You are on page 1of 7

Carbon 99 (2016) 658e664

Contents lists available at ScienceDirect

Carbon
journal homepage: www.elsevier.com/locate/carbon

Crumpled graphene paper for high power sodium battery anode


Young Soo Yun a, Young-Uk Park a, Sung-Jin Chang b, Byung Hoon Kim c, Jaewon Choi d,
Junjie Wang d, Ding Zhang a, Paul V. Braun d, Hyoung-Joon Jin e, Kisuk Kang a, f, *
a

Department of Materials Science and Engineering, Seoul National University, Seoul 151-742, South Korea
Department of Chemistry, Chung-Ang University, Seoul 156-756, South Korea
c
Department of Physics, Incheon National University, Incheon 406-772, South Korea
d
Department of Materials Science and Engineering, University of Illinois at Urbana-Champaign, Urbana, IL 61801, United States
e
Department of Polymer Science and Engineering, Inha University, Incheon 402-751, South Korea
f
Center for Nanoparticles Research, Institute for Basic Science (IBS), Seoul National University, Seoul 151-742, South Korea
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 7 July 2015
Received in revised form
15 December 2015
Accepted 16 December 2015
Available online 29 December 2015

Graphene-based electrodes typically form a compact uniaxially oriented stacked structure during electrode preparation due to the highly anisotropic morphology. This leads to limited diffusion paths for the
insertion of Li or Na when used as electrodes in rechargeable batteries. Here, we demonstrate that selfstanding electrodes formed of randomly folded and/or crumpled graphene nanosheets can be obtained
via a simple modied reduction process, and that the crumpled structure can signicantly increase the
power capability of graphene-based anodes of sodium-ion batteries. These electrodes can deliver a
power density of approximately 20,000 W kg1, which surpasses the Li storage capability of conventional
graphene paper electrodes. Moreover, the specic capacity was stably maintained without a binder,
conductive agent, or substrate for more than 500 charge/discharge cycles and 1000 cycles of repeated
bending.
2015 Elsevier Ltd. All rights reserved.

1. Introduction
Sodium-ion batteries (NIBs) are gaining more attention as alternatives to lithium-ion batteries (LIBs), due to the abundance of
sodium salts, as well as the similar chemistry to LIBs [1]. Recent
studies have identied promising electrode materials that exhibit
comparable energy and power densities to those used in LiBs
[2e16]. Nevertheless, intrinsic differences between lithium and
sodium in terms of ionic radius, electropotential, and solvation
properties often result in inferior electrochemical properties of
NIBs [9,11,15,16]. For example, sodium ion is ca. 55% larger, 330%
heavier, and with an electropotential that is 0.33 eV higher (vs. the
elemental metal); for these reasons NIBs generally have a lower
energy density and exhibit slower kinetic characteristics for intercalation [9,11,15,16]. One particular disadvantage of NIBs is that in
contrast to lithium ions, sodium ions cannot intercalate into the
carbon-based graphitic layers [17] that serve as cheap and robust
electrodes for LIBs. Even intercalation into well-stacked graphene

* Corresponding author. Department of Materials Science and Engineering, Seoul


National University, Seoul 151-742, South Korea.
E-mail address: matlgen1@snu.ac.kr (K. Kang).
http://dx.doi.org/10.1016/j.carbon.2015.12.047
0008-6223/ 2015 Elsevier Ltd. All rights reserved.

layers is typically not feasible.


Recently, graphene-based paper has successfully been applied
as an anode material for LIBs [18e24]. These freestanding paper
electrodes are advantageous compared with conventional electrodes in terms of energy density, because they do not require a
binder, conductive agent, or substrate. Additionally, these highly
exible paper electrodes are applicable to exible electronics and
wearable devices. Abouimrane et al. reported that reduced graphene oxide paper, when used as an anode for LIBs, exhibited
reversible capacities of 214, 151, and 84 mAh g1 at current densities of 10, 20, and 50 mA g1, respectively [20]. The folded or
porous structure of the stacked graphene components enables
enhanced electrochemical performance [22,23]. More recently,
graphene has been studied as an electrode material for NIBs
[25,26], and has been shown to deliver a capacity of as much as
95.6 mAh g1 at a current rate of 1 A g1 when the electrode was
fabricated using conventional graphene powder [25]. This result
suggest the pseudocapacitive sodium-ion storage on graphene has
a possibility to achieve high rate performances. Nanohybrids
[7e12] or nanostructured graphene powder can further improve
the electrochemical performance [13e15]. Nevertheless, a few
report about a exible and self-standing graphene electrode for

Y.S. Yun et al. / Carbon 99 (2016) 658e664

659

boosting energy has been reported for NIBs [27]. This is primarily
because conventional fabricated graphene paper contains compact
graphene sheets with a strong preferential stacking orientation, in
which sodium ions cannot be efciently transported (in contrast to
lithium ions). Also, the conventional fabrication process with vacuum ltering needs much time, because the stacked graphene
sheet blocks a pathway of solvent, limiting the thickness of the
graphene paper electrodes as several micrometres. Furthermore, it
is not clear how the electrochemical properties of sodium ions
would depend on the stacking nature of the graphene sheets.
In this work, we prepared randomly oriented graphene paper
(ROGP) with high exibility, which contained permanently folded
and/or crumpled graphene, by thionyl chloride treatment of graphene oxide (G-O) and compared the electrochemical activity with
that of conventional uniaxially oriented paper. Simple vacuum
ltering of the crumpled graphene induced a thick and robust paper electrode in a brief space of time, with no binder, conducting
carbon and metal substrate as well as additional process. Also, we
found that the morphological difference between ROGP and
conventionally stacked graphene paper leads to a signicant
change in the sodium storage performance, which demonstrates
the importance of the interplay between the structure of the graphene sheets and the kinetic characteristics of the electrode with
NIBs.

mode. Raman measurements were performed using NTEGRA


Spectra spectrometer (NT-MDT, Russia) equipped with a 473 nm
(2.62 eV) laser in backscattering conguration. The spectral resolution was ~2 cm1 with a 600-grooves/mm grating. A 100
objective (N.A. 0.9) provided a laser spot size about 330 nm. The
laser power was kept well below 0.3 mW for nondestructive Raman
measurements. X-ray photoelectron spectroscopic analysis (XPS,
PHI 5700 ESCA) was performed using monochromated Al Ka radiation (hn 1486.6 eV). Fourier Transform-Infrared spectroscopy
(FT-IR) was performed using VERTEX 80v (Bruker Optics, Germany). Temperature-dependent IeV characteristics were determined by the conventional four-probe method in the Janis
cryogenic system with a semiconductor characterization system
(4200-SCS, Keithley). The electrical measurements were performed
after vacuum degassing for 12 h (<2  106 Torr). The tensile
properties were tested using an Instron 4665 ultimate tensile
testing machine (UTM) at 25  C and a humidity of 30%. The tensile
tests were performed using ten specimens. The BrunauereEmmetteTeller (BET) specic surface area and differential
pore volumes were determined from nitrogen adsorption/desorption isotherms (ASAP 2020, Micromeritics, Norcross, GA, USA)
at 196  C. The BET surface areas were calculated according to BET
theory, and the micropore surface areas were obtained using t-plot
theory.

2. Experimental

2.4. Electrochemical characterization

2.1. Preparation of ROGPs

The electrochemical performance of ROGPs-600 was evaluated


using a Wonatec automatic battery cycler and CR2032-type coin
cells. The coin cells were assembled in a glove box lled with argon,
employing a composite electrode with metallic sodium and 1 M
NaClO4 (Aldrich 99.99%) dissolved in a solution of propylene carbonate as the electrolyte. The cells were galvanostatically cycled
between 0.005 and 2.5 V ns. Na/Na at various current densities.
The mass loading of ROGPs-600 electrode was ~1.1 mg cm2.

G-O was prepared according to the method described in


Refs. [28,29], which was prepared from natural graphite (SigmaeAldrich) using the Hummers method. Aqueous G-O suspensions
were frozen in liquid nitrogen and then freeze-dried using a
lyophilizer (LP3, Jouan, France) at 50  C and 0.045 mbar for 72 h.
After lyophilization, low-density and loosely packed G-O powders
were obtained. 100 mg of the as-obtained lyophilized G-O powder
was dispersed in 50 ml of thionyl chloride (99%, Sigma-aldrich) by
ultrasound treatment. The G-O dispersion was then treated at 60  C
for 12 h. The resultant product was diluted in dimethylformamide
and then vacuum-ltered on an Anodisc membrane lter (47 mm
diameter, 0.2 mm pore size; Whatman) to obtain ROGPs. The ROGPs
were heated in a tubular furnace from room temperature to 600  C,
using a heating rate of 10  C/min and an Ar ow rate
of 200 ml min1. The ROGPs were held at this temperature
(600  C) for 2 h. The products (ROGPs-600) were stored in a vacuum
oven at 30  C.
2.2. Preparation of UOGPs
The G-O powder was dispersed in distilled water, and hydrazine
monohydrate (hydrazine: G-O 1:2) was added to reduce the G-O.
After stirring at 90  C for 1 h, the reduced G-O nanosheets (H2N2GNSs) were vacuum-ltered using an Anodisc membrane lter
(47 mm in diameter, with 0.2-mm pores; Whatman), yielding uniaxially oriented stacked graphene paper (UOGP) as a control
sample.
2.3. Characterization
The morphology of the prepared samples was examined by
using eld emission scanning electron microscopy (FE-SEM, S4300, Hitachi, Japan) and eld emission transmission electron microscopy (FE-TEM, JEM2100F, JEOL, Japan). The topographical image
was obtained using atomic force microscope (NT-MDT, Russia) with
a NSG-10 cantilever (NT-MDT, Russia) in semicontact operation

3. Results and discussion


Fig. 1(a) shows the cross-sectional morphology of the UOGP
control sample, which exhibited a compact layer-by-layer stacking,
which is consistent with previous reports [18,19]. A preliminary test
of the sodium storage capability revealed that only a small
reversible capacity of 10 mAh g1 was obtained at a current density
of 100 mA g1, as shown in Fig. 1(b). This behaviour contrasts with
that of the lithium storage capability, which exhibited a reversible
lithium capacity of 170 mAh g1 at the same current density (see
the inset of Fig. 1(b)) [18]. This indicates that the insertion and
extraction of sodium ions in the UOGPs is much more difcult than
that for lithium ions with the same morphological graphene
structure. As part of our strategy to provide a more suitable structure for the insertion and extraction of sodium ions, we attempted
to randomise the orientation of the graphene building blocks. In a
modied process, thionyl chloride (99%, SigmaeAldrich) was
used as a reducing agent instead of hydrazine monohydrate. Thionyl chloride interacts in a nucleophilic fashion with a carboxylic
acid, and acyl chloride is formed at the edges and basal planes of
the G-O [30e32]. The acyl chloride moiety on the G-O is also
reactive to nucleophilic attack by the carboxylic acid and hydroxyl
groups. As a result of the intersheet and intrasheet reactions of G-O,
the G-O sheet becomes permanently folded or crumpled. The crosssectional morphology of the ROGP shown in Fig. 1(c) and (d) reveals
a randomly oriented structure of the graphene-based lms. Fig. 2
also shows the morphology of the building block, the thionyl
chloride derived-GNS (T-GNS), which had a corrugated surface with
a height of ~10 nm and a lateral size of several micrometres.

660

Y.S. Yun et al. / Carbon 99 (2016) 658e664

Fig. 1. (a) SEM image of the cross-sectional morphology of the UOGP. (b) Galvanostatic charge/discharge proles of UOGPs in the potential range 0.01e2.5 V vs. Na/Na at a current
density of 100 mA g1. The inset shows galvanostatic charge/discharge proles vs. Li/Li. (c) and (d) SEM images of the cross-sectional morphology of ROGP in different magnications. The inset of (c) shows an optical image of the ROGP. (A colour version of this gure can be viewed online).

Fig. 2. AFM data (left) and TEM image (right) of T-GNS. (A colour version of this gure can be viewed online).

Fig. 3 shows X-ray photoelectron spectroscopy (XPS) O 1s


spectra of G-O, H2N2-GNS and T-GNS. Two distinct peaks appeared
in the O 1s spectra, which reveal the presence of oxygen atoms in
the carbonyl groups, as well as various other oxygen groups. The
oxygen content of G-O was lowered following treatment with hydrazine or thionyl chloride. Additionally, the T-GNS was partially
doped with chlorine. In the XPS Cl 2p spectra of T-GNS, two peaks
were found with binding energies of 199.0 eV and 200.9 eV, indicating the presence of CeCl and O]CeCl bonds, respectively.
Fourier transform infrared (FT-IR) spectra of the T-GNS support the
chemical bonding between graphene and Cl (see Fig. S1).
Fig. 4(a) shows XPS C 1s spectra of ROGP following thermal
treatment at 600  C (denoted ROGP-600). The main sp2 C]C peak
was observed at 284.2 eV and an sp3 CeC peak was observed at
285.1 eV [33]. A remaining C(O)O oxygen peak was found at
290.1 eV. This result is distinct from the XPS C 1s spectra of G-O,
which exhibited a large sp3 CeC peak at 285.1 eV and CeO peak at

287.5 eV (see Fig. S2). The chemical conguration of the oxygen


groups in ROGP-600 was similar to that of T-GNS; however, the
content of carbonyl groups was higher in ROGP-600, as shown in
Fig. 4(b). The C/O ratio was 11.5, which is higher than that observed
in T-GNS (i.e., 6.9) or H2N2-GNS (7.9). Raman spectroscopy was used
to investigate the carbon structure of ROGP-600, as shown in
Fig. 4(c), and D and G peaks were observed at ~1367 and
~1599 cm1, respectively, indicating relatively symmetrical
breaking of the innite carbon honeycomb lattice [33,34]. The ratio
of the intensity of the D peak to that of the G peak was ID/IG ~ 1.07,
which is lower than that observed for UOGP (~1.23) or ROGP (~1.11),
indicating a lower defect density in the ROGP-600. A similar trend
was observed in the Raman spectra of H2N2-GNS, T-GNS and TGNS-600 nanosheets (see Fig. S3). Fig. 4(d) shows the temperaturedependent c-axis (out-of-plane) and a-axis (in-plane) resistivities,
i.e., rc and ra, respectively, of the ROGP-600. Those of UOGP were
also obtained for comparison. Both samples exhibited nonmetallic

Y.S. Yun et al. / Carbon 99 (2016) 658e664

661

Fig. 3. XPS O 1s spectra of (a) G-O, (b) H2N2-GNS, and (c) T-GNS. (d) XPS Cl 2p spectra of T-GNS. (A colour version of this gure can be viewed online).

Fig. 4. (a) C 1s XPS spectra and (b) O 1s XPS spectra of ROGP-600. (c) Raman spectra of ROGP-600. (d) Temperature-dependent resistivity curves of ROGP-600 and UOGP. (A colour
version of this gure can be viewed online).

behaviour over the entire temperature range, i.e., dr(T)/dT was


negative. As expected from the measurement of the oxygen content
and carbon structure, the resistivity of UOGP was higher than that
of ROGP-600. Interestingly, the difference between rc and ra, i.e.,
Dr (ra  rc)/ra, was signicantly smaller for ROGP-600
(Dr 0.769) than for UOGP (Dr 2.928). This indicates that efcient conduction channels were present in the c-axis direction in
ROGP-600, in contrast to UOGP. This is due to the randomly oriented structure of ROGP-600, as shown in Fig. 4(d).
Pore structures of both ROGP-600 and UOGP were characterized
by nitrogen adsorption and desorption tests, as shown in Fig. 5.
ROGP-600 showed a specic surface area of 47.4 m2 g1 of which tplot micropore was 7.6 m2 g1 and t-plot external surface area was
39.8 m2 g1. The nitrogen adsorption and desorption isotherm

curves of ROGP-600 exhibited IUPAC type-IV mesoporous structure.


Pore size distribution data (Inset of Fig. 5) reveals that ROGPs-600
have wide range of pores, including numerous micropores
(<2 nm) and mesopores (2e50 nm). In contrast, UOGP had a specic surface area of several metres per unit gram (~5 m2 g1),
indicating UOGP has a poor micropore and/or mesopore structure.
The well-stacked graphene layers of UOGP are similar to a microstructure of graphite which has no micropore for nitrogen
adsorption/desorption, leading to the low specic surface area. In
contrast, the randomized graphene layers contain an open void for
all-around internal sides of ROGP-600. This difference in microstructure between ROGP-600 and UOGP could induce dramatic
effects for their electrochemical performances. Additionally, the
mechanical property of ROGP-600 was characterized by UTM, as

662

Y.S. Yun et al. / Carbon 99 (2016) 658e664

Fig. 5. Nitrogen adsorption and desorption isotherm curves of ROGP-600 and inset of
pore size distribution data of ROGP-600. (A colour version of this gure can be viewed
online).

shown in Fig. S4. The ROGP-600 showed a Young's modulus of 720


Mpa, tensile strength of 24.2 3.7 Mpa and strain of 2.5%, indicating its potential applicability to a exible electrode. Although the
tensile strength of ROGP-600 is lower than graphene oxide paper of
~81.9 MPa [35] and graphene paper of ~150 MPa [36], the value is
higher than tensile strengths of the previously reported exible
paper electrodes such as graphene/polyaniline composite paper of
12.6 MPa [37] and carbon nanotube buckypaper electrode of
14.3 MPa [38] as well as the graphene form electrode of approximately ~0.13 MPa [39].
The electrochemical activity of ROGP-600 in the Na cells was
investigated via constant-current charge/discharge cycling over the
potential range of 0.005e2.5 V with various current densities. The
charge/discharge curves (sodium extraction/insertion) of ROGP-

600 did not exhibit any distinct potential plateaus, as shown in


Fig. 6(a). This is probably related to sodium-ion storage via the
surface reactions at electrochemically nonequivalent storage sites
[13,15,25,26]. The relatively large capacity between 0.5 and 0.7 V
may be attributed to the formation of a solid electrolyte interphase
(SEI) lm on the electrode surface. Note that the reversible capacity
was 183 mAh g1, at a current density of 100 mA g1, which is
approximately 18 times greater than that of UOGPs (10 mAh g1).
This nding strongly suggests that a simple process of creating
disorder in the graphene orientation signicantly enhances the
sodium storage capability [25].
Fig. 6(b) shows the rate performances of ROGP-600 from
100 mA g1 to 8 A g1. Even at a current rate of 8 A g1, a capacity of
approximately 61 mAh g1 was observed. This performance is
obviously greater than that of UOGP [Fig. S5] and comparable to
that of the conventional non-exible graphene powder electrodes,
and surpasses the power capability of graphene-based paper
electrodes for Li-ion storage [13,19e21,25]. Fig. 6(c) compares the
current density as a function of capacity of ROGP-600 (navy square)
with the available previously reported data. Note that a capacity of
61 mAh g1 at a current density of 8 A g1 corresponds to a power
density of approximately 20,000 W kg1 with 4-V cathode counterparts in a full cell. The cycle stability of ROGP-600 was investigated at a current density of 1 A g1 (Fig. 6(d)). Stable cycles over
500 charge/discharge cycles were maintained, and the Coulombic
efciency was almost 100% during all cycles (except for the initial
cycles).
Electrochemical performance did not degrade after 1000 cycles
of bending. The bending tests for ROGP-600 were performed using
custom-made equipment, as shown in Fig. 7(a) and (b). The charge/
discharge proles shown in Fig. 7(c) and (d) indicate that the performance of the ROGP-600 was almost identical to that prior to the
bending tests. Moreover, the cycle stability was maintained over
the 500 cycles, implying the ROGP-600 electrodes are suitable for
exible NIB applications.
The feasibility of ROGP-600 as a exible battery was demonstrated by a home-made pouch cell, as shown in Fig. 8(a). The

Fig. 6. (a) Galvanostatic charge/discharge proles of ROGP-600 over the potential range 0.01e2.5 V vs. Na/Na at a current density of 100 mA g1. (b) The rate capability of ROGP600 at a various current densities in the range 100e8000 mA g1. (c) The current density vs. capacity of ROGP-600 and reference materials. (d) The cycling performance and
Coulombic efciency of ROGP-600 over 500 cycles at a current density of 1000 mA g1. (A colour version of this gure can be viewed online).

Y.S. Yun et al. / Carbon 99 (2016) 658e664

663

Fig. 7. Optical images (a) before and (b) after bending of ROGP-600 using custom-made equipment. (c) Galvanostatic charge/discharge proles of ROGP-600 following 1000 cycles of
bending in the potential range 0.01e2.5 V vs. Na/Na at a current density of 100 mA g1, and (d) cycling performance of ROGP-600 for 500 cycles at a current density of 1000 mA g1,
as well as the Coulombic efciency. (A colour version of this gure can be viewed online).

Fig. 8. (a) Photograph of a home-made pouch cell and (b) galvanostatic discharge/charge proles of the pouch cell at a current density of 200 mA g1 under its at and bent
conditions. (A colour version of this gure can be viewed online).

pouch cell was repeatedly folded and spread out into the initial
state under the electric circuit being connected, which can be
conrmed in supporting movie 1. The electrochemical performances of the ROGP-600-based pouch cell were tested in a bent
condition, as shown in Fig. 8(b) [40,41]. The galvanostatic discharge
charge proles in the bent condition are similar to the at condition, indicating a exibility and stability of the pouch cell.
Supplementary data related to this article can be found online at
http://dx.doi.org/10.1016/j.carbon.2015.12.047.

structure, leading to an anisotropic structure. The unique open


structure for sodium ion diffusion as well as the well-developed
carbon structure led to signicantly improved electrochemical
performance compared with that of uniaxially oriented graphene
paper electrodes, indicating the importance of the interplay between the graphene micro-morphology and the electrochemical
activity. The high power capability, excellent cycle stability, and the
expeditiousness and simplicity of the fabrication process used to
form the ROGP electrodes suggest that it may be a promising
exible electrode for NIBs.

4. Conclusion
Acknowledgements
Flexible graphene paper with random orientation was prepared
via a simple modied reduction process, and the resulting materials
were used as self-standing electrodes for sodium ion batteries. The
GNS building blocks had a permanently folded and crumpled

This work was supported by the World Premier Materials grant


funded by the Korea government Ministry of Trade, Industry and
Energy and by Basic Science Research Program through the

664

Y.S. Yun et al. / Carbon 99 (2016) 658e664

National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2013R1A1A2A10008534). This work was
also supported by Project Code (IBS-R006-G1).
Appendix A. Supplementary data
Supplementary data related to this article can be found at http://
dx.doi.org/10.1016/j.carbon.2015.12.047.
References
[1] S.-W. Kim, D.-H. Seo, X. Ma, G. Ceder, K. Kang, Electrode materials for
rechargeable sodium-ion batteries: potential alternatives to current lithiumion batteries, Adv. Energy Mater. 2 (7) (2012) 710e721.
[2] Y. Sun, L. Zhao, H. Pan, X. Lu, L. Gu, Y.-S. Hu, H. Li, M. Armand, Y. Ikuhara,
L. Chen, X. Huang, Direct atomic-scale conrmation of three-phase storage
mechanism in Li4Ti5O12 anodes for room-temperature sodium-ion batteries,
Nat. Commun. 4 (2013) 1870.
[3] B. Koo, S. Chattopadhyay, T. Shibata, V.B. Prakapenka, C.S. Johnson, T. Rajh,
E.V. Shevchenko, Intercalation of sodium ions into hollow iron oxide nanoparticles, Chem. Mater. 25 (2) (2013) 245e252.
[4] Y. Kim, Y. Kim, A. Choi, S. Woo, D. Mok, N.-S. Choi, Y.S. Jung, J.H. Ryu, S.M. Oh,
K.T. Lee, Tin phosphide as a promising anode material for Na-ion batteries,
Adv. Mater. 26 (24) (2014) 4139e4144.
[5] Y.U. Park, D.-H. Seo, H.-S. Kwon, B. Kim, J. Kim, H. Kim, I. Kim, H.-I. Yoo,
K. Kang, A new high-energy cathode for a Na-ion battery with ultrahigh stability, J. Am. Chem. Soc. 135 (37) (2013) 13870e13878.
[6] K. He, Y. Zhou, P. Gao, L. Wang, N. Pereira, G.G. Amatucci, K.-W. Nam, X.Q. Yang, Y. Zhu, F. Wang, D. Su, Sodium via heterogeneous disproportionation
in FeF2 electrodes for sodium-ion batteries, ACS Nano 8 (7) (2014)
7251e7259.
[7] D.Y.W. Yu, P.V. Prikhodchenko, C.W. Mason, S.K. Batabyal, J. Gun,
S. Sladkevich, A.G. Medvedev, O. Lev, High-capacity antimony sulphide
nanoparticle-decorated graphene composite as anode for sodium-ion batteries, Nat. Commun. 4 (2013) 2922.
[8] D. Su, H.-J. Ahn, G. Wang, SnO2@graphene nanocomposites as anode materials
for Na-ion batteries with superior electrochemical performance, Chem.
Commun. 49 (30) (2013) 3131e3133.
[9] D. Wu, X. Li, B. Xu, N. Twu, L. Liu, G. Ceder, NaTiO2: a layered anode material
for sodium-ion batteries, Energy Environ. Sci. 8 (1) (2015) 195e202.
[10] Y. Zhu, X. Han, Y. Xu, Y. Liu, S. Zheng, K. Xu, L. Hu, C. Wang, Electrospun Sb/C
bers for a stable and fast sodium-ion battery anode, ACS Nano 7 (7) (2013)
6378e6386.
[11] W. Li, S.-L. Chou, J.-Z. Wang, J.H. Kim, H.-K. Liu, S.-X. Dou, Sn4xP3@amorphous SneP composites as anodes for sodium-ion batteries with low cost,
high capacity long life, and superior rate capability, Adv. Mater. 26 (24) (2014)
4037e4042.
[12] Y. Kim, Y. Park, A. Choi, N.-S. Choi, J. Kim, J. Lee, J.H. Ryu, S.M. Oh, K.T. Lee, An
amorphous red phosphorus/carbon composite as a promising anode material
for sodium ion batteries, Adv. Mater. 25 (22) (2013) 3045e3049.
[13] K. Tang, L. Fu, R.J. White, L. YU, M.-M. Titirici, M. Antonietti, J. Maier, Hollow
carbon nanospheres with superior rate capability for sodium-based batteries,
Adv. Energy Mater. 2 (7) (2012) 873e877.
[14] J. Ding, H. Wang, Z. Li, A. Kohandehghan, K. Cui, Z. Xu, B. Zahiri, X. Tan,
E.M. Lotfabad, B.C. Olsen, D. Mitlin, ACS Nano 7 (12) (2013) 11004e11015.
[15] Z. Wang, L. Qie, L. Yuan, W. Zhang, X. Hu, Y. Huang, Functionalized N-doped
interconnected carbon nanobers as an anode material for sodium-ion storage with excellent performance, Carbon 55 (2013) 328e334.
[16] H. Kim, J. Hong, Y.-U. Park, J. Kim, I. Hwang, K. Kang, Sodium storage behavior
in natural graphite using ether-based electrolyte system, Adv. Funct. Mater.
25 (4) (2015) 534e541.
[17] D.A. Stevensa, J.R. Dahn, The mechanisms of lithium and sodium insertion in
carbon materials, J. Electrochem. Soc. 148 (8) (2001) A803eA811.
[18] H. Gwon, H.-S. Kim, K.U. Lee, D.-H. Seo, Y.C. Park, Y.-S. Lee, B.T. Ahna, K. Kang,
Flexible energy storage devices based on graphene paper, Energy Environ. Sci.
4 (4) (2011) 1277e1283.

[19] C. Wang, D. Li, C.O. Too, G.G. Wallace, Electrochemical properties of graphene
paper electrode used in lithium batteries, Chem. Mater. 21 (13) (2009)
2604e2606.
[20] A. Abouimrane, O.C. Compton, K. Amine, S.B.T. Nguyen, Non-annealed graphene paper as a binder-free anode for lithium-ion batteries, J. Phys. Chem. C
114 (29) (2010) 12800e12804.
[21] J.-Z. Wang, C. Zhong, S.-L. Chou, H.-K. Liu, Flexible free-standing graphenesilicon composite lm for lithium-ion batteries, Electrochem. Commun. 12
(11) (2010) 1467e1470.
[22] F. Liu, S. Song, D. Xue, H. Zhang, Folded structured graphene paper for high
performance electrode materials, Adv. Mater. 24 (8) (2012) 1089e1094.
[23] X. Zhao, C.M. Hayner, M.C. Kung, H.H. Kung, Flexible holey graphene paper
electrodes with enhanced rate capability for energy storage applications, ACS
Nano 5 (11) (2011) 8739e8749.
[24] R. Mukherjee, A.V. Thomas, A. Krishnamurthy, N. Koratkar, Photo-thermally
reduced graphene as high power anodes for lithium ion batteries, ACS Nano 6
(9) (2012) 7867e7878.
[25] Y.-X. Wang, S.-L. Chou, H.-K. Liu, S.-X. Dou, Reduced graphene oxide with
superior cycling stability and rate capability for sodium storage, Carbon 57
(2013) 202e208.
[26] H.-G. Wang, Z. Wu, F.-L. Meng, D.-L. Ma, X.-L. Huang, L.-M. Wang, X.-B. Zhang,
Nitrogen-doped porous carbon nanosheets as low-cost, high-performance
anode material for sodium-ion batteries, ChemSusChem 6 (1) (2013) 56e60.
[27] L. David, G. Singh, Reduced graphene oxide paper electrode: opposing effect of
thermal annealing on Li and Na cyclability, J. Phys. Chem. C 118 (49) (2014)
28401e28408.
[28] Y.S. Yun, V.-D. Le, H. Kim, S.-J. Chang, S.J. Baek, S. Park, B.H. Kim, Y.-H. Kim,
K. Kang, H.-J. Jin, Effects of sulfur doping on graphene-based nanosheets for
use as anode materials in lithium-ion batteries, J. Power Sources 262 (2014)
79e85.
[29] Y.S. Yun, Y.H. Bae, D.H. Kim, J.Y. Lee, I.-J. Chin, H.-J. Jin, Reinforcing effects of
adding alkylated graphene oxide to polypropylene, Carbon 49 (11) (2011)
3553e3559.
[30] J.K. Wassei, K.C. Cha, V.C. Tung, Y. Yang, R.B. Kaner, The effects of thionyl
chloride on the properties of graphene and graphene-carbon nanotube
composites, J. Mater. Chem. 21 (10) (2011) 3391e3396.
[31] R.L.D. Whitby, A. Korobeinyk, K.V. Glevatska, Morphological changes and covalent reactivity assessment of single-layer graphene oxides under carboxylic
group-targeted chemistry, Carbon 49 (2) (2011) 725.
[32] D. Yu, Y. Yang, M. Durstock, J.-B. Baek, L. Dai, Soluble P3HT-grafted graphene
for efcient bilayer-heterojunction photovoltaic devices, ACS Nano 4 (10)
(2010) 5633e5640.
[33] Y.S. Yun, J.M. Kim, H.H. Park, J. Lee, Y.S. Huh, H.-J. Jin, Free-standing heterogeneous hybrid papers based on mesoporous g-MnO2 particles and carbon
nanotubes for lithium-ion battery anodes, J. Power Sources 244 (SI) (2013)
747e751.
[34] Y.S. Yun, G. Yoon, K. Kang, H.-J. Jin, High-performance supercapacitors based
on defect-engineered carbon nanotubes, Carbon 80 (2014) 246e254.
[35] S. Park, K.-S. Lee, G. Bozoklu, W. Cai, S.T. Nguyen, R.S. Ruoff, Graphene oxide
papers modied by divalent ionsdenhancing mechanical properties via
chemical cross-linking, ACS Nano 2 (3) (2014) 572e578.
[36] H. Chen, M.B. Mller, K.J. Gilmore, G.G. Wallace, D. Li, Mechanically strong,
electrically conductive, and biocompatible graphene paper, Adv. Mater. 20
(18) (2008) 3557e3561.
[37] 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, Fabrication of graphene/polyaniline composite paper via in situ
anodic electropolymerization for high-performance exible electrode, ACS
Nano 3 (7) (2009) 1745e1752.
[38] J. Che, P. Chena, M.B. Chan-Park, High-strength carbon nanotube buckypaper
composites as applied to free-standing electrodes for supercapacitors,
J. Mater. Chem. A 1 (2013) 4057e4066.
[39] G. Zhou, L. Li, C. Ma, S. Wang, Y. Shi, N. Koratkar, W. Ren, F. Li, H.-M. Cheng,
A graphene foam electrode with high sulfur loading for exible and high
energy LieS batteries, Nano Energy 11 (2015) 356e365.
[40] G. Zhou, F. Li, H.-M. Cheng, Progress in exible lithium batteries and future
prospects, Energy Environ. Sci. 7 (4) (2014) 1307e1338.
[41] G. Zhou, L. Li, D.-W. Wang, X.-Y. Shan, S. Pei, F. Li, H.-M. Cheng, A exible
sulfur-graphene-polypropylene separator integrated electrode for advanced
LieS batteries, Adv. Mater. 27 (4) (2015) 641e647.

You might also like