] Q. Ding, B. Liu, Q. Zhang, Q. He, B. Hu, J. Shen, Polym. Int. 2006,
55, 500504.
[71] X. Dai, J. Xu, X. Guo, Y. Lu, D. Shen, N. Zhao, X. Luo, X. Zhang,
Macromolecules 2004, 37, 56155623.
[72] D. Li, R. B. Kaner, Science 2008, 320, 11701171.
[73] A. K. Geim, Science 2009, 324, 15301534.
[74] K. A. Worsley, P. Ramesh, S. K. Mandal, S. Niyogi, M. E. Itkis,
R. C. Haddon, Chem. Phys. Lett. 2007, 445, 5156.
[75] Y. Hernandez, V. Nicolosi, M. Lotya, F. M. Blighe, Z. Y. Sun, S. De,
I. T. McGovern, B. Holland, M. Byrne, Y. K. Gunko, J. J. Boland,
P. Niraj, G. Duesberg, S. Krishnamurthy, R. Goodhue, J. Hutchison,
V. Scardaci, A. C. Ferrari, J. N. Coleman, Nat. Nanotechnol. 2008,
3, 563568.
[76] H. C. Schniepp, J. L. Li, M. J. McAllister, H. Sai, M. Herrera-Alonso,
D. H. Adamson, R. K. Prudhomme, R. Car, D. A. Saville, I. A. Aksay,
J. Phys. Chem. B 2006, 110, 85358539.
[77] K. N. Kudin, B. Ozbas, H. C. Schniepp, R. K. Prudhomme,
I. A. Aksay, R. Car, Nano Lett. 2007, 8, 3641.
[78] A. B. Bourlinos, V. Georgakilas, R. Zboril, T. A. Steriotis,
A. K. Stubos, Small 2009, 5, 18411845.
[79] P. Blake, P. D. Brimicombe, R. R. Nair, T. J. Booth, D. Jiang,
F. Schedin, L. A. Ponomarenko, S. V. Morozov, H. F. Gleeson,
E. W. Hill, A. K. Geim, K. S. Novoselov, Nano Lett. 2008, 8, 1704
1708.
[80] S. Stankovich, R. D. Piner, X. Q. Chen, N. Q. Wu, S. T. Nguyen,
R. S. Ruoff, J. Mater. Chem. 2006, 16, 155158.
[81] S. Niyogi, E. Bekyarova, M. E. Itkis, J. L. McWilliams, M. A. Hamon,
R. C. Haddon, J. Am. Chem. Soc. 2006, 128, 77207721.
[82] Y. Y. Liang, D. Q. Wu, X. L. Feng, K. Mullen, Adv. Mater. 2009, 21,
16791683.
[83] J. R. Lomeda, C. D. Doyle, D. V. Kosynkin, W.-F. Hwang, J. M. Tour,
J. Am. Chem. Soc. 2008, 130, 1620116206.
[84] Y. Si, E. T. Samulski, Nano Lett. 2008, 8, 16791682.
[85] H. F. Yang, C. S. Shan, F. H. Li, D. X. Han, Q. X. Zhang, L. Niu, Chem.
Commun. 2009, 38803882.
[86] A. J. Patil, J. L. Vickery, T. B. Scott, S. Mann, Adv. Mater. 2009, 21,
31593164.
[87] J. Shen, Y. Hu, C. Li, C. Qin, M. Ye, Small 2009, 5, 8285.
[88] D. H. Everett, Basic Principles of Colloid Science, Royal Society of
Chemistry, Cambridge, UK 1988.
[89] V. C. Tung, M. J. Allen, Y. Yang, R. B. Kaner, Nat. Nanotechnol.
2009, 4, 2529.
[90] C. M. Hansen, Hansen Solubility Parameters: A Users Handbook,
CRC Press, Boca Raton, FL 2000.
[91] X. Li, X. Wang, L. Zhang, S. Lee, H. Dai, Science 2008, 319, 1229
1232.
[92] X. Li, G. Zhang, X. Bai, X. Sun, X. Wang, E. Wang, H. Dai, Nat.
Nanotechnol. 2008, 3, 538542.
[93] C. A. Furtado, U. J. Kim, H. R. Gutierrez, L. Pan, E. C. Dickey,
P. C. Eklund, J. Am. Chem. Soc. 2004, 126, 60956105.
[94] G. Chen, W. Weng, D. Wu, C. Wu, J. Lu, P. Wang, X. Chen, Carbon
2004, 42, 753759.
reviews O. C. Compton and S. T. Nguyen
722 www.small-journal.com 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim small 2010, 6, No. 6, 711723
[95] F. M. Fowkes, F. L. Riddle, Jr, W. E. Pastore, A. A. Weber, Colloids
Surf. 1990, 43, 367387.
[96] C. E. Hamilton, J. R. Lomeda, Z. Sun, J. M. Tour, A. R. Barron, Nano
Lett. 2009, 9, 34603462.
[97] Z. Liu, Z. Wang, X. Yang, K. Ooi, Langmuir 2002, 18, 4926
4932.
[98] Y. Xu, H. Bai, G. Lu, C. Li, G. Shi, J. Am. Chem. Soc. 2008, 130,
58565857.
[99] A. N. Obraztsov, Nat. Nanotechnol. 2009, 4, 212213.
[100] 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,
A. K. Schmid, D. Waldmann, H. B. Weber, T. Seyller, Nat. Mater.
2009, 8, 203207.
[101] X. Wang, L. Zhi, K. Mullen, Nano Lett. 2008, 8, 323327.
[102] J. B. Wu, H. A. Becerril, Z. N. Bao, Z. F. Liu, Y. S. Chen, P. Peumans,
Appl. Phys. Lett. 2008, 92, 263302/13.
[103] R. A. Vaia, H. D. Wagner, Mater. Today 2004, 7, 3237.
[104] N. A. Luechinger, N. Booth, G. Heness, S. Bandyopadhyay,
R. N. Grass, W. J. Stark, Adv. Mater. 2008, 20, 30443049.
[105] 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, 282286.
[106] S. Ansari, E. P. Giannelis, J. Polym. Sci. Part B: Polym. Phys. 2009,
47, 888897.
[107] D. A. Nguyen, Y. R. Lee, A. V. Raghu, H. M. Jeong, C. M. Shin,
B. K. Kim, Polym. Int. 2009, 58, 412417.
[108] T. Ramanathan, A. A. Abdala, S. Stankovich, D. A. Dikin,
M. Herrera-Alonso, R. D. Piner, D. H. Adamson, H. C. Schniepp,
X. Chen, R. S. Ruoff, S. T. Nguyen, I. A. Aksay, R. K. Prudhomme,
L. C. Brinson, Nat. Nanotechnol. 2008, 3, 327331.
[109] T. Ramanathan, S. Stankovich, D. A. Dikin, H. Liu, H. Shen, S. T.
Nguyen, L. C. Brinson, J. Polym. Sci. Part B: Polym. Phys. 2007, 45,
20972112.
[110] M. Fang, K. G. Wang, H. B. Lu, Y. L. Yang, S. Nutt, J. Mater. Chem.
2009, 19, 70987105.
[111] H. Yang, F. Li, C. Shan, D. Han, Q. Zhang, L. Niu, A. Ivaska, J. Mater.
Chem. 2009, 19, 46324638.
[112] R. Verdejo, F. Barroso-Bujans, M. A. Rodriguez-Perez, J. A. de
Saja, M. A. Lopez-Manchado, J. Mater. Chem. 2008, 18, 2221
2226.
[113] D.-W. Wang, F. Li, J. Zhao, W. Ren, Z.-G. Chen, J. Tan, Z.-S. Wu,
I. Gentle, G. Q. Lu, H.-M. Cheng, ACS Nano 2009, 3, 17451752.
[114] P. Asuri, S. S. Karajanagi, R. S. Kane, J. S. Dordick, Small 2007, 3,
5053.
[115] F. H. Li, H. F. Yang, C. S. Shan, Q. X. Zhang, D. X. Han, A. Ivaska,
L. Niu, J. Mater. Chem. 2009, 19, 40224025.
[116] G. M. Scheuermann, L. Rumi, P. Steurer, W. Bannwarth,
R. Mulhaupt, J. Am. Chem. Soc. 2009, 131, 82628270.
[117] B. Seger, P. V. Kamat, J. Phys. Chem. C 2009, 113, 79907995.
[118] Z. Lee, K.-J. Jeon, A. Dato, R. Erni, T. J. Richardson, M. Frenklach,
V. Radmilovic, Nano Lett. 2009, 9, 33653369.
[119] M. Hirata, T. Gotou, S. Horiuchi, M. Fujiwara, M. Ohba, Carbon
2004, 42, 29292937.
[120] S. Gilje, S. Han, M. Wang, K. L. Wang, R. B. Kaner, Nano Lett. 2007,
7, 33943398.
[121] A. A. Green, M. C. Hersam, Nano Lett. 2009, 9, 40314036.
Received: October 12, 2009
Revised: November 26, 2009
Published online: March 11, 2010
Graphene Oxide- and Graphene-Based Materials
small 2010, 6, No. 6, 711723 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.small-journal.com 723

Você também pode gostar