You are on page 1of 53

Journal Pre-proofs

Review

Graphene oxide: An emerging electromaterial for energy storage and conver‐


sion

Yuheng Tian, Zhichun Yu, Liuyue Cao, Xiao Li Zhang, Chenghua Sun, Da-
Wei Wang

PII: S2095-4956(20)30494-0
DOI: https://doi.org/10.1016/j.jechem.2020.07.006
Reference: JECHEM 1458

To appear in: Journal of Energy Chemistry

Received Date: 28 May 2020


Revised Date: 7 July 2020
Accepted Date: 7 July 2020

Please cite this article as: Y. Tian, Z. Yu, L. Cao, X. Li Zhang, C. Sun, D-W. Wang, Graphene oxide: An
emerging electromaterial for energy storage and conversion, Journal of Energy Chemistry (2020), doi: https://
doi.org/10.1016/j.jechem.2020.07.006

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Published by ELSEVIER B.V. and Science Press on behalf of Science Press and Dalian Institute of
Chemical Physics, Chinese Academy of Sciences.
Review

Graphene oxide: An emerging electromaterial for energy storage


and conversion
Yuheng Tiana, Zhichun Yua, Liuyue Caoa,*, Xiao Li Zhangb, Chenghua Sunc,d, Da-Wei
Wanga,*
aSchool of Chemical Engineering, The University of New South Wales, Sydney 2052, NSW,
Australia
bSchool of Materials Science and Engineering, Zhengzhou University, Zhengzhou 450001,
Henan, China
cDepartment of Chemistry and Biotechnology, and Centre for Translation Atomaterials,
Swinburne University of Technology, Hawthorn 3122, VIC, Australia
dSchool of Chemical Engineering and energy Technology, Dongguan University of
Technology, Dongguan 523808, Guangzhou, China
*Corresponding authors.

Email addresses: da-wei.wang@unsw.edu.au (D. Wang), l.cao@unsw.edu.au (L. Cao)

Abstract

This paper gives a comprehensive review of the recent progress on electrochemical energy
storage devices using graphene oxide (GO). GO, a single sheet of graphite oxide, is a
functionalised graphene, carrying many oxygen-containing groups. This endows GO with
various unique features for versatile applications in batteries, capacitors and fuel cells.
Specific applications are considered principally including use in electrodes as the active
materials to enhance the performance or as substrates to diversify the structures, in solid-state
electrolytes and membranes to improve the ionic conductivity and mechanical properties, and
in interlayers to protect the electrodes, membranes or current collectors. Furthermore, the
challenges and future prospects are discussed in the paper for encouraging further research
and development of GO applications.

Keywords: Graphene oxide; Electrochemical storage; Batteries; Capacitors; Fuel cells

1
Yuheng Tian received her Bachelor’s degree in chemical engineering from the
Shenyang University of Chemical Technology in 2012. She received her
Master’s degree in sustainability engineering from Heriot-Watt University in
2013. She obtained her Master’s degree by research (2016) and Ph.D. degree
(2019) in chemical engineering from the University of New South Wales.

Zhichun Yu received his Bachelor’s degree from the Beijing University


of Technology in 2017. He is studying for his master’s degree in the field
of electrochemical engineering at the University of New South Wales. His
research mainly focuses on high-performance zinc-based batteries.

Liuyue Cao is a Research Associate at School of Chemical Engineering,


University of New South Wales (UNSW). She received her Bachelor’s
(2011) and Master’s degree (2014) from Hunan University. She obtained
her Ph.D, degree from the University of New South Wales in 2018. Her
current research interests are in renewable energy storage and conversion,
including redox flow batteries and zinc-based batteries.

Xiao Li Zhang received her Bachelor’s degree (2001)


in Chemistry at Shandong University, P. R. China, both Master’s
degree (2005) and PhD (2008) in Material Chemistry under
Professor Young Soo Kang at Pukyong National University, South
Korea. Then she worked as a Research Fellow with Professor Yi-Bing
Cheng and Professor Udo Bach at Monash University, Australia.
After shortly trained in the University of Melbourne as a Research
Fellow, she joined the University of New South Wales, Australia,
as a Vice-Chancellor’s Research Fellow (2012) working with Professor Rose Amal. In 2016,
she took a full Professor position at Zhengzhou University, P.R. China.

Chenghua Sun received her Bachelor’s degree (2000) in Materials Physics


at Central South University and PhD (2008) in Materials Science under
Professor Hui-Ming Cheng at Institute Of Metal Research, Chinese
Academy of Sciences, China. Dr Sun joined the University of Queensland
as a Postdoctoral Fellow in 2007 and got the Faculty Position (Lecturer) in
Monash University in 2013. Since 2017, Dr Sun have been as an associate
Professor at Swinburne University of Technology, Australia.

Da-Wei Wang is currently an Associate Professor, an ARC Future Fellow


and a UNSW Scientia Fellow in the School of Chemical Engineering,

2
UNSW Sydney. His research interest spreads from the fundamentals of the chemistry and
interface mechanisms of two-dimensional energy materials, to the development of advanced
electrochemical energy devices, including supercapacitors, rechargeable batteries, and
electrolysers. As a Chief Investigator, Da-Wei has attracted numerous competitive external
grants. He has contributed 2 book chapters, >100 journal publications, 8 patents and over 20
keynote/invited presentations, which received >20,000 citations with an H-index of 52. Da-
Wei has won some prestigious awards including the 2018, 2019 Clarivate Analytics HCR
Awards, the Finalist of 2018 AMP Tomorrow Maker, and the 2013 Scopus Young
Researcher Award in Engineering and Technology.

1. Introduction
Since 2004, graphene, which comprises a 2D honeycomb network of sp2-hybridised
carbon, has been considered to be a novel material as a building block for carbonaceous
materials [1–3]. It has a profound impact in the field of electrochemistry, due to its
exceptional physicochemical properties including a high specific surface area, strong
mechanical strength, excellent thermal conductivity, extraordinary optical transmittance and
electron transport capabilities [1,2,4,5]. It is not surprising, therefore, that intense research
covering graphene and related derivatives for revealing their potential applications in the
areas of science and technology has been motivated in recent years [6]. One remarkable
branch of graphene research is given to graphene oxide (GO), which is a precursor for
graphene synthesis by either chemical or thermal reduction processes [7]. GO is considered a
monolayer of graphite oxide, obtained by the oxidation treatment of graphite with subsequent
dispersion and delamination in water or appropriate organic solvents, though it can be used or
stored in the dried, agglomerated forms.

The chemical structure of GO is identified as mainly the carboxylic groups at the sheet
edges and the abundant hydroxyl and epoxide groups on the basal plane [8]. Its precise
atomic structure, however, still remains to be fully determined, due to the nature and various
distribution of the oxygenated groups, and the complicated nonstoichiometric atomic
composition.[7–9] The existence of oxygen functional groups endow GO with many
intriguing physicochemical properties, which include electronic, optical, thermal, mechanical,
and electrochemical properties, as well as chemical reactivity [10,11]. These oxygen
functional groups can serve as effective sites to immobilise various electroactive species for
chemical modification or functionalisation of GO [8,12–14]. Therefore, the physicochemical
properties of GO can be tuned chemically, thermally, or electrochemically [7,12,15,16].

3
Since GO forms stable dispersion in aqueous or non-aqueous solvents, conventional
techniques through solution processing such as spin coating, drop casting, dip coating or
spray coating are generally used to fabricate GO-based thin films [17]. GO can be virtually
applied as an amphiphilic substance because of the hydrophilic –COOH groups and the
hydrophobic polyaromatic network of unoxidized benzene rings [18]. The amphiphilicity
enables self-assembly of GO at various interfaces in which aqueous GO dispersion is
normally employed as one phase (liquid phase) to pair with another phase (air, liquid or solid)
[19]. The self-assembly methods provide simple and effective strategies for preparing various
GO-based materials with controlled microstructures and tuned properties [19], which enable
the widespread applications for transparent conductors, polymer composites, catalytic support,
and electrochemical energy storage [17].

Noticeably, substantial effort has been devoted to the development of renewable and
green energy sources and carriers over the past decades, because of the global concerns in the
ever-increasing environmental problems and the up-coming depletion of fossil fuels [20]. GO
and its derivatives as well as composites, with the excellent properties and structural diversity,
are being studied in versatile applications of electrochemical energy storage, including
batteries, capacitors and fuel cells, as shown in Fig. 1. For example, the oxygen functional
groups of GO can be reduced and act as oxidant, thus forming closely contacted composites
with the properties of rGO (e.g. rGO/S cathodes, rGO/alkali metal anodes, etc.). The negative
charged groups on GO can not only attract positive charged species (e.g. protons, metal
cations, etc.) for faster ion transport or work as anchoring sites for the growth of metal or
metal oxide nanoparticles, but also repulse ions of the same charge (e.g. anions of the
electrolyte, polysulfide, etc.) to prevent corrosion as a current collector protector or hinder the
diffusion of polysulfides as a membrane material. Furthermore, these functional groups,
defects and edges exhibit great electro-activity which can facilitate the electrochemical
reaction kinetics. In the meantime, the high specific surface area as well as the adjustable
interlayer distance of the thin-layered structures of GO allow the occurrence of
electrochemical reactions while confining the size or accommodating volume variation of the
products. In another aspect, the unoxidized polyaromatic rings can function as the
carbonaceous source, enable hydrophobic interactions (π-π stacking) and ensure mechanical
strength. Additionally, depending on the reduction extent, GO can be used as either insulating
dielectric spacers or electrically conductive substrates with the aid of the free π electrons on
the surface. Due to the versatile functions derived from its unique structure, GO has shown

4
great potential in energy systems by contributing to electrodes, electrocatalysts, protection
layers, printing inks, fillers and membranes, etc. Herein, we aim to summarise the
applications of GO in the energy storage and conversion systems and discuss the roles and
advantages of GO in detail.

Fig. 1. Schematic illustration of applications of GO in electrochemical energy storage.

2. Application of graphene oxide in batteries

2.1. Lithium ion batteries

2.1.1. Contributions to electrodes


Rechargeable lithium ion batteries (LIBs), normally using graphite anode and lithiated
transition metal oxide cathode, power today’s portable, entertainment, computing and

5
telecommunication equipment [21]. Nano-sized cathode materials, beneficial to large surface
area and short diffusion distance, have been successfully obtained through a GO assisted
facile hydrothermal synthesis method. GO significantly affects the morphology and particle
size of the electrode materials. LiMnPO4 nanoparticles were synthesised according to the
illustration in Fig. 2(a), during which Mn2+ ions bonded by the oxygen functional groups on
GO sheets via coordination interactions and the LiMnPO4 was finally formed by a
solvothermal process [22]. Adding GO additive in a solvothermal system was effective to
reduce size of LiMnPO4. Likewise, assisted by GO, LiMn0.6Fe0.4PO4 nanoparticles and
V3O7·H2O nanobelts were prepared as the cathodes in LIBs [23,24]. Especially noteworthy is
that GO, in these cases, plays a role on restricting the particle size of the materials, while the
reduction of GO happens in the subsequent solvo-/hydrothermal process.

Furthermore, GO can be coated on the cathode materials so as to construct the


advantageous architectures to improve the performance of LIBs. GO-coated vanadium oxide
nanomaterials, where the functional groups of GO played an anchoring role for vanadium
oxide nanostructures, was reported as the cathode for LIBs [25]. The GO-coated cathode
exhibited a discharge capacity of 185 mA h g−1 at 100 mA g−1, which was higher than that of
the carbon-coated vanadium oxide electrode (135 mA h g−1). The aggregation of active
materials occurred in the vanadium oxide nanostructures during the charge-discharge process,
resulting in the capacity loss. The agglomeration and deformation of the electrode
nanostructures, however, was suppressed by the GO, which ensured the electrode to endure
the large volume variation and simultaneously maintains good electrical contact during the
cycling, thereby enhancing the electrochemical performance [25]. Similarly, 3D porous
LiFePO4 microspheres, coating by in-situ microwave exfoliated GO, were presented as
cathode in the LIBs, which provided a specific capacity of 158.1 mA h g−1 at 0.1 C and a
good specific energy of 518.1 W h kg−1 and energy efficiency of 89.8% [26].Besides, a GO-
coated LiNi0.5Co0.2Mn0.3O3 cathode and a double-layer coated LiNi0.5Co0.2Mn0.3O3 cathode
using a V2O5 inner layer and a GO outer layer were proposed in LIB system [27,28]. The GO
layer not only retained the Li ion conductivity but also stabilised the interface between the
electrolyte and the electrode during cycles, thereby improving the battery performance.

Significantly, GO can be used as an active cathode material in the LIBs. A theoretical


calculation shows that Li ions can attach to GO either by forming bonds with oxygen-
containing groups that is prevalent at high and medium oxygen coverages or by forming LiC6
rings that is dominant at low coverage [29]. This suggests a new application of GO as a

6
cathode material in lithium storage. The epoxide-enriched GO, without being reduced, was
found as a sustainable carbonaceous cathode material for rechargeable lithium storage, which
delivered a high capacity of 360.5 mA h g−1 at 50 mA g−1 and a good cycling stability [30].
Its performance was significantly better than those of many polymer cathodes and
conventional Li-intercalated transition metal oxide cathodes. Additionally, the n-
Butyllithium-promoted carbonylated/hydroxylated GO was studied as a cathode material for
lithium storage, whose lithiation/delithiation process between Li+ and oxygen groups is
illustrated in Fig. 2(b) [31]. Because this chemically functionalised GO cathode was able to
easily contact with the electrolyte, promoting fast Faradaic reaction of the electroactive
functional groups with Li-ions and rapid charge transfer reaction, it showed an enhanced Li
storage properties in comparison to conventional LiCoO2 and LiFePO4 cathodes. The charge–
discharge curves of the functionalised GO electrode at various current densities are presented
in Fig. 2(c).

Fig. 2. (a) Schematic presenting the preparative steps in the formation of LiMnPO4, where
introduction of GO helped to confine the electrode nano-size [22]; (b) schematic illustration
of the Faradaic reactions on the n-Butyllithium-promoted carbonylated/hydroxylated GO
cathode [31]; (c) charge–discharge curves of the n-Butyllithium-promoted
carbonylated/hydroxylated GO electrode at various current densities [31].

GO has also been extensively used in the production of anode materials for LIBs.
Although the GO is reduced eventually in the most cases, the functions of the pristine GO
still cannot be negligible. For instance, various SnO2 anode structures were fabricated with
the assistance of GO in forms of nanoparticles, quantum dots, hollow spheres and nanopillars,

7
where GO provided interfacial interactions between oxygen groups and metal ions and/or
functioned as a template to form the advantageous architectures and/or prohibited the
aggregation of active materials during the cycling [32–35]. GO can also act as a reducing
agent during the heat treatment for the growth of the active anode materials, such as oxygen-
deficient Li4Ti5O12−x powders (a discharge capacity of 172.4 mA h g–1 at 0.5 C) [36],
exfoliated GO/FeO composites (a capacity of 857 mA h g–1 at 50 mA g–1) [37], and GeO2
thin film (a capacity of about 1200 mA h g–1 at 250 mA g–1) [38]. Besides, GO-coated
Li4Ti5O12 microspheres, Cu2O microspheres and CuO microspheres exhibited discharge
capacities of 160.1 mA h g–1 at 0.5 C, 440 mA h g–1 at 100 mA g–1 and 592 mA h g–1 at 100
mA g–1, respectively [39–41]. GO, acting as coating layers, not only played a confined effect
for the formation of effective 3D networks, but also reduced the diffusion paths for both Li
ions and electrons and inhibited the large volume change of anode materials during cycles.
Other GO-supported anode materials, e.g. nanostructured Fe3O4 [42], ultrafine Co3O4
nanocrystallites [43], Fe2O3 microspheres [44], GaN nanosheets [45], have also been
developed for high-performance LIBs, where GO served as a powerful template to facilitate,
guide and control the growth of active materials.

More importantly, GO can be directly used in the anodes without being reduced, as it
can provide additional Li ion storage through surface oxygen-containing functional groups
and nano-cavity and also provide LiC2 storage through the graphite layers and aromatic ring
occurring in the GO [46–48]. This endowed GO/Li4Ti5O12 composites, GO/graphite
composites and GO/graphite/CNTs composites with discharge capacities of 201, 690 and
1172.5 mA h g–1 at 0.5 C, respectively [46–48]. The outstanding performance of
GO/graphite/CNTs composites was because the graphite and CNTs efficiently improved the
electric conductivity of the electrodes, greatly enhancing the utilisation of the active materials
[48]. Furthermore, a sandwich composite of polyaniline/amorphous TiO2-GO/polyaniline
[49], GO/polydopamine (PDA)-coated Si nanocomposites [50], GO-wrapped Fe2O3
interleaved hybrid [51], and GO-wrapped porous NiCo2O4 microspheres [52] were
synthesised as the anodes for LIBs. Although, in these anode materials, GO was not used to
enhance the capacity via oxygen functional groups, it functioned as a flexible and
mechanically strong substrate to accommodate the volume variation and also improve the
mechanical integrity and hinder the aggregation of the electrodes during cycling. Moreover, a
covalently linked GO-ionic liquid nanocomposite, which grafted the ionic liquids on the GO
surface via the carboxyl-amino bond, was explored as the anode material for LIBs. The as-

8
prepared nanocomposite provided a good electrolyte-electrode interfacial interaction and a
high ionic conductivity, thus exhibiting an excellent reversible capacity of 903.9 mA h g–1
with a long cycle performance [53].

2.1.2. Contributions to interfaces


Continuing interest for the use of GO in LIBs is further related to its strong affinities to
polar components, capability to be reduced and repulse against negative charges. The
electrolyte-electrode or current collector interfaces can thus be modified with the assistance
of GO to improve the battery performance.

An example of enlarging the electrolyte-electrode interface involved the utilization of


GO dispersion to bridge nanocomposite electrolyte containing redox species with electrodes
[54]. As most LIBs generate energy by the Li+ intercalation/de-intercalation of the electrodes,
the energy capacity is always restricted by the electrode materials. A new strategy is to utilize
the energy stored in the redox species in the electrolyte as an additional energy supplier. To
extend the electrolyte-electrode interface from the electrode side only further to electrolyte
side, GO was first well dispersed into the redox electrolyte and then through casting and
evaporating, the mixture was coated on the electrode surface, as shown in Fig. 3(a) [54]. The
function of GO here is not only to anchor the redox-species or form a molecule-level
dispersion, but also to conduct electrically between the redox species and the electrode after
being electrochemically reduced during the process. The obtained capacity of LIBs with
GO/redox solid electrolyte was 3.5 times of that with conventional liquid electrolyte (Fig.
3(b)) [54].

9
Fig. 3. (a) Schematics of preparation of GO and redox electrolyte on the electrode surface
[54]; (b) discharging performances of the LIBs using 0.1 wt% GO and redox electrolyte,
redox electrolyte without the GO and conventional liquid electrolyte [54]; (c) discharge
capacity and Coulombic efficiencies at 1 C and (d) Rate capability of the LIBs using GO-
coated Li (SR-G-Li) and bare Li [55].

Furthermore, the feature that GO was able to be directly reduced by alkali metal at room
temperature was employed to build a dendrite-free Li metal anode with an interface
engineering. Specifically, a spontaneous GO reduction process is facilitated when GO is
homogeneously coated on the Li metal [55]. The as-obtained reduced GO layer on the Li
anode played a significant role to suppress the Li dendrite and stabilise the solid-state
electrolyte interface. The better cycling stability and rate capability of the LIB of LiFePO4
and GO-coated Li were observed than the counterpart using the bare Li, as shown in Fig. 3(c
and d) [55]. Besides, GO was used as a protective coating layer to prevent the corrosion of Al
current collector in LIBs [56]. The GO coating was able to (1) serve as a charge transport
barrier to pitting-causing PF6- due to the electrostatic repulsion; (2) prevent the native Al2O3

10
passive layer from rapidly degrading at high voltage conditions, and (3) reduce the contact
resistance due to the strong adhesion between the GO and the Al [56].

2.1.3. Contributions to solid-state electrolytes


Solid-state electrolyte with low flammability, dimensional stability and high safety is
expected to enhance the energy density, cycling life and safety of LIBs. Solid-state polymer
electrolytes have attracted the main attention due to their facile processability, flexibility,
lightweight and wide electrochemical stability. However, LIBs with these solid-state
electrolytes are still challenged by interface failures originated chemically, electrochemically,
electrically and mechanically.

GO has been explored as a filler material for various polymer matrices to develop the
solid-state electrolytes of LIBs. The incorporation of GO into the polymers, on one hand, can
enhance the mechanical stability of the polymer electrolytes [57,58]. On the other hand, GO
can improve the ionic conductivity of the electrolytes, because the interactions between the
functional groups of GO and Li ions can accelerate the dissociation of the lithium salt [57,58].
Also, the interface of GO and polymer matrix creates a new pathway for fast Li ion transport.
Based on these advantages, GO/polyethylene oxide (PEO) solid electrolyte was prepared by a
solution casting method [57,58]. The presence of GO inhibited the PEO crystallisation and
enhance the PEO crystal orientation, which ensured the polymer chain vertical to the film
surface [58]. GO also helped to guide the ion transport, which caused a highly anisotropic
ionic conductivity of the solid electrolyte membrane [58]. More significantly, a poly(ethylene
glycol)-grafted GO was reported as a filler material to functionalise a composite solid
polymer electrolyte using an organic/inorganic hybrid branched-graft copolymer for all-solid-
state LIB applications [59]. The ionic conductivity of the functionalised solid electrolyte was
2.1×10−4 S cm−1, which was one order of magnitude higher than that without adding
poly(ethylene glycol)-grafted GO fillers. This was attributed to the strong interactions
between the poly(ethylene glycol)-grafted GO and Li salts [59]. Also, the fillers improved the
thermal and mechanical stabilities of the composite solid electrolyte, whose performance was
maintained at the temperature of up to 150oC [59].

2.1.4. Contributions to 3D printing ink


GO was discovered as a promising ink for 3D-printed LIBs [60]. The GO with a high
concentration provided the prerequisite viscosity to bind the active nanomaterials together,
facilitating the 3D printing. The digital image of Fig. 4(a) exhibits the 3D-printed GO-based

11
Li4Ti5O12 anode and LiFePO4 cathode on a glass substrate [60]. Subsequently, freeze-drying
and thermal treatment were conducted to remove the water in the electrodes and to reduce the
GO materials. A gel polymer electrolyte was injected in the channels between the pretreated
electrodes (Fig. 4(b)) [60]. The 3D-printed full cell gave a capacity of about 100 mA h g–1 at
50 mA g–1 in 10 cycles with a nearly 100% Coulombic efficiency, as exhibited in Fig. 4(c).
The GO-based ink showed high viscosity and optimum viscoelastic properties, which enabled
the fabrication of 3D-printed batteries [60].

Fig. 4. (a) Image of 3D-printed GO-based anode and cathode for LIB; (b) schematic of 3D
printing of solid-state electrolyte; (c) charge and discharge behaviours of the 3D-printed LIB
[60].

2.2. Lithium sulphur batteries


Inexpensive, abundant, and nontoxic elemental sulphur (S) is a very promising cathode
material to construct a lithium-sulphur battery [61,62]. However, the poor electrical
conductivity of S or Li2S and the formation of soluble lithium polysulphides as intermediate
products and irreversible solid sulphides (Li2S2/Li2S) in liquid electrolyte during the
discharge-charge processes result in low utilisation of active materials, low coulombic
efficiency, and rapid capacity decay of the Li-S batteries [61,62]. Therefore, several strategies
have been developed to address these issues with respect to electrodes, separators and solid-
state-electrolytes. As a 2D sp2-hybridized carbonaceous material with high capacity and
specific surface area as well as abundant oxygen functional groups, Graphene oxides has
played a significant role in the construction of (1) cathodes with high S loading and restricted

12
dissolution, (2) polysulfide regulation interlayers with high Li+ ion conduction and (3)
dendrite-inhibitor for Li anodes.

2.2.1. Contributions to electrodes


GO has been used to accommodate and immobilise S to moderate Li polysulphide
dissolution and enhance the electrochemical stability of Li-S batteries. GO/S nanocomposite
can be prepared by an electrostatic self-assembly approach (Fig. 5(a)), in which the
negatively charged GO sheets are coated on the surface of positively charged S
microparticles [63–65], and the GO-S cathode for the Li-S cell exhibited a high reversible
capacity of 950 ― 1400 mA h g–1 and stable cycling performance [66]. This is because (1) the
incorporation of sulphur can partially reduce the GO, thereby increasing the conductivity of
the cathode; (2) the carbon rings and oxygen functional groups of GO form chemical bonds
with sulphur and thus stabilise S atoms/polysulfide anions on the GO surface and prevent the
diffusion of Li polysulphides [67]. The electrochemical performances of S and GO/S
electrodes are compared in Fig. 5(b) [64]. It is worth noting that although GO can enhance
the initial cycling performance of Li-S batteries, the unexpected strong chemical reactions
between Li+ and the oxygen groups of GO and immobilised S (that is, the formation and
accumulation of Li2CO3, Li2SO3, Li2SO4 and COSO2Li insulating layers) occur on the GO/S
cathode surface during the long-term cycling process (Fig. 5(c)) [68].

13
Fig. 5. (a) Scheme representation of the electrostatic self-assembly of GO-S composites [63];
(b) specific capacity at different current rates of S and GO-S electrodes [64]; (c) schematic
illustration of the discharge–charge process for the Li/GO–S battery [68].

Various GO-based S or Li2S electrodes have been developed for Li-S batteries. For
example, GO-wrapped hierarchical porous carbon–sulphur composite cathode in which the
sulphur was melt-infused into the porous carbon and GO sheets were then coated on the
surface via electrostatic interaction delivered a capacity fading rate of 0.12% per cycle up to
400 cycles [69]. A 3D cross-linked amylopectin-wrapped GO/S composite was constructed
for stabilising Li-S batteries, as the cross-linked structure confined the sulphur particles
effectively among the GO layers [70]. A GO coated hollow sulphur sphere structure with the
assistance of polyvinylpyrrolidone produced a 73% retention at 1 C after 150 cycles [71].
GO-wrapped bowl-like sulphur composite with void space inside was reported as the cathode
material for Li–S cells and the bowl-like S particles provided the void space to accommodate
volume expansion during the cycling process [72]. More GO-based composite cathode
materials including an MnO2-GO double-shell sulfur cathode [73], a GO-wrapped, multi-
cavity, walnut-like carbon sphere cathode [74], a cetyltrimethylammonium bromide-modified
sulfur-GO-CNT nanocomposite cathode [75], a microspherical sulfur/GO composite cathode
[76] and etc. have been proposed for the Li-S batteries for improving the rate capability and
cycling performance. GO played a synergistic effect for formation of the attractive cathode
structure, increase of active material loading, guarantee of structural integrity during cycling
and capture of polysulfide species.

Furthermore, a core−shell nanostructure comprising Li2S nanospheres with an


embedded GO sheet as a core material and a conformal carbon layer as a shell was proposed
for a high-rate and long-life Li-S cell, which delivered a very low capacity decay rate of only
0.046% per cycle with a high Coulombic efficiency of up to 99.7% for 1500 cycles at 2 C
[77]. The conformal carbon coating prohibits the polysulfide dissolution, improves the
electrical conductivity and provides void space to accommodate the volume expansion, and
the GO acts as a second inhibitor for polysulfide dissolution, and also the spherical shape of
the particles shortens the solid-state Li diffusion pathway [77].

Another interesting work deployed the oxidizing ability of GO to react with pollutant
H2S gas. This reaction simultaneously reduced GO and reformed the H2S gas to S, which is
homogeneously distributed on the reduced GO network. Compared to the graphene/S cathode

14
prepared by diffusing S melt into low-temperature exfoliated graphene, the in-situ reduced
GO/S hybrid show higher cycling stability and superior power performance [78].

2.2.2. Contributions to polysulfide regulating layers


GO surfaces are negatively charged and the oxygen functional groups can bind with Li
cations that facilitate the cation hopping. In the meantime, the negatively charged surfaces
repel the polysulfide anions, endowing the GO with a cation-selective feature. These are the
basic properties that support the use of GO as interlayers in Li-S cells. GO can be applied as
an electroactive interlayer between cathode and separator in a Li-S cell, as illustrated in Fig.
6(a), which is found to improve both cathode capacity owing to a higher utilisation efficiency
of polysulphides and the extra lithium storage [79]. The extra capacity was resulted from the
reversible lithium uptake by the oxygen functional groups in the GO interlayers and the
irreversible formation of lithium hydroxide [79]. Similarly, a unique Li-S battery
configuration with an ultrathin permselective GO membrane between the sulfur cathode and
the separator was also presented, which afforded a reduced cyclic capacity decay of
0.23%/cycle [80]. Additionally, the electrochemical performance of Li-S batteries can be
enhanced by simply coating a thin GO-based barrier layer on the separator. It was reported
that GO, which was coated on the cathode side of the separator through a tape casting method,
functioned as an effective polysulfide shuttle inhibitor for improving the Li-S battery [81].
The performances of the batteries with the GO-coated separator and with pristine separator
are compared in Fig. 6(b and c) [81]. The performance enhancement demonstrated that the
GO effectively prevented the polysulfulfide cross-transport, thus maximising the active
material utilisation [81].

15
Fig. 6. (a) Schematic illustrating a cell configuration with GO interlayer-cathode structure
[79]; charge-discharge profiles of (b) the cell with pristine separator and (c) the cell with GO-
coated separator [81].

Moreover, GO mixed with other materials has been used to functionalise the separator
or as an interlayer in the Li-S batteries. A tri-layered separator based on a macroporous
polypropylene (PP) supporting layer, a GO barrier layer and a Nafion retarding layer was
designed for the Li-S batteries [82]. It was also reported that the synergistic effects of the
metal-organic framework (MOF) particles and GO laminates as the separator helped to
improve the cycling stability of the Li-S batteries [83]. The MOF materials with a large
surface area and highly ordered micropores served as an effective ionic sieves to inhibit the
polysulfide ions shuttle and the GO laminates with the mechanically flexible property acted
as a robust substrate to support the functional membrane. Besides, a highly porous
polyacrylonitrile/GO (PAN/GO) nanofiber membrane separator was synthesised for high-
performance Li-S system, taking advantage of the relatively high energy binding between the
−C ≡ N groups of PAN and the polysulfide ions as well as the electrostatic repulsion between
the negatively charge surface of GO and the Sn2− species [84]. More significantly, a new
strategy via coating a highly Li ion conductive lithium fluoride/GO solid electrolyte on the
separator was presented recently, which efficiently alleviated the polysulfide shuttle
behaviour and meantime allowed Li ion transport [85]. The outstanding performance was

16
achieved with a 1205 mA h g–1 discharge capacity at 0.2 C and a 0.043% capacity decay rate
per cycle over 400 cycles. Regarding the polysulfide-trapping interlayer, a GO/CNT
composite interlayer and an ultrathin interlayer by loading MnO2 nanoparticles and GO on
CNT films were proposed to effectively suppress polysulfide cross-transport [86,87].

2.2.3. Contributions to solid electrolyte interphase


It is well known that dendrite growth on the Li metal can be effectively hindered by
fabrication of a stable solid electrolyte interphase. GO, due to its 2D structure, high Li ion
conductivity and good mechanical strength, was used to suppress the formation of lithium
dendrite, which was evenly coated on metallic lithium foil through an automatic spreading
method using DMC as the solvent [88]. The Li-S battery using GO-coated Li anode exhibited
a smaller charge transfer resistance and a higher Li ion conductivity after the initial charge-
discharge process. This was because the GO protective layer suppressed the side reactions on
the Li anode surface that inhibited the transport of electrons and Li ions. Its improved
performance with better cyclic stability was demonstrated in Fig. 7 [88].

Fig. 7. (a) Charge-discharge profiles and (b) cycling performance for 200 cycles for Li-S
cells using the GO-protected anode and the pristine anode [88].

2.3. Lithium air batteries


Lithium air batteries have attracted worldwide attention, because of the ultrahigh
theoretical energy density of ~11,000 W h kg−1 [89,90]. A typical Li-air battery is composed
of a metallic Li anode, an organic electrolyte and a porous air-breathing cathode. However,
the cathode materials of Li-air cells have met a huge challenge, which includes transport of
oxygen through the pores and the deposition of insulating products on active sites for oxygen
reduction and evolution [90]. GO shows some potential in the Li-air batteries. A free-
standing GO paper (Fig. 8(a)) was used as a O2-breathing cathode without binder or other

17
additives [91]. The Li-air battery using the GO cathode was assembled according to Fig. 8(b)
and its electrochemical performances are shown in Fig. 8(c and d). It exhibited a 612 mA h g–
1 discharge capacity and a 585 mA h g–1 charge capacity after 10 cycles. The unique structure
of the GO paper provided various porosity for O2 diffusion, thus improving the efficiency of
the formation and decomposition of Li2O or Li2O2. Moreover, a free-standing, hierarchically
porous carbon derived from GO gel in nickel foam without any additional binder was
produced by a sol-gel method, wherein the GO not only served as a special carbon source, but
also provided the 3D framework [92]. The specific capacity of the Li-air cell using the GO-
derived porous carbon cathode reached as high as 11,060 mA h g–1 at a current density of 280
mA g–1, as the hierarchically porous structure promoted a continuous oxygen flow in the O2-
breathing electrode as well as provided sufficient void space for Li2O2 deposition [92].

Fig. 8. (a) SEM images of GO paper; (b) schematic illustration of the Li-air battery assembles
with the GO cathode; (c) charge-discharge curves of GO paper in the Li-air cell; (d) Cyclic
performance of the GO paper until tenth cycle in the Li-air cell [91].

18
2.4. Redox flow batteries

2.4.1. Contributions to electrocatalytic electrodes


Redox flow batteries (RFBs) represent a class of large-scale energy storage techniques,
which consists of positive and negative electrodes immersing in soluble electroactive
electrolytes, separated by an ion exchange membrane [93–95]. The vanadium RFB system,
which has been commercialised in many countries, applies VO2+/VO2+ and V2+/V3+ redox
couples in sulphuric acid as the positive and the negative half-cell electrolytes, respectively
[96]. GO nanoplatelets were directly used as electrochemical active materials for VO2+/VO2+
and V2+/V3+ redox couples [97–102]. The experimental results showed that although the
reversibility towards the positive VO2+/VO2+ was not improved, the addition of GO was able
to significantly enhance the reversibility between V2+ and V3+ and increase the peak current
for both VO2+/VO2+ and V2+/V3+ redox reactions, compared with graphite [96]. This was
because the relatively large active surface area and abundant oxygen groups allowed the GO
to play an electrocatalytic function during the redox reaction of vanadium species, that is, to
facilitate the adsorption of the reactive ions onto the electrode surface. However, electrodes
using the pure GO exhibited a low rate capability due to the inferior electrical conductivity,
thus delivering a limited energy efficiency and power capability [103]. Consequently, the GO
sheets were assembled onto the multiwalled CNT surfaces to form an electrocatalytic hybrid
with an effective mixed conducting network. The hybrid not only provided a rapid electron
and ion transport capability but also delivered a much better electrocatalytic redox
reversibility of VO2+/VO2+ couple, especially for the reduction reaction of VO2+ to VO2+
[103]. Moreover, phosphonated GO showed an enhanced electrocatalytic performance to
redox reactions of vanadium ions and when it decorated on carbon felt electrodes, an
improved performance with stable efficiencies of the vanadium RFB was obtained [104].

In addition to the all-vanadium RFBs, application of GO as the electrocatalytic


electrode material in other RFBs was also reported. For example, the GO-based electrodes
mixing with polymer binder in a vanadium bromide RFB presented a better electrochemical
performance than the pristine graphite, including lower overvoltage for both Br−/Br3− and
V3+/V2+ redox couple reactions and higher peak currents for the V3+/V2+ redox couple [105].
The study also demonstrated that the oxygen-containing groups on the GO sheets generated
ample active sites to electrochemically catalyse the redox reactions. Moreover, a nanoporous
GO electrode was investigated as an active anode material in cerium vanadium RFB and the
nanopores containing oxygen functional groups was found to provide a remarkable

19
electrocatalytic activity towards Ce4+/Ce3+ redox couples [106]. A zinc cerium RFB using a
graphite/GO composite positive electrode also exhibited a significantly improved
electrochemical performance [107]. Because of the presence of oxygen groups and large
surface area introduced by the GO, the composite electrode possessed a much larger redox
peak current of Ce3+/Ce4+ and a reduced charge transfer resistance.

2.4.2. Contributions to membranes


One of the key components of the RFBs is the ion exchange membrane, which is
responsible for ion transport and preventing the cross-mixing of the positive and negative
electrolytes that are stored in separate tanks [93,108]. Importantly, the incorporation of GO in
a polymer membrane is determined to increase the proton conductivity and mechanical
strength [109]. When GO is added in the Nafion membrane for the vanadium RFBs, the GO
can act as effective barriers to inhibit the transport of vanadium ion, thus significantly
decreasing the vanadium ion permeability [110–112]. For instance, GO-incorporated recasted
Nafion membrane (rNafion/GO) showed the higher ion selectivity and lower vanadium
permeation than rNafion and rNafion/graphene nanoplate (rNafion/GNP), whose
performances were compared in Fig. 9(a) [113].

Furthermore, the loading of GO into the sulfonated poly(ether ether ketone) (SPEEK)
membrane is found to improve the membrane performance, which are due to the following
reasons [108]. (1) GO sheets can prevent the transport of vanadium ions because of the
increase in tortuosity; (2) the oxygen functional groups of GO can form hydrogen bonds with
polymer chains, which is beneficial for the formation of the hydrophobic/hydrophilic
separation structure in the acquired composite membranes, and hence the enhancement of ion
selectivity; (3) the strong interaction between GO and the polymer can improve the
mechanical stability of the membrane; (4) GO itself possesses a high proton conductivity.
Likewise, sulfonated GO (SGO)-doped SPEEK composite membrane and p-phenylene
diamine-functionalized GO (PPD-GO)/SPEEK hybrid membrane were also reported for the
vanadium RFBs and their improved battery performances are demonstrated in Fig. 9 (b and c)
[114,115]. Regarding the polyvinylpyrrolidone (PSF-PVP) membrane for the vanadium
RFBs, although the addition of GO decreased the proton conductivity, the membrane with
0.05 wt% GO loading is endowed with optimal ion selective capability (see Fig. 9(d)), which
was achieved by size sieving effect, tortuosity in migration paths and electrochemical
exclusive behaviours on vanadium ions [116].

20
Fig. 9. (a) Proton conductivity, VO2+ ion permeability and ion selectivity of the membranes
[113]; charge-discharge profiles of vanadium RFBs using (b) sulfonated GO-doped SPEEK
composite membrane [114] and (c) p-phenylene diamine-functionalized GO/SPEEK hybrid
membrane [115]; (d) ion selectivity of the PSF-PVP membranes with different amounts of
GO loading [116].

2.5. Other batteries


Sodium-ion batteries (SIBs) have emerged as an attractive alternative for electrochemical
energy storage applications [117–119]. Noteworthy is that GO makes a nonnegligible
contribution to developing the SIB. Carbonyl group-based croconic acid disodium salt as an
anode material in the SIB presented a high capacity of 246.7 mA h g–1 but with a fast
capacity decay, which was ascribed to the particle pulverisation, triggered by a serial phase
transformation during charge/discharge process [120]. The GO-wrapped croconic acid
disodium salt particles with smaller size provided a higher capacity of 293 mA h g–1 and
improved cycling stability, as the pulverisation was effectively suppressed by minimising the
particle size. A sandwiched composite based on GO and Fe2O3 nanoparticles was also
presented as an anode in SIBs, which utilised the flexibility of GO as a buffer to
accommodate the large volume expansion and prevent the aggregation of active materials
[121]. The enhanced Na storage capacity, cyclic stability and rate capability were obtained, as
exhibited in Fig. 10 (a and b). Furthermore, although hard carbon is an attractive anode
candidate material in SIBs, it suffers a low first-cycle Coulombic efficiency and fast capacity
fading [122]. As the first-cycle Coulombic efficiency for carbon anodes can be increased by

21
lowering the surface area, an effective approach to reduce the surface area of sucrose-derived
hard carbon by introducing GO dispersion into the sucrose solution was reported in the SIB
system [122]. The GO-assisted hard carbon possessed a lower specific surface area of 5.4 m2
g–1 and an improved first-cycle Coulombic efficiency from 74% to 83%.
Various functionalised GO has also exclusively designed for SIBs. For example, an
innovative porphyrin interspersed GO framework, as schematically displayed in Fig. 10 (c),
contained considerable cavities and free space at the interlayers, which was favourable for Na
ion insertion and extraction upon charging and discharging [123]. Besides, a simple and
scalable process by using alkali metal ions to functionalise GO was reported in the SIB
system [124]. It was found that a reaction between the oxygen atom of GO and alkali metal
hydroxides was able to affect the porosity, morphology, disorder degree, specific surface area
and electrical conductivity , thereby crucially improving the sodium ion storage capability of
the functionalised GO [124].

Fig. 10. (a) Cycling performance and (b) rate performance of the Fe2O3/GO composite and
bare Fe2O3 anode [121]; (c) schematic illustration of formation of porphyrin interspersed GO
framework [123].

Moreover, as GO has a high proton conductivity and sulphuric acid affinity, the GO
paper was used an electrolyte substitute for sulphuric acid to fabricate a solid-state lead acid
battery [125]. Fig. 11(a) illustrates the models of the overall electrochemical reactions
occurring in the small-sized lead acid cell. The cycle performance of GO lead battery is

22
shown in Fig. 11(b) [125]. The capacity (discharging time) was observed to increase initially,
but then it decreased after approximately four cycles. The initial increase was due to the
formation of the electrochemical active species and the subsequent decrease was due to their
change to nonactive forms after four cycles [125]. However, the performance could be
recovered by adding small amounts of a H2SO4 solution. Besides, the addition of lignin into
the active mass in the negative electrode brought about an automatic recovery after a similar
nonactivation at around four cycles. In summary, the lead acid battery using the GO
membrane, as a dry cell, had no risk of electrolyte leakage and delivered good performance at
the initial cycles, whereas further improvements were necessary for practical use [125].

Fig. 11. (a) Illustration of the models of the total reactions in the GO lead battery [125]; (b)
Cycle performance of GO lead battery (Open and closed symbols correspond to GO lead
battery and conventional lead-acid battery respectively. The arrow indicates the addition of 6
M H2SO4 into both Pb paste active masses. Open square symbol is the case of the lignin
added cell.) [125]; (c) Schematic illustration of metal/GO battery [126]; (d) Schematic
illustration of GO in contact with the metal [126].

An innovative design of metal/GO batteries, in light of the spontaneous redox reaction


between the metals and the GO, was proposed.[126] As displayed in Fig. 11(c), the metals,
including Li, Na, Zn, Fe and Cu, acted as anode and the GO functioned as both cathode and
separator. When the GO contacted the metal, the reduction of GO happened along with the
formation of metal oxides (see Fig. 11(d)) [126]. A specific capacity of about 1604 mA h g−1
was observed for the Li/GO battery. And the performance of the Cu/GO battery was

23
improved by using 3D Cu foam with flowing GO/ionic-liquid catholyte, because the
increased contact area achieved enhancement of energy density [126]. A Zn/3D-GO battery,
characterised by compressibility, all solid state and pressure-response, was also constructed,
which allowed for precise control of the energy output responding to simulated pressures in
the absence of traditional battery management systems [126]. Cycling performance of such
metal/GO batteries, however, still required to be further developed.

In addition, the oxygen functional groups of GO show great affinity to metal oxides and
the negative charge when GO is dispersed in water can attract positively charged metal
cations by electrostatic force. These features enable GO as burgeoning substrate to disperse
and stabilized metal or metal oxides nanoparticles which present novel catalytic effects
towards versatile reactions. Moreover, the large surface area of GO can also provide enough
space for the decomposition of the intermediate products during these reactions. In fact, many
emerging energy systems such as Na-S and Li/Na-CO2 batteries suffer from severe capacity
fading issues due to the intermediate products accumulation during discharge (e.g. Na2S4,
Li2CO3) [127–132]. GO-supported catalysts were proposed to incorporate into the electrode
preparation to improve the battery capacity retention. For example, by using GO supported-
Co single atoms in the cathode of Li-CO2 battery, the discharge capacity show no obvious
fading over 100 cycles at 100mAg-1 and this was attributed to the storage space offered by
GO and the efficient LiCO3 decomposition catalytic effect of the Co atoms anchored by GO
[132].

2.6. Brief summary

As exhibited in Fig. 12, the contributions of GO in batteries can be divided into four
main parts, including electrodes, interlayers, solid-state electrolytes and membranes. In the
field of the electrodes, GO can work as the substrate to support, guide and control the growth
of active materials, which endows the electrodes with various attractive nano-sized structures.
In Li-S batteries, GO, coated on the S particles, can accommodate and immobilise S to
moderate Li polysulphide dissolution and enhance the electrochemical stability. GO with
large active surface area and abundant oxygen groups also plays an electrocatalytic role for
RFBs to facilitate the redox reactions. In Li-air batteries, GO can serve as the O2-breathing
cathode, due to the various porous in GO for efficient O2 diffusion. As for the interlayers,
first, GO mixing with redox species on the electrode surface can be considered as additional
electrode-electrolyte solid interface, thus increasing the utilisation of the redox species so as

24
to improving battery performance. Second, GO interlayer between the electrodes and the
membranes in Li-S batteries can inhibit the polysulfide transport and meantime allow the Li
ion diffusion. Third, GO, when coating on the Li metal anode, can suppress the formation of
lithium dendrites as well as when coating on the current collector, can prevent the corrosion
of Al. In terms of solid-state electrolytes, GO can be used as the filler material for various
polymer matrices to enhance the ionic conductivity and the thermal and mechanical stabilities
of the solid-state electrolytes. Regarding the membranes, the addition of GO into the
polymer-based membrane can not only increase the ionic conductivity and mechanical
strength, but also improve ion selective capability.

Fig. 12. Schematic illustration of applications of GO in batteries.

3. Application of graphene oxide in capacitors


3.1. Dielectric Capacitors

Dielectric capacitors are the traditional electrical capacitors that store energy through
potential field-induced polarity of dielectric medium [133]. A dielectric capacitor is
composed of two conducting electrodes separated by an insulating dielectric spacer [134].
Such dielectric medium is required to have a high dielectric constant and low dielectric loss
for the maximum capacitance. It was presented that a GO-encapsulated CNT hybrid,
synthesised by π-π interactions between negatively charged GO and positively charged CNTs,
possessed not only a high dielectric constant and low dielectric loss, but also a highly
enhanced breakdown strength and maximum energy storage density [135]. The GO shells (1)
effectively enhanced the dispersion of CNTs; (2) served as insulation barriers to decrease

25
dielectric loss; and (3) increased the charge scattering due to their ultrathin structure and
insulating property, resulting in an increase of breakdown strength [135,136]. More GO-
based polymer-functionalised nanocomposite materials were fabricated with high dielectric
constant and low dielectric loss, such as polyvinyl alcohol (PVA)/polyethylene glycol
(PEG)/GO composite [137], poly (4-styrenesulfonic acid) (PSSA)/PVA/GO composite [138],
polydimethylsiloxane (PDMS)/poly(glycidyl methacrylate) (PGMA)/GO composite paper
[139]. They all showed great potential as dielectric materials for dielectric capacitors.

A free-standing hydrated GO film was used as a dielectric spacer between two stainless
steel electrodes to construct a prototype water-dielectric capacitor, as show in Fig. 13 (a and b)
[134]. When water was confined between GO sheets by hydrogen bonding, it retained its
insulating nature and behaved as a dielectric. The dielectric capacitance of hydrated GO films
not only depended on the water content but also was sensitive to the voltage applied. Fig. 13c
shows the gradually narrowed dielectric capacitive current of GO as the annealing
temperature increased from 100 °C to 120 °C. This was because the significant loss of water
and the decomposition of oxygen functional groups caused weakened dielectric behaviour
and improved electrical conductivity of the GO [134,140]. Fig. 13d displays the cyclic
voltammograms (CV) of the water-dielectric capacitor in which the hydrated GO film was
used as the dielectric spacer within different voltage windows [134]. It was observed that the
capacitance of the hydrated GO film changed as the voltage window changed. However,
when the voltage is higher than ± 2 V, the sharp capacitance increase indicated the dramatic
increment of leakage current through the hydrated GO film, which was a signature of the
reduced resistivity. This was attributed to the voltage-induced dehydration and the voltage-
induced deoxygenation (reduction) [134].

26
Fig. 13. (a) Cross sectional SEM image of a hydrated GO film with the inset showing the
hydrogen bonding between water molecules and GO sheets; (b) scheme illustrating the
assembly of a water-dielectric capacitor with a hydrated GO film as a dielectric spacer; (c)
cyclic voltammetry of the hydrated GO film and films annealed at 100, 110 and 120 °C; (d)
relationship between dielectric capacitance and voltage of the water-dielectric capacitor. (The
dashed rectangular curves are CV profiles for 1: ideal dielectric capacitor and 2: dielectric
capacitor with resistance) [134]

3.2. Supercapacitors
Supercapacitors (SCs), also called electrochemical capacitors, are the high-performance
energy storage devices with excellent power capability, short charge-discharge time, and long
cyclic life [141]. Charge storage in SCs is principally based on either the electrostatic charge
accumulation at the electrode-electrolyte interface, i.e. electrical double layer capacitance, or
the fast and reversible Faradaic processes on the electrode surface, i.e. pseudo-capacitance
[141,142]. GO, in virtue of its unique physicochemical properties, has been applied to
enhance the performances of SCs. The use of GO as both electrodes and solid-state
electrolytes in the SCs will be discussed below.

3.2.1. Contributions to electrodes


GO possesses abundant oxygen functional groups that can provide a large additional
pseudo-capacitance and improved wettability of the electrodes in electrolytes [143]. As
shown in Fig. 14(a), GO-based electrode exhibits a larger capacitance than graphene (both

27
GO and graphene electrodes were prepared by pressing a mixture of 87 wt% of sample, 10 wt%
of acetylene black and 3 wt% of binder into pellets). The pseudo-capacitance of GO possibly
originates from electrochemical reactions, e.g. >C–OH ↔ >C=O+H+ + e− etc., at the
electrode/electrolyte interfaces. The GO electrode delivered a maximum specific capacitance
of 146 F g−1 and an energy density of 20.39 W h kg−1 [144]. Other studies, e.g. an asymmetric
SC using LiCoO2 as a positive electrode and GO as a negative electrode and a symmetric SC
using the porous GO on conductive stainless steel, were also conducted, which showed the
great potential of GO without any additives as the SC electrodes [145,146].

Nonetheless, the inferior electrical conductivity of GO, to some extent, limits the
electrode performances. Various GO-based composites incorporating conventional electrode
materials of SCs are studied with a purpose to take advantage of the synergistic effect of GO
with other electrode components. Electrode materials of the SCs can mainly be divided into
three categories: carbonaceous materials, transition metal oxides/nitrides/sulphides, and
conducting polymers [147]. GO, on one hand, plays a significant role in the formation of
porous electrode structures, despite subsequent annealing treatment [148–150]. The porous
architecture benefits to increase the specific surface area with more active sites. On the other
hand, GO can function as one of active materials in SCs. For instance, cooperating with
carbon materials, the GO and single-walled CNT composite ink electrode delivered a highest
specific capacitance of 295 F g−1 at a current density of 0.5 A g−1 and a capacitance retention
of 85% after 60,000 cycles [151]. A core–shell heterostructure with multi-walled CNTs as the
core and GO nanoribbons as the shell was fabricated as a SC electrode material in which GO
shell provided more active sites for electron accumulation and the inner CNT core
contributed highly conductive paths for electron transport [152]. Moreover, a flexible solid-
state symmetric SC including the GO-rGO (reduced GO) patterned electrodes showed a
capacitance of 141.2 F g−1 at 1 A g−1 and a good stability over 2000 cycles [153]. An
asymmetric SC using a GO/cobalt(II) tetrapyrazinoporphyrazine composite as the positive
electrode and GO/carbon black as the negative electrode in a Na2SO4 electrolyte provided
good specific capacitance, energy and power densities of ~500 F g−1, 44 Wh kg−1 and 31 kW
kg−1, respectively [154].

28
Fig. 14. (a) The electrochemical capacitive behaviour of GO and graphene in cyclic
voltammetry and charge–discharge curves[143]; (b) illustration of GO/MnO2 hybrid
formation [155]; (c) schematic for the synthesis process of Co-Al/GO composite [156].

With respect to transition metal oxide electrodes, GO can serve as a template to layered
MnO2 nanosheets for increasing the surface area [157] and also be incorporated into the
MnO2 electrode for enhancing the capacitance [155,158–160]. The oxygen groups of GO,
functioning as anchor sites, enable the in-situ formation of MnO2 nanoparticles attaching on
the surfaces and edges of GO sheets (Fig. 14(b)), which not only enlarge the specific surface
area but also prevent the aggregation of GO and MnO2 [155]. Likewise, the GO/Mn3O4
composite thin film prepared by a layer-by-layer technique exhibited a specific capacitance of
344 F g−1 at 5 mV s−1 [161]. Additionally, by means of an electrophoretic deposition or
hydrothermal process or electrostatic co-precipitation, nickel oxide, nickel sulphide and
copper oxide could be attached on the surfaces of GO nanosheets for high-performance SCs,

29
which achieved the specific capacitances of 569 F g−1 at 5 A g−1 [162], 800 F g−1 at 1 A g−
1[163], 245 F g−1 at 0.1 A g−1 [164], respectively. The interaction of the GO and the metal
oxides/sulphides was believed to play a synergistic effect on the performance of SCs. It was
also reported that both NiCo2O4-based nanowires and ultrafine nanoparticles were deposited
on the surface of GO substrate under microwave irradiation and the presence of GO
significantly enhanced the retention at even higher current densities [165]. Besides, GO can
promote the nucleation and in-situ growth of Co-Al layered double hydroxide nanosheets. Fig.
14(c) illustrates the synthesis process of the Co-Al/GO composite during which positively
charged Co–Al hydroxides and negatively charged GO sheets are self-assembled via
electrostatic interactions [156,166]. GO inhibited the aggregation of layered hydroxide
nanosheets and the layer-by-layer assembly optimised the contact area of GO and Co-Al
sheets, both of which were favourable for effective electron transport between active
materials and the current collector [156].

Electroactive polymers, though they can provide a large pseudo-capacitance, exhibit a


swelling behaviour during the redox processes that degrades the cycling stability [167]. In
order to overcome this drawback, GO is widely used to functionalise the conducting
polymers for high-performance SCs. The preparation of conducting polymer–GO sheets can
normally be achieved by an electrostatic interactions between negatively charged GO sheets
and positively charged surfactant micelles [168]. When the surfactant micelles are
electrostatically adsorbed on the GO surfaces, the addition of polymer monomer enables the
monomer to be predominately solubilised in the hydrophobic cores of the surfactant micelles
and subsequently the polymerisation can undergo predominately in the micelle cores. For
instance, the GO/polyaniline (PANI) nanocomposite was synthesised by in situ
polymerisation [169–172]. Combining GO and PANI composite is possibly based on three
modes: (1) π-π stacking, (2) electrostatic interactions, and (3) hydrogen bonding [173]. The
GO/PANI hybrids possessed higher surface area and larger pore volume compared to the
pure PANI, which increased the contact area of PANI nanoparticles with the electrolyte,
facilitating the fast transport of electrolyte ions and electrons into the active sites of the
composite electrode [170,174,175]. Besides, the oxygen functional groups on GO not only
enable easy ion transportation but also produce additional pseudo-capacitance [174,176].
These features endowed the GO/PANI electrodes with an increased capacitance [175–177].
The synergistic effect of GO and PANI also improve the cycling stability, as GO sheets

30
endure some mechanical deformation during the Faradaic process of PANI, preventing the
destruction of the electrode material [175].

Similarly, GO/polypyrrole (PPy) nanocomposites can be produced via either an in situ


chemical polymerisation method [178–182] or an electrochemical co-deposition method
[183,184]. The GO/PPy composite electrodes show a high specific capacitance, good rate
capability and cycling stability due to their synergistic effects, e.g., a specific capacitance of
633 F g−1 at 1 A g−1 and a 6% capacitance attenuation after 100 cycles were reported, in
contrast to the pure PPy with 227 F g−1 and 68% retention [185]. Moreover, poly (3,4-
ethylenedioxythipohene) (PEDOT), as one of the conducting polymers, has also been
explored with GO in SCs. For example, the PEDOT/GO hybrid electrodes delivered a
specific capacitance of 115.15 F g−1 at 0.3 A g−1 and retained 86% initial capacitance after
1000 cycles at 100 mV s−1 [186]. Furthermore, other polymers, such as poly [(thiophene-2,5-
diyl)-co-(benzylidene)] (PTCB) and poly(o-methoxyaniline) (POMA), were proposed to
decorate the functionalised GO (sulfur-doped GO and amine-grafted GO, respectively) for
high-performance SCs [187,188].

The superior electrochemical performances of the GO/conducting polymer composites


can be attributed to the following points: (1) GO provides a large accessible surface for the
conducting polymers; (2) the interface binding in the 3D or layered structures enhances the
mechanical strength and stabilises the polymers during the cycling process; (3) the
GO/polymer nanostructure effectively reduces the dynamic resistance of electrolyte ions; and
(4) the oxygen functional groups of the GO sheets contribute a large additional pseudo-
capacitance to the overall energy storage [168].

Furthermore, multi-component electrode materials, which normally comprise (1) GO


with metal oxides and polymers [189–192], (2) GO with metal oxides and carbon materials
[193,194], (3) GO with polymers and carbon materials [195–199] and (4) GO with metal
oxides, polymers and carbon materials [200–202], have attracted considerable attention in
terms of improving the performance of the SCs. The high capacitance and good cycling
stability of the multi-component electrodes can be summarised in the following reasons. First,
excellent water solubility and surface activity of the highly oxidised GO, which acts as a
surfactant to directly disperse other active materials thus simplifying the process [193].
Second, the oxygen-containing functional groups enriched on the GO that can enhance the
total capacitance through additional faradaic reactions (pseudocapacitance effects) in addition

31
to improving the wettability of the electrodes with electrolytes for fast ion transport [193].
Third, the conducting polymer or carbon materials, which preserve the electrical conductivity
of the electrodes, and prevent the GO sheets from stacking/aggregating, and facilitate metal
oxide particles anchor onto the surface of GO [189]. Fourth, the formation of 3D or porous
structure with the aids of GO and/or other materials, which increases the specific surface area
of the composite electrodes [195].

3.2.2. Contributions to solid-state electrolytes


It is generally acknowledged that GO possesses a superior proton conductivity, due to the
presence of the hydrophilic sites, i.e., −O−, −OH, and −COOH functional groups, which
attract the protons and propagate through hydrogen-bonding networks along the water
intercalate layer [203,204]. The high proton conductivity allows GO to be used as a solid-
state electrolyte in the SCs. A thin, flexible, all-solid-state all-GO SC was hence developed,
where both surfaces of a GO film were treated by photoirradiation to form the rGO along
with retaining a GO interlayer in the middle [205]. Such rGO/GO/rGO device took advantage
of the remarkable ionic conductivity and electric insulating property of the GO as both solid-
state electrolyte and separator.

An all-solid-state SC using the crosslinked PVA/graphene hydrogel electrodes and a GO-


doped ion gel as a gel polymer electrolyte and separator was reported, which delivered a
specific capacitance of 190 F g−1 and energy density of 76 W h Kg−1 at 1 A g−1 [206]. The gel
membrane was prepared by mixing copolymer and ionic liquid with GO and the function of
GO is explained in Fig. 15(a) [206,207]. Without adding the GO, the ions can only
transported through the disordered amorphous pathway of the gel electrolyte (Fig. 15(a),
1).[207] The addition of GO decreased the degree of crystalline in the ion gel, as the
abundant oxygen-containing groups on the GO sheets interconnected with the copolymer to
form a “highway” for ion transport, which resulted in higher ionic conductivity, as illustrated
in Fig. 15(a), 2. However, the more addition of GO was expected to cause the restacking of
GO sheets and the blocking effect by the excessive GO as shown in panel 3 in Fig. 15(a), thus
deteriorating the ionic conductivity. Furthermore, a new family of boron cross-linked
GO/PVA (GO-B-PVA) nanocomposite gel electrolyte was synthesised through a freeze-
thaw/boron cross-linking method and then assembled into an electric double layer capacitor,
as presented in Fig. 15(b) [208]. GO functioned as a promoter for the ionic conductivity and
the boron atoms played a role as physical or chemical cross-linkers of the PVA matrix to
improve the thermal and mechanical stabilities of GO-PVA hybrid. As a result, the SC using

32
20 wt% GO-B-PVA/KOH gel electrolyte exhibited a stable cycling performance up to 1000
cycles and a higher discharge capacitance of 190 F g−1 at 0.1 A g−1 compared to that using the
aqueous KOH electrolyte (109.4 F g−1).

Fig. 15. (a) Proposed schematic structures of GO-doped ion gels with different GO doping
mass fraction: (1) pure ion gel; (2) ion gel with optimised GO doping; and 3) ion gel with
excess GO doping [207]; (b) schematic for assembly of GO-B-PVA nanocomposite gel
electrolyte [208].

3.3. Brief summary

As illustrated in Fig. 16, due to the insulating nature, GO can behave as a dielectric
spacer between two conductive plates for dielectric capacitors. As for the SCs, GO, on one
hand, can function as one of active electrode materials to improve the capacitance, because its

33
abundant oxygen functional groups can provide additional pseudo-capacitance. On the other
hand, GO with other conducting materials, e.g. carbonaceous materials, transition metal
oxides/nitrides/sulphides, and conducting polymers, play a synergistic effect to electrodes,
which include increasing the surface area, providing anchor sites for formation of nano-sized
composite electrodes, inhibiting the aggregation of active materials during cycling and etc.
Additionally, the hydrophilic functional groups endow GO with a highly ionic conducting
property, which facilitates its use as the solid-state electrolyte for all-solid-state SCs.

Fig. 16. Schematic illustration of applications of GO in capacitors.

4. Application of graphene oxide in fuel cells


A fuel cell is an electrochemical conversion device that generates electricity via chemical
redox reactions when supplied with a fuel, e.g., hydrogen, natural gas, or methanol, and an
oxidant, e.g., oxygen, air, or hydrogen peroxide [209–213]. Noteworthy is that GO can make
great contribution to the enhancement of electrocatalytic activities and the fabrication of the
membranes in the fuel cells, e.g. proton exchange membrane fuel cell (PEMFC) and direct
methanol/ethanol fuel cell (DMFC/DEFC), which are discussed below.

4.1. Contributions to electrocatalysts


Platinum (Pt) has been widely used as an effective electrocatalyst in fuel cell technology
for the electro-oxidation of fuel at the anode and electro-reduction of oxygen at the cathode
[214]. Pt catalyst is normally supported on the high surface area carbon material such as
carbon black. However, the degradation of the carbon black occurs at the potential of oxygen
reduction reaction, which hence causes the agglomeration and loss of Pt nanoparticles,
decreasing the electrochemical surface area and the electrocatalytic activity. Importantly, the
addition of small amount of GO greatly increases the durability of commercial Pt/C catalyst

34
without sacrificing initial electrochemical active surface area [214]. This is because the GO
can provide the anchoring sites of eluted metal ions, thereby inhibiting the Pt agglomeration
in Pt/C catalyst.

Furthermore, as the Fermi level of GO is above the reduction potentials of most noble
metal ions, GO can be used as the electron donors for reducing noble metal ions without
adding any additional reductants and surfactants, leading to the growth of the metal
nanoparticles on the GO surfaces with excellent monodispersity, uniformity and purity [215–
220]. The noble metals with ultrafine sizes, possessing increased surface area and more edges
and corner atoms, are found to be able to express a superior electrocatalytic ability in the fuel
cells [215]. For example, gold nanoparticles monodispersed on GO nanosheets were
successfully synthesised by the redox reaction between AuCl42- and GO, which exhibited a
high electrocatalytic activity and a dominant four-electron pathway toward oxygen reduction
reaction [218]. The spontaneous electron transfer from the GO to the AuCl42− was due to their
relative potential levels. The Fermi level of the GO is +0.24 V vs. standard hydrogen
electrode (SHE), which is well above the reduction potential of AuCl42− (+1.002 V vs. SHE).
More interestingly, a 3D GO/carbon sphere supported silver composite was produced using
GO as the reductant, which presented a significantly improved activity towards the oxygen
reduction reaction in alkaline media [219]. The 3D architecture was advantageous, as it
enforced the mass transport in the catalyst layer, facilitating the reactants access to the active
sites.

Nevertheless, the high cost of the noble metal catalysts is a major barrier for the
commercialisation of fuel cells [221]. Electrically conductive π−conjugated polymers, thus,
have attracted interest as an acceptable candidate in terms of their low cost and high
performance. The embedding of GO into the polymer matrix can increase its conductivity and
reduce the fouling of the membrane [221]. The GO/polymer composite can be formed
through hydrogen bonding and a ring-opening reaction of the epoxide groups with the amine
to form a new C−N bond. Such C−N bonds and the sulphur atoms in the polymer backbone
play a function of the active sites for the electrocatalytic reduction of O2 to H2O. Therefore,
the combination of GO with the polymer provided sufficient electroactive sites, resulting in
more favourable electron transfer kinetics and greatly enhanced electrocatalytic performance
towards oxygen reduction on the composite electrode [221].

35
4.2. Contributions to membranes
Owing to a high proton conduction, GO has been widely explored as a promising
membrane/separator in various fuel cells [222–226]. It is demonstrated that GO possesses a
slightly higher water uptake, which possibly results in a better water retention for high
temperature operation, and higher tensile strength and elastic modulus than Nafion membrane
[223]. The free-standing GO membrane in a PEMFC presented a maximum power density of
~34 mW cm−2 within the temperature range of 30–80 °C, which was close to that of a Nafion
based fuel cell under the similar conditions [223]. Furthermore, functionalised GO sheets,
such as ozonated GO and sulfonic acid treated GO, show an enhanced proton conductivity
owing to the higher quantity of oxygenated functional groups on the GO sheets, which thus
significantly improve the performance of the fuel cells [227–229].

Additionally, Nafion/GO composite membranes were substantially investigated in order


to improve the physical and chemical properties of the Nafion membrane [230–235]. The
addition of GO into the Nafion membrane can lead to further enhancement in power density
of the fuel cells, which is because GO can retain more water within the membrane to increase
the proton conductivity [232,235]. Besides, sulfonated GO/Nafion membrane in a PEMFC
was found to display a higher power density of 150 mW cm-2 than the pure Nafion membrane
(42 mW cm−2) [234]. The addition of phosphonic acid-functionalised GO into Nafion matrix
provided extra proton conductive pathways and improved water uptake capability, thereby
enhancing the performance of the PEMFC [236]. Moreover, the GO or functionalised
GO/Nafion composite membranes can also be used in DMFCs, which greatly decreased the
methanol permeability and simultaneously maintaining ionic conductivity [237–242]. In
particular, the exfoliated and orderly GO sheets showed an effective reduction of the
diffusion of methanol, which was attributed to an increase in the tortuosity of the pathway
[239–242].

As the Nafion is an expensive material, the embedding of GO or functionalised GO into


other various polymers as the membranes is also studied and exhibits a competitive
performance in the fuel cells [243,244,253,254,245–252]. For instance, poly (ethylene oxide)
(PEO) cross-linked GO membrane [246] and the sulfonated GO/SPEEK composite
membrane [252] were reported in the PEMFC system. The SPEEK/GO membrane and the
SPEEK/PVA/sulfonated GO/Fe3O4 nanocomposite membrane were reported in the DMFCs
with a low methanol permeability, excellent tensile strength and superior proton conductivity
[244,255]. Other composite membranes, e.g. a SPEEK/sulfonated GO/SPEEK membrane

36
with a sandwiched architecture [256], and a similar sandwiched membrane based on
SPEEK/sulfonated holey GO paper/SPEEK [257], sulfonated organosilane functionalised
GO/SPEEK membrane [250], were designed for high-performance DMFCs, whose improved
power densities were mainly benefited from increasing the proton transport paths and
retarding the methanol permeability. Moreover, a membrane based on polybenzimidazole
(PBI) and ionic liquid functionalised GO with excellent alkaline stability was developed for
AFCs [258] as well as a membrane based on phosphoric acid-doped PBI/GO with increased
acid doping and acid leaching properties was proposed for high-temperature PEMFCs [259].

The introduction of GO into the membranes not only increases the ion-exchange capacity,
water uptake, proton conductivity and mechanical properties, e.g. in PEMFCs and DMFCs,
but also reduces the methanol permeability and improves the selectivity of the membrane, e.g.
in DMFCs [247,249]. More importantly, the size of GO sheets was found to have an effect on
the performance of the fuel cells [260]. The smaller the size of the GO, the better the
performance for the polymer-GO composites. This was because GO sheets with the smaller
size had a stronger hydrogen bonding interaction with the polymer matrix, thus promoting the
formation of the well-defined microstructure and well-connected proton transport pathways.

5. Conclusions and Outlook


GO, produced from natural graphite by chemical oxidation and subsequent exfoliation, is
a form of a monolayer graphene platelet that mainly decorates the basal plane with epoxide
and hydroxyl groups and the edges with carboxylic acid. The solution processability of GO
provides a simple and practical strategy to thin film, paper, or carbon-based composite
preparation, which can be achieved by the self-assembly of GO at various interfaces, e.g.
liquid-liquid, liquid-air, or liquid-solid interfaces. Benefited from the oxygenated groups, GO
possesses intriguing chemistry and architectures, including its facile synthesis and
macroscopic assembly, large specific surface area, excellent solubility and processability,
adjustable electrical conductivity, high flexibility and elasticity, remarkable platform for
chemical modification and functionalisation, and satisfactory electrochemical performance.

The energy storage applications of GO presented in this review include batteries,


capacitors and fuel cells. For lithium ion batteries, GO play various functions of confining the
particle size of the electrode materials, providing larger surface area, suppressing the
agglomeration and deformation of the electrode nanostructures during cycling, acting as
coating layers for the formation of 3D network for extra storage of Li+ ions, and/or increasing

37
the capacity via oxygen groups and nano-cavity. GO can be incorporated in solid/gel polymer
electrolyte membranes for enhancing ionic conductivity and accelerating the ion transfer
pathways. Besides, GO is used as a protective coating to inhibit the corrosion of the current
collector in lithium ion batteries. For lithium sulphur batteries, GO, on one hand, plays the
role of accommodating and immobilising S to moderate lithium polysulphide dissolution; one
the other hand, is used as either a permselective or an electroactive interlayer to reduce the
polysulphide shuttle. For lithium air batteries, GO provides abundant porous structure for O2
diffusion, which enhances the efficiency of the formation and decomposition of Li2O or
Li2O2. For redox flow batteries, the oxygen groups of GO play an electrocatalytic function to
facilitate the adsorption of the reactive ions onto the electrode surface. For supercapacitors,
GO, owing to its large surface area and abundant oxygen functional groups, can provide the
electrodes with large pseudo-capacitance and better wettability. Besides, the addition of GO
in the gel polymer electrolyte and separator results in higher ionic conductivity for solid-state
supercapacitors. For fuel cells, GO contributes greatly to the fabrication of the electrolyte
membranes by enhancing the proton conductivity and/or improving the membrane stability.

Despite the comprehensive study of GO in electrochemical energy storage, there are


several challenging aspects remaining unsolved. First, more studies on precisely controlling
the microstructures (size and surface chemistry) of GO are needed. Depending not only on
the particular oxidants used but also on the carbon sources and reaction conditions, the size of
GO and the amount and distribution of oxygen groups show strong variance, which further
determine the properties of GO including the electrical and electronic conductivity as well as
mechanical strength. The properties of GO are thus controllable, but how to tailor the
properties consciously should be studied extensively to improve the application performance
of the GO.

Second, fundamental study on the interaction mechanism and the potential synergistic
effects between the GO and the additional material in a composite should attract more interest
for future energy storage applications. To date, the research of GO-based composite has
mainly focused on the preparations of the electrodes, solid-state electrolytes and separators in
the energy storage devices, which have demonstrated the improved electrochemical
performance. Based on these achievements, in situ nano-characterisation techniques should
be developed for analyse the mechanism and the synergistic effects of the GO-based
composite, which can provide a significant shortcut to design the composite materials with
optimal performance.

38
Lastly, more studies on next-generation energy storage devices using GO or GO-based
composites should be developed. GO-based materials deliver unique advantages, particularly
for fabricating the lithium-ion batteries and supercapacitors. Based on these achievements,
lithium-sulfur batteries, lithium-oxygen batteries, sodium-ion batteries, metal-ion
supercapacitors and so on should attract more attention to explore the feasibility of GO-based
materials on enhancing the specific capacity, rate capability, and cycling stability of these
devices. Nonetheless, it is certain that GO and GO-based materials show advantageous
potentials in electrochemical applications and undoubtedly, they have a very bright and
exciting future.

Acknowledgment

This work was financially supported by the Australian Research Council Discovery Projects
Discovery Project (DP190101008), Future Fellowship (FT190100058), ARC ITRP
(IH180100020), the UNSW Scientia Program, and the UNSW-Tsinghua joint grant. This
research was partially supported by funding from the UNSW Digital Grid Futures Institute,
UNSW, Sydney, under a cross disciplinary fund scheme; the views expressed herein are
those of the authors and are not necessarily those of the institute. C.S. appreciates the
financial support from Guangdong Innovation Research Team for Higher Education
(2017KCXTD030), and High-Level Talents Project of Dongguan University of Technology
(KCYKYQD2017017).

References

[1] M.J. Allen, V.C. Tung, R.B. Kaner, Chem. Rev. 110 (2010) 132–145.
[2] W. Choi, I. Lahiri, R. Seelaboyina, Y.S. Kang, Crit. Rev. Solid State Mater. Sci. 35
(2010) 52–71.
[3] K.S. Novoselov, A.K. Geim, S. V Morozov, D. Jiang, Y. Zhang, S. V Dubonos, I. V
Grigorieva, A.A. Firsov, Science 306 (2004) 666–669.
[4] D. a C. Brownson, D.K. Kampouris, C.E. Banks, J. Power Sources 196 (2011) 4873–
4885.
[5] R. Raccichini, A. Varzi, S. Passerini, B. Scrosati, Nat. Mater. 14 (2015) 271–279.
[6] X. Cai, L. Lai, Z. Shen, J. Lin, J. Mater. Chem. A 5 (2017) 15423–15446.
[7] D. Chen, H. Feng, J. Li, Chem. Rev. 112 (2012) 6027–53.
[8] D.R. Dreyer, S. Park, C.W. Bielawski, R.S. Ruoff, Chem. Soc. Rev. 39 (2010) 228–40.

39
[9] O.C. Compton, S.T. Nguyen, Small 6 (2010) 711–23.
[10] O.C. Compton, S.W. Cranford, K.W. Putz, Z. An, L.C. Brinson, M.J. Buehler, S.T.
Nguyen, ACS Nano 6 (2012) 2008–19.
[11] S. Saxena, T.A. Tyson, S. Shukla, E. Negusse, H. Chen, J. Bai, Appl. Phys. Lett. 99
(2011) 2009–2012.
[12] A. Ganguly, S. Sharma, P. Papakonstantinou, J. Hamilton, J. Phys. Chem. 115 (2011)
17009–17019.
[13] S. Wang, P.J. Chia, L.L. Chua, L.H. Zhao, R.Q. Png, S. Sivaramakrishnan, M. Zhou,
R.G.S. Goh, R.H. Friend, A.T.S. Wee, P.K.H. Ho, Adv. Mater. 20 (2008) 3440–3446.
[14] Y. Yang, J. Wang, J. Zhang, J. Liu, X. Yang, H. Zhao, Langmuir 25 (2009) 11808–
11814.
[15] G. Eda, G. Fanchini, M. Chhowalla, Nat. Nanotechnol. 3 (2008) 270–274.
[16] C. Mattevi, G. Eda, S. Agnoli, S. Miller, K.A. Mkhoyan, O. Celik, D. Mastrogiovanni,
C. Cranozzi, E. Carfunkel, M. Chhowalla, Adv. Funct. Mater. 19 (2009) 2577–2583.
[17] F. Kim, L.J. Cote, J. Huang, Adv. Mater. 22 (2010) 1954–8.
[18] J. Kim, L.J. Cote, F. Kim, W. Yuan, K.R. Shull, J. Huang, J. Am. Chem. Soc. 132
(2010) 8180–6.
[19] J.-J. Shao, W. Lv, Q.-H. Yang, Adv. Mater. 26 (2014) 5586-5612.
[20] F. Li, X. Jiang, J. Zhao, S. Zhang, Nano Energy 16 (2015) 488–515.
[21] Z. Song, H. Zhou, Energy Environ. Sci. 6 (2013) 2280.
[22] Y.X. Ke Wang, Yonggang Wang, Congxiao Wang, Electrochim. Acta 146 (2013) 8–
14.
[23] C. Xu, L. Li, F. Qiu, C. An, Y. Xu, Y. Wang, Y. Wang, L. Jiao, H. Yuan, J. Energy
Chem. 23 (2014) 397–402.
[24] J. Xu, Zhaolong Li, X. Zhang, S. Huang, S. Jian, Q. Zhu, H. Sun, Galina S. Zakharova,
Int. J. Nanotechnol. 11 (2014) 808–818.
[25] V.S. Reddy Channu, D. Ravichandran, B. Rambabu, R. Holze, Appl. Surf. Sci. 305
(2014) 596–602.
[26] F. Yu, L. Zhang, L. Lai, M. Zhu, Y. Guo, L. Xia, P. Qi, G. Wang, B. Dai, Electrochim.
Acta 151 (2015) 240–248.
[27] W. Luo, B. Zheng, J. He, J. Alloys Compd. 705 (2017) 405–412.
[28] W. Luo, B. Zheng, Appl. Surf. Sci. 404 (2017) 310–317.
[29] M.E. Stournara, V.B. Shenoy, J. Power Sources 196 (2011) 5697–5703.
[30] D.-W. Wang, C. Sun, G. Zhou, F. Li, L. Wen, B.C. Donose, G.Q. (Max) Lu, H.-M.
Cheng, I.R. Gentle, J. Mater. Chem. A 1 (2013) 3607.

40
[31] W. Ai, Z. Du, Z. Fan, J. Jiang, Y. Wang, H. Zhang, L. Xie, W. Huang, T. Yu, Carbon
N. Y. 76 (2014) 148–154.
[32] Y.G. Zhu, Y. Wang, J. Xie, G.-S. Cao, T.-J. Zhu, X. Zhao, H.Y. Yang, Electrochim.
Acta 154 (2015) 338–344.
[33] H. Song, N. Li, C. Wang, J. Mater. Chem. A 1 (2013) 7558–7562.
[34] S. Information, A. Bhaskar, M. Deepa, M. Ramakrishna, T.N. Rao, J. Phys. Chem. C
118 (2014) 7296–7306.
[35] M.J.K. Reddy, S.H. Ryu, A.M. Shanmugharaj, Nanoscale 8 (2016) 471–482.
[36] K.H. Yunwei Shi, Dongyun Zhang, Chengkang Chang, R. Holze, J. Alloys Compd.
639 (2015) 274–279.
[37] S. Petnikota, S.K. Marka, A. Banerjee, M.V. Reddy, V.V.S.S. Srikanth, B.V.R.
Chowdari, J. Power Sources 293 (2015) 253–263.
[38] A.G. Medvedev, A.A. Mikhaylov, D.A. Grishanov, D.Y.W. Yu, J. Gun, S. Sladkevich,
O. Lev, P. V. Prikhodchenko, ACS Appl. Mater. Interfaces 9 (2017) 9152–9160.
[39] J. Zhang, Y. Cai, J. Wu, J. Yao, Electrochim. Acta 165 (2015) 422–429.
[40] Y.-T. Xu, Y. Guo, C. Li, X.-Y. Zhou, M.C. Tucker, X.-Z. Fu, R. Sun, C.-P. Wong,
Nano Energy 11 (2015) 38–47.
[41] Y.-T. Xu, Y. Guo, H. Jiang, X.-B. Xie, B. Zhao, P.-L. Zhu, X.-Z. Fu, R. Sun, C.-P.
Wong, Energy Technol. 3 (2015) 488–495.
[42] X. Meng, Y. Xu, X. Sun, J. Wang, L. Xiong, X. Du, S. Mao, J. Mater. Chem. A 3
(2015) 12938–12946.
[43] Q. Qu, T. Gao, H. Zheng, X. Li, H. Liu, M. Shen, J. Shao, H. Zheng, Carbon N. Y. 92
(2015) 119–125.
[44] D. Kong, J.-J. Shao, Q.-H. Yang, G. Zhou, C. Luo, S. Wu, W. Lv, F. Kang, B. Li, Z.
Li, Energy Storage Mater. 6 (2016) 98–103.
[45] C. Sun, M. Yang, T. Wang, Y. Shao, Y. Wu, X. Hao, ACS Appl. Mater. Interfaces 9
(2017) 26631–26636.
[46] Z. Xie, X. Li, W. Li, M. Chen, M. Qu, J. Power Sources 273 (2015) 754–760.
[47] J. Zhang, H. Cao, X. Tang, W. Fan, G. Peng, M. Qu, J. Power Sources 241 (2013)
619–626.
[48] J. Zhang, Z. Xie, W. Li, S. Dong, M. Qu, Carbon N. Y. 74 (2014) 153–162.
[49] Y. Ye, P. Wang, H. Sun, Z. Tian, J. Liu, C. Liang, RSC Adv. 5 (2015) 45038–45043.
[50] C. Fang, Y. Deng, Y. Xie, J. Su, G. Chen, J. Phys. Chem. C 119 (2015) 1720–1728.
[51] H. Li, X. Zhu, H. Sitinamaluwa, K. Wasalathilake, L. Xu, S. Zhang, C. Yan, J. Alloys
Compd. 714 (2017) 425–432.

41
[52] Z. Jiang, Z.-J. Jiang, Y. Qin, H. Rong, M. Liu, J. Alloys Compd. 731 (2017) 1095–
1102.
[53] J. Hu, H. Diao, W. Luo, Y.F. Song, Chem. - A Eur. J. 23 (2017) 8729–8735.
[54] Y.F. Huang, W.H. Ruan, D.L. Lin, M.Q. Zhang, ACS Appl. Mater. Interfaces 9 (2017)
909–918.
[55] M. Bai, K. Xie, K. Yuan, K. Zhang, N. Li, C. Shen, Y. Lai, R. Vajtai, P. Ajayan, B.
Wei, Adv. Mater. 30 (2018) 1801213.
[56] S.J. Richard Prabakar, Y.H. Hwang, E.G. Bae, D.K. Lee, M. Pyo, Carbon 52 (2013)
128–136.
[57] S. Gao, J. Zhong, G. Xue, B. Wang, J. Memb. Sci. 470 (2014) 316–322.
[58] S. Cheng, D.M. Smith, C.Y. Li, Macromolecules 48 (2015) 4503–4510.
[59] J. Shim, D.-G. Kim, H.J. Kim, J.H. Lee, J.-H. Baik, J.-C. Lee, J. Mater. Chem. A
Mater. Energy Sustain. 2 (2014) 13873–13883.
[60] K. Fu, Y. Wang, C. Yan, Y. Yao, Y. Chen, J. Dai, S. Lacey, Y. Wang, J. Wan, T. Li, Z.
Wang, Y. Xu, L. Hu, Adv. Mater. 28 (2016) 2587–2594.
[61] A. Manthiram, Y. Fu, Y.-S. Su, Acc. Chem. Res. 46 (2013) 1125–1134.
[62] E.S. Shin, K. Kim, S.H. Oh, W. Il Cho, Chem. Commun. (Camb). 49 (2013) 2004–6.
[63] H. Wu, Y. Huang, M. Zong, X. Ding, J. Ding, X. Sun, Mater. Res. Bull. 64 (2015) 12–
16.
[64] J. Rong, M. Ge, X. Fang, C. Zhou, Nano Lett. 14 (2014) 473–479.
[65] M. Xiao, M. Huang, S. Zeng, D. Han, S. Wang, L. Sun, Y. Meng, RSC Adv. 3 (2013)
4914.
[66] L.S. Cells, L. Ji, M. Rao, H. Zheng, L. Zhang, O.Y. Li, W. Duan, J. Am. Chem. Soc.
133 (2011) 18522–18525.
[67] L. Zhang, L. Ji, P.-A. Glans, Y. Zhang, J. Zhu, J. Guo, Phys. Chem. Chem. Phys. 14
(2012) 13670.
[68] X. Feng, M.-K. Song, W.C. Stolte, D. Gardenghi, D. Zhang, X. Sun, J. Zhu, E.J.
Cairns, J. Guo, Phys. Chem. Chem. Phys. 16 (2014) 16931-16940.
[69] S. Liu, K. Xie, Y. Li, Z. Chen, X. Hong, L. Zhou, J. Yuan, C. Zheng, RSC Adv. 5
(2015) 5516–5522.
[70] W. Zhou, H. Chen, Y. Yu, D. Wang, Z. Cui, F.J. Disalvo, C. Biology, N. York, U.
States, ACS Nano 7 (2013) 8801–8808.
[71] J. Zhang, N. Yang, X. Yang, S. Li, J. Yao, Y. Cai, J. Alloys Compd. 650 (2015) 604–
609.
[72] C. Sun, L. Shi, C. Fan, X. Fu, Z. Ren, G. Qian, Z. Wang, RSC Adv. 5 (2015) 28832–
28835.

42
[73] X. Huang, K. Shi, J. Yang, G. Mao, J. Chen, J. Power Sources 356 (2017) 72–79.
[74] X. Du, X. Zhang, J. Guo, S. Zhao, F. Zhang, J. Alloys Compd. 714 (2017) 311–317.
[75] Y. Hwa, H.K. Seo, J.M. Yuk, E.J. Cairns, Nano Lett. 17 (2017) 7086–7094.
[76] X. Wang, Y. Zhang, Y. Tian, F. Yin, Z. Sun, Z. Bakenov, Nanomaterials 8 (2018) 50.
[77] Y. Hwa, J. Zhao, E.J. Cairns, Nano Lett. (2015) 150430103348003.
[78] C. Zhang, W. Lv, W. Zhang, X. Zheng, M.B. Wu, W. Wei, Y. Tao, Z. Li, Q.H. Yang,
Adv. Energy Mater. 4 (2014) 1–5.
[79] Q. (Ray) Zeng, X. Leng, K.-H. Wu, I.R. Gentle, D.-W. Wang, Carbon 93 (2015) 611–
619.
[80] J.-Q. Huang, T.-Z. Zhuang, Q. Zhang, H.-J. Peng, C.-M. Chen, F. Wei, ACS Nano 9
(2015) 3002–3011.
[81] F. Chen, Z. Jiao, B. Zhao, Z. Chen, Y. Jiang, S. Wang, Y. Gao, Q. Gao, Y. Wang, J.
Power Sources 342 (2017) 929–938.
[82] T.-Z. Zhuang, J.-Q. Huang, H.-J. Peng, L.-Y. He, X.-B. Cheng, C.-M. Chen, Q. Zhang,
Small 12 (2015) 381–389.
[83] S. Bai, X. Liu, K. Zhu, S. Wu, H. Zhou, Nat. Energy 1 (2016) 16094.
[84] J. Zhu, C. Chen, Y. Lu, J. Zang, M. Jiang, D. Kim, X. Zhang, Carbon 101 (2016) 272–
280.
[85] X. Ni, T. Qian, X. Liu, N. Xu, J. Liu, C. Yan, Adv. Funct. Mater. 28 (2018) 1–8.
[86] W. Kong, L. Yan, Y. Luo, D. Wang, K. Jiang, Q. Li, S. Fan, J. Wang, Adv. Funct.
Mater. 27 (2017).
[87] J.Q. Huang, Z.L. Xu, S. Abouali, M. Akbari Garakani, J.K. Kim, Carbon 99 (2016)
624–632.
[88] Y.J. Zhang, X.H. Xia, X.L. Wang, C.D. Gu, J.P. Tu, RSC Adv. 6 (2016) 66161–66168.
[89] V. Chabot, D. Higgins, A. Yu, X. Xiao, Z. Chen, J. Zhang, Energy Environ. Sci. 7
(2014) 1564.
[90] M.K. Song, S. Park, F.M. Alamgir, J. Cho, M. Liu, Mater. Sci. Eng. R Reports 72
(2011) 203–252.
[91] T. Cetinkaya, S. Ozcan, M. Uysal, M.O. Guler, H. Akbulut, J. Power Sources 267
(2014) 140–147.
[92] Z.L. Wang, D. Xu, J.J. Xu, L.L. Zhang, X.B. Zhang, Adv. Funct. Mater. 22 (2012)
3699–3705.
[93] P.K. Leung, X. Li, C. Ponce de León, L. Berlouis, C.T.J. Low, F.C. Walsh, RSC Adv.
2 (2012) 10125–10156.
[94] W. Wang, Q. Luo, B. Li, X. Wei, L. Li, Z. Yang, Adv. Funct. Mater. 23 (2013) 970–

43
986.
[95] A.Z. Weber, M.M. Mench, J.P. Meyers, P.N. Ross, J.T. Gostick, Q. Liu, J. Appl.
Electrochem. 41 (2011) 1137–1164.
[96] P. Han, H. Wang, Z. Liu, X. Chen, W. Ma, J. Yao, Y. Zhu, G. Cui, Carbon N. Y. 49
(2011) 693–700.
[97] O. Di Blasi, N. Briguglio, C. Busacca, M. Ferraro, V. Antonucci, a. Di Blasi, Appl.
Energy 147 (2015) 74–81.
[98] Z. González, C. Botas, C. Blanco, R. Santamaría, M. Granda, P. Álvarez, R. Menéndez,
Nano Energy 2 (2013) 1322–1328.
[99] Z. González, C. Botas, C. Blanco, R. Santamaría, M. Granda, P. Álvarez, R. Menéndez,
J. Power Sources 241 (2013) 349–354.
[100] W. Li, J. Liu, C. Yan, Carbon N. Y. 55 (2013) 313–320.
[101] M. Park, I.-Y. Jeon, J. Ryu, J.-B. Baek, J. Cho, Adv. Energy Mater. 5 (2015) 1401550.
[102] H.M. Tsai, S.Y. Yang, C.C.M. Ma, X. Xie, Electroanalysis 23 (2011) 2139–2143.
[103] P. Han, Y. Yue, Z. Liu, W. Xu, L. Zhang, H. Xu, S. Dong, G. Cui, Energy Environ.
Sci. 4 (2011) 4710.
[104] A. Saharkhiz, E. Abouzari-Lotf, A. Ahmad, T.M. Ting, M. Etesami, M. Mahmoud
Nasef, A. Ripin, Int. J. Hydrogen Energy 43 (2017) 189–197.
[105] X. Rui, M.O. Oo, D.H. Sim, S.C. Raghu, Q. Yan, T.M. Lim, M. Skyllas-Kazacos,
Electrochim. Acta 85 (2012) 175–181.
[106] M. Govindan, I.S. Moon, Int. J. Electrochem. Sci. 8 (2013) 12172–12183.
[107] Z. Xie, B. Yang, L. Yang, X. Xu, D. Cai, J. Chen, Y. Chen, Y. He, Y. Li, X. Zhou, J.
Solid State Electrochem. 19 (2015) 3339-3345.
[108] W. Dai, Y. Shen, Z. Li, L. Yu, X. Qiu, Journals Mater. Chem. A 2 (2014) 12423–
12432.
[109] C.Y. Tseng, Y.S. Ye, M.Y. Cheng, K.Y. Kao, W.C. Shen, J. Rick, J.C. Chen, B.J.
Hwang, Adv. Energy Mater. 1 (2011) 1220–1224.
[110] B.G. Kim, T.H. Han, C.G. Cho, J. Nanosci. Nanotechnol. 14 (2014) 9073–9077.
[111] K.J. Lee, Y.H. Chu, Vacuum 107 (2014) 269–276.
[112] D. Zhang, S. Peng, Y. Luo, G. He, L. Su, X. Wu, Int. J. Hydrogen Energy 42 (2017)
21806–21816.
[113] L. Yu, F. Lin, L. Xu, J. Xi, RSC Adv. 6 (2016) 3756–3763.
[114] L. Zheng, H. Wang, R. Niu, Y. Zhang, H. Shi, Electrochim. Acta 282 (2018) 437–447.
[115] H. Shi, H. Wang, L. Kong, R. Niu, L. Zheng, RSC Adv. 6 (2016) 100262–100270.

44
[116] C. Wu, H. Bai, Y. Lv, Z. Lv, Y. Xiang, S. Lu, Electrochim. Acta 248 (2017) 454–461.
[117] M.H. Han, E. Gonzalo, G. Singh, T. Rojo, Energy Environ. Sci. 8 (2015) 81–102.
[118] D. Kundu, E. Talaie, V. Duffort, L.F. Nazar, Angew. Chemie Int. Ed. 54 (2015) 3431–
3448.
[119] V. Palomares, P. Serras, I. Villaluenga, K.B. Hueso, J. Carretero-González, T. Rojo,
Energy Environ. Sci. 5 (2012) 5884.
[120] C. Luo, Y. Zhu, Y. Xu, Y. Liu, T. Gao, J. Wang, C. Wang, J. Power Sources 250
(2014) 372–378.
[121] C. Yan, H. Li, K. Wasalathilake, L. Xu, H. Sitinamaluwa, Compos. Commun. 1 (2016)
48–53.
[122] W. Luo, C. Bommier, Z. Jian, X. Li, R. Carter, S. Vail, Y. Lu, J.-J. Lee, X. Ji, ACS
Appl. Mater. Interfaces 7 (2015) 2626–2631.
[123] R.R. Gaddam, S.R. Varanasi, N.A. Kumar, M. Suresh, X.S. Zhao, D. Yang, S.K.
Bhatia, J. Mater. Chem. A 5 (2017) 13204–13211.
[124] F. Wan, Y.H. Li, D.H. Liu, J.Z. Guo, H.Z. Sun, J.P. Zhang, X.L. Wu, Chem. - A Eur. J.
22 (2016) 8152–8157.
[125] H. Tateishi, T. Koga, K. Hatakeyama, a. Funatsu, M. Koinuma, T. Taniguchi, Y.
Matsumoto, ECS Electrochem. Lett. 3 (2014) A19–A21.
[126] J. Gao, T. Xu, Y. Xiao, M. Ye, L. Qu, Y. Zhao, Carbon N. Y. 125 (2017) 299–307.
[127] X. Yu, A. Manthiram, J. Phys. Chem. C 118 (2014) 22952–22959.
[128] X. Hu, Z. Li, Y. Zhao, J. Sun, Q. Zhao, J. Wang, Z. Tao, J. Chen, Sci. Adv. 3 (2017)
1–8.
[129] B.W. Zhang, T. Sheng, Y.X. Wang, S. Chou, K. Davey, S.X. Dou, S.Z. Qiao, Angew.
Chemie - Int. Ed. 58 (2019) 1484–1488.
[130] B.W. Zhang, T. Sheng, Y.D. Liu, Y.X. Wang, L. Zhang, W.H. Lai, L. Wang, J. Yang,
Q.F. Gu, S.L. Chou, H.K. Liu, S.X. Dou, Nat. Commun. 9 (2018) 4082.
[131] S. Li, Y. Liu, J. Zhou, S. Hong, Y. Dong, J. Wang, X. Gao, P. Qi, Y. Han, B. Wang,
Energy Environ. Sci. 12 (2019) 1046–1054.
[132] B.W. Zhang, Y. Jiao, D.L. Chao, C. Ye, Y.X. Wang, K. Davey, H.K. Liu, S.X. Dou,
S.Z. Qiao, Adv. Funct. Mater. 29 (2019) 1–7.
[133] Q. Zhang, K. Scrafford, M. Li, Z. Cao, Z. Xia, P.M. Ajayan, B. Wei, Nano Lett. 14
(2014) 1938-1943.
[134] D.-W. Wang, A. Du, E. Taran, G.Q. (Max) Lu, I.R. Gentle, J. Mater. Chem. 22 (2012)
21085.
[135] C. Wu, X. Huang, X. Wu, L. Xie, K. Yang, P. Jiang, Nanoscale 5 (2013) 3847–55.
[136] T. Kavinkumar, S. Manivannan, J. Mater. Sci. Mater. Electron. 28 (2017) 344–353.

45
[137] K. Deshmukh, M.B. Ahamed, K.K. Sadasivuni, D. Ponnamma, R.R. Deshmukh,
S.K.K. Pasha, M.A.A. AlMaadeed, K. Chidambaram, J. Polym. Res. 23 (2016) 159.
[138] K. Deshmukh, M.B. Ahamed, K.K. Sadasivuni, D. Ponnamma, M.A.A. AlMaadeed,
S.K. Khadheer Pasha, R.R. Deshmukh, K. Chidambaram, Mater. Chem. Phys. 186
(2017) 188–201.
[139] S. Song, Y. Zhai, Y. Zhang, ACS Appl. Mater. Interfaces 8 (2016) 31264–31272.
[140] X. Hong, W. Yu, D.D.L. Chung, Carbon N. Y. 119 (2017) 403–418.
[141] L.L. Zhang, R. Zhou, X.S. Zhao, J. Mater. Chem. 20 (2010) 5983.
[142] P. Yang, W. Mai, Nano Energy 8 (2014) 274–290.
[143] B. Xu, S. Yue, Z. Sui, X. Zhang, S. Hou, G. Cao, Y. Yang, Energy Environ. Sci. 4
(2011) 2826–2830.
[144] P. Karthika, Soft Nanosci. Lett. 02 (2012) 59–66.
[145] A.B. Dighe, D.P. Dubal, R. Holze, Zeitschrift Für Anorg. Und Allg. Chemie 640 (2014)
2852–2857.
[146] Y. Wang, J. Zhu, Nanotechnology 26 (2015) 055401.
[147] Y. Zhang, H. Feng, X. Wu, L. Wang, A. Zhang, T. Xia, H. Dong, X. Li, L. Zhang, Int.
J. Hydrogen Energy 34 (2009) 4889–4899.
[148] Y. Ma, C. Ma, J. Sheng, H. Zhang, R. Wang, Z. Xie, J. Shi, J. Colloid Interface Sci.
461 (2016) 96–103.
[149] G. Chen, J. Xu, K. Ni, Y. Zhao, W. Zeng, Z. Tan, Y. Zhu, Z. Tao, H. Ji, M. Ikram, S.
Wu, Adv. Mater. 28 (2016) 5222–5228.
[150] A. Kafy, A. Akther, L. Zhai, H.C. Kim, J. Kim, Synth. Met. 223 (2017) 94–100.
[151] S.R. Guo, W. Wang, C.S. Ozkan, M. Ozkan, J. Mater. Res. 28 (2013) 918–926.
[152] L.-Y. Lin, M.-H. Yeh, J.-T. Tsai, Y.-H. Huang, C.-L. Sun, K.-C. Ho, J. Mater. Chem.
A 1 (2013) 11237.
[153] Y. Xue, L. Zhu, H. Chen, J. Qu, L. Dai, Carbon N. Y. 92 (2015) 305–310.
[154] J.N. Lekitima, K.I. Ozoemena, C.J. Jafta, N. Kobayashi, Y. Song, D. Tong, S. Chen, M.
Oyama, J. Mater. Chem. A 1 (2013) 2821.
[155] K. Dai, L. Lu, C. Liang, J. Dai, Q. Liu, Y. Zhang, G. Zhu, Z. Liu, Electrochim. Acta
116 (2014) 111–117.
[156] J. Fang, M. Li, Q. Li, W. Zhang, Q. Shou, F. Liu, X. Zhang, J. Cheng, Electrochim.
Acta 85 (2012) 248–255.
[157] G. Zhao, J. Li, L. Jiang, H. Dong, X. Wang, W. Hu, Chem. Sci. 3 (2012) 433.
[158] S. Chen, J. Zhu, X. Wu, Q. Han, X. Wang, ACS Nano 4 (2010) 2822–2830.

46
[159] C.J. Jafta, F. Nkosi, L. Le Roux, M.K. Mathe, M. Kebede, K. Makgopa, Y. Song, D.
Tong, M. Oyama, N. Manyala, S. Chen, K.I. Ozoemena, Electrochim. Acta 110 (2013)
228–233.
[160] Y. Liu, D. Yan, Y. Li, Z. Wu, R. Zhuo, S. Li, J. Feng, J. Wang, P. Yan, Z. Geng,
Electrochim. Acta 117 (2014) 528–533.
[161] G.S. Gund, D.P. Dubal, B.H. Patil, S.S. Shinde, C.D. Lokhande, Electrochim. Acta 92
(2013) 205–215.
[162] M.-S. Wu, Y.-P. Lin, C.-H. Lin, J.-T. Lee, J. Mater. Chem. 22 (2012) 2442.
[163] A. Wang, H. Wang, S. Zhang, C. Mao, J. Song, H. Niu, B. Jin, Y. Tian, Appl. Surf. Sci.
282 (2013) 704–708.
[164] A. Pendashteh, M.F. Mousavi, M.S. Rahmanifar, Electrochim. Acta 88 (2013) 347–
357.
[165] D. Carriazo, J. Patiño, M.C. Gutiérrez, M.L. Ferrer, F. del Monte, RSC Adv. 3 (2013)
13690.
[166] L. Wang, D. Wang, X.Y. Dong, Z.J. Zhang, X.F. Pei, X.J. Chen, B. Chen, J. Jin, Chem.
Commun. (Camb). 47 (2011) 3556–3558.
[167] P. Simon, Y. Gogotsi, Nat. Mater. 7 (2008) 845–854.
[168] L.L. Zhang, S. Zhao, X.N. Tian, X.S. Zhao, Langmuir 26 (2010) 17624–8.
[169] Y. Liu, R. Deng, Z. Wang, H. Liu, J. Mater. Chem. 22 (2012) 13619.
[170] G. Xu, N. Wang, J. Wei, L. Lv, J. Zhang, Z. Chen, Q. Xu, Ind. Eng. Chem. Res. 51
(2012) 14390–14398.
[171] H. Wang, Q. Hao, X. Yang, L. Lu, X. Wang, Electrochem. Commun. 11 (2009) 1158–
1161.
[172] Z. Luo, L. Zhu, H. Zhang, H. Tang, Mater. Chem. Phys. 139 (2013) 572–579.
[173] H. Wang, Q. Hao, X. Yang, L. Lu, X. Wang, ACS Appl. Mater. Interfaces 2 (2010)
821–8.
[174] Q. Zhang, Y. Li, Y. Feng, W. Feng, Electrochim. Acta 90 (2013) 95–100.
[175] J. Xu, K. Wang, S.-Z. Zu, B.-H. Han, Z. Wei, ACS Nano 4 (2010) 5019–26.
[176] E. Mitchell, J. Candler, F. De Souza, R.K. Gupta, B.K. Gupta, L.F. Dong, Synth. Met.
199 (2015) 214–218.
[177] Z.F. Li, H. Zhang, Q. Liu, Y. Liu, L. Stanciu, J. Xie, Carbon 71 (2014) 257–267.
[178] S. Konwer, R. Boruah, S.K. Dolui, J. Electron. Mater. 40 (2011) 2248–2255.
[179] J. Li, H. Xie, Y. Li, J. Power Sources 241 (2013) 388–395.
[180] F.-H. Hsu, T.-M. Wu, Synth. Met. 198 (2014) 188–195.

47
[181] C. Bora, S.K. Dolui, Polym. (United Kingdom) 53 (2012) 923–932.
[182] L.Q. Fan, G.J. Liu, J.H. Wu, L. Liu, J.M. Lin, Y.L. Wei, Electrochim. Acta 137 (2014)
26–33.
[183] H. Zhou, G. Han, Y. Xiao, Y. Chang, H.J. Zhai, J. Power Sources 263 (2014) 259–267.
[184] C. Zhu, J. Zhai, D. Wen, S. Dong, J. Mater. Chem. 22 (2012) 6300.
[185] J. Li, H. Xie, Mater. Lett. 78 (2012) 106–109.
[186] N.H. Nabilah Azman, H.N. Lim, Y. Sulaiman, Electrochim. Acta 188 (2016) 785–792.
[187] A. Alabadi, S. Razzaque, Z. Dong, W. Wang, B. Tan, J. Power Sources 306 (2016)
241–247.
[188] A. Mohammadi, S.J. Peighambardoust, A.A. Entezami, N. Arsalani, J. Mater. Sci.
Mater. Electron. 28 (2017) 5776–5787.
[189] G. Han, Y. Liu, L. Zhang, E. Kan, S. Zhang, J. Tang, W. Tang, Sci. Rep. 4 (2014)
4824.
[190] H. Su, T. Wang, S. Zhang, J. Song, C. Mao, H. Niu, B. Jin, J. Wu, Y. Tian, Solid State
Sci. 14 (2012) 677–681.
[191] W.K. Chee, H.N. Lim, I. Harrison, K.F. Chong, Z. Zainal, C.H. Ng, N.M. Huang,
Electrochim. Acta 157 (2015) 88–94.
[192] Y. Zou, Q. Wang, C. Xiang, Z. She, H. Chu, S. Qiu, F. Xu, S. Liu, C. Tang, L. Sun,
Electrochim. Acta 188 (2016) 126–134.
[193] B. You, N. Li, H. Zhu, X. Zhu, J. Yang, ChemSusChem 6 (2013) 474–480.
[194] B. Yuan, C. Xu, D. Deng, Y. Xing, L. Liu, H. Pang, D. Zhang, Electrochim. Acta 88
(2013) 708–712.
[195] D. Xu, Q. Xu, K. Wang, J. Chen, Z. Chen, ACS Appl. Mater. Interfaces, 6 (2013) 200–
209.
[196] Y. Zhang, L. Si, B. Zhou, B. Zhao, Y. Zhu, L. Zhu, X. Jiang, Chem. Eng. J. 288 (2016)
689–700.
[197] H. Zhou, G. Han, Electrochim. Acta 192 (2016) 448–455.
[198] H. Zhou, H.J. Zhai, Org. Electron. Physics, Mater. Appl. 37 (2016) 197–206.
[199] H. Zhou, H.J. Zhai, G. Han, J. Power Sources 323 (2016) 125–133.
[200] M.G. Hosseini, E. Shahryari, J. Mater. Sci. Technol. 32 (2016) 763–773.
[201] M. Ghasem Hosseini, E. Shahryari, J. Colloid Interface Sci. 496 (2017) 371–381.
[202] G. Han, H. Zhou, H. Li, D. Fu, Y. Chang, X.-M. Zhang, Mater. Chem. Phys. 179
(2016) 166–173.
[203] M.R. Karim, K. Hatakeyama, T. Matsui, H. Takehira, T. Taniguchi, M. Koinuma, Y.

48
Matsumoto, T. Akutagawa, T. Nakamura, S. Noro, T. Yamada, H. Kitagawa, S.
Hayami, J. Am. Chem. Soc. 135 (2013) 8097–100.
[204] K. Hatakeyama, M.R. Karim, C. Ogata, H. Tateishi, A. Funatsu, T. Taniguchi, M.
Koinuma, S. Hayami, Y. Matsumoto, Angew. Chem. Int. Ed. Engl. 53 (2014) 6997–
7000.
[205] C. Ogata, R. Kurogi, K. Hatakeyama, T. Taniguchi, M. Koinuma, Y. Matsumoto,
Chem. Commun. 52 (2016) 3919–3922.
[206] X. Yang, L. Zhang, F. Zhang, T. Zhang, Y. Huang, Y. Chen, Carbon 72 (2014) 381–
386.
[207] X. Yang, F. Zhang, L. Zhang, T. Zhang, Y. Huang, Y. Chen, Adv. Funct. Mater. 23
(2013) 3353–3360.
[208] Y.-F. Huang, P.-F. Wu, M.-Q. Zhang, W.-H. Ruan, E.P. Giannelis, Electrochim. Acta
132 (2014) 103–111.
[209] A. Chandan, M. Hattenberger, A. El-kharouf, S. Du, A. Dhir, V. Self, B.G. Pollet, A.
Ingram, W. Bujalski, J. Power Sources 231 (2013) 264–278.
[210] S. Giddey, S.P.S. Badwal, a. Kulkarni, C. Munnings, Prog. Energy Combust. Sci. 38
(2012) 360–399.
[211] S. Mekhilef, R. Saidur, a. Safari, Renew. Sustain. Energy Rev. 16 (2012) 981–989.
[212] Y.-J. Wang, J. Qiao, R. Baker, J. Zhang, Chem. Soc. Rev. 42 (2013) 5768–87.
[213] Y. Wang, K.S. Chen, J. Mishler, S.C. Cho, X.C. Adroher, Appl. Energy 88 (2011)
981–1007.
[214] J.H. Jung, H.J. Park, J. Kim, S.H. Hur, J. Power Sources 248 (2014) 1156–1162.
[215] X. Chen, G. Wu, J. Chen, X. Chen, Z. Xie, X. Wang, J. Am. Chem. Soc. 133 (2011)
3693–3695.
[216] K. Kakaei, Electrochim. Acta 165 (2015) 330–337.
[217] M. Khan, A. Bin Yousaf, M. Chen, C. Wei, X. Wu, N. Huang, Z. Qi, L. Li, J. Power
Sources 282 (2015) 520–528.
[218] X.-R. Li, X.-L. Li, M.-C. Xu, J.-J. Xu, H.-Y. Chen, J. Mater. Chem. A 2 (2014) 1697.
[219] L. Yuan, L. Jiang, J. Liu, Z. Xia, S. Wang, G. Sun, Electrochim. Acta 135 (2014) 168–
174.
[220] V.K. Gupta, M.L. Yola, N. Atar, Z. Üstündaǧ, A.O. Solak, J. Mol. Liq. 191 (2014)
172–176.
[221] M.H. Naveen, H.-B. Noh, M.S. Al Hossain, J.H. Kim, Y.-B. Shim, J. Mater. Chem. A
3 (2015) 5426–5433.
[222] H. Tateishi, K. Hatakeyama, C. Ogata, K. Gezuhara, J. Kuroda, a. Funatsu, M.
Koinuma, T. Taniguchi, S. Hayami, Y. Matsumoto, J. Electrochem. Soc. 160 (2013)

49
F1175–F1178.
[223] T. Bayer, S.R. Bishop, M. Nishihara, K. Sasaki, S.M. Lyth, J. Power Sources 272
(2014) 239–247.
[224] E.C.S. Transactions, T.E. Society, ECS Trans. 64 (2014) 441–448.
[225] U.R. Farooqui, A.L. Ahmad, N.A. Hamid, Renew. Sustain. Energy Rev. 82 (2018)
714–733.
[226] R.P. Pandey, G. Shukla, M. Manohar, V.K. Shahi, Adv. Colloid Interface Sci. 240
(2017) 15–30.
[227] W. Gao, G. Wu, M.T. Janicke, D. a. Cullen, R. Mukundan, J.K. Baldwin, E.L. Brosha,
C. Galande, P.M. Ajayan, K.L. More, A.M. Dattelbaum, P. Zelenay, Angew. Chemie -
Int. Ed. 53 (2014) 3588–3593.
[228] K. Scott, Chem. Commun. 48 (2012) 5584.
[229] Z. Jiang, Y. Shi, Z.-J. Jiang, X. Tian, L. Luo, W. Chen, J. Mater. Chem. A 2 (2014)
6494.
[230] K. Feng, B. Tang, P. Wu, ACS Appl. Mater. Interfaces 5 (2013) 1481–1488.
[231] Y. Kim, K. Ketpang, S. Jaritphun, J.S. Park, S. Shanmugam, J. Mater. Chem. A 3
(2015) 8148–8155.
[232] D.C. Lee, H.N. Yang, S.H. Park, W.J. Kim, J. Memb. Sci. 452 (2014) 20–28.
[233] L. Wang, J. Kang, J.-D. Nam, J. Suhr, a. K. Prasad, S.G. Advani, ECS Electrochem.
Lett. 4 (2014) F1–F4.
[234] H. Zarrin, D. Higgins, Y. Jun, Z. Chen, M. Fowler, J. Phys. Chem. C 115 (2011)
20774–20781.
[235] K.J. Peng, J.Y. Lai, Y.L. Liu, J. Memb. Sci. 514 (2016) 86–94.
[236] B. Zhang, Y. Cao, S. Jiang, Z. Li, G. He, H. Wu, J. Memb. Sci. 518 (2016) 243–253.
[237] H.C. Chien, L.D. Tsai, C.P. Huang, C.Y. Kang, J.N. Lin, F.C. Chang, Int. J. Hydrogen
Energy 38 (2013) 13792–13801.
[238] B.G. Choi, Y.S. Huh, Y.C. Park, D.H. Jung, W.H. Hong, H. Park, Carbon 50 (2012)
5395–5402.
[239] C.W. Lin, Y.S. Lu, J. Power Sources 237 (2013) 187–194.
[240] S.J. Lue, Y.-L. Pai, C.-M. Shih, M.-C. Wu, S.-M. Lai, J. Memb. Sci. 493 (2015) 212–
223.
[241] I. Nicotera, C. Simari, L. Coppola, P. Zygouri, D. Gournis, S. Brutti, F.D. Minuto, A.S.
Arico, D. Sebastian, V. Baglio, J. Phys. Chem. C 118 (2014) 24357–24368.
[242] L. Sha Wang, A. Nan Lai, C. Xiao Lin, Q. Gen Zhang, A. Mei Zhu, Q. Lin Liu, J.
Memb. Sci. 492 (2015) 58–66.

50
[243] H. Beydaghi, M. Javanbakht, Ind. Eng. Chem. Res. 54 (2015) 7028–7037.
[244] H. Beydaghi, M. Javanbakht, A. Bagheri, P. Salarizadeh, H.G.- Zahmatkesh, S.
Kashefi, E. Kowsari, RSC Adv. 5 (2015) 74054–74064.
[245] H. Beydaghi, M. Javanbakht, E. Kowsari, Ind. Eng. Chem. Res. 53 (2014) 16621–
16632.
[246] Y.-C. Cao, C. Xu, X. Wu, X. Wang, L. Xing, K. Scott, J. Power Sources 196 (2011)
8377–8382.
[247] Y. Heo, H. Im, J. Kim, J. Memb. Sci. 425–426 (2013) 11–22.
[248] Y. He, J. Wang, H. Zhang, T. Zhang, B. Zhang, S. Cao, J. Liu, J. Mater. Chem. A 2
(2014) 9548.
[249] Z. Jiang, X. Zhao, Y. Fu, A. Manthiram, J. Mater. Chem. (2012) 24862–24869.
[250] Z. Jiang, X. Zhao, A. Manthiram, Int. J. Hydrogen Energy 38 (2013) 5875–5884.
[251] T. Ko, K. Kim, M.-Y. Lim, S.Y. Nam, T.-H. Kim, S.-K. Kim, J.-C. Lee, J. Mater.
Chem. A (2015).
[252] R. Kumar, M. Mamlouk, K. Scott, RSC Adv. 4 (2014) 617.
[253] T. Yuan, L. Pu, Q. Huang, H. Zhang, X. Li, H. Yang, Electrochim. Acta 117 (2014)
393–397.
[254] C. Xue, J. Zou, Z. Sun, F. Wang, K. Han, H. Zhu, Int. J. Hydrogen Energy 39 (2014)
7931–7939.
[255] H. Wang, H. Wu, Z. Li, M. Gang, Y. Yin, Z. Li, C. Wang, L. Cao, P. Zhang, Z. Jiang,
Electrochim. Acta 203 (2016) 178–188.
[256] C. Li, H. Yang, W. Liu, L. Xiao, Z.-L. Xu, L. Luo, X. Tian, Z. Jiang, Z.-J. Jiang, Int. J.
Hydrogen Energy 42 (2017) 16731–16740.
[257] C. Li, N. Huang, Z. Jiang, X. Tian, X. Zhao, Z.L. Xu, H. Yang, Z.J. Jiang, Electrochim.
Acta 250 (2017) 68–76.
[258] C. Wang, B. Lin, G. Qiao, L. Wang, L. Zhu, F. Chu, T. Feng, N. Yuan, J. Ding, Mater.
Lett. 173 (2016) 219–222.
[259] N. Üregen, K. Pehlivanoğlu, Y. Özdemir, Y. Devrim, Int. J. Hydrogen Energy 42
(2017) 2636–2647.
[260] Y. He, C. Tong, L. Geng, L. Liu, C. Lü, J. Memb. Sci. 458 (2014) 36–46.

Graphical abstract

51
Graphene oxide with exceptional physical, chemical and electrochemical properties
has shown great potential in energy storage devices. Here is an overview of its application in
batteries, capacitors and fuel cells.

Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:

52

You might also like