You are on page 1of 7

J Mater Sci: Mater Electron

DOI 10.1007/s10854-017-7705-9

Facile and scalable synthesis of hierarchically porous graphene


architecture for hydrogen storage and high-rate supercapacitors
Zhiyuan Huang1 · Kaisheng Xia1   · Lin Zheng1 · Bo Han1 · Qiang Gao2 ·
Hongquan Wang2 · Zhen Li2 · Chenggang Zhou1 

Received: 18 April 2017 / Accepted: 12 August 2017


© Springer Science+Business Media, LLC 2017

Abstract  Porous graphene materials have been developed 1 Introduction


for many energy applications. However, a facile and scalable
synthetic approach is still a challenge. Herein, we prepare The growing energy crisis and environmental pollution
graphene-based carbons with hierarchically porous struc- associated with fossil fuels has stimulated rapid research
tures via a simple steam activation of oxygen-functionalized on sustainable energy storage and conversion systems such
graphene nanosheets. The as-prepared carbons show inter- as lithium-ion batteries, supercapacitors, fuel cells and so
connected micro-meso-macroporous structures, high specific forth [1, 2]. Hydrogen fuel cells are fast emerging as an
surface areas up to 1238 m2 g−1 and ultra-large pore volumes alternative candidate for electric vehicles (EVs) and other
up to 4.28 cm3 g−1. The highest hydrogen storage on the mobile applications, however, safe, efficient and revers-
carbon sorbents is found to be 1.47 wt% at 77 K and 1 bar. ible hydrogen storage still remains a key challenge [3, 4].
As electrode materials for supercapacitor, one of the carbons Among the developed techniques to store hydrogen, phy-
delivers a specific capacitance of 121 F g−1 at 0.5 A g−1 sisorption on carbon adsorbents such as activated carbons,
with 85% retention at 20 A g−1. The enhanced surface area, carbon nanotubes and carbide derived carbons has attracted
unique porous structure combined with the inherent prop- considerable interest due to their lightweight, high surface
erties of graphene are responsible for the excellent perfor- area, relatively low cost, excellent stability and fast kinetics
mances in hydrogen storage and high-rate supercapacitors. [5, 6]. Meanwhile, supercapacitors represent an important
energy storage device which possess high power density,
superior rate capability and long operation life. Electrode
materials are the key components that determine a superca-
pacitor’s performance. Over the past decades, carbon mate-
rials, including activated carbons and templated carbons,
have been widely investigated as the entity of supercapaci-
tor electrodes [6, 7]. Although carbon-based materials have
Electronic supplementary material  The online version of this exhibited great potential for applications in hydrogen storage
article (doi:10.1007/s10854-017-7705-9) contains supplementary
material, which is available to authorized users. and supercapacitors, however, the limited energy densities
of these materials have constrained the commercialization.
* Kaisheng Xia To resolve this problem, design and development of novel
xks_mail@126.com carbon materials with more controlled structural and chemi-
Chenggang Zhou cal characteristics is absolutely necessary.
cgzhou@cug.edu.cn Graphene as a dazzling carbon material has aroused much
1
Sustainable Energy Laboratory, Faculty of Material
attention in the field of hydrogen storage and supercapaci-
Science and Chemistry, China University of Geosciences, tors [6–8]. Unluckily, graphene sheets prefer restacking
Wuhan 430074, People’s Republic of China or aggregating to form non-porous nanostructures during
2
Faculty of Material Science and Chemistry, China University preparation, leading to dramatic decrement of surface area
of Geosciences, Wuhan 430074, People’s Republic of China [9]. As a consequence, the hydrogen adsorption capacities

13
Vol.:(0123456789)
J Mater Sci: Mater Electron

and electrochemical capacitances of solely graphene are investigated. This work would help to achieve a better
far below expectations. Many attempts have been made in understanding on the structural change of the graphene
recent years for tuning the textural properties of graphene oxide subjected to steam activation, and may be further
to improve its energy storage, where chemical or physical extrapolated to benefit other applications of steam-acti-
activation have been demonstrated to be effective approaches vated graphene.
[10–16]. In general, chemical activation can be implemented
using different chemical reagents (typically KOH, ­H3PO4
or ­ZnCl2), while physical activation is conducted in C ­ O 2,
air or steam atmosphere. It is recognized that chemical 2 Experimental section
activation can greatly increase the microporosity and spe-
cific surface area and thus the energy density of graphene 2.1 Materials synthesis
material. Nevertheless, chemical activation is a complicated
process and may bring about secondary pollution, which Graphite oxide (GO) was synthesized by the modi-
will heavily hinder its application in large-scale production. fied Hummer’s method [25]. The starting materials and
In comparison, the physical activation (such as ­CO2, air or chemicals were analytical grade and without any further
steam activation) offers a more desirable option because the purification. First, Graphite (5 g) was added to concen-
process of activation involves a simple gas–solid reaction trated ­H2SO4 (20 mL) in which ­P2O5 (2.5 g) and ­K2S2O8
without introducing any impurities [17–20]. Among the (2.5 g) were completely dissolved at 80 °C. The mixture
commonly used physical activation methods, steam activa- was kept at 80 °C for 6 h using an oil bath. After cool-
tion can be regarded as a facile and environmentally friendly ing down and diluted with deionized water (400 mL), the
approach [21, 22]. A few recent studies have prepared steam pre-oxidized graphite was filtered and washed until the
activated graphene materials with high specific surface area filtrate was neutral and dried at 65 °C for 24 h. Next, 2 g
(SSA) and porosity, which exhibited potential applications of it was dispersed into chilled H­ 2SO4 (48 mL), and then
in adsorption [13], sensing [14, 23] and lithium-ion battery ­KMnO4 (6 g) was slowly added with vigorous stirring. The
[24]. In particular, Huang and co-workers prepared steam rate of addition was carefully controlled to keep the reac-
etched porous graphene oxide by hydrothermal steaming tion temperature below 5 °C. After stirring for 1.5 h, the
of graphene oxide at 200 °C, and the products showed a obtained dark green mixture was transferred to 35 °C oil
­ O2 detection [14]. Neverthe-
great increased sensitivity for N bath and stirred for another 1.5 h followed by addition of
less, the graphene oxide sheets was severely disrupted by 100 mL deionized water. After additional stirring at 35 °C
the steam activation because the graphene oxide contains for 1 h, the mixture was heated to 80 °C and stirred for 1 h,
abundant defect sites, leading to a possible violent and fast and then a brown turbid liquid was obtained. Then, H ­ 2O 2
reaction between graphene oxide and steam. Alternatively, (30%, 20 mL) were added to get a bright yellow mixture.
Han and co-workers synthesized carbon-based porous mate- The mixture was allowed to settle overnight. The superna-
rials through steam activation of graphene aerogel, and the tant was decanted, and the remaining product was further
products exhibited high SSA (830–1230 m2 g−1) and large washed with diluted HCl and deionized water until the pH
pore volume (2.2–3.6 cm3 g−1) [13]. However, the synthesis of the solution became neutral. The obtained product was
of graphene aerogel suffered from a lengthy and complicated dried at 50 °C.
procedure (including hydrothermal treatment and freeze- To prepare steam-activated exfoliated graphene (SAEG),
drying), which is unsuitable for large-scale preparation. thermally exfoliated graphene oxide (TEG) needs to be pre-
In order to simplify the synthesis procedure and pared beforehand. Typically, 1.0 g of dried GO was placed
increase the productivity of steam activated porous gra- in a quartz tube and purged with N ­ 2 for 30 min to remove
phene materials, we employed our previously reported air, and then the tube was rapidly transferred to a vertical
ultra-fast thermal exfoliation technique [20] to prepare furnace preheated to 800 °C. The TEG was obtained after
thermally exfoliated graphene nanosheets (TEG), and thermal exfoliation at 800 °C for 1 min followed by gradual
then used it as a precursor for steam activation in the cooling to room temperature under ­N2 flow. Subsequently,
present work. The TEG possesses highly loose structure the TEG powder was placed in the centre of a tube fur-
and moderate defects, which could facilitate the steam nace and activated at different temperature (800, 850 and
activation. The as-prepared carbons show interconnected 900 °C, respectively) for 1 h under a flowing N ­ 2 and steam
−1
micro-meso-macroporous structures, high specific surface atmosphere, with a ­N2 flow of 40 mL min . The heating
areas up to 1238 m2 g−1 and ultra-large pore volumes up rate of the tube furnace was set at 10 °C min−1. After cool-
to 4.28 cm 3 g −1. Furthermore, the effects of activation ing to room temperature under N ­ 2, the obtained sample was
temperature on the hydrogen adsorption and superca- collected and denoted as SAEG-x, where x represents the
pacitive properties of the resulting carbon materials are activation temperature.

13
J Mater Sci: Mater Electron

2.2 Materials characterizations 10 MPa for 1 min. Approximately, 1.5 mg of active materials


was loaded in each electrode. The electrochemical meas-
Powder X-ray diffraction (XRD) patterns of the prepared urements were carried out by cyclic voltammetry (CV),
samples were recorded on a Bruker AXS D8-FOCUS dif- galvanostatic charge–discharge (GCD) and electrochemical
fractometer using monochromatized Cu-Kα radiation (40 kV, impedance spectroscopy (EIS) methods. The CV and GCD
20  mA; λ = 0.1540598  nm). Fourier transform infrared were recorded in the −1.0 to 0 V potential window. EIS
(FT-IR) spectroscopy was carried on a Nicolet6700 system measurements were carried out in the frequency range of
using the KBr pellet method. The microscopic feature and 10 kHz to 10 mHz. The specific capacitance was calculated
morphology of the samples were observed by field scanning from the GCD measurement using the following equation:
electron microscopy (FESEM, Hitachi SU8010) and trans-
i⋅t
mission electron microscopy (TEM, Philips CM12TEM/ C=
m⋅V
STEM) operated at 120 kV. Specific surface area and pore
distribution measurements were carried out by ­N2 adsorp- where i is the current, t is the time of discharge, m is the
tion–desorption analysis at 77  K using a Micromeritics mass of the active material, and V is the potential window.
ASAP 2020 HD88 system. All the samples were degassed
at 200 °C for 6 h before adsorption experiments. The Raman
spectra were recorded using a Horiba Jobin-Yvon LabRAM 3 Results and discussion
HR800 Raman microspectrometer with an excitation laser
of 532 nm. The specific surface area (SSA) was calculated The strategy to prepare SAEGs is schematically shown in
from the adsorption data in the relative pressure range Fig. 1. First, TEG was obtained by rapid thermal exfoliation
between 0.05 and 0.35 using Brunauer–Emmett–Teller of GO at 800 °C. During this process, the gas ­(CO2 or O ­ 2)
(BET) method. The total pore volume was determined from escaping can exfoliate GO into numerous curved graphene
the amount of ­N2 adsorbed at a relative pressure of 0.995. nanosheets, which may stack together to form mesopores and
The micropore volume (­ VMicro) and micropore surface area macropores [26]. Due to the rapid thermal exfoliation, abun-
­(SMicro) were estimated using a t-plot method. The pore size dant oxygen-containing groups are still preserved on the gra-
distribution was analysed by Barrett-Joyner-Halenda (BJH) phene nanosheets, which endows the TEG with high reactiv-
method using the data of the adsorption branch. ity. In the following activation procedure, active carbon atoms
in graphene layer would react with ­H2O to produce CO and
2.3 Performance measurements ­H2, leaving behind carbon vacancies that gradually grow into
micropores or mesopores [14, 27]. Thus, the SAEGs with large
The hydrogen adsorption capacities of the samples were specific surface area and hierarchical porosity were obtained.
measured by the volumetric method at 77 K in the pressure The SEM and TEM measurements were first performed to
range of 0–1 bar using the Micromeritics ASAP 2020 HD88 examine the morphology and porous structure of the prepared
system. Before the adsorption measurements, samples were samples. It is seen that the TEG (Fig. 2a) is composed of thin
degassed at 350 °C for 6 h. sheets with curly and wrinkly morphology. In contrast, there
The electrochemical capacitive properties of the samples are a number of nanopores on the surface of the SAEG-850 as
were tested in an electrochemical workstation (VMP3, Bio- highlighted by the dashed boxes in Fig. 2b, which are believed
logic) using a conventional three-electrode configuration to be created by steam activation. Figure 2c, d show the TEM
with 6.0 M aqueous KOH solution as electrolyte. A Pt wire images of SAEG-850, which display a typical wrinkled layer-
and an Hg/HgO electrode were used as counter and reference like structure and numerous nanopores in the graphene sheets.
electrodes, respectively. The active materials were prepared It is anticipated that the SAEGs with such architecture might
by mixing the SAEG, acetylene black and polytetrafluoroeth- have large surface area and high porosity, which are beneficial
ylene (PTFE) in a weight ratio of 80:10:10 with ethanol. And for both hydrogen storage and supercapacitive applications.
then the pastes were pressed into Ni foam and then dried The XRD patterns of the samples are exhibited in Fig. 3a.
overnight at 80 °C. The dried electrode was then pressed at It is observed that the TEG and SAEGs show a very weak

Fig. 1  Schematic diagram
of the preparation process of
SAEGs via steam activation of
TEG

13
J Mater Sci: Mater Electron

Fig. 2  SEM images of TEG (a) and SAEG-850 (b); TEM images of the SAEG-850 (c and d)

broad (002) graphitic peak, while GO exhibits a sharp (001) from 432 m2 g−1 and 1.45 cm3 g−1 for TEG to 1238 m2 g−1
peak at 2θ of 11.4°. It indicates that the GO was converted and 4.28 cm3 g−1 for SAEG-850, respectively. The remark-
into many single- and several-layer nanosheets during the able increments are ascribed to the porosity developments
exfoliation process and the exfoliated structure is main- of SAEGs caused by steam activation. In our experiments,
tained after steam activation. The FT-IR spectra shown in 850 °C seems to be the optimal activation temperature. A
Fig. 3b demonstrate the weakening of bands attributed to higher temperature such as 900 °C may lead to overreact
C–O and C=O. It suggests the reduction of GO after exfolia- between graphene and steam, which may results in the
tion and steam activation. The textural properties of SAEGs destruction of the porous structure and low yield [13]. Sig-
are investigated by nitrogen adsorption–desorption analy- nificantly, the high SSA of 1238 m2 g−1 and the ultra-large
ses. As shown in Fig. 3c, the SAEGs show increased N ­ 2 pore volume of 4.28 cm3 g−1 in SAEG-850 are higher than
adsorption after steam activation and exhibit a combination that of most physical activated graphene-based carbons,
of type IV and type II isotherms, indicating the presence of such as ­CO2 activated graphene framework (829 m2 g−1 and
mesopores and macropores. Moreover, the SAEGs exhibit 2.83 cm3 g−1) [15], air activated graphene (646 m2 g−1 and
distinct ­N2 adsorption at very low relative pressure, which 2.39 cm3 g−1) [12] and steam activated graphene aerogels
implies the existence of abundant micropores. The results (1230 m2 g−1 and 3.6 cm3 g−1) [13].
are consistent with the SEM and TEM measurements and To determine the quality and layer number of the as-
clearly show the hierarchically micro-meso-macroporous prepared graphene, Raman measurement was performed
structures in the SAEGs. The pore size distributions based as shown in Fig. 4. All the samples have two pronounced
on the BJH method (Fig. 3d) show a narrow distribution of peaks at around 1350 and 1580 cm−1, which corresponds
small mesopores centered at about 2.5 nm and a broad dis- to the D and G bands of graphene, respectively. It is known
tribution of pores in the range of 5–50 nm. As summarized that the intensity ratio of D to G peak ­(ID/IG) is related to
in Table 1, the calculated SSAs and pore volumes increase the crystalline quality of graphene layers: the smaller the

13
J Mater Sci: Mater Electron

Fig. 3  XRD patterns (a), FT-IR spectra (b), nitrogen adsorption–desorption isotherms (c) and BJH pore size distributions (d) of the prepared
samples

Table 1  The textural parameters and hydrogen uptake capacities of


the TEG and SAEGs

Samples SBET ­(m2 g−1) VTotal SMicro ­(m2 g−1) H2


­(cm3 g−1) ­uptakea
(wt%)

TEG 432 1.45 70 0.58


SAEG-800 838 2.34 262 0.96
SAEG-850 1238 4.28 184 1.47
SAEG-900 996 3.31 172 1.25
a
 H2 uptake at 77 K and 1 bar

value, the better the crystalline quality. TEG has an ­ID/


IG value of 0.92, indicating the presence of defects due
to rapid heating. After steam activation, the value of ­ID/
IG increases to 1.04 for SAEG-800, and 1.31 for SAEG- Fig. 4  Raman spectra of TEG, SAEG-800, SAEG-850 and SAEG-
900
850, and then drops to 1.14 for SAEG-900, respectively.
This illustrates that steam activation bring more defects
and edges to the as-prepared samples. The 2D band CVD grown graphene, which is different from bulk-pro-
(∼2750 cm−1) has been used to estimate the layer num- duced graphene materials. Therefore, we choose another
ber of the graphene because the shape and intensity of calculating method, which is based on the value of I­ 2D/
the 2D band changes with number of layers [28]. How- IG and empirical fitting, to estimate the mean number of
ever, traditional method is often carried out on very few layers of the as-prepared graphene: Layer number = 1.04
nanosheets of micromechanically-cleaved graphene or ­(I2D/IG)−2.32 [29, 30]. The calculated layer number of TEG

13
J Mater Sci: Mater Electron

is about 6.6, which is consistent with its specific surface The supercapacitive performance of SAEGs are shown
area (432 m2 g−1), ∼1/6 of the theoretical surface area of in Fig. 5b, c. It can be found that all the cyclic voltammetry
graphene [31]. In contrast, the calculated layer numbers of (CV) curves exhibit the nearly rectangular shape (Fig. 5b),
the SAEGs are all greater than 10, which may be ascribed showing the standard electrochemical double layer capaci-
to the irreversible stacking of graphene during thermal tive behavior. The specific capacitances of SAEG-800,
annealing process. Although the graphene nanosheets SAEG-850 and SAEG-900 calculated under the current
stack to some extent, the steam activation can introduce density of 0.5 A g−1 are 121, 111 and 96 F g−1, respectively.
a lot of micro- and mesopores into the graphene-based Although this value is higher than air activated graphene
carbon, and thus increase its SSA greatly. (102 F g−1 at 0.1 A g−1) [12], it is still moderate compared
Figure 5a shows the hydrogen adsorption isotherms of with other physical activated carbons such as steam activated
TEG and SAEGs at 77 K, which display broad knees and carbon nanofibers [35] and ­CO2 activated graphene frame-
gradually increased H­ 2 uptake with the increment of pres- works [15] as shown in Table S2. The possible reason is that
sure. The highest hydrogen adsorption capacity of 1.47 wt% the functional groups introduced by steam activation could
at 77 K and 1 bar is found in sample SAEG-850 (Table 1). degrade the supercapacitive performance of SAEGs because
The value is higher than those of ­CO2-activated graphene of their instability or polarity mismatch with the electrolyte
(0.75 wt%) [25], GO-based frameworks (1.2 wt%) [32], GO- [35]. Remarkably, all the SAEGs display an excellent capac-
based network materials (1.24 wt%) [33] and many other ity retention rate up to 85% at 20 A g−1 (compared with
carbon materials under the same testing conditions as shown 0.5 A g−1), indicating their potential in high-rate superca-
in Table S1. Besides, there is a linear relationship between pacitors (Fig. 5c). The high-rate capability is attributed to
the SSAs and hydrogen adsorption capacities as illustrated the hierarchical porous structure of the SAEGs, in which
in Fig. S1, which indicates that the ­H2 uptake in SAEGs is macropores and large mesopores can minimize the distances
mainly physisorption mode [34]. Thus, the obvious higher between electrolytes and the interior surfaces of the pores
SSA of SAEG-850 could be used to explain its superior and thus enable rapid ion transport, and small mesopores
hydrogen adsorption performance. and micropores can facilitate the charge storage [36–39].

Fig.  5  a Hydrogen adsorption isotherms of TEG and SAEGs; b CV curves of SAEGs in 6 M KOH at scan rate of 50 mV s−1; c specific capaci-
tance change of SAEGs at different current densities; d Nyquist plots impedance of SAEGs

13
J Mater Sci: Mater Electron

The electrochemical impedance spectroscopy (EIS) were 10. Y. Zhu, S. Murali, M.D. Stoller, K.J. Ganesh, W. Cai, P.J. Ferreira,
performed in a frequency range from 100 kHz to 10 mHz, A. Pirkle, R.M. Wallace, K.A. Cychosz, M. Thommes, D. Su, E.A.
Stach, R.S. Ruoff, Science 332, 1537–1541 (2011)
and the corresponding Nyquist impedance plots of the 11. M. Myglovets, O.I. Poddubnaya, O. Sevastyanova, M.E. Lind-
SAEGs electrodes are shown in Fig. 5d. All EIS curves show ström, B. Gawdzik, M. Sobiesiak, M.M. Tsyba, V.I. Sapsay, D.O.
straight line in low-frequency region and small semicircle in Klymchuk, A.M. Puziy, Carbon 80, 771–783 (2014)
the high-frequency region. It suggests the fast formation of 12. Y. Lin, X. Han, C.J. Campbell, J.-W. Kim, B. Zhao, W. Luo, J. Dai,
L. Hu, J.W. Connell, Adv. Funct. Mater. 25, 2920–2927 (2015)
electric double layer capacitance and small charge transfer 13. Z.Y. Sui, Q.H. Meng, J.T. Li, J.H. Zhu, Y. Cui, B.H. Han, J. Mater.
resistance in SAEGs, which is also in consistent with the CV Chem. A 2, 9891–9898 (2014)
results. By comparing the patterns in the insets of Fig. 5d, 14. T.H. Han, Y.K. Huang, A.T. Tan, V.P. Dravid, J. Huang, J. Am.
SAEG-800 exhibits the smallest Z′-axis intercept and semi- Chem. Soc. 133, 15264–15267 (2011)
15. S. Yun, S.O. Kang, S. Park, H.S. Park, Nanoscale 6, 5296–5302
circle at high frequency, indicating the lowest charge transfer (2014)
resistance and internal resistance in it. This may explain why 16. H. Xuan, Y. Wang, G. Lin, F. Wang, L. Zhou, X. Dong, Z. Chen,
the SAEG-800 has the best supercapacitive performance. RSC Adv. 6, 15313–15319 (2016)
17. K. Xia, Q. Gao, J. Jiang, J. Hu, Carbon 46, 1718–1726 (2008)
18. K. Xia, Q. Gao, S. Song, C. Wu, J. Jiang, J. Hu, L. Gao, Int. J.
Hydrog. Energy 33, 116–123 (2008)
4 Conclusions 19. K. Xia, Q. Gao, C. Wu, S. Song, M. Ruan, Carbon 45, 1989–1996
(2007)
A facile and scalable steam activation was employed to 20. K. Xia, Q. Li, L. Zheng, K. You, X. Tian, B. Han, Q. Gao, Z.
Huang, G. Chen, C. Zhou, Microporous Mesoporous Mater. 237,
treat oxygen-functionalized graphene nanosheets for super- 228–236 (2017)
capacitors and hydrogen storage. As a result, the obtained 21. K. Fu, Q. Yue, B. Gao, Y. Sun, L. Zhu, Chem. Eng. J. 228, 1074–
graphene-based carbons exhibited a hierarchically micro- 1082 (2013)
meso-macroporous architecture built by holey graphene 22. M. Molina-Sabio, M.T. Gonzalez, F. Rodriguez-Reinoso, A.
Sepúlveda-Escribano, Carbon 34, 505–509 (1996)
sheets. The specific surface area and pore volume of the 23. D.H. Wang, Y. Hu, J.J. Zhao, L.L. Zeng, X.M. Tao, W. Chen, J.
graphene-based carbons could be tuned in the range of Mater. Chem. A 2, 17415–17420 (2014)
838–1238 m2 g−1 and 2.34–4.28 cm3 g−1, respectively, by 24. H. Tang, J. Zhang, Y.J. Zhang, Q.Q. Xiong, Y.Y. Tong, Y. Li, X.L.
varying the steam activation temperature. The carbons were Wang, C.D. Gu, J.P. Tu, J. Power Sources 286, 431–437 (2015)
25. K. Xia, X. Tian, S. Fei, K. You, Int. J. Hydrog. Energy 39, 11047–
found to have impressive hydrogen adsorption capacities as 11054 (2014)
high as 1.47 wt% at 77K and 1 bar and high capacitance 26. H.C. Schniepp, J.L. Li, M.J. McAllister, H. Sai, M. Herrera-
retentions up to 85% from 0.5 to 20 A g−1 using as super- Alonso, D.H. Adamson, R.K. Prud’homme, R. Car, D.A. Saville,
capacitor electrodes. Our results demonstrated the potential I.A. Aksay, J.Phys. Chem. B 110, 8535–8539 (2006)
27. R.T. Yang, K.L. Yang, Carbon 23, 537–547 (1985)
applications of steam-activated graphene materials in hydro- 28. Z. Peng, Z. Yan, Z. Sun, J.M. Tour, ACS Nano 5, 8241–8247
gen storage and high-rate supercapacitors. (2011)
29. C. Backes, K.R. Paton, D. Hanlon, S. Yuan, M.I. Katsnelson, J.
Acknowledgements  This work was supported by the Natural Sci- Houston, R.J. Smith, D. McCloskey, J.F. Donegan, J.N. Coleman,
ence Foundation of Hubei Province (No. 2015CFB187), National Nanoscale 8, 4311–4323 (2016)
Natural Science Foundation of China (No. 21303170), and the Fun- 30. J. Phiri, P. Gane, T.C. Maloney, J. Mater. Sci. 52, 8321–8337
damental Research Funds for the Central Universities, China Univer- (2017)
sity of Geosciences (Wuhan) (No. CUGL150414, CUGL140413 and 31. X. Huang, K. Qian, J. Yang, J. Zhang, L. Li, C. Yu, D. Zhao, Adv.
CUG120115). Mater. 24, 4419–4423 (2012)
32. G. Srinivas, J.W. Burress, J. Ford, T. Yildirim, J. Mater. Chem.
21, 11323–11329 (2011)
33. Y. Cui, Q.Y. Cheng, H. Wu, Z. Wei, B.H. Han, Nanoscale 5,
References 8367–8374 (2013)
34. R. Ströbel, J. Garche, P.T. Moseley, L. Jörissen, G. Wolf, J. Power
1. P. Simon, Y. Gogotsi, Nat. Mater. 7, 845–854 (2008) Sources 159, 781–801 (2006)
2. L. Schlapbach, A. Zuttel, Nature 414, 353–358 (2001) 35. K.-H. Jung, W. Deng, D.W. Smith, J.P. Ferraris, Electrochem.
3. P. Jena, J. Phys. Chem. Lett. 2, 206–211 (2011) Commun. 23, 149–152 (2012)
4. D. Mori, K. Hirose, Int. J. Hydrogen Energy 34, 4569–4574 36. J. Chmiola, G. Yushin, Y. Gogotsi, C. Portet, P. Simon, P.L. Tab-
(2009) erna, Science 313, 1760–1763 (2006)
5. Y. Xia, Z. Yang, Y. Zhu, J. Mater. Chem. A 1, 9365–9381 (2013) 37. Z.S. Wu, Y. Sun, Y.Z. Tan, S. Yang, X. Feng, K. Mullen, J. Am.
6. M. Sevilla, R. Mokaya, Energy Environ. Sci. 7, 1250–1280 (2014) Chem. Soc. 134, 19532–19535 (2012)
7. Y. Huang, J. Liang, Y. Chen, Small 8, 1805–1834 (2012) 38. C. Largeot, C. Portet, J. Chmiola, P.L. Taberna, Y. Gogotsi, P.
8. A. Ghosh, K.S. Subrahmanyam, K.S. Krishna, S. Datta, A. Govin- Simon, J. Am. Chem. Soc. 130, 2730–2731 (2008)
daraj, S.K. Pati, C.N.R. Rao, J. Phys. Chem. C 112, 15704–15707 39. D. Zhou, H. Wang, N. Mao, Y. Chen, Y. Zhou, T. Yin, H. Xie,
(2008) W. Liu, S. Chen, X. Wang, Microporous Mesoporous Mater. 241,
9. F. Akbar, M. Kolahdouz, S. Larimian, B. Radfar, H.H. Radamson, 202–209 (2017)
J. Mater. Sci. Mater. Electron. 26, 4347–4379 (2015)

13

You might also like