You are on page 1of 5

Materials Transactions, Vol. 51, No. 8 (2010) pp.

1455 to 1459
#2010 The Mining and Materials Processing Institute of Japan

EXPRESS REGULAR ARTICLE

Production of Graphite Nanosheets by Low-Current Plasma Discharge


in Liquid Ethanol
Sunghoon Kim1;2 , Ruslan Sergiienko2 , Etsuro Shibata2; * ,
Yuichiro Hayasaka3 and Takashi Nakamura2
1

Samsung Electro-Mechanics Co. Ltd., 581 Myunghak-Li, Dong-Myon, Yeongi-Gun, Chungcheongham-Do, 339-702, Korea
Institute of Multidisciplinary Research for Advanced Materials (IMRAM), Tohoku University, Sendai 980-8577, Japan
3
High Voltage Electron Microscope laboratory, Tohoku University, Sendai 980-8577, Japan
2

Graphite nanosheets were produced by low-current plasma discharge in ultrasonically cavitated liquid ethanol. The microstructure,
morphology and thickness of the graphite nanosheets were characterized by scanning and transmission electron microscopy, X-ray diraction,
dynamic force microscopy and Raman spectroscopy. The results indicated that the synthesized nanosheets have many folds, curled edges and are
up to 11 mm in extent. The graphite nanosheets typically ranged in thickness from 6.7 to 23.5 nm. The proposed method is substrate-free, does not
require expensive vacuum equipment and nor does it consume large amounts of electricity. [doi:10.2320/matertrans.M-M2010813]
(Received December 4, 2009; Accepted April 20, 2010; Published July 14, 2010)
Keywords: plasma discharge, ultrasonic cavitation, graphite nanosheets, transmission electron microscopy, dynamic force microscopy, X-ray
diraction, Raman spectroscopy

1.

Introduction

Graphite nanosheets (GNSs)13) (sometimes called carbon


nanosheets, nanoakes,4,5) or nanowalls6,7)) have high surface-to-volume ratio and exceptional electrical and mechanical properties, in addition to high thermal and chemical
stability. At the time of writing, carbon nanosheets are being
considered as electrode materials and catalyst supports in
electrochemical supercapacitors and fuel cells,8,9) lithium-ion
batteries10,11) and in eld emission5,12) perfomance, and as
ller in polymer composites.13,14) In graphite nanosheets, the
weak van der Waals bonding between adjacent graphene
layers make them promising candidates for exfoliation into
single-graphene sheets. The synthesis of graphite nanosheets
and single-graphene sheets has been previously demonstrated
by using radio-frequency or microwave plasma-enhanced
chemical vapor deposition (PECVD),47) arc-discharge
methods,15,16) chemical reduction of exfoliated graphite
oxide17,18) and a solvothermal synthesis.19,20) However, from
a technical standpoint, chemical vapor deposition (CVD) and
PECVD methods are low yielding and require sophisticated
apparatus, controlled atmosphere, gas ow adjustments, and
ammable gaseous mixtures. Equally, conventional arcdischarge methods require expensive vacuum equipment and
the electric power requirement for arc-discharge usually
exceeds 1 kW.15,16) The process of graphite oxidation has
been suggested as the most generally eective way to
produce GNSs and/or single-graphene sheets in a large
quantities and at low cost. However, the GNSs and singlegraphene sheets obtained by this method are usually of
comparatively poor quality, mainly due to the introduction
of oxygen-containing functional groups during the synthesis,
which consequently limits further application, especially
as electrically conductive composites and electronic nanodevices. In addition, the time-consuming process of graphite
oxidation in the presence of strong acids and oxidants
*Corresponding

author, E-mail: etsuro@tagen.tohoku.ac.jp

requires specic precautions to minimize the risk of


explosion.17)
In this study, we propose a new approach for the
milligram-scale production of graphite nanosheets using
low-current plasma discharge in a liquid ethanol ultrasonic
cavitation eld. This method is safe, simple, and substratefree, and in contrast to PECVD does not require expensive
vacuum equipment. A plasma discharge produced under
conditions of ultrasonication is relatively safe, even in an
organic liquid because a low electric current is used
compared with arc-discharge methods. This method has
already been described in the previous paper, but only as a
route for the manufacture of carbon nanocapsules.21)
2.

Experimental Procedures

Low-current plasma discharge was generated between a


consumed 50 at% Fe-50 at% Platinum alloy cylindrical anode
(3 mm) and the bottom of a titanium ultrasonic horn which
serves as a cathode (Fig. 1(a)). The voltage between the
anode and cathode was held at 55 V and the upper limit of
the current on the electrodes was set at 3.0 A throughout
the experiment. The eervescent ultrasonic cavitation eld
may enhance electrical conductivity due to the high-energy
speciesradicals, atoms, ions and free electrons that form
within it. Generation of the plasma discharge begins with the
process of ultrasonic cavitation and pointed end geometry of
the consumed anode assists the emission of electrons (e )
from the cathode (Fig. 1(a)).
A large amount of graphite nanosheets and Fe-Pt alloy
lled carbon nanocapsules were found using a consumed
50 at% Fe-50 at% Pt alloy anode, whilst using a pure Fe
anode21) produced a signicantly lower number of graphite
nanosheets. The nature of the inuence of Pt on GNSs
synthesis has not yet been resolved. It is possible that Pt plays
a key catalytic role in the reduction of liquid ethanol, but in
this short paper we are not examining the mechanism of
graphite nanosheet formation.

1456

S. Kim, R. Sergiienko, E. Shibata, Y. Hayasaka and T. Nakamura


(b)
(a)
Ultrasonic generator

Ti ultrasonic
horn cathode

Plasma
discharge
3 A, 55 V

e-

Me

Fe tip

Ethanol

gap 100 m

C2

Ethanol decomposition

DC power
supply

Metallic vapor

(c)

Ultrasonic cavitation
(tiny bubbles)

(Fe50Pt50 alloy
anode)

Fig. 1 (a) Schematic view of experimental apparatus. (b) and (c) show typical SEM images of the synthesized graphite nanosheets freely
suspended on a carbon TEM microgrid. 1 (b) Curled edges (c) folded central regions indicated by arrows.

After the synthesis of the graphite nanosheets, the material


was reuxed in 1015% hydrogen peroxide aqueous solution
at 90 C for 12 hours to remove impurities of amorphous
carbon by selective oxidation. The carbon powder sample
was then etched in aqua regia for 24 at 40 C to remove
exposed Fe-Pt alloy nanoparticles. Finally, the Fe-Pt alloy
lled carbon nanocapsules were removed by permanent
magnets,21) leaving the dispersed graphite nanosheets in the
liquid ethanol.
The microstructure, morphology and thickness of graphite
nanosheets were characterized by eld emission scanning
electron microscopy (FE-SEM) (JEOL, JSM-7000F) and
transmission electron microscopy (TEM) (80 kV and 300 kV,
JEOL), X-ray diraction (XRD) (Rigaku, RINT2000),
dynamic force microscopy (DFM) (NanoNavi/L-trace II)
and visible Raman spectroscopy (HoloLab 5000/Raman
Rxn1) using a second harmonic of Nd:YAG laser at 532 nm
wavelength as an excitation.
3.

Experimental Results and Discussion

Typical SEM images of the synthesized GNSs, freely


suspended on a TEM carbon grid, are shown in Figs. 1(b),
(c). We measured the average lateral extent of 171 graphite
nanosheets using SEM images, and found that the sizes of the
GNSs ranged from several hundred nanometers to 11 mm
(Fig. 2(f)) with average size of 45 mm. The nanosheets are
also transparent to an electron beam (Figs. 1(b), (c)) and

irregularly shaped. We frequently observed curled edges


(Fig. 1(b), Figs. 2(a), (c) and (d)) and folds (Fig. 1(c)) in the
central regions of the synthesized GNSs along with plane
featureless regions. As reported previously by Meyer et al.,22)
corrugation and scrolling are intrinsic to nanosheets.
Selected area electron diraction (SAED) (Fig. 2(b))
along the [111] zone axis (marked with a circle in Fig. 2(a))
clearly shows the typical hexagonally arranged carbon lattice
in the nanosheet, i.e. hexagonal closed packed structure
(HCP). The well-dened diraction spots (as seen in
Fig. 2(b)) and layered structures (Figs. 2(c), (d)) conrm
the crystalline structure of the graphite nanosheets obtained by low-current plasma discharge in liquid ethanol.
Figures 2(c) and (d) shows high-resolution TEM images of
the curled edges of two dierent graphite nanosheets.
Aligned graphene layers with an interlayer spacing of about
0.34 nm are clearly visible, corresponding to 002 graphite
crystal spacing (from ASTM card # 41-1487). Curled edges
can provide a clear TEM signature for the number of
graphene layers by direct visualization, since at a curled edge
the sheet is locally parallel to the electron beam. Therefore,
the thickness of graphite nanosheets can be estimated from
high-resolution TEM images (HRTEM).22) Counting the
number of graphene layers in the curled edges in the HRTEM
images (Figs. 2(c), (d)), revealed 36, and 13 layers and
thence the respective thicknesses of graphite nanosheets can
be evaluated at about 12.2, and 4.4 nm at an interlayer
distance of 0.34 nm.

Production of Graphite Nanosheets by Low-Current Plasma Discharge in Liquid Ethanol

(c)

40
Number of measurements

(a)

1457

Thickness of nanosheets

35

(e)

Height of folds and curled


edges

30
25
20
15
10
5
0
0

(b)

20

40

60

80

100

120

140

160

Thickness (nm)

(d)
16

(f)

Number of nanosheets

14
12
10
8
6
4
2
0
2

10

12

Lateral size (m)

Fig. 2 (a) shows typical TEM image of the transparent graphite nanosheet. Curled edges, indicated by arrows, exhibit multiple dark lines.
(b) shows a SAED pattern of the nanosheet taken from the position marked by a circle in (a). The diameter of the circle is equal to size of a
selected area aperture, 950 nm (a). (c), (d) High-resolution TEM images of graphene layers in the curled edges of graphite nanosheets.
The number of graphene layers is about 36 and 13, respectively. (e) Histogram of graphite nanosheets thicknesses and the heights of folds
and curled edges above substrate level taken from height proles of DFM images of 70 sheets. (f) Histogram of graphite nanosheets
lateral extents taken from SEM images of 171 sheets.

Dynamic force microscopy measurements were employed


to quantitatively analyze the thickness of the graphite nanosheets, and some results are shown in Fig. 3. Figures 3(a) and
(b) exhibit typical topography images of dierent graphite
nanosheets and height proles (Figs. 3(c) and (d)) through
those graphite nanosheets along the white line as shown in
Figs. 3(a) and (b). The minimum step heights measured
between the surface of the graphite nanosheets and the silicon
substrates were found to be 5.3 and 7.4 nm, which most
probably correspond to the actual thicknesses of the GNSs
with about 15 and 21 respective graphene layers. However,
it cannot be unambiguously shown that the entire graphite
nanosheet has uniform thickness because folds in the central
regions, and curled edges signicantly increase the height
prole of sheet above the substrate level. The white areas
of folds and curled edges in Figs. 3(a) and (b) depict the
highest places above substrate level. The thicknesses of 70
nanosheets were analyzed using height proles and the
histogram is presented in Fig. 2(e). Each sheet was measured
in two or three places, and the minimum measured step
height of each sheet was considered as the thickness. The
most frequently observed minimum step heights ranged from
6.7 to 23.5 nm, and about 50% of measurements showed
minimum step heights in the range of 12.518 nm. The
heights of folds and curled edges above substrate level are
also shown in histogram form (Fig. 2(e)).
Figure 4 shows the X-ray diraction (XRD) pattern of the
as-prepared carbon powder sample and the puried graphite

nanosheets. The as-prepared carbon powder exhibited the


characteristic graphite and face-centered cubic (FCC) (Fe, Pt) diraction peaks at 26.34 , 40.67 , 47.32 , 69.42 ,
marked by their indices (002), (111), (200), (220) (top prole
in Fig. 4). It was observed that carbon nanocapsules with face
centered cubic -(Fe, Pt) core structures were eectively
removed by magnetic separation, although some contaminants due to the Pt rich carbon nanocapsules (diraction
peaks at 39.83 and 46.34 , respectively, for Pt(111) and
Pt(200)) remained on the surface of the graphite nanosheets
(bottom prole in Fig. 4). The XRD pattern of puried
graphite nanosheets gives a distinguishable (002) graphite
peak at 26.34 .
The Raman spectrum may provide more supporting
evidence about the nature of the structure and morphology,
in particular to determine the defects and the ordered and
disordered structures of carbon nanomaterials. Figure 5
compares the Raman spectra of reference graphite powder
with synthesized GNSs. It is obvious from this comparison
that nanosheets have a graphitic structure. In particular,
Fig. 5 shows the D peak (1349 cm1 ), the G peak
(1583 cm1 ) and the D0 peak (a shoulder at 1620 cm1 ),
which are also seen in microcrystalline graphite,23) indicating
that the synthesized nanosheets have a crystalline graphite
structure which contains defects. Second-order modes in
the range of 20003000 cm1 are also present in Fig. 5.
The strong peak at 2698 cm1 , the so-called G0 peak, is an
overtone of the D peak (2  1349 cm1 ). The small peak at

1458

S. Kim, R. Sergiienko, E. Shibata, Y. Hayasaka and T. Nakamura

(a)

(b)

25

(d)

30
20
10

20

15

10
5.3nm

44.4nm

40

18.6nm

Height profile (nm)

(c)

7.4nm

Height profile (nm)

50

0
0

500

1000

1500

2000

2500

500

1000

1500

2000

2500

3000

Lateral size (nm)

Lateral size (nm)

Fig. 3 (a), (b) Tapping-mode DFM image of graphite nanosheets deposited on silicon substrates. White areas depict folds and curled
edges with peak heights up to 18.6 and 44.4 nm. (c), (d) Height proles through the white lines shown in (a) and (b), respectively.

G-1583

( Fe, Pt) (111)

Graphite nanosheets

As-prepared carbon powder

Intensity (a.u.)

( Fe, Pt) (200)


G (002)

( Fe, Pt) (220)

G (002)

Intensity (a.u.)

Reference graphite

G (2D)-2698

D-1349
I

D -1620
D+G
(2932)

Graphite nanosheets

2450

Pt (111)
Pt (200)

10

20

30

40

50

60

70

80

Angle, 2 /

Fig. 4 X-ray diraction patterns of the as-prepared carbon powder sample


(top prole) and puried graphite nanosheets (bottom prole).

2932 cm1 is attributed to the sum of the D and G peaks.


From the position and shape of the G0 (2D) peak as well
as the intensity ratio of the 2D peak to the G peak it could
be concluded that the synthesized graphite nanosheets are
not single-layer graphene sheets.24,25) Single-layer graphene
sheets have a single, sharp 2D peak distinctly below
2700 cm1 (at  26722679 cm1 )26,27) and the second-order
2D peak is more intense than the G peak.24,25) In our results
the intensity of 2D peak in synthesized graphite sheets is
about 1.7 times less than the intensity of G peak, and the

1000

1250

1500

1750

2000

2250

2500

2750

3000

-1

Raman shift (cm )

Fig. 5 Raman spectra of the puried graphite nanosheets and reference


graphite powder (12 mm, Aldrich). Measurements were done from the
aggregation of graphite nanosheets deposited on the silicon substrate.
Raman spectra were excited with a 532 nm laser using a laser spot size of
about 10 mm.

position is at 2698 cm1 . The shape of the 2D peak coincides


with those observed in the reference graphite powder
(Fig. 5). Graphite sheets with more than 10 graphene layers
and bulk graphite exhibit similar 2D peaks.24,25) In conclusion we can say that the synthesized graphite sheets are
multi-layer graphene sheets and most probably contain more
than 10 graphene layers.

Production of Graphite Nanosheets by Low-Current Plasma Discharge in Liquid Ethanol

The appearance of a D peak at 1349 cm1 is attributed to


defects or structural disorder in graphite nanosheets, which
are not observed in single crystals of graphite28) and in
perfect graphene sheets.24) The defects or structural disorder
may be attributed to the curled edges and folded regions of
the graphite nanosheets which contribute to the increasing
of the D peak signal.27) Our graphite nanosheets possessed
considerable curved and corrugated regions (the afore
mentioned curled edges and folds in the central regions), as
is shown by SEM (Figs. 1(b), (c)), TEM (Figs. 2(a), (c) and
(d)) and DFM (Figs. 3(a), (b)) examination. Nevertheless, the
observed intensity ratio of the D-to-G peaks (ID =IG 0:5) in
the Raman spectrum of synthesized graphite nanosheets is
smaller than that observed for carbon nanosheets produced
by using hot lament CVD,5) PECVD (ID =IG 0:99)6) and
solvothermal synthesis (ID =IG 1:5),19) indicating that our
graphite nanosheets have better crystallinity, an observation
supported by the sharp SAED pattern shown in Fig. 2(b).
4.

Summary

This paper oers a novel and ecient way of preparing


graphite nanosheets using low-current plasma discharge in
ultrasonically cavitated liquid ethanol. This method does not
need high electric current, nor does it require expensive
vacuum equipment and is capable of being scaled-up. The
by-products of our method are carbon nanocapsules and
amorphous carbon that is easily removed by magnetic
separation, hydrogen peroxide and aqua regia acid treatment.
The crystalline graphite structure of the nanosheets was
conrmed by TEM, XRD and Raman spectroscopy. The
crystallinity of the graphite nanosheets produced was
superior to that of carbon nanosheets prepared by other
methods. The graphite nanosheets range in lateral extent
from few hundred nanometers up to 11 mm. Dynamic force
microscopy (DFM) measurements showed that the most
frequently observed minimum step heights between the
surface of graphite nanosheets and silicon substrates ranged
from 6.7 to 23.5 nm. The minimum measured step height of
each sheet was considered as a measurement of the thickness.
The results of the DFM measurements were supported by the
HRTEM investigations of curled edges. The synthesized
graphite nanosheets can be used as a precursor for production
of single- and few-layer graphene sheets.
Acknowledgment
This work was nancially supported by a Grant-in-Aid for
Exploratory Research (No. 17656243) and Young Scientists

1459

(A) (No. 20686051) from the Ministry of Education, Culture,


Sports, Science and Technology, Japan.
REFERENCES
1) G. Chen, W. Weng, D. Wu, C. Wu, J. Lu, P. Wang and X. Chen:
Carbon 42 (2004) 753759.
2) M. V. Antisari, A. Montone, N. Jovic, E. Piscopiello, C. Alvani and L.
Pilloni: Scr. Mater. 55 (2006) 10471050.
3) G. Sun, X. Li, Y. Qu, X. Wang, H. Yan and Y. Zhang: Mater. Lett. 62
(2008) 703706.
4) H. Zhang, N. Kikuchi, T. Kogure and E. Kusano: Vacuum 82 (2008)
754759.
5) N. G. Shang, F. C. K. Au, X. M. Meng, C. S. Lee, I. Bello and S. T. Lee:
Chem. Phys. Lett. 358 (2002) 187191.
6) A. T. H. Chuang, B. O. Boskovic and J. Robertson: Diamond Relat.
Mater. 15 (2006) 11031106.
7) M. Zhu, J. Wang, R. A. Outlaw, K. Hou, D. M. Manos and B. C.
Holloway: Diamond Relat. Mater. 16 (2007) 196201.
8) X. Zhao, H. Tian, M. Zhu, K. Tian, J. J. Wang, F. Kang and R. A.
Outlaw: J. Power Sources 194 (2009) 12081212.
9) Y. Si and E. T. Samulski: Chem. Mater. 20 (2008) 67926797.
10) E. Yoo, J. Kim, E. Hosono, H.-S. Zhou, T. Kudo and I. Honma: Nano
Lett. 8 (2008) 22772282.
11) G. Wang, X. Shen, J. Yao and J. Park: Carbon 47 (2009) 20492053.
12) M. Y. Chen, C. M. Yeh, J. S. Syu, J. Hwang and C. S. Kou:
Nanotechnology 18 (2007) 185706 (4pp).
13) J. Li, J.-K. Kim and M. L. Sham: Scr. Mater. 53 (2005) 235240.
14) L. Chen, G. Chen and L. Lu: Adv. Funct. Mater. 17 (2007) 898904.
15) Z.-S. Wu, W. Ren, L. Gao, J. Zhao, Z. Chen, B. Liu, D. Tang, B. Yu, C.
Jiang and H.-M. Cheng: ASC Nano 3 (2009) 411417.
16) K. S. Subrahmanyam, L. S. Panchakarla, A. Govindaraj and C. N. R.
Rao: J. Phys. Chem. C 113 (2009) 42574259.
17) 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 and I. A. Aksay: Chem. Mater. 19 (2007) 43964404.
18) S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas, A.
Kleinhammes, Y. Jia, Y. Wu, SonBinh T. Nguyen and R. S. Ruo:
Carbon 45 (2007) 15581565.
19) Q. Kuang, S.-Y. Xie, Z.-Y. Jiang, X.-H. Zhang, Z.-X. Xie, R.-B. Huang
and L.-S. Zheng: Carbon 42 (2004) 17371741.
20) M. Choucair, P. Thordarson and J. A. Stride: Nature Nanotechnol. 4
(2009) 3033.
21) S. Kim, E. Shibata, R. Sergiienko and T. Nakamura: Carbon 46 (2008)
15231529.
22) J. C. Meyer, A. K. Geim, M. I. Katsnelson, K. S. Novoselov, T. J. Booth
and S. Roth: Nature 446 (2007) 6063.
23) R. J. Nemanich and S. A. Solin: Phys. Rev. B 20 (1979) 392401.
24) 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. 97 (2006) 187401.
25) A. Gupta, G. Chen, P. Joshi, S. Tadigadapa and P. C. Eklund: Nano
Lett. 6 (2006) 26672673.
26) Y. Y. Wang, Z. H. Ni, T. Yu, Z. X. Shen, H. M. Wang, Y. H. Wu, W.
Chen and A. T. S. Wee: J. Phys. Chem. C 112 (2008) 1063710640.
27) D. Graf, F. Molitor, K. Ensslin, C. Stampfer, A. Jungen, C. Hierold and
L. Wirtz: Nano Lett. 7 (2007) 238242.
28) F. Tuinstra and J. L. Koenig: J. Chem. Phys. 53 (1970) 11261130.

You might also like