You are on page 1of 3

5470 Chem. Commun.

, 2011, 47, 54705472 This journal is c The Royal Society of Chemistry 2011
Cite this: Chem. Commun., 2011, 47, 54705472
Graphene solutionsw
Ame lie Catheline,
a
Cristina Valle s,z
a
Carlos Drummond,
a
Luca Ortolani,
b
Vittorio Morandi,
b
Massimo Marcaccio,
c
Matteo Iurlo,
c
Francesco Paolucci
c
and Alain Pe nicaud*
a
Received 24th February 2011, Accepted 24th March 2011
DOI: 10.1039/c1cc11100k
Thermodynamics drive the spontaneous dissolution of a graphite
intercalation compound (GIC) KC
8
in NMP to form stable
solutions. Reduction potential of graphene is measured at
+22 mV vs. SCE. Single layer graphene akes (ca. 1 lm
2
)
have been unambiguously identied by electron diraction.
Since 2004 and the experimental rise of monolayer graphite as
graphene,
1
there has been a need for a synthesis of large
quantities of graphene, for large scale applications.
2
Several
up-to-date reviews have appeared recently on graphene
35
and chemical routes to graphene.
68
The ultimate goal of
chemical graphene or solution route graphene is to obtain
large quantities of liquid formulations of graphene for
composite formulation, surface patterning with graphene
akes, for sensors, transparent electrodes, self-standing lms,
etc. The most explored route has been the chemical reduction
of graphite oxide.
8,9
An alternative approach consists in
dispersing graphite itself by sonication-induced dispersion in
organic solvents,
1015
or in water with surfactants,
16,17
or
p-stacking pyrenic molecules.
18
Oleum-
19
and ClF
3
-intercalated
20
graphite have also been dispersed using sonication. It was
shown recently that the ternary GIC K(THF)C
24
leads to
solutions of graphene in N-methyl pyrrolidone (NMP).
21
We
now report that more concentrated and chemically simpler
solutions can be obtained by mild, spontaneous, dissolution of
the parent GIC KC
8
22
in NMP. The solutions only contain
negatively charged graphene sheets and K
+
ions. A rationale
in terms of negative free energy of dissolution is provided.
Stage 1 GICs, such as KC
8
, are made up of negatively
charged graphene layers, separated by layers of ordered
countercations.
23
KC
8
was prepared according to literature
procedures (cf. ESIw). Exposure to dry, freshly distilled NMP
leads to spontaneous dissolution, i.e. true solubilization,
rather than energy-aided metastable dispersions. After stirring
overnight (to accelerate the process) and mild centrifugation to
remove insoluble material, the dry extract of the solution gave
a concentration of 0.7 mg ml
1
(yield = 35% of the starting
material, cf. ESIw). The solutions were found to be stable
under an inert atmosphere. Their absorption spectrum is
characterized by a band in the UV range at 300 nm (4.14 eV)
with no tailing into higher wavelength region (full width at
half maximum (FWHM) = 50 nm). From series of dilutions
of two dierent solutions, an extinction coecient e
300
of
25 L g
1
cm
1
was obtained. Expressed in moles of carbon
atoms, e
300
= 0.18 L mol
1
cm
1
. It is remarkable that two
dierent solutions from dierent graphite sources (natural and
expanded) gave the same value for the extinction coecient.
This 300 nm band was associated to the presence of charged
(reduced) graphene akes in solution because of its disappearance
upon both electrochemical and air oxidations. Upon exposure
to air, its intensity dramatically decreases within ca. 1 hour
(Fig. 1, top). Electrochemical oxidation under an inert atmosphere
gave similar results (Fig. 1, bottom): close to full disappearance
Fig. 1 Evolution of absorption spectra of KC
8
solutions in NMP
upon air oxidation (top) and electrochemical oxidation (bottom). OCP
stands for open-circuit potential. Inset: non-linear t of the 300 nm
peak intensity vs. applied potential according to Nernst law (see text).
The inexion point yields graphene reduction potential, E
1/2
.
a
Universite de Bordeaux, Centre de Recherche Paul Pascal-CNRS,
115 av. Schweitzer, 33600, Pessac, France.
E-mail: penicaud@crpp-bordeaux.cnrs.fr; Fax: +33 55684 5600;
Tel: +33 55684 3028
b
CNR IMM-Bologna, Via Gobetti 101, 40129 Bologna, Italy
c
INSTM, Unit of Bologna, Dipartimento di Chimica,
Universita` di Bologna, Via Selmi 2, I-40126, Bologna, Italy
w Electronic supplementary information (ESI) available: KC
8
synthesis
and dissolution procedure, graphene deposition on wafers, XPS analysis,
TEM and absorption technical data, calculation of average ake size
from solution conductivity measurements. See DOI: 10.1039/c1cc11100k
z Present address: Instituto de Carboquimica (CSIC), Department of
Nanotechnology, C/Miguel Luesma Castan 4, E-50018 Zaragoza,
Spain.
ChemComm
Dynamic Article Links
www.rsc.org/chemcomm COMMUNICATION
D
o
w
n
l
o
a
d
e
d

b
y

C
N
R

B
o
l
o
g
n
a

o
n

0
1

J
u
l
y

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

1
1

A
p
r
i
l

2
0
1
1

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
C
1
C
C
1
1
1
0
0
K
View Online
This journal is c The Royal Society of Chemistry 2011 Chem. Commun., 2011, 47, 54705472 5471
of the 300 nm peak was observed upon application of poten-
tials up to B0.1 V with the concomitant growth of a novel
band (shoulder) at 340 nm. This behaviour is reversible (i.e.,
the original intensity is recovered by applying potentials
r0.5 V). By contrast, the application of higher potential
leads to irreversible behaviour manifested, e.g., by the sponta-
neous spectral evolution displayed in Fig. 1, bottom. Such an
evolution leads to the formation, at the expense of the band at
340 nm, of a narrow band at 292 nm and a very broad one
centered around 400410 nm (with two well-dened isosbestic
points at 315 and 358 nm respectively). Importantly, such a
spectrum retains most of the features observed in the case of
air oxidation (Fig. 1, top). The analysis of the spectroelectro-
chemical results shown in Fig. 1 (limited to E r0.15 V because
of the above irreversible behaviour) allowed us to obtain the
reduction potential of graphene: the absorbance of the sample is
described as a function of electrochemical potential (inset, Fig. 1
bottom) according to Nernst equation and assuming that the
absorbance at 300 nm is only due to reduced graphene
(and according to LambertBeer law).
24
A standard potential
of +22 mV vs. SCE was obtained, i.e., 4.7 eV in energy units
(SCE = 4.68 eV vs. vacuum),
25
that compares extremely well
with a graphene Fermi energy of 4.6 eV.
26
The presence of charged species in the KC
8
solutions is also
evidenced by low frequency conductivity measurements
(Fig. 2). The theoretical conductivity of potassium ions
(dashed blue line) is similar to the measured conductivity of
KCl solutions, indicating that signicant formation of ionic
pairs occurs in this case. In contrast, the conductivity of KC
8
solutions greatly surpasses the possible contribution of potassium
ions in solution, as a consequence of the contribution of
charged graphenic species. The conductivity of the solutions
saturates at the largest concentrations investigated. A similar
behaviour has been reported before for polyelectrolyte solutions
in semi-diluted and concentrated regimes,
27,28
and has been
attributed to the overlap of the poly-ion chains and the
increased fractions of condensed counterions. Assuming graphene
akes overlap at B1 mM [K] and are square shaped, a typical
lateral ake size of 5 mm can be estimated (cf. ESIw). Light
scattering (inset, Fig. 2) also indicates the presence of large
dissolved species through Tyndall eect: a laser going through
the solution has a visible path due to scattering by the
particles.
Deposits from KC
8
solutions were performed on several
substrates/grids, dried and air-exposed. As for carbon nano-
tubes salts,
29
we expect the highly reducing negatively charged
graphene to reduce oxygen to superoxide that further evolves
into potassium oxide or hydroxide. Further rinsing with
acetone, water and isopropanol removes NMP traces and
potassium oxide/hydroxide. After drying under vacuum at
200 1C for 2 days, XPS analysis on silicon wafers shows an
unfunctionalized carbon C1S peak and absence of potassium
and nitrogen peaks (cf. ESIw). Graphene akes were deposited
on TEM grids by tension-assisted dip-coating of the grids into
the solution of KC
8
. Fig. 3a shows the electron micrograph of
graphene akes. At large defocus, the akes appear as darker
regions over the amorphous carbon support of the grid. A
large number of akes of typical sizes up to one micron, either
single layer or re-aggregated few layers akes have been
observed. In Fig. 3a the borders of an individual layer ake
have been highlighted. Discrimination between mono- and
multi-layer akes was done analyzing their electron diraction
patterns. Fig. 3b shows the diraction intensity prole, along
the line indicated in the pattern in the inset, obtained from the
ake highlighted in Fig. 3a (electron beam perpendicular to
the ake surface). The intensity of the diraction spots of the
inner hexagon is greater than that of the outer hexagon. A
specic ngerprint of genuine single layer crystals is that in the
diraction pattern the intensity of the inner reections is
always stronger than that of the outer ones, irrespectively of
the orientation of the electron beam with the surface of the
ake.
30
For all the akes analyzed, diraction patterns were
acquired at dierent tilt angles, and the intensity of reections
was compared. Fig. 3c shows the intensity prole from the
diraction pattern in the inset, acquired over the same ake
with the electron beam tilted by 101. The intensity of the inner
hexagon reection is still more intense than the outer ones,
conrming the ake single layer nature.
Fig. 2 Conductivity vs. K
+
concentration of solutions of KC
8
(black
squares) and KCl (red circles) in NMP (n = 1 kHz, similar results were
obtained between 50 Hz and 4 kHz). The blue line represents an
estimation of the contribution of potassium ions to the conductivity in
this solvent.
35
Inset: the presence of large dissolved particles in the
KC
8
solutions is evidenced by intense light scattering of a laser beam,
which is not detected in NMP alone.
Fig. 3 TEM and electron diraction (ED) of graphene deposits. (a)
TEM micrograph showing graphene and few layer akes onto a holey
carbon lm. The borders of one graphene (mono-layer) crystal have
been highlighted. (b) ED intensity prole for perpendicular electrons
incidence over a graphene ake. The prole is taken along the white
line in the ED pattern shown in the inset. (c) Same as (b) for tilted
beam incidence (101) over the same graphene ake.
D
o
w
n
l
o
a
d
e
d

b
y

C
N
R

B
o
l
o
g
n
a

o
n

0
1

J
u
l
y

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

1
1

A
p
r
i
l

2
0
1
1

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
C
1
C
C
1
1
1
0
0
K
View Online
5472 Chem. Commun., 2011, 47, 54705472 This journal is c The Royal Society of Chemistry 2011
The well known anity of NMP for carbon nanotubes or
graphite has recently been rationalized in terms of solvent
solubility parameters,
31
showing a minimum enthalpy loss
upon dispersion. Nonetheless, graphite remains insoluble in
NMP. A fraction can be dispersed at the cost of extensive
sonication.
1114
On the other hand, we showed a few years ago
that alkali metal salts of nanotubes spontaneously dissolve in
polar solvents.
32
Counterions entropy gain has long been
known to promote dissolution of polyelectrolytes.
33
A complete
description of KC
8
dissolution process seems out of reach
because of the complexity and polydispersity of the system.
However, we can estimate the most relevant factors. The free
energy change is given by DG
dis
= DH
dis
TDS
dis
, where T is
the absolute temperature, and DG
dis
, DH
dis
and DS
dis
are the
free energy, enthalpy and entropy of dissolution. A negative
change of free energy implies a thermodynamically stable
solution. We need to consider the cohesive lattice term vs.
the solvation energy of the potassium counterions and of
the charged graphene planes. Doyen-Lang et al. reported a
Madelung energy value of 239 kJ mol
1
for KC
8
and an
interaction energy between the graphite layers of 26.5 kJ mol
1
KC
8
.
34
The contribution of the solvation of the potassium
counterions in NMP, which includes the exothermic solvation
process and the entropy gain of the solvated counterions,
amounts to DG
solv
= 335 kJ per mol of K.
35
This energy
gain largely prevails over the lattice contributions, and is the
main driving force for the stability of the KC
8
solutions. Finally
we have to account for the solvation of the charged graphene
planes. Coleman and coworkers
14
used the Hansen solubility
parameters to evaluate the possibility of dispersing graphene in
dierent solvents. Using their results for the case of NMP, a
small positive free energy change of the order of 0.1 kJ mol
1
graphene is obtained. This small value indicates that the NMP
graphene interaction energy is very similar to the interaction
between graphene planes. However, this estimate involves the
dissolution of neutral graphene, which entails the penalty of
separating closely interacting graphene layers. The unfavourable
enthalpic contribution from the separation of the neighbouring
graphene layers in KC
8
is obviously much smaller: 3.3 kJ mol
1
C
vs. 22.8 kJ mol
1
C in graphite.
34
This dierence already
argues in the direction of a thermodynamically stable dissolution
of KC
8
: the dispersion of the graphene planes would amount
to a favourable free energy change of 155 kJ mol
1
KC
8
.
However, it is the presence of the charges in the reduced
graphene and the solvated counterions that act against the
reaggregation of the graphene planes, stabilizing the solution.
A.C. thanks Arkema and l
0
Association Nationale pour la
Recherche et la Technologie (ANRT) for a PhD grant. Support
from the Agence Nationale de la Recherche (GRAAL), Italian
MIUR (PRIN) and University of Bologna is acknowledged.
This work has been performed within the framework of the
GDR-I 3217 graphene and nanotubes.
Notes and references
1 K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang,
S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004,
306, 666.
2 A. K. Geim and K. S. Novoselov, Nat. Mater., 2007, 6, 183.
3 A. K. Geim, Science, 2009, 324, 1530.
4 M. J. Allen, V. C. Tung and R. B. Kaner, Chem. Rev., 2010, 110,
132.
5 C. Soldano, A. Mahmood and E. Dujardin, Carbon, 2010, 48,
2127.
6 R. Y. N. Gengler, K. Spyrou and P. Rudolf, J. Phys. D: Appl.
Phys., 2010, 43, 374015.
7 R. Ruo, Nat. Nanotechnol., 2008, 3, 10.
8 S. Park and R. S. Ruo, Nat. Nanotechnol., 2009, 4, 217.
9 S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas,
A. Kleinhammes, Y. Jia, Y. Wu, S. T. Nguyen and R. S. Ruo,
Carbon, 2007, 45, 1558.
10 P. Blake, P. D. Brimicombe, R. R. Nair, T. J. Booth, D. Jiang,
F. Schedin, L. A. Ponomarenko, S. V. Morozov, H. F. Gleeson
and E. W. Hill, et al., Nano Lett., 2008, 8, 1704.
11 Y. Hernandez, V. Nicolosi, M. Lotya, F. M. Blighe, Z. Sun, S. De,
I. McGovern, B. Holland, M. Byrne and Y. Gunko, et al., Nat.
Nanotechnol., 2008, 3, 563.
12 C. E. Hamilton, J. R. Lomeda, Z. Sun, J. M. Tour and
A. R. Barron, Nano Lett., 2009, 9, 3460.
13 A. B. Bourlinos, V. Georgakilas, R. Zboril, T. A. Steriotis and A. K.
Stubos, Small, 2009, 5, 1841.
14 Y. Hernandez, M. Lotya, D. Rickard, S. D. Bergin and
J. N. Coleman, Langmuir, 2010, 26, 3208.
15 X. Li, X. Wang, L. Zhang, S. Lee and H. Dai, Science, 2008, 319,
1229.
16 M. Lotya, Y. Hernandez, P. J. King, R. J. Smith, V. Nicolosi,
L. S. Karlsson, F. M. Blighe, S. De, Z. Wang and I. T. McGovern,
et al., J. Am. Chem. Soc., 2009, 131, 3611.
17 Z. Sun, T. Hasan, F. Torrisi, D. Popa, G. Privitera, F. Wang,
F. Bonaccorso, D. M. Basko and A. C. Ferrari, ACS Nano, 2010,
4, 803.
18 X. An, T. Simmons, R. Shah, C. Wolfe, K. M. Lewis,
M. Washington, S. K. Nayak, S. Talapatra and S. Kar, Nano
Lett., 2010, 10, 4295.
19 X. Li, G. Zhang, X. Bai, X. Sun, X. Wang, E. Wang and H. Dai,
Nat. Nanotechnol., 2008, 3, 538.
20 J. H. Lee, D. W. Shin, V. G. Makotchenko, A. S. Nazarov,
V. E. Fedorov, Y. H. Kim, J.-Y. Choi, J. M. Kim and
J.-B. Yoo, Adv. Mater., 2009, 21, 4383.
21 C. Valle s, C. Drummond, H. Saadaoui, C. A. Furtado, M. He,
O. Roubeau, L. Ortolani, M. Monthioux and A. Pe nicaud, J. Am.
Chem. Soc., 2008, 130, 15802.
22 E. Nixon and S. Parry, Nature, 1967, 216, 909.
23 M. S. Dresselhaus and G. Dresselhaus, Adv. Phys., 1981, 30,
139326.
24 D. Paolucci, M. Melle Franco, M. Iurlo, M. Marcaccio, M. Prato,
F. Zerbetto, A. Pe nicaud and F. Paolucci, J. Am. Chem. Soc., 2008,
130, 7393.
25 S. Trasatti, Pure Appl. Chem., 1986, 58, 955.
26 D. Marchand, C. Fre tigny, M. Lague s, F. Batallan, C. Simon,
I. Rosenman and R. Pinchaux, Phys. Rev. B, 1984, 30, 4788.
27 F. Bordi, R. H. Colby, C. Cametti, L. De Lorenzo and T. Gili,
J. Phys. Chem. B, 2002, 106, 6887.
28 C. Wandrey, Langmuir, 1999, 15, 4069.
29 A. Pe nicaud, L. Valat, A. Derre , P. Poulin, C. Zakri, O. Roubeau,
M. Maugey, P. Miaudet, E. Anglaret and P. Petit, et al., Compos.
Sci. Technol., 2007, 67, 795.
30 J. C. Meyer, A. K. Geim, M. I. Katsnelson, K. S. Novoselov,
T. J. Booth and S. Roth, Nature, 2007, 446, 60.
31 S. D. Bergin, Z. Sun, D. Rickard, P. V. Streich, J. P. Hamilton and
J. N. Coleman, ACS Nano, 2009, 3, 2340.
32 A. Pe nicaud, P. Poulin, A. Derre , E. Anglaret and P. Petit, J. Am.
Chem. Soc., 2005, 127, 8.
33 R. R. Netz and D. Andelman, Polyelectrolytes in solution and at
surfaces, in Encyclopedia of Electrochemistry, ed. M. Urbakh
and E. Giladi, Wiley-VCH, Weinheim, 2002, vol. I, ch. 2.7,
pp. 282.
34 S. Doyen-Lang, A. Charlier, L. Lang and M. F. Charlier, Synth.
Met., 1993, 58, 95.
35 K. Izutsu, Electrochemistry in Nonaqueous Solutions, Wiley, 2002.
D
o
w
n
l
o
a
d
e
d

b
y

C
N
R

B
o
l
o
g
n
a

o
n

0
1

J
u
l
y

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

1
1

A
p
r
i
l

2
0
1
1

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
C
1
C
C
1
1
1
0
0
K
View Online

You might also like