You are on page 1of 3

C

Journal of
Materials Chemistry

Dynamic Article Links <

Cite this: J. Mater. Chem., 2011, 21, 9491

COMMUNICATION

www.rsc.org/materials

Conversion of carbon dioxide to few-layer graphene


Amartya Chakrabarti,ab Jun Lu,c Jennifer C. Skrabutenas,a Tao Xu,*a Zhili Xiao,*c John A. Maguireb
and Narayan S. Hosmane*a
Received 22nd March 2011, Accepted 19th May 2011
DOI: 10.1039/c1jm11227a

Burning magnesium metal in dry ice resulted in few-layer nanosheets


of graphene in high yields. These carbon nanomaterials were
characterized by Raman spectroscopy, energy-dispersive X-ray
analysis, X-ray powder diffraction and transmission electron
microscopy. This work provides an innovative route for producing
one of the most promising carbon nanostructures by capturing
carbon dioxide that is popularly known as the greenhouse gas.
Within the last few years, graphene has received special attention
from the scientific world due to its unique mechanical and electrical
properties.1 The widespread application of graphene is not just
limited to the fields of sensors,2 nanoelectronics,3 composites,4
hydrogen storage,5 lithium-ion batteries,6 but also shown promise in
medicine as antibacterial materials.7 The diversity of the technological
applications of graphene materials drives the search for facile routes
to produce graphenes in high yields. Recent research for synthesis of
such materials involved either chemical or electrochemical reduction
of exfoliated graphite oxide.8 Most of these techniques require the use
of strong oxidizing agents, e.g., H2SO4/KMnO4, while some of them
claimed to have utilized greener chemicals. A recent report described
the production of graphene via reduction of CO using Al2S3.9
Nonetheless, a well-controlled large scale production protocol for
graphene structures is still in great demand and thus we embarked on
research involving graphene synthesis with a particular emphasis on
green chemistry. Consequently, our work described here provides
a new methodology that can be potentially scaled up in directly
capturing CO2 to produce few-layer graphene as useful solid carbon
materials.
Thus, burning magnesium metal in a CO2 environment produces
carbon materials as shown in eqn. (1). Although the metalCO2
propulsion system for Mars missions has been explored,10 the
conversion of CO2 into solid nanostructured carbon materials has
not been reported. Therefore, our approach involving combustion of
magnesium metal in carbon dioxide to form few-layer graphene is

a
Department of Chemistry and Biochemistry, Northern Illinois University,
DeKalb, IL, 60115, USA. E-mail: hosmane@niu.edu; txu@niu.edu
b
Department of Chemistry, Southern Methodist University, Dallas, TX,
75275, USA
c
Department of Physics, Northern Illinois University, DeKalb, IL, 60115,
USA. E-mail: zxiao@niu.edu
Authors contributed equally to the work.

This journal is The Royal Society of Chemistry 2011

unprecedented and provides further incentives for exploration of


a number of environmentally friendly ways for capturing carbon
dioxide.11
2Mgs CO2 g/2MgOs Cs

(1)

In a typical experiment, 3 g of Mg ribbon was ignited inside a dry


ice bowl, covered up by another dry ice slab. After the combustion
of Mg in CO2 was completed, the black products were collected
and transferred to a beaker containing 100 ml of 1 M HCl. The
product was stirred in the acid solution at room temperature
overnight to remove the MgO formed and any remaining Mg
metal. Both Mg and MgO react with hydrochloric acid to form
MgCl2 which is soluble in water and thus it can easily be removed
to get pure carbon material as the product. The mixture was then
filtered, washed with deionized water several times until the filtrate
turned out to be of neutral pH. Finally, the isolated solid carbon
product was dried under high vacuum overnight at 100  C and the
yield was 680 mg (92%). This product was characterized by
transmission electron microscope (TEM, high-resolution TEM was
done in a Philip CM-30 with an acceleration voltage of 250 kV and
low-resolution TEM was performed in a Hitachi H-600
transmission electron microscope at an acceleration voltage of
75 kV), Raman spectroscopy (Renishaw system 2000 microfocus
Raman spectrometer), Energy-dispersive X-ray analysis (EDX,
studies done using field emission scanning electron microscope
Hitachi S-4700-II) and X-ray powder diffraction (XRD, Rigaku
MiniFlex, Cu, 30 kV, 15 mA X-ray) techniques.
Raman spectroscopy is considered to be an effective tool for
characterization of mono- or few-layer graphenes, and several
theoretical and experimental studies have been reported
recently.1216 Raman spectrum of the nanostructured carbon species
obtained during our experiments is depicted in Fig. 1. The two
major components of the spectrum consisted of peaks at 1570 cm1
and 2645 cm1, which are commonly designated as the G-band and
the G0 -band or 2D-band respectively. In a recent study on the
structure of graphene, Ferrari et al. demonstrated clearly that the
number of layers in a graphene structure can be revealed from the
Raman peaks, and thus, graphite can be easily distinguished from
graphenes.12,14 The position and shape of the G0 band in the
Raman spectrum identify the presence and the number of layers of
the graphene structures respectively. With a 633 nm Raman, the G0
band peak of graphene was found at about 2645 cm1,12 which
closely matches our finding as shown in Fig. 1. In the case of
J. Mater. Chem., 2011, 21, 94919493 | 9491

Fig. 1 633 nm Raman spectrum of the carbon species exhibiting


a G0 -band at 2645 cm1, a G-band at 1570 cm1 and a D-band at
1325 cm1. Inset: the expanded view of the G0 -band showing the splitting
pattern of the peak.

monolayer graphene, the G0 band is a sharp single peak; while in


the case of bi- or multi-layer graphenes, there are splittings
generated either from the phonon branches or the electronic bands.
The G0 band shifted more towards 2700 cm1 in the case of
graphenes with more than 710 layers, which is indistinguishable
from graphite.12 From the observed spectrum, splitting of the 2D
band (inset of Fig. 1) and its position provided a strong evidence in
favor of few-layer graphene as the major product. Moreover, the
peak intensities of the G-band and the G0 -band are also related to
the number of layers of the graphene structures. Gupta et al.
compared the peak intensities of different layer structures of
graphene and found that with the number of layers of five or more
the G-band grows higher with respect to the intensity than that of
the G0 -band,13 which again was a confirmation of our product to be
few-layer graphene. The other band found in the spectrum is the
D-band at 1325 cm1, which is of significantly lower intensity and
represents some lattice defects present in the structure.17
Fig. 2a and b show TEM images of the few-layer graphene,
prepared by our new method described above, in which graphene
sheets with varying lengths between 50 nm and 300 nm have been
found. While large crystalline graphene structures are evidently
observed in Fig. 2c, high-resolution TEM (Fig. 2d) clearly exhibits the
signature image of the few-layer graphene with the number of layers
ranging from 37. The measured lattice space of this material is about
 which is in good agreement with the thickness of a mono-layer
3.5 A,
 The inset image in Fig. 2d, corresponding to the
graphene (3.4 A).
diffraction pattern of few-layer graphene, is indicative of the
crystallization.
An X-ray diffraction pattern of our sample is described in Fig. 3.
The prominent 002 peak at 26.3 degree is observed along with the
101 peak at 44.6 degree. The other characteristic peak for graphene
structure is 100 peak located at 2q of 43.2 degree, which is overlapped
with one of MgO peaks. While we examined the purity of the product
via EDX spectroscopy, the absence of any impurity other than a trace
amount of Mg and O in the product was confirmed (Fig. 4). The trace
amount of Mg (2.38 atomic weight percent) and O (7.30 atomic
weight percent) is mainly due to the trapped MgO and some
absorbed O2. Therefore, the contribution of MgO to the peak at 2q of
43.2 degree is small, and this peak can be readily assigned to
100 diffraction of graphenes.
9492 | J. Mater. Chem., 2011, 21, 94919493

Fig. 2 TEM images of few-layer graphene. (a) Graphenes with an


average length of 50100 nm. (b) Larger size graphene sheets, average
length 300 nm. (c) Crystalline graphenes with an average length of
200 nm. (d) High-resolution TEM image of few-layer graphenes, the
number of layers ranging from 37. Inset: the electron diffraction pattern
of graphenes.

Although the exact mechanism of the formation of graphene is still


under investigation, the high temperature generated during the
burning of magnesium metal undoubtedly plays a crucial role. It is
possible that the combustion of the solid magnesium in gaseous CO2
favors the rapid flee of the solid product from the reaction center. As
such, the retention time of the sp2 carbon atoms in the reaction core is
not long enough to form multi-layer graphene (graphite). Instead,
only few-layer graphene is kinetically favored. Experiments on the
effect of catalysts on formation of monolayer graphenes with higher
yields using similar techniques are presently underway in our
laboratories.
In conclusion, the current methodology produces few-layer
graphene captured directly from CO2 by burning Mg in it. The
structure of few-layer graphene product was confirmed by TEM,
Raman spectroscopy and XRD and they are all consistent with the

Fig. 3 XRD pattern of the carbon species.

This journal is The Royal Society of Chemistry 2011

Notes and references

Fig. 4 EDX spectrum of the graphene nanosheets, C 90.32 (atm wt%);


Mg 2.38 (atm wt%); O 7.30 (atm wt%).

data available in the literature. The synthetic process is cost effective


and can be used to produce few-layer graphene in large quantities.
Furthermore, the use of non-toxic chemicals and recyclable
materials during the synthesis constitutes this work as part of green
chemistry.

Acknowledgements
This work was supported by multiple grants from the National
Science Foundation (CHE-0840504 and CHE-0906179 to NSH),
Petroleum Research Fund administered by the American Chemical
Society (Type G 46374-G10 to TX), the Department of Energy
(DE-FG02-06ER46334 to ZX), and Robert A. Welch Foundation
(N-1322 to JAM). NSH also acknowledges the Inaugural Board of
Trustees Professorship from Northern Illinois University and the
Alexander von Humboldt Foundations Research Prize for Senior
Scientists. Structural and compositional analyses were performed at
the Electron Microscopy Center (EMC) of Argonne National
Laboratory which is funded by DOE BES under the contract
DE-AC02-06CH11357. The timely help of Lori Bross at NIU
Biology Department in obtaining TEM images is hereby gratefully
acknowledged.

This journal is The Royal Society of Chemistry 2011

1 K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang,


Y. S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004,
306, 666; K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang,
M. I. Katsnelson, I. V. Grigorieva, S. V. Dubonos and
A. A. Firsov, Nature, 2005, 438, 197; C. N. R. Rao, K. Biswas,
K. S. Subrahmanyam and A. Govindaraj, J. Mater. Chem., 2009,
19, 2457.
2 F. Schedin, A. K. Geim, S. V. Morozov, E. M. Hill, P. Blake,
M. I. Katsnelson and K. Novoselov, Nat. Mater., 2007, 6, 652;
A. Sakhaee-Pour, M. T. Ahmadian and A. Vafai, Solid State
Commun., 2008, 147, 336.
3 Y. W. Son, M. L. Cohen and S. G. Louie, Nature, 2006, 444, 347.
4 S. Stankovich, D. A. Dikin, G. H. B. Dommett, K. M. Kohlhaas,
E. J. Zimney, E. A. Stach, R. D. Pinen, S. T. Nguyen and
R. S. Ruoff, Nature, 2006, 442, 282; D. A. Dikin, S. Stankovich,
E. J. Zimney, R. D. Piner, G. H. B. Dommett, G. Evmenenko,
T. Nguyen and R. S. Ruoff, Nature, 2007, 448, 457.
5 K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich,
S. V. Morozov and A. K. Geim, Proc. Natl. Acad. Sci. U. S. A., 2005,
102, 10451.
6 T. Takamura, K. Endo, L. Fu, Y. P. Wu, K. J. Lee and
T. Matsumoto, Electrochim. Acta, 2007, 53, 1055.
7 W. Hu, C. Peng, W. Luo, X. Li, D. Li, Q. Huang and C. Fan, ACS
Nano, 2010, 4, 4317.
8 G. Wang, J. Yang, J. Park, X. Gou, B. Wang, H. Liu and J. Yao, J.
Phys. Chem. C, 2008, 112, 8192; S. J. Park and R. S. Ruoff, Nat.
Nanotechnol., 2009, 4, 217; H.-L. Guo, X.-F. Wang, Q.-Y. Qian,
F.-B. Wang and X.-H. Xia, ACS Nano, 2009, 3, 2653.
9 C.-D. Kim, B.-K. Min and W.-S. Jung, Carbon, 2009, 47, 1605.
10 E. Y. Shafirovich, A. A. Shiryaev and U. I. Goldshleger, J. Propul.
Power, 1993, 9, 197.
11 (a) D. Monroe, Capturing carbon dioxide. Sci. Am., Feb. 14, 2005.
http://www.scientificamerican.com/article.cfm?idcapturing-carbondioxide; (b) R. Banerjee, A. Phan, B. Wang, C. Knobler,
H. Furukawa, M. OKeeffe and O. M. Yaghi, Science, 2008, 319,
939.
12 A. C. Ferrari, J. C. Meyer, V. Scardaci, C. Casiraghi, M. Lazzeri,
F. Mauri, S. Piscanec, D. Jiang, K. S. Novoselov, S. Roth and
A. K. Geim, Phys. Rev. Lett., 2006, 97, 187401.
13 A. Gupta, G. Chen, P. Joshi, S. Tadigadapa and P. C. Eklund, Nano
Lett., 2006, 6, 2667.
14 A. C. Ferrari, Solid State Commun., 2007, 143, 47.
15 Z. Ni, Y. Wang, T. Yu and Z. Shen, Nano Res., 2008, 1, 273.
16 C. Casiraghi, A. Hartschuh, H. Qian, S. Piscanec, C. Georgi,
A. Fasoli, K. S. Novoselov, D. M. Basko and A. C. Ferrari, Nano
Lett., 2009, 9, 1433.
17 J.-H. Chen, W. G. Cullen, C. Jang, M. S. Fuhrer and E. D. Williams,
Phys. Rev. Lett., 2009, 102, 236805.

J. Mater. Chem., 2011, 21, 94919493 | 9493

You might also like