You are on page 1of 6

Journal of Alloys and Compounds 615 (2014) 582587

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: www.elsevier.com/locate/jalcom

Effects of calcination on the preparation of carbon-coated SnO2/graphene


as anode material for lithium-ion batteries q
Guiliang Wu, Zhongtao Li, Wenting Wu, Mingbo Wu
State Key Laboratory of Heavy Oil Processing, China University of Petroleum, Qingdao 266580, China

a r t i c l e

i n f o

Article history:
Received 15 May 2014
Received in revised form 19 June 2014
Accepted 19 June 2014
Available online 5 July 2014
Keywords:
Carbon-coated SnO2/graphene composite
Calcination
Cycling stability
Conjugated system
Lithium-ion batteries

a b s t r a c t
Carbon-coated SnO2/graphene (SnO2-C/GN) composites with high electrochemical performance are
prepared via hydrothermal method followed by a facile calcination at different temperatures. The effects
of calcination temperature on the structure and morphology as well as electrochemical behavior of
synthesized SnO2-C/GNs are systematically investigated by X-ray diffraction, Fourier transform infrared
spectrometry, transmission electron microscopy, and related electrochemical performance tests. The
results show that the as-made SnO2-C/GNs calcined at 500 C has the smallest charge transfer resistance,
the largest lithium ion diffusion coefcient and the best cycling stability, exhibiting a stable capacity of
460 mAh g1 even after 120 cycles at a current density of 200 mA g1. The simple calcination process
with proper temperature can effectively promote the uniformity of dispersed SnO2 nanoparticles and
the formation of an excellent conjugated system with high electrical conductivity by graphene and
homogeneous carbon, which contribute to the improved electrochemical properties of as-made
SnO2-C/GNs.
2014 Elsevier B.V. All rights reserved.

1. Introduction
Lithium-ion batteries (LIBs), with high electromotive force, high
energy density, and long cycle life, have attracted considerable
research interest during the past decades [13]. Among numerous
anode materials of LIBs, tin oxide (SnO2) is considered to be one of
the most promising anode materials because of its nontoxicity, low
cost, good safety and high theoretical capacity of 782 mAh g1
based on an alloying mechanism [47]. Unfortunately, the practical
applications of SnO2 in LIBs are restricted owing to severe volume
variation (around 300%) during the alloying/de-alloying processes,
which results in critical mechanical damage to Sn-based electrodes, and leads to a drastic loss of capacity [810].
In order to overcome these obstacles, various methods have
been proposed [1116]. Hybridizing SnO2 with homogeneous carbon at nanoscale is an effective method to accommodate the strain
of volume change during cyclic process, and the carbon coating
layer also can enhance the conductivity of anodes [11,1719].
Graphene (GN), an atomic single layer of honeycomb carbon
lattice, has been considered as an ideal supporting material to
functionalize the carbon-coated SnO2/graphene (SnO2-C/GN)
q

Electronic supplementary information (ESI) available.

Corresponding author. Tel.: +86 532 8698 3452.


E-mail address: wumb@upc.edu.cn (M. Wu).
http://dx.doi.org/10.1016/j.jallcom.2014.06.143
0925-8388/ 2014 Elsevier B.V. All rights reserved.

[17,20,21] or SnO2/graphene [2224] due to its outstanding


electrical conductivity, excellent mechanical exibility, large
specic surface area, and high thermal and chemical stabilities.
Calcination also has a signicant effect on the structure, particle
size, and morphology of carbon materials as well as the formation
of conductivity networks [2527]. It is noted that calcination temperature plays important role on the structure and performance of
nal samples. On one hand, the reduction of graphene oxide (GO)
and construction of covalent frameworks are hard to occur at
low calcined temperature. On the other hand, severe agglomeration of SnO2 nanoparticles (NPs) distributed on carbon substrates
would happen at high calcined temperature. The both factors have
a great inuence on the electrochemical performances of SnO2-C/
GNs as LIBs anode. Therefore, the calcination temperature needs
to be systematically explored in order to obtain the optimal
electrode material with excellent electrochemical property. To
the best of our knowledge, the effect of calcination, especially the
role of calcined temperature on the structure and electrochemical
performance of SnO2-C/GNs as LIBs anode has not been sufciently
discussed.
In this work, polyvinyl alcohol (PVA) is adopted as carbon
precursor of the conductive matrix and graphene as coating carbon
to synthesize SnO2-C/GNs followed by calcination at various
temperatures. Herein, the effects of varied calcined temperature
on the structure and electrochemical property of SnO2-C/GNs are

G. Wu et al. / Journal of Alloys and Compounds 615 (2014) 582587

systematically explored, aiming at precise control of the structure


and conformation of SnO2-C/GNs, thereby obtaining SnO2-C/GNs
with high electrochemical performance via facile calcination
technique.
2. Experimental
2.1. Preparation of SnO2-C/GN composites
All chemicals were of analytical grade and used as received, without further
purication. SnO2-C/GN nanocomposites were synthesized by hydrothermal and
subsequent calcination processes. In a typical process, graphite oxide was rstly
synthesized from graphite powder by a modied Hummers method [28]. 100 mg
graphite oxide was added to 100 mL distilled water followed by sonication for
2 h. The obtained suspension was labelled as solution A. Solution B was prepared
by dissolving 3.0 g PVA in 100 mL of distilled water after 30 min of stirring.
12.5 g SnCl45H2O was dissolved to the solution of 60 mL distilled water and 1 mL
appropriate concentrated hydrochloric acid under stirring. The resultant solution
was marked as solution C. Solution A was slowly injected into the mixture of solutions B and C, and its pH value was adjusted among 910 by injecting amounts of
ammonium hydroxide (25.038.0 wt%). The above mixture was stirred sufciently
in a water bath for 1 h at 85 C and then dried at 100 C for 24 h. The samples were
collected for the further calcination at 400 C, 500 C, and 600 C under nitrogen
atmosphere, and the resulting samples were denoted as SnO2-C/GN-400, SnO2-C/
GN-500 and SnO2-C/GN-600, respectively. For a comparison, pure SnO2 (without
adding GO and PVA) was synthesized by the same method as mentioned above.

2.2. Samples characterization


The crystal structures of samples were characterized by X-ray powder diffraction (XRD) (XPert PRO MPD, Holland). Functional groups of samples were detected
by Fourier transform infrared spectrometry (FTIR) (Thermo Nicolet NEXUS 670,
USA). The morphology of the as-prepared samples was investigated by transmission
electron microscopy (TEM) (JEM-2100UHR, Japan). The thermal properties of the
samples were characterized by thermogravimetric analysis (TGA) (STA 409 PC Luxx,
Germany).

2.3. Electrochemical measurements


To measure the electrochemical performance, the electrodes were prepared by
coating the slurry of the active material powders, acetylene black, and poly(vinylidene) uoride (PVDF) binder (80:10:10 in weight ratio) dissolved in N-methyl-2pyrrolidinone (NMP) onto a Cu foil substrate. The electrodes were dried in a vacuum
at 100 C for 10 h, and then assembled into a half-battery (CR2032 coin type) with
lithium foil as a counter electrode in an Ar-lled glove box. The electrolyte used was
1 M LiPF6 in a 1:1 weight ratio ethylene carbonate (EC):dimethyl carbonate (DMC)
solvent. The cells were galvanostatically chargeddischarged in the potential range
0.0052.5 V vs. Li/Li+ at the current density of 200 mA g1 on a Land CT2001A cycler.
Cyclic voltammograms (CV) were performed using an Ametek PARSTAT4000
electrochemistry workstation at 0.25 mV s1 within the potential range of
0.0052.5 V. Electrochemical impedance spectroscopy (EIS) tests were also
measured on an Ametek PARSTAT4000 electrochemistry workstation in the
frequency range of 100 kHz to 10 mHz with AC voltage amplitude of 10 mV.

583

3. Results and discussion


Fig. 1 illustrates the synthesis procedure for SnO2-C/GNs. When
solutions B and C are completely mixed, the injection of Sn4+ can
intensely interact with hydroxide radicals on the PVA molecules,
which is regarded as an excellent surface-active and dispersing
agent. Then, with the assistance of PVA chains, the homogeneous
dispersion of Sn4+ can be obtained. The introduced graphene oxide,
with many oxygen-containing functional groups (e.g. ACOOH,
AOH), can induce the PVA chains regular arrangement on the plane
of graphene oxide via the interaction between functional groups.
SnO2-C/GNs with different structure and morphology can be
obtained after calcination at various temperatures.
The XRD patterns of the as-prepared samples at various temperatures and the standard XRD pattern of SnO2 are shown in Fig. 2
(JCPDS No. 41-1445). The sharp peak characteristic of GO at about
10.4 (Fig. S1) represents the weak van der Waals force between
layers of GO [29], which could not be identied in the SnO2-C/GN
samples (Fig. 2). Furthermore, the appearance of a broad peak
centered at around 26 in the SnO2-C/GN samples, attributed to
diffraction of the (0 0 2) plane of graphite, indicates that graphene
oxide is reduced by calcination [22,30]. As shown in Fig. 2, the
XRD patterns of all three SnO2-C/GN samples exhibit four major
diffraction peaks (1 1 0), (1 0 1), (2 1 1), and (3 0 1), which are attributed to the tetragonal rutile SnO2 phase (JCPDS No. 41-1445). The
broad peaks of three SnO2-C/GN samples shown in Fig. 2 indicate
that the as-formed SnO2 consists of very small nanocrystallines.
However, with the increase of calcination temperature, the peak
width becomes smaller from SnO2-C/GN-400 to SnO2-C/GN-600.
Based on the Scherrer equation, the average crystal size of
SnO2-C/GN-400, SnO2-C/GN-500 and SnO2-C/GN-600 is 12.1 nm,
22.6 nm and 44.9 nm, respectively. It should be noted that the
particle size of SnO2-C/GN-600 sample is much larger than the
other samples in accordance with the TEM images, which could
be ascribed to slight agglomeration of SnO2 nanoparticles at higher
temperature.
Fourier transform infrared spectrometry (FT-IR) is employed to
analyze graphene oxide and as-prepared SnO2-C/GN composites.
FT-IR spectrum of GO (Fig. S2(b)) shows the ACOOH peak at
1740 cm1, a weak peak at 1401 cm1 ascribed to OAH deformation, a peak at 1220 cm1 assigned to CAOH stretching. The peaks
at about 827 and 1052 cm1 could be attributed to OAC@O and
CAO stretching vibrations, respectively. A peak at 1620 cm1 corresponding to the remaining sp2 character is also shown [31].
Fig. 3 exhibits the FT-IR spectra of SnO2-C/GN-400, SnO2-C/GN500, and SnO2-C/GN-600. An obvious band at 540 cm1ascribed

Fig. 1. Synthesis schematic of SnO2-C/GNs: (a) mixing solutions; (b) the calcination of mixture in nitrogen gas.

584

G. Wu et al. / Journal of Alloys and Compounds 615 (2014) 582587

Fig. 2. XRD patterns of (a) SnO2-C/GN-600, (b) SnO2-C/GN-500 and (c) SnO2-C/GN-400, and (d) the standard XRD pattern of SnO2.

small peaks from 900 cm1 to 600 cm1 on the SnO2-C/GN-600


curve become more distinct than other two samples, which could
be signed as nger print region of CAH on aromatic cycle
[32,33]. Therefore, we deduce that the samples calcined at higher
temperature could form a conductive framework and the additive
GO has been further reduced to graphene.
To investigate the thermal property and the composition of
SnO2-C/GNs as well as the carbon content in the products, thermogravimetric analysis (TGA) was carried out in air. As shown in
Fig. 4, two obvious weight losses over two different temperature
ranges can be found in all samples. For SnO2-C/GN-500, a major
weight loss step occurs in the range of 35250 C, corresponding
to the removal of H2O. Another weight decrease is revealed in
the range of 250600 C, which should be ascribed to the combustion reaction of amorphous carbon and graphene. At last, only SnO2
NPs could be reserved. The same trend also can be observed in the
curves of SnO2-C/GN-400 and SnO2-C/GN-600. According to the
TGA curves of SnO2-C/GNs, the weight losses of SnO2-C/GN-400,
SnO2-C/GN-500, and SnO2-C/GN-600 are about 57%, 53%, and
52%, respectively. The highest weight loss of SnO2-C/GN-400 is
attributed to its more functional groups existing at low heat
calcination, which has also been proved by FT-IR.

Fig. 3. FT-IR spectra of (a) SnO2-C/GN-400, (b) SnO2-C/GN-500, and (c) SnO2-C/GN600.

to SnAO can be found on the curves of all three SnO2-C/GN


samples, and the calcination removes most of oxygen-containing
groups on GO such as COOH peaks (1735 cm1), CAO stretching
vibrations (1052 cm1), and OAC@O peak (827 cm1), which indicates that GO has been reduced to graphene in different degrees.
However, the OAH deformation peak (1401 cm1) of SnO2-C/GN400 sample still maintains at a higher intensity due to the
relatively low heat calcination. In addition, a new peak at
630 cm1 is found in SnO2-C/GN-500 and SnO2-C/GN-600, which
should originate in the aromatic cycles. As revealed in Fig. 3, the

Fig. 4. TGA curves of SnO2-C/GN-400, SnO2-C/GN-500 and SnO2-C/GN-600 in air.

G. Wu et al. / Journal of Alloys and Compounds 615 (2014) 582587

The morphology of all SnO2-C/GN samples is studied by eld


emission transmission electronic microscopy (TEM), as shown in
Fig. 5. A large number of SnO2 nanoparticles (darker dots) dispersed on the surface of amorphous carbon are clearly observed
in Fig. 5(a), (b), and (d), which suggest that the morphology of
SnO2-C/GNs samples can keep well after calcination. As calcination
temperature rises, more wrinkles are emerged on the surface of
amorphous carbon. These phenomena may be well explained that
calcination at elevated temperature can increase the conjugated
degree of carbonaceous polymer, which can improve the electrical
conductivity of materials. Furthermore, some large particles with a
diameter of about 50 nm can be clearly observed in Fig. 5(d) due to
the agglomeration SnO2 NPs induced by the calcination at higher
temperatures. However, the SnO2 NPs deposited on the SnO2-C/
GN-400 and SnO2-C/GN-500 samples are much smaller than those
on the sample of SnO2-C/GN-600, which is about 10 nm and 20 nm
in accordance with the results of XRD patterns.
The detailed nanostructure of SnO2-C/GN-500 is further investigated by high-resolution transmission electron microscopy (HRTEM). As seen in Fig. 5(b) and (c), it can be found that SnO2 NPs
are homogeneously loaded on the carbon surface. The selected area
electron diffraction (SAED) pattern (inset of Fig. 5(b)) exhibits four
sharp diffraction rings corresponding to the (1 1 0), (1 0 1), (2 1 1),
and (3 0 1) crystalline planes of SnO2, conrming the formation of
polycrystalline SnO2 nanocrystal. In addition, the lattice fringes of
SnO2 NPs are clearly visible in Fig. 5(c) and its inset. The interlayer
spacings of about 0.332 nm and 0.264 nm respectively correspond
to (1 1 0) and (1 0 1) planes of SnO2. Thin layers of carbon in form of
stacked graphene sheets can be found on the edges of SnO2
nanoparticles (cf. yellow arrows).
The electrochemical performances of as-made SnO2-C/GN
samples and pure SnO2 are evaluated by galvanostatic charge
discharge cycling between 0.005 V and 2.5 V at a current density
of 200 mA g-1. Fig. 6(a) shows the charge/discharge proles of
SnO2-C/GN-500. A classical plateau around 0.8 V corresponds to
the formation of the solid electrolyte interphase (SEI) layer and
the decomposition of SnO2 to Sn [34], as shown in Eq. (1):

SnO2 4Li 4e ! 2Li2 O Sn

The peak below 0.5 V corresponds to alloying of Sn with Li


(forward reaction), as given in Eq. (2):

Sn xLi xe

! Lix Sn 0  x  4:4

585

Specically, SnO2-C/GN-500 electrode delivers a discharge


capacity of 1108 mAh g1 and a charge capacity of 785 mAh g1
in the rst cycle. The coulombic efciency of the rst cycle is
higher (about 71%) than 49% of pure SnO2 in Fig. S3, which could
be ascribed to the introduction of graphene [35,36]. The discharge/charge capacities of SnO2-C/GN-500 reach almost the same
values after the rst cycle, indicating that SnO2-C/GN-500 has a
high capacity retention. A comparison of the charge/discharge cyclic performance for SnO2-C/GN-400, SnO2-C/GN-500, and SnO2-C/
GN-600 is shown in Fig. 6(b). The rst discharge capacities of above
mentioned samples are almost identical, which are 1049, 1108,
and 1142 mAh g1, respectively. However, all three samples have
large irreversible capacities in the rst cycle. The initial capacity
loss is about 35.9%, 29.1%, and 56.3% for SnO2-C/GN-400, SnO2-C/
GN-500, and SnO2-C/GN-600, respectively. The reduction of SnO2
and the formation of SEI lm contribute to major capacities loss
for the rst time. The discharge capacities of SnO2-C/GN-400,
SnO2-C/GN-500, and SnO2-C/GN-600 after 50 cycles are about
270, 462 and 240 mAh g1, respectively. The corresponding capacity retention (compared with the second discharge capacity) are
around 50.7%, 90.8% and 68.9%, respectively, which is much higher
than pure SnO2 (only 33.8%, 210 mAh g1 after 50 cycles) in Fig. S3.
Compared with pure SnO2, the enhanced electrochemical performances in SnO2-C/GNs can be ascribed to the carbon coating layers
in SnO2-C/GNs, which can improve the conductivity, meanwhile,
mitigate the volume expansion of SnO2. It should be noted that
the capacity of SnO2-C/GN-500 is almost no decrease, even after
120 cycles, which still delivers a high capacity 460 mAh g1. As
revealed in Fig. 6(b), the relevant coulombic efciencies of SnO2C/GN-500 stay around 100% after several cycles, indicating the
good electrochemical reversibility of SnO2-C/GN-500. The excellent
electrochemical performance of SnO2-C/GN-500 can be attributed
to the SnO2 NPs uniformly distributed on graphene and amorphous
carbon layer, which has been proved by XRD and TEM results. In
addition, appropriate calcination temperature could not only help
to build a better conductive network, but also increase the reduction degree of graphene oxide, both of which can signicantly
improve the conductivity. In contrast, SnO2-C/GN-400 shows a
drastic decrease of discharge capacity after 10th cycle at the rate

Fig. 5. TEM images of (a) SnO2-C/GN-400, (b) SnO2-C/GN-500, (d) SnO2-C/GN-600 and HRTEM images of (c) SnO2-C/GN-500. The insets of (b) and (c) are the corresponding
SAED pattern and magnied image of the region by a circle.

586

G. Wu et al. / Journal of Alloys and Compounds 615 (2014) 582587

Fig. 6. (a) Charge/discharge proles of SnO2-C/GN-500, (b) the cyclic performance


of SnO2-C/GN-400, SnO2-C/GN-500, and SnO2-C/GN-600 at the current density of
200 mA g1.

of 200 mA g1. It is evident that the calcination at 400 C is insufcient to completely reduce the GO and to carbonize the carbonaceous polymer lm, which also has been demonstrated by FT-IR.
The drastic capacity loss and poor cycle performance of SnO2-C/
GN-600 could be due to the pulverization of agglomerated SnO2
NPs after the rst discharging process, which could lead to the
separation of nanoparticles from the carbon support and a poor
contact between them [5,37].
Fig. 7(a) shows the cyclic voltammograms of three samples at
the 5th cycle with the scan rate of 0.25 mV s1. The cathodic peak
around 0.20 V and the anodic peak around 0.56 V are clearly
observed from all CV curves, corresponding to the lithium alloying
reaction with Sn and dealloying of LixSn, respectively, as depicted
in Eq. (2). In particular, SnO2-C/GN-500 displays a pair of welldened and symmetrical oxidationreduction peaks compared
with the other two samples, suggesting SnO2-C/GN-500 has the
fastest chargedischarge process characteristic and the most
excellent reversibility [38]. The oxidation peak of SnO2-C/GN-400
slightly moves to higher voltage compared to SnO2-C/GN-500,
which is attributed to that the low conductivity of SnO2-C/GN400 is bad for charge transfer through coating carbon. The higher
treating temperature (600 C) could help the formation of conjugated system which leads to better conductivity. However, the
oxidation peak of SnO2-C/GN-600 also presents at higher potential
compared to SnO2-C/GN-500, which could be ascribed to the
greater charge transfer resistance induced by agglomeration of
SnO2 nanoparticles. Besides, the high peak intensity and integral
area on the curve SnO2-C/GN-500 reveal a good electrochemical
activity, while the other two samples calcined at 400 C and
600 C become lower, implying the inferior electrochemical performance of latter samples. These results accord with those of the
discharge/charge measurements shown in Fig. 6.

Fig. 7. (a) The cyclic voltammograms of the 5th cycle and (b) electrochemical
impedance spectra of SnO2-C/GN-400, SnO2-C/GN-500, and SnO2-C/GN-600,
respectively; the inset is the corresponding magnied curve in the high frequency
region.

Electrochemical impedance spectroscopy (EIS) tests are performed to explore the effect of the carbon coating on the electronic
conductivity. As shown in Fig. 7(b), Nyquist plots of SnO2-C/GN400, SnO2-C/GN-500, and SnO2-C/GN-600 consist of one depressed
semicircle at high frequencies and a straight line at low frequencies, corresponding to ohmic resistance (total resistance of the
electrolyte, separator, and electrical contacts) and the charge transfer resistance, respectively. Compared to other two samples, the
semicircle diameter of SnO2-C/GN-500 is much smaller and its line
slope is much bigger. The smaller resistance of SnO2-C/GN-500 can
be attributed to the high reduction of GO, uniform distribution of
SnO2 NPs, and the highly conjugated carbonaceous polymer. Furthermore, the facile charge transfer contributes to the electrochemical performance improvement in SnO2-C/GN-500.
4. Conclusion
In this work, carbon-coated SnO2/graphene composites with
high electrochemical performance via hydrothermal method followed by facile calcination at various temperatures are successfully synthesized. The effects of calcined temperature on SnO2-C/
GNs are systematically explored. The results reveal that SnO2-C/
GN-500 calcined at 500 C exhibits the best electrochemical performance, which can keep a stable capacity of 460 mAh g1 even after
120 cycles at a current density of 200 mA g1. The remarkable electrochemical performance of SnO2-C/GN-500 could be attributed to

G. Wu et al. / Journal of Alloys and Compounds 615 (2014) 582587

the homogeneous distribution of SnO2 NPs on the carbon matrix,


the formation of conductivity framework, and the high reduction
of GO under the appropriate temperature.
Acknowledgements
This work is supported by the National Natural Science Foundation of China (Nos. 51303212, 51172285, 51372277); National Natural Science Foundation of Shandong Province (ZR2013EMQ013);
The Fundamental Research Fund for the Central Universities
(14CX02060A).
Appendix A. Supplementary material
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.jallcom.2014.
06.143.
References
[1] C.H. Jiang, E. Hosono, H.S. Zhou, Nano Today 1 (2006) 2833.
[2] J.M. Tarascon, M. Armand, Nature 414 (2001) 359367.
[3] X. Li, X. Meng, J. Liu, D. Geng, Y. Zhang, M.N. Banis, Y. Li, J. Yang, R. Li, X. Sun, M.
Cai, M.W. Verbrugge, Adv. Funct. Mater. 22 (2012) 16471654.
[4] Y. Chen, B.H. Song, R.M. Chen, L. Lu, J.M. Xue, J. Mater. Chem. A 2 (2014)
56885695.
[5] Y. Idota, T. Kubota, A. Matsufuji, Y. Maekawa, T. Miyasaka, Science 276 (1997)
13951397.
[6] M.V. Reddy, G.V. Subba Rao, B.V.R. Chowdari, Chem. Rev. 113 (2013)
53645457.
[7] J.F. Liang, W. Wei, D. Zhong, Q.L. Yang, L.D. Li, L. Guo, ACS Appl. Mater.
Interfaces 4 (2012) 454459.
[8] J. Fan, T. Wang, C.Z. Yu, B. Tu, Z.Y. Jiang, D.Y. Zhao, Adv. Mater. 16 (2004)
14321436.
[9] Z.Y. Wang, L. Zhou, X.W. Lou, Adv. Mater. 24 (2012) 19031911.
[10] D. Larcher, S. Beattie, M. Morcrette, K. Edstroem, J.C. Jumas, J.M. Tarascon,
J. Mater. Chem. 17 (2007) 37593772.
[11] J.S. Chen, X.W. Lou, Small 9 (2013) 18771893.
[12] L.L. Xing, Y.Y. Zhao, J. Zhao, Y.X. Nie, P. Deng, Q. Wang, X.Y. Xue, J. Alloys Comp.
586 (2014) 2833.

587

[13] M.S. Park, G.X. Wang, Y.M. Kang, D. Wexler, S.X. Dou, H.K. Liu, Angew. Chem.
Int. Ed. 46 (2007) 750753.
[14] Y.H. Hwang, E.G. Bae, K.S. Sohn, S. Shim, X.K. Song, M.S. Lah, M. Pyo, J. Power
Sources 240 (2013) 683690.
[15] B.B. Zhang, C.Y. Wang, Q. Ru, S.J. Hu, D.W. Sun, X. Song, J. Li, J. Alloys Comp. 581
(2013) 15.
[16] L. Zhang, G.Q. Zhang, H.B. Wu, L. Yu, X.W. Lou, Adv. Mater. 25 (2013) 2589
2593.
[17] C.F. Zhang, X. Peng, Z.P. Guo, C.B. Cai, Z.X. Chen, D. Wexler, S. Li, H.K. Liu,
Carbon 50 (2012) 18971903.
[18] H.X. Zhang, C. Feng, Y.C. Zhai, K.L. Jiang, Q.Q. Li, S.S. Fan, Adv. Mater. 21 (2009)
22992304.
[19] B. Lee, S.C. Han, M. Oh, M.S. Lah, K.S. Sohn, M. Pyo, Electrochim. Acta 113
(2013) 149155.
[20] J.L. Cheng, H.L. Xin, H.M. Zheng, B. Wang, J. Power Sources 232 (2013) 152
158.
[21] B.J. Li, H.Q. Cao, J.X. Zhang, M.Z. Qu, F. Lian, X.H. Kong, J. Mater. Chem. 22
(2012) 28512854.
[22] J. Yao, X.P. Shen, B. Wang, H.K. Liu, G.X. Wang, Electrochem. Commun. 11
(2009) 18491852.
[23] S.M. Paek, E. Yoo, I. Honma, Nano Lett. 9 (2009) 7275.
[24] S.J. Ding, D.Y. Luan, F.Y.C. Boey, J.S. Chen, X.W. Lou, Chem. Commun. 47 (2011)
71557157.
[25] Y.S. Shang, J.L. Liu, T. Huang, A.S. Yu, Electrochim. Acta 113 (2013) 248255.
[26] Z.Q. Huo, Y.T. Cui, D. Wang, Y. Dong, L. Chen, J. Power Sources 245 (2014)
331336.
[27] X.Z. Liu, H.Q. Li, E. Yoo, M. Ishida, H.S. Zhou, Electrochim. Acta 83 (2012)
253258.
[28] W.S. Hummers, R.E. Offeman, J. Am. Chem. Soc. 80 (1958). 13391339.
[29] C. Hontoria-Lucas, A.J. Lopez-Peinado, J.D.D. Lopez-Gonzalez, M.L. RojasCervantes, R.M. Martin-Aranda, Carbon 33 (1995) 15851592.
[30] S.K. Park, S.H. Yu, N. Pinna, S. Woo, B. Jang, Y.H. Chung, Y.H. Cho, Y.E. Sung, Y.Z.
Piao, J. Mater. Chem. 22 (2012) 25202525.
[31] Y.M. Li, X.J. Lv, J. Lu, J.H. Li, J. Phys. Chem. C 114 (2010) 2177021774.
[32] R. Bissessur, P.K.Y. Liu, W. White, S.F. Scully, Langmuir 22 (2006) 17291734.
[33] J. Peckett, P. Trens, R. Gougeon, A. Poppl, R. Harris, Carbon 38 (2000) 345353.
[34] G.D. Du, C. Zhong, P. Zhang, Z.P. Guo, Z.X. Chen, H.K. Liu, Electrochim. Acta 55
(2010) 25822586.
[35] Z.T. Li, G.L. Wu, D. Liu, W.T. Wu, B. Jiang, J.T. Zheng, Y.P. Li, J.H. Li, M.B. Wu, J.
Mater. Chem. A 2 (2014) 74717477.
[36] Q. Guo, Z. Zheng, H.L. Gao, J. Ma, X. Qin, J. Power Sources 240 (2013) 149154.
[37] W.J. Zhang, J. Power Sources 196 (2011) 1324.
[38] C.C. Peng, H.L. Bai, M.W. Xiang, C.W. Su, G.Y. Liu, J.M. Guo, Int. J. Electrochem.
Sci. 9 (2014) 17911798.

You might also like