NiCo2O4 Reffer

Download as pdf or txt
Download as pdf or txt
You are on page 1of 87

Accepted Manuscript

Review and prospect of NiCo2 O4 -based composite materials for


supercapacitor electrodes http://www.journals.elsevier.com/
journal-of-energy-chemistry/
Yanmei Li , Xiao Han , Tingfeng Yi , Yanbing He , Xifei Li

PII: S2095-4956(18)30228-6
DOI: 10.1016/j.jechem.2018.05.010
Reference: JECHEM 606

To appear in: Journal of Energy Chemistry

Received date: 15 March 2018


Revised date: 18 April 2018
Accepted date: 15 May 2018

Please cite this article as: Yanmei Li , Xiao Han , Tingfeng Yi , Yanbing He , Xifei Li , Review and
prospect of NiCo2 O4 -based composite materials for supercapacitor electrodes, Journal of Energy
Chemistry (2018), doi: 10.1016/j.jechem.2018.05.010

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service
to our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. 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.
ACCEPTED MANUSCRIPT

Highlight
 Progress in understanding and modifying NCO from various aspects is

summarized.

 Key strategies for improving the electrochemical property of NCO is discussed.

 Strategies developing NCO-based supercapacitors are discussed.

T
IP
CR
US
AN
M
ED
PT
CE
AC

1
ACCEPTED MANUSCRIPT

Review
Review and prospect of NiCo2O4-based composite materials
for supercapacitor electrodes
Yanmei Lia, Xiao Hana, Tingfeng Yia,c,*, Yanbing Heb,*, Xifei Lid

a
School of Chemistry and Chemical Engineering, Anhui University of Technology, Maanshan

T
IP
243002, Anhui, China

CR
Engineering Laboratory for the Next Generation Power and Energy Storage Batteries, Graduate

School at Shenzhen, Tsinghua University, Shenzhen 518055, Guangdong, China

US
School of Resources and Materials, Northeastern University at Qinhuangdao, Qinhuangdao

066004, Hebei, China


AN
d
Institute of Advanced Electrochemical Energy, Xi'an University of Technology, Xi'an 710048,
M

Shaanxi, China
ED


Corresponding authors.
PT

E-mail address: [email protected] (T. Yi); [email protected] (Y. He).


CE
AC

2
ACCEPTED MANUSCRIPT

Abstract

Supercapacitors known as typical electrochemical capacitors have been considered as

one of the most promising candidates of energy storage systems owing to their

advantages such as high-power density, long life span and lower production cost. The

electrode materials play a crucial role on properties of supercapacitors. Hence, many

T
researches have been focused on the development of novel electrode materials for

IP
high-performance supercapacitors. NiCo2O4 as supercapacitor electrode material has

CR
drawn more and more attentions in recent years due to its outstanding advantages,

US
such as high theoretical capacity, low cost, natural abundance and easy of synthesis.

However, the NiCo2O4 always suffer from severe capacity deterioration because of
AN
the low electrical conductivity and small surface area. Hence, it is necessary to
M

systematically and comprehensively summarize the progress in understanding and

modifying NiCo2O4-based materials from various aspects. In this review, the structure
ED

and synthesis method of NiCo2O4-based materials are discussed in detail. And then,
PT

the major goal of this review is to highlight new progress in using proposed strategies

to improve the cycling stability and rate capacity of NiCo2O4-based materials,


CE

including synthesis, control of special morphologies and design of composite


AC

materials. Finally, an insight into the future research and development of

NiCo2O4-based materials for supercapacitors is prospected.

Key words: NiCo2O4; Electrochemical performance; Synthesis; Morphology;

Supercapacitors

3
ACCEPTED MANUSCRIPT

1. Introduction

The ever-increasing attention on the world energy crisis and environmental

pollution has greatly boosted the development of electrochemical energy storage

devices, such as rechargeable batteries, fuel cells and supercapacitors as energy

storage systems for electric vehicles (EVs), hybrid electric vehicles (HEVs), plug-in

T
hybrid electric vehicles (PHEVs), and smart grids [1–7]. For example, Ni/MH, lead

IP
acid batteries and lithium ion batteries as typical rechargeable batteries have been

CR
widely used as electrochemical energy storage devices and systems [8,9]. However,

US
most of batteries suffer from low power delivery, and cannot satisfy the faster and

higher-power energy requirements. In this situation, supercapacitors (SCs) were


AN
exploited to reserve and deliver energy with high rate capability, which has been well
M

adapted to provide the electricity demand for electric vehicles, tramways,

diesel-engine starting, wind turbines, computers, lasers, and cranes [10,11].


ED

SCs, also known as ultracapacitors or electrochemical capacitors, can be fully


PT

charge-discharge only in a few seconds, leading to very higher charge/discharge

power density (10 kW kg-1) [12]. The Ragone plot as given in Fig. 1(a) exhibits the
CE

relationship of specific power density (W kg-1) vs. specific energy density (Wh kg-1)
AC

for miscellaneous electrical energy storage devices [13]. Obviously, the SCs are

capable of perfectly fill the power-energy gap between conventional dielectric

capacitors with great power output and batteries owning high energy density [14,15].

In short, SCs have the greatest potential in the field of energy storage devices [16].

According to the species of the electrode material and the mechanism of energy

4
ACCEPTED MANUSCRIPT

storage, SCs can originally be divided into three major categories [17–19], including

electrical double-layer capacitors (EDLCs), pseudocapacitors (PCs) and hybrid

capacitors in which the EDLC and PC coupling work together. In principle, the

energy storage mechanism of EDLCs is on the basis of the electrostatic interaction

between ions on the large specific surface area of active electrode materials and

T
electrolytes. As shown in Fig. 1(b), a space charge layer in an electrode, a diffusive

IP
layer in an electrolyte and a tight Helmholtz layer are included in an EDLC, whose

CR
thickness is around 1 nm. A fast charge-discharge process occurring within seconds,

US
and over 100,000 cycles can endure in EDLC systems. Carbon based materials are

regarded as ideal electrode materials for EDLCs owing to their predominant


AN
characteristics such as controllable porous structures, flexible and stretchable, high
M

conductivity, large surface area [20–22]. For instance, the activated carbon (AC),

carbon aerogel, carbon fiber, carbon nanotube (CNT), porous carbon, and graphene
ED

manifest rapid charge storage performance [23]. Zhang et al. [24] discussed the
PT

advances and challenges of stretchable supercapacitors based on carbon materials and

gave an overview of recent advances towards the development of carbon-based SSCs.


CE

Pseudocapacitors mainly store charges by the rapid and reversible redox


AC

reactions on the surface of electrode materials, as shown in the right of Fig. 1(b).

Typically, the transition metal oxides/hydroxides such as RuO2, MnO2, Ni(OH)2,

electrically conducting polymers (polyaniline (PANI), polypyrrole (PPy), etc.) and

their composites could deliver high specific capacitance, but present poor cycle

stability [25]. To improve their energy density, hybrid supercapacitors incorporating

5
ACCEPTED MANUSCRIPT

the favorable merits of EDLCs and pseudocapactiors have been developed, which

sustain high power feature and meanwhile possess high energy density. According to

the difference of two electrode materials adopted, hybrid supercapacitors could be

generally divided into three classes including composite symmetric supercapacitors,

battery-type supercapacitors and asymmetric supercapacitors (ASCs) [26]. Notably,

T
ASCs consist of two different electrodes and are becoming a more popular kind of

IP
supercapacitors, which can achieve high energy density close to batteries and

CR
high-power density like supercapacitors (Fig. 1c) [27].

US
The performance of supercapacitors mainly includes five aspects, including work

voltage, specific capacitance, specific energy (energy density), specific power (power
AN
density) and cycle life [28–31]. It is well reported that the energy storage performance
M

of supercapacitors is directly related to the electrode materials. The preparation of

electrode materials with stable performance and good electrical conductivity has been
ED

a hot research topic [32–35]. To date, grounded on the charge storage mechanism
PT

three kinds of electrode materials have been explored to fabricate the supercapacitors,

including carbonaceous materials, high oxidation state transition metal oxides and
CE

conducting polymers [36–38].


AC

6
ACCEPTED MANUSCRIPT

(a
)

T
IP
CR
(b
)

US
AN
M
ED

(c
)
PT
CE
AC

Figure 1. (a) Ragone plot for various electrical energy storage devices [13]; (b)

Aschematic diagram illustrating the operation principles of: (left) electric double layer

capacitor (EDLC); (right) pseudo supercapacitor (PC) [13]; and (c) Ragone plot of

7
ACCEPTED MANUSCRIPT

ASCs with various typical electrochemical energy storage devices. Times given in the

plot are the time constants of the devices, obtained by dividing the energy density by

the power [27]. Reproduced from Refs. [14,27] with permission from Elsevier.

With the desirable structural features and electrochemical activity, metal oxides

have received more attention as supercapacitor electrodes. Meanwhile, metal oxide

T
could overcome the poor capacity of carbon materials and avoid inferior cycle

IP
stability. In the past, pseudocapacitive materials of various transition metal oxides,

CR
such as MnO2 [39], NiO [40], Co3O4 [41], V2O5 [42] and MoO3 [43] are widely used

US
because of their increasing specific capacitance, both energy and power densities.

Recently, binary or ternary transition metal oxides have attracted extensive attention
AN
because of their outstanding electrochemical performance and electrical conductivity
M

compared with the single-metal oxides. Binary metal oxides such as NiCo2O4,

MnCo2O4, CuCo2O4, and ZnCo2O4 are of the origin of Co3O4. They have provided
ED

excellent performance in supercapacitors [44,45]. It is noteworthy that the electrical


PT

conductivity has the following order: NiCo2O4 > NiO > Co3O4, and many researchers

have also verified that NiCo2O4 possessed a much lower resistance than nickel oxides
CE

and cobalt monoxides. The reaction mechanism for M3O4 (M=Co or Ni) during the
AC

charge storage process is as follows.

M3O4+OH-+H2O↔3MOOH+e- (1)

Benefiting from richer redox reaction from both nickel and cobalt ions than

single component nickel oxides or cobalt oxides, the spinel nickel cobaltite (NiCo2O4)

exhibits wonderful electrochemical performance such as much better electrical

8
ACCEPTED MANUSCRIPT

conductivity and superior electrochemical activity [46,47]. More importantly,

NiCo2O4 has some advantages such as low cost, natural abundance, environmental

safety and low toxicity. Therefore, NiCo2O4 has become a more promising and

flexible alternative electrode material for application in high-performance

supercapacitors and has drawn more and more attention in recent years. Various

T
NiCo2O4 materials with diverse morphologies, such as nanoparticles [48] and

IP
nanowires [49–51] have been reported. NiCo2O4 has multiple redox states and good

CR
electrical conductivity as a P-type semiconductor with a band gap of 2.1 eV [52].

US
Zhang et al. [53] prepared novel ordered meso-NiCo2O4 nanospheres and present the

XRD pattern of meso-NiCo2O4 nanospheres and bulk-NiCo2O4. Obviously, there is no


AN
difference in wide-angle XRD patterns (Fig. 2a). However, in the low-angle XRD
M

patterns (Fig. 2b), a sharp diffraction peak (2θ = 1°) accompanied with a weak peak

(2θ = 1.8°) was founded in the meso-NiCo2O4 pattern, corresponding to the (211) and
ED

(332), which reflects 3D cubic Ia3̅d ordered mesoporous architecture. As shown in Fig.
PT

2(c and d), the NiCo2O4 possesses a face-centered cubic (fcc) spinel structure with

Fd3̅m symmetry with a = 8.11 Å and a strong (100) orientation, where all the nickel
CE

cations take up the octahedral interstices and the cobalt cations are occupied among
AC

the tetrahedral and octahedral gap [54]. In other cases, NiCo2O4 may be properly

regarded as a nickel atom replacing one of the cobalt atoms in Co3O4. Obviously, the

difference in physical properties and electrochemical performance between NiCo 2O4

and Co3O4 is ascribed to the introduce of the nickel atom whose grain size is similar

to that of cobalt atom that causes a little change in the crystal structure, in which the

9
ACCEPTED MANUSCRIPT

defects have some unexpected effects on electrochemical performance. An

investigation of the cation valences in NiCo2O4 has been performed by means of

X-ray absorption near edge structure (XANES), X-ray absorption fine structure

61
(EXAFS) and Ni Mössbauer spectroscopy. The oxidation state distributions of

nickel cobaltite are uncertain, and the ionic configuration of NiCo2O4 is as the

T
following formula: Co1-x2+Cox3+[Co3+Nix2+Ni1-x3+]O4 (0≤x≤1). The cations before the

IP
square brackets are considered to be in the tetrahedral interstices and those within the

CR
square brackets are in the octahedral interstices [55,56]. As reported in some

US
researches [57–59], in an alkaline solution, NiCo2O4 exhibits a potential window

range of 0–0.55 V and its pseudo-capacitive behavior occurred in the alkaline


AN
electrolyte can be presented in the Fig. 2(e).
M
ED
PT
CE
AC

Fig. 2 (a) Wide-angle XRD, (b) low-angle XRD patterns of meso-NiCo2O4

nanospheres (red curve) and bulk-NiCo2O4 (blue curve), (c, d) crystal structure of
10
ACCEPTED MANUSCRIPT

NiCo2O4 (ICSD No. 24211) [53] and (e) schematic of the charge storage mechanism

of NiCo2O4 [57]. Reproduced from Ref. [53] with permission from American

Chemical Society. Reproduced from Ref. [57] with permission from The Royal

Society of Chemistry.

During the charge-discharge process, fast and reversible faradaic reactions occur

T
on the surface of the electrode materials, where the valence state changes of

IP
Co3+/Co4+, as well as M2+/M3+ (M = Co or Ni) exist. Interestingly, the specific

CR
capacitance of NiCo2O4 increased to some extent after several hundred cycles in some

US
research, which was ascribed to their peculiar morphologies and the process of

electrode activation [60,61]. However, the rate capability and electrochemical


AN
reversibility of NiCo2O4 electrodes are poor due to the nature of redox reaction. To
M

resolve this problem, many effective strategies have been reported. Here, the structure

and synthesis method of NiCo2O4-based materials are discussed in detail. And then,
ED

the major goal of this review is to highlight new progress in using proposed strategies
PT

to improve the cycling stability and rate capacity of NiCo2O4-based materials,

including synthesis, control of special morphologies and design of composite


CE

materials. Finally, an insight into the future research and development of


AC

NiCo2O4-based materials for supercapacitors is prospected.

2. Preparation of NiCo2O4 materials

Advanced NiCo2O4 electrode material with unique nanostructure and high

performance is expected by researcher and industry, while it is very difficult to find a

universal and cost-effective routine to synthesize all types of NiCo2O4 nanomaterials

11
ACCEPTED MANUSCRIPT

by a one-pot approach [62–64]. In general, several methods have been used to prepare

nanoscale NiCo2O4 oxide materials with variety of microstructures, including sol-gel,

bio-template, electrodeposition process, electrospinning and hydrothermal. These

synthesis methods are summarized and discussed in detail below.

2.1. Sol-gel method

T
Sol-gel method is a kind of facile and cheap method to produce ideal

IP
nanostructures with homogeneity, porosity and high purity. The sol-gel method

CR
generally covers four steps: the fabrication of the precursor solutions, the formation of

US
intermediate products as sol, the transformation from sol to gel and the calcination

subsequently. In this process, solvents, reaction temperature, reaction time, surfactants


AN
have key influences to get the desired structures with extraordinary electrochemical
M

performances [65]. For instance, in Wu’s [66] work, NiCo2O4 coral-like porous

crystals, nanoparticles and submicron-sized particles (Fig 3b–d) have been


ED

methodically synthesized through a simple sol-gel method, employing H2O


PT

(H2O-DMF) as solvent and citric acid as the chelating ligand. The results indicate that

submicron-sized NiCo2O4 particles show the prime capacitive performance with high
CE

specific capacitance and good cycle stability. Wei et al. [67] reported the synthesis of
AC

NiCo2O4 aerogels through an epoxide-driven sol–gel method for the first time, which

exhibited high specific capacitance of 1400 F g-1 after additional 500 fully activate

cycles of the Ni–Co–O–200 (see Fig. 3f), well reversibility and excellent cycle

stability under a mass loading of 0.4 mg cm-2. The majority of the pores sizes in the

range of 2 to 5 nm were appropriate for supercapacitor applications. Liu and

12
ACCEPTED MANUSCRIPT

co-workers [68] fabricated NiCo2O4 crystals by a sol-gel method, which presents

specific capacitance of 1128 F g−1 at a current density of 5 mA cm−2 and the specific

capacitance retention is as high as 92.5% after 1000 cycles. In succession,

mesoporous NiO/NiCo2O4/Co3O4 composite with high specific capacity (Fig. 3e) and

high performance is fabricated by a sol-gel process and then calcined at 250 °C. The

T
typical synthesize process of a sol-gel process is clarified in Fig 3(a) and the

IP
mechanism of gelation can be described by the following reaction equation where M

CR
represents nickel or cobalt ions, and A- represents Cl- or NO3- ions [69].

US (2)
AN
Moreover, the submicron-sized NiCo2O4 particles with various morphologies and
M

excellent properties were prepared using chelating ligand, different surfactants and

organic-water binary solvent via a facile sol-gel process. Extensive researches are
ED

on-going through a sol–gel method to fabricate a better modality of NiCo2O4 so as to


PT

achieve a higher specific capacitance and excellent cycle stability. The sol–gel process

has been confirmed to be a significant way to prepare NiCo2O4 electrode materials


CE

with low-cost and high electrochemical performance.


AC

13
ACCEPTED MANUSCRIPT

(a)

T
(b) (c) (d)

IP
CR
(e)
US (f)
AN
M
ED
PT

Figure 3 (a) Synthesis procedure for mesoporous NiO/NiCo2O4/Co3O4 composite [69],


CE

FESEM images of as-obtained (b) sample (I), (c) sample (III) and (d) sample (II) [66],

(e) specific capacitances of Co3O4, NiO, NiCo2O4, and NiO/ NiCo2O4/Co3O4


AC

composite at a controlled current density [69] and (f) stability test in terms of specific

capacitance for sample Ni–Co–O–200 [67]. Reproduced from Refs. [66,67] with

permission from Wiley-VCH. Reproduced from Ref. [69] with permission from

American Chemical Society.

14
ACCEPTED MANUSCRIPT

2.2. Template method

Mesostructured transition metal oxides have dramatic advantages, such as high

specific surface areas, controllable pore sizes and structures. These characteristics are

significant for electrode materials because the particular morphologies can further

accelerate the transportation of charge/ions in the electrode surface and electrolyte.

T
Furthermore, the mesopores allow for the fast-redox reaction [70]. However, it is not

IP
easy to prepare mesoporous binary metal oxides due to the thermal instability. As the

CR
advancement of novel methods for the synthesis of mesoporous binary metal oxides,

US
the template method is widely adopted combined with solvothermal/hydrothermal

method for the fabrication of mesostructured binary metal oxides [71]. For instance,
AN
Yuan et al. [72] synthesized the uniform mesoporous hollow NiCo2O4
M

sub-microspheres using silica spheres as a hard-template (Fig. 4a). The obtained

mesoporous hollow NiCo2O4 sub-microspheres self-assembled thoroughly by


ED

ultrathin nanosheets with a thickness of a few nanometers. They are very uniform in
PT

size and structurally robust by taking advantage of the in situ template removal (Fig.

4b). What’s more, the hierarchical hollow NiCo2O4 sub-microspheres exhibit high
CE

specific capacitance of 678 F g-1 and excellent cycling stability of 87% (10 A g-1) after
AC

3500 successive cycles. In recent years, biological templates are abundant reserves,

renewable, cheap, readily available and environmentally benign compared to artificial

templates, such as graphene and carbon nanotubes (CNTs). There are many kinds of

templates with inherently hierarchical and complex structures. Furthermore, even the same

biology may also have different multilevel structures and morphologies. Therefore,

15
ACCEPTED MANUSCRIPT

bio-based carbon templates attract numerous researchers due to their natural highly

ordered architecture with superior physicochemical properties, such as large surface

area, great flexibility, and good mechanical strength. In addition, the bio-based carbon

templates can be easily obtained without intricate chemical/physical conditions.

Especially, it is propitious to electrolyte penetration and ion diffusion when they serve

T
as the conductive substrate [73–76]. Thus, synthesis of simple and low-cost electrode

IP
materials with delicate framework derived from bio-template have obtained extensive

CR
exploration in recent years [77]. As shown in Fig. 4(c), Xiong et al. [78] employed

US
mollusc shell as template and the NiCo2O4 nanowires grew directly on it by a facile

hydrothermal strategy (Fig. 4d and e). Meanwhile, the shell based macroporous
AN
carbon material (MSBPC) with well-aligned microstructure and large number of pores
M

honeycomb-like architecture was served as the conducting substrate. The

as-synthesized NiCo2O4/MSBPC electrode presented large specific capacitance of


ED

1696 F g-1 (1 A g-1), super rate capability (even 421.8 F g-1 at 50 A g-1), and excellent
PT

cycling stability (88% retention after 2000 cycles), revealing that the gained 3D

carbon skeleton consisting of hexangular and highly ordered channels can efficiently
CE

improve electrolyte penetration and rapid electron transfer. In addition, Shen et al. [79]
AC

exploited a novel and efficient self-template method for the preparation of uniform

NiCo2O4 hollow spheres with a core-in-double-shell interior structure.

16
ACCEPTED MANUSCRIPT

(a) (b)

(d) (c)

(e)

T
IP
CR
(f) (g)

US
AN

Figure 4. (a) Schematic illustration of the formation of the mesoporous hollow


M

NiCo2O4 sub-microspheres, (b) FESEM images for the hollow NiCo2O4


ED

sub-microspheres [72], (c) preparation process of NiCo2O4/MSBPC composites, (d)

SEM and (e) high-magnification SEM images of the NiCo2O4 nanowires [75], (f)
PT

specific capacitance at different current density and (g) cycling performance of the
CE

NiCo2O4 electrode at a current density of 5 A g-1 [76]. Reproduced from Ref. [72]

with permission from Royal Society of Chemistry. Reproduced from Refs. [75,76]
AC

with permission from Weilly-VCH.

As exhibited in Fig. 4(f and g), the NiCo2O4 hollow spheres electrode possess

excellent high-rate capability. At current densities of 1, 2, 5, 10, and 15 A g-1, the

electrode delivers high capacitance of 1141, 1048, 965, 862, and 784 F g-1,

17
ACCEPTED MANUSCRIPT

respectively. Notably, the specific capacitance is about 914 F g-1 at 5 A g-1 after 4000

cycles with a high capacity retention of 94.7%. These works demonstrate that

template-assisted method has specific advantage in preparing high-performance

electrode materials [80]. Therefore, it is an effective way to prepare high-performance

transition metal oxide electrode materials with high theoretical capacity for

T
supercapacitors by using template-assisted method because of the advantages of the

IP
particular natural structure of template.

CR
2.3. Electrospinning

US
Electrospinning is a comparatively complicated but multifunctional and

cost-efficient way to fabricate well-established nanotubes and nanofibers with size in


AN
the range from several micrometers to tens of nanometers. During the preparation
M

process, many important factors can influence the morphologies, size distributions,

and diameters of the nanotubes and nanofibers, such as the concentrations of metal
ED

precursor, viscosity of precursor solution, type of polymer, and electrospinning


PT

parameters (applied voltage, working distance, and solution feed rate). The composite

electrode materials with high capacitances and long cycle stability are achieved by the
CE

novel electrospinning method using the synergistic effect of different factors. For
AC

example, Li et al. [81] synthesized one-dimensional porous NiCo2O4 nanotubes

(NCO-NTs) with porous architecture and hollow inside via electrospinning method.

As shown in Fig. 5(b and c), the prepared nanofibers possess a mean diameter of

about 89 nm and a length up to several micrometers (NCO-NFs). Because of their

unique structure, the fabricated NCO-NT electrode displays a high specific

18
ACCEPTED MANUSCRIPT

capacitance of 1647 F g-1 at 1 A g-1, excellent rate capability of 77.3% capacity

retention at 25 A g-1, and outstanding cycling stability of 93.6% after 3000 cycles (see

Fig. 5e).

Furthermore, electrospinning has been regarded as a valid approach to prepare

continuous fiber membranes as a binder-free carbonaceous template with micronsized

T
or even nano-sized diameter, large specific surface area and excellent mechanical

IP
performances [82–84]. Owning to these prominent characteristics, electrospun carbon

CR
nanofiber (CNF) membranes are widely used as the electrode materials of

US
supercapacitors [85–87]. In Addition, the electrospun CNF can remarkably enhance

the conductivity of transition metal oxides or conductive polymers, which is favorable


AN
to maintain high specific capacitances of transition metal oxides even at larger current
M

density [88,89]. The CNFs suffer a closed attachment for the NiCo2O4 particles,

which could undertake good mechanical adhesion and excellent cycling stability. Lai
ED

et al. [90] developed a facile approach of growing δ-phase and γ-phase MnO2 with
PT

distinctly different morphologies on highly conductive NiCo2O4-doped carbon

nanofibers (NCCNFs) through the combination of electrospinning, solution


CE

co-deposition, and redox deposition methods to form core-sheath nanostructures with


AC

high performance. The formation process of NCCNF@MnO2 nanosheet and nanorod

is illustrated in Fig. 5(a). The relationship between scan rate and the specific

capacitance shows an excellent rate stability (Fig. 5d). For the NCCNF@MnO2 NR

hybrid membrane, the specific capacitance reaches 146 F g−1 (based on whole

materials) even at a scan rate of 100 mV s−1.

19
ACCEPTED MANUSCRIPT

(a)

(b) (c)

T
IP
CR
(d) (e)

US
AN
M

Figure 5. (a) Schematic illustration for the preparation of NCCNF@MnO2 NS and NR


ED

hybrid membranes [90], SEM images at high and low (inset) magnifications for (b)

NCONFs and (c) NCO-NBs [81], (d) rate stability of bare CNF, NCCNF,
PT

CNF@MnO2 nanorod, and NCCNF@MnO2 nanorod membranes at different scan


CE

rates [90] and (e) cycling stability of NCO-NTs at a current density of 10 A g-1. The

inset shows the first ten charge/discharge curves of NCO-NT [81]. Reproduced from
AC

Ref. [81] with permission from Weilly-VCH. Reproduced from Ref. [90] with

permission from American Chemical Society.

2.4. Electrodeposition

The electrodeposition process occurs in the electrolysis of related plating

solutions where micron- or submicron-size molecules are dissolved, and large


20
ACCEPTED MANUSCRIPT

amounts of particles according to the stoichiometric ratio are embedded in the

electrochemically produced composite with special performance [91]. In principle, the

electrodeposition process can be classified into three steps, including preparation of a

precursor solution, co-electrodeposition of precursor, and thermal decomposition

transformation. Facile electrodeposition synthesis methods and subsequent thermal

T
treatment have been diffusely adopted to prepare 2D metal oxide nanosheets. Recently,

IP
some conductive substrates also have been served as current collector to further

CR
improve electrochemical performance of the electrode materials, such as

US
single-walled carbon nanotube, carbon fabric and nickel foam. For instance, the

carbon fabric (CF) is an ideal conductive substrate for directly growing active
AN
material because of the excellent electronic conductibility, good flexibility, advanced
M

mechanical strength and inexpensive [92]. Therefore, some researchers have used CF

as conductive substrate to support active material. For example, An et al. [93] reported
ED

the facile co-electrodeposition of NiCo2O4 radially on carbon cloth with good


PT

electrochemical properties. The schematic illustration of the synthesis of

honeycomb-shaped NiCo2O4/CC (HSNC) is displayed in Fig. 6(a). With deposition


CE

time goes on 120 s, NiCo2O4 nanoparticles would voluntarily be attached to the


AC

generated sheets because of the minimization of facet free energy (Fig. 6b) [94].

21
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Figure 6. (a) Schematic illustration of electrodeposition process for forming the


PT

NiCo2O4 samples, (b) FESEM images of samples for 120 s [93], (c)

high-magnification SEM image of the N-CNT/NiCo2O4 nanosheets [95], (d)


CE

corresponding specific capacitance at different current densities and (e) cycling


AC

performance at a current density of 10 A g-1 [92]. Reproduced from Refs. [92,93] with

permission from American Chemical Society. Reproduced from Ref. [95] with

permission from Springer.

Wu et al. [95] prepared the homogeneous thin NiCo2O4 nanosheets onto

nitrogen-doped carbon nanotubes (N-CNTs) through a facile electrochemical

22
ACCEPTED MANUSCRIPT

deposition method. The electrodeposited cubical- and hexagonal-shaped NiCo2O4

particles with some micro porous space were shown in Fig. 6(c). The electrochemical

performances were displayed in Fig. 6(d and e). The SCs are 2658, 2200, 2170, 2000,

1866 F g−1 at different current densities of 2, 5, 10, 15, and 20 A g−1, respectively. The

result reveals about 70% of the initial SC is remained even at a high current density of

T
20 A g−1, indicating that the obtained NiCo2O4/CF composites possess extraordinary

IP
rate performance. Moreover, the SC of the electrode decays approximately 20% after

CR
3000 cycles, indicating high electrochemical stability for long cycle life. Besides, the

US
electrodeposition is also one of the most common methods in preparing various

nanomaterials composites. Several composite electrodes such as Co3O4@NiCo2O4 [96]


AN
and Ni(OH)2@NiCo2O4 [97], have also been synthesized by the electrodeposition
M

route. However, the precise relationship between the composite structure and

electrochemical properties has not been verified yet.


ED

2.5. Hydrothermal method


PT

From the discussion mentioned above, it can be found that the synthesis method

plays a key role on the preparation of nanomaterials and improvement of


CE

pseudocapacitive performance. In recent years, hydrothermal reaction also has been


AC

considered as a promising approach to synthesize nanostructures due to its

environmentally benign, and good controllability of reaction condition and controlled

production of various nanostructures at high pressure and moderate temperature

including nanowires, nanosheets, nanoflowers and so on [98]. The NiCo2O4

precursors were also prepared using hydrothermal reaction by a mixture of related

23
ACCEPTED MANUSCRIPT

metal salts reagent, and then were heated in a sealed Teflon-lined stainless-steel

autoclave. As a wide-used method of preparing NiCo2O4, it is necessary to synthesis

highly monodispersed nanoparticles with controllable size and unique morphology

[99]. In addition, many groups have made great efforts to optimize the reaction

conditions to form special morphologies with enhanced electrochemical performance

T
of NiCo2O4 electrodes [100–104]. Wang et al. [105] successfully prepared 3D

IP
hierarchical NiCo2O4 with tremella-like (T-NiCo2O4), flower-like (F-NiCo2O4),

CR
urchin-like (U-NiCo2O4) and pine-needle-like (P-NiCo2O4) structures through a facile

US
hydrothermal method. As shown in Fig. 7(a), the morphologies of NiCo2O4 can be

controlled by varying the categories of alkali sources and adding NH4F, and the
AN
electrochemical performance can be affected. The important roles of alkali source and
M

NH4F in fabrication of these nanostructures have been researched. The as-prepared

four NiCo2O4 electrode materials were designed to continuous galvanostatic


ED

charge–discharge (GCD) measurements at a large current density (10 A g-1) for 1000
PT

cycles. As displayed in Fig. 7(b), there was no obvious capacitance decline observed

for all the four materials, revealing their excellent cycling ability.
CE
AC

Figure 7 (a) Schematic diagram illustrating the hydrothermal process of fabricating


24
ACCEPTED MANUSCRIPT

F-NiCo2O4, T-NiCo2O4, U-NiCo2O4 and P-NiCo2O4 and (b) cycling performance of

the different NiCo2O4 electrodes at high current density of 10 A g-1 [104]. Reproduced

from Ref. [104] with permission from The Royal Society of Chemistry.

Very recently, some surfactants, such as polyvinylpyrrolidone (PVP), polyvinyl

alcohol (PVA) and hexadecyltrimethyl-ammoniumbromide (CTAB) [106–108], have

T
also been used as the structure-directing agent to prepare electrode materials with

IP
unique morphologies and high electrochemical performance. However, the use of

CR
surfactants and growth-directing agents usually reduces the charge transfer in the

US
target products and thus weakens the electrode material performance. Fortunately, a

unique laser ablation in liquid (LAL) method could conduce to prepare high activity
AN
precursors by prefabricating ternary oxide nanocrystals (TONs) with a clean surface,
M

which can acquire enhanced charge transport in the electrochemical energy storage

devices [109]. It is necessary to develop other feasible technology to improve the


ED

performance of electrode materials, such as ion-doping and transition metal oxides


PT

modification. For example, a porous vanadium doped-zinc–nickel–cobalt ternary

oxide (VZnNiCo) nanostructure has been used to high performance flexible


CE

supercapacitor [3]. Moreover, dandelion-like Fe-doped NiCo2O4


AC

microspheres@nanomeshes (Fe-NCO-M@N-1h) prepared by a facile hydrothermal

method shows high specific capacitance, excellent rate performance and remarkable

cyclic stability [5]. Successively, self-supported 3D ultrathin porous quaternary

Zn-Ni-Al-Co oxide nanosheets with hierarchical, porous structures, are regarded as

auspicious electrodes for high-performance hybrid supercapacitors [7]. In summary,

25
ACCEPTED MANUSCRIPT

various nanostructured types NiCo2O4-based materials with 1D, 2D or 3D

morphology can be obtained by using facile hydrothermal method.

2.6. Summary of preparation methods

In short, this section suggests five means to synthesize NiCo2O4-based electrode

materials, including sol-gel, bio-template, electrodeposition, electrospinning and

T
hydrothermal methods. Each method has advantages and drawbacks. The sol-gel

IP
method is inclined to prepare 0 D structures such as nanoparticles with homogeneity,

CR
porosity and high purity. The template method, especial biological templates with

US
abundant reserves, renewable and environmentally benign, can be fabricated highly

ordered architecture electrode materials with special shapes. Electrospinning is usually


AN
used in preparing one-dimensional (1D) fiber with a diameter in the range from several
M

micrometers to tens of nanometers. Electrodeposition process possesses some significant

advantages to deposit NiCo2O4 onto conductive substrates and to combine NiCo2O4 with
ED

some other high-performance materials with favorable synergistic effect, which can be
PT

controlled by changing concentration and pH value of solution and applying overpotential

or current density. By comparison, the hydrothermal process is a direct and effective


CE

way to synthesize one-dimensional (1D), two-dimensional (2D) and


AC

three-dimensional (3D) nanostructured NiCo2O4-based materials with unique

architectures and superior electrochemical performances by adjusting temperature

parameters, the kinds and concentration of surfactants. Many researchers have

explored to prepare different morphologies of NiCo2O4 including 1D, 2D and 3D

nanostructures. Compared with 1D and 2D nanostructures, 3D nanostructure has more

26
ACCEPTED MANUSCRIPT

advantages in energy storage materials. 3D structural morphology not only forms the

desired hierarchical porous continuous channels for electrolyte ions by serving as an

ion buffering reservoir, but also has higher electrical conductivity and excellent

structural mechanical stability, which can alleviate the volume expansion effect during

charge/discharge process.

T
3. Nanostructured NiCo2O4 electrodes with different morphologies

IP
3.1 Pristine NiCo2O4

CR
As we know, the physical and chemical properties of materials are dramatically

US
changed when the particle size is decreased to the nanoscale. Nanomaterials are

expected to bring important improvements for energy storage devices [110]. The
AN
cross-linked network in nanostructures includes nanowires, nanoneedles, nanosheets,
M

nanoarrays and nanoflowers. 1D nanostructures can provide direct current pathways,

shorten the ion diffusion length, reduce the charge–discharge time, increase the
ED

electrolyte–electrode contact area, limit mechanical degradation, and accommodate


PT

volume expansion [111]. Due to the inherently high specific surface area and

distinctive electronic properties of ultrathin sheets, 2D nanomaterials are popular for


CE

energy storage and conversion system. Besides the above advantages, 2D materials
AC

also show other special properties such as flexibility, high packing densities,

mechanical stability and transparency, which prompt them fit for the development of

flexible and wearable electrochemical capacitors [112]. Compared with 1D and 2D

nanostructures, the complex 3D structures contributed by 1D nanostructures possess

unique properties because they incorporated the advantages of all related nanoscale

27
ACCEPTED MANUSCRIPT

building-blocks [113]. In addition, homogeneous hierarchical 3D nanostructures

assembled from nanotubes, nanoflakes, nanowires, or nanorods with several levels of

structure, also have exhibited super capacitive property because of the large SSA

(specific surface area) and high density of defects arising from nanoscale building

blocks and have numerous promising applications in many professional fields such as

T
chemical sensors [114], catalysis [115], batteries [116,117], electrochromic devices

IP
[118] and supercapacitors. Generally, it is important for SCs to achieve large specific

CR
surface area and fitting pore size distribution suitable for decreasing the consumption

US
of electrolyte by regulating porous structure of electrode materials. For example,

Liang et al. [119] successfully prepared NiCo2O4 nanosheets@halloysite nanotubes


AN
(Fig. 8a and b) by a facile coprecipitation route followed by a thermal annealing

treatment. The specific capacitance can reach 1728 F g−1 at the end of 8600 cycles at a
M

current density of 10 A g−1 and a capacity loss of 5.26%. As shown in Fig. 8(c and d),
ED

novel hierarchical NiCo2O4 nanosheets@hollow microrod arrays (NSs@HMRAs)


PT

were prepared by a simple and environment-friendly template-assisted

electrodeposition followed through thermal annealing [120]. Owing to their unique


CE

nanostructures, the NiCo2O4 NSs@HMRAs showed a high specific capacitance (Csp)


AC

(678 F g-1 at 6 A g-1) and excellent cycle stability (Csp retention of 96.06% after 1500

cycles). Additionally, a kind of NiCo2O4 pseudocapacitive materials with an

urchin-like hollow micro-sphere structure (UHMS) ware designed and fabricated (Fig.

8e and f). The as-synthesized NiCo2O4 UHMS sample exhibited an improved specific

capacitance of about 942.2 F g-1 at a current density of 0.5 A g-1 and excellent rate

28
ACCEPTED MANUSCRIPT

capability [121].

T
IP
CR
US
AN
M
ED
PT
CE
AC

Fig. 8 (a) SEM and (b) TEM image of the halloysite nanotubes [120], (c) and (d) TEM

images of NiCo2O4 NS@HMRAs [119], (e) SEM and (f) TEM images of the final

NiCo2O4 UHMS [121]. Reproduced from Refs. [119,121] with permission from The

Royal Society of Chemistry. Reproduced from Ref. [120] with permission from

American Chemical Society.

29
ACCEPTED MANUSCRIPT

Traditional dielectric capacitors and batteries usually include a sandwiched

separator between two electrodes in an ionic liquid electrolyte. The electrodes usually

include binders, and the binders often affect the performance of the electrodes, and

then affect the practical engineering applications of the device, such as wearable

flexible SC. Consequently, it is necessary to develop a new technique, by which

T
nanostructured electrode materials can be in-situ formed on the surface of conductive

IP
substrates, especially on 3D porous substrates. This is a binder/additive-free electrode

CR
for electrochemical capacitors. These binder/additive-free electrodes avoid the

US
“non-activated surface” in traditional slurry-derived electrodes. In addition, it also

brings high charge-discharge rate capability due to the shortened transport path length
AN
for both electrons and ions and is beneficial to the charge transfer. And it also
M

provides effective interface contact between current collector and active materials,

resulting in highly enhanced electrochemical performance. Compared with powdery


ED

nanomaterials, self-supported nanoelectrodes offer particular benefits over the


PT

slurry-casting electrodes such as enhanced utilization of electroactive materials and

improved charge/ion transport efficiency. One of the most common ways of getting
CE

nanoelectrodes is to directly fabricate electroactive materials into the forms of


AC

free-standing either 1D nanoarrays, 2D nanoarrays, or 3D nanoporous architectures on

conductive substrates [122–124]. Usually, these nanoelectrodes are called as

self-supported uniform nanoelectrodes (Fig. 9), because the nanoelectrodes are

nanoarrays of only single electroactive material. Additionally, there are also

self-supported heterogenous nanoelectrodes for supercapacitors that mean the

30
ACCEPTED MANUSCRIPT

nanoarrays include either electroactive substances with nanostructured current

collectors/supporters or more than one type of electroactive substances in single

nanoelectrodes (Fig. 9). Generally, self-supported heterogeneous nanoelectrodes with

a core-shell structure consists of a shell of active materials uniformly deposited on a

core of independent nanostructured current collectors or another kind of active

T
materials that are directly grown on a conductive substrate [125].

IP
CR
US
AN
M
ED
PT
CE
AC

Fig. 9 Schematically illustrating the typical geometries of self-supported

homogeneous and heterogeneous nanoelectrodes for supercapacitors [125].

Reproduced from Ref. [125] with permission from Wiley-VCH.

Chen and co-workers directly grew NiCo2O4 nanosheets on flexible carbon fabric

31
ACCEPTED MANUSCRIPT

through an electrochemical deposition method and the specific capacitances (SCs) of

composites are as high as 2658 F g-1 (2 A g-1) [126]. Zou et al. reported a 3D

micro-sphere composed of radial chain-like NiCo2O4 nanowires with a SC of 1284 F

g-1 (2 A g-1) by a facile hydrothermal route and thermal decomposition [127]. To

improve the electrochemical performance of NiCo2O4, the nanostructured NiCo2O4

T
(such as nanowires, nanosheets, nanotubes) with high specific capacitance and

IP
excellent rate capabilities can be used as substrates for NiCo2O4-based composite.

CR
Other compounds directly grow on these NiCo2O4 substrates, and can form

US
NiCo2O4-based composite, which has more excellent electrochemical performance

than pristine nanostructured NiCo2O4.


AN
3.2. NiCo2O4-based composite materials
M

Unfortunately, single NiCo2O4 often exhibits limited kinetics during the redox

reaction because of the low electrical conductivity or easily agglomerate in substrates,


ED

leading to the poor capacitance, inferior cycling stability and bad rate performance. In
PT

addition, the active material near the bottom is difficult to participate in the

electrochemical charge storage process because of the limited electrolyte diffusion in


CE

supercapacitor electrodes, which results in an unsatisfactory ASC (areal specific


AC

capacitance). An efficient solution is to design the nanocomposite architectures.

Currently, it was discovered that NiCo2O4 with well-organized nanostructures, such as

nanowires, nanosheets, nanorods and nanotubes, can combine with other kinds of

materials like carbon, metal oxides/hydroxide, conducting polymers, metal-organic

frameworks (MOFs) and covalent-organic frameworks (COFs) to improve the

32
ACCEPTED MANUSCRIPT

performance of NiCo2O4. The resulting functional hierarchical nanostructures or

“core-shell” configurations contribute to a more high-level of electrochemical

performance [128]. The electrode material design with “core” and “shell”structure

can achieve a desired synergistic effect [129]. This effective approach to fabricate

hybrid materials composites not only possess the advantages of all the constituents,

T
but also overcome the disadvantages of individual components. Hence, construction

IP
of different structures with the combination of two different materials has been proved

CR
to be a promising strategy to boost the electrochemical performance of metal oxides

[130].

3.2.1 NiCo2O4-metal oxide composites


US
AN
Recently developed hierarchical nanostructures obtained by metal oxide directly
M

grown on an active material (current collector) show remarkable merits [131].

Generally, they can enlarge the contact area between active materials and the
ED

electrolyte, increase the porosity, and provide more active materials in a given unit
PT

area without any auxiliary components, resulting in higher electrochemical properties

[132,133]. Recently, significant efforts have been taken to controllable synthesis of


CE

perfected core/shell nanowire heterostructures by combination of two types of


AC

materials, which show improved electrochemical performances. For example, Chen et

al. [134] obtained a heterostructures composite by coating a NiWO4 nanosheets on

self-supported NiCo2O4 nanowire arrays, which presents a high enhancement of

specific capacitance of 1384 F g-1 at a current density of 1 A g-1 and superior cycling

stability (87.6% retention over 6000 cycles at a current density of 5 A g-1). However,

33
ACCEPTED MANUSCRIPT

the pristine NiCo2O4 electrode only shows a specific capacitance of 865.3 F g-1 at the

corresponding current density. NiCo2O4 has relatively high electrical conductivity but

low capacitance. Conversely, MnMoO4 possesses high capacitance but inferior

electrical conductivity. The combination between MnMoO4 and NiCo2O4 can reach

their full advantages of both materials. For example, Wang and co-workers [135]

T
synthesized the NiCo2O4@CoMoO4 nanowire/nanoplate arrays, and the capacitance

IP
could reach 14.67 F cm-2 at the current density of 10 mA cm-2, with a capacitance

CR
decrease of 25.9% after 1000 cycles. Besides, many groups [136–138] grew MnMoO4

US
nanowires onto the backbone material NiCo2O4 nanowire arrays in the aqueous

solution system to prepare hierarchical 3D heterostructured NiCo2O4@MnMoO4


AN
nanocolumn arrays (NCAs) on Ni foam. The crystal growth method during the

complicated nano-architecture process was “self-assembly” and “oriented attachment”.


M

As shown in Fig. 10(a, b), firstly, the NiCo2O4 nanowire array was directly grown on
ED

nickel foam. In the hydrothermal process, the as-prepared NiCo2O4 nanowires on


PT

nickel foam acted as templates and then MnMoO4 nanosheets were uniformly

wrapped in the surface of the NiCo2O4 nanowires, forming NiCo2O4@MnMoO4


CE

composite nanocolumn arrays. The core-shell structure could supply short


AC

electron/ion transfer path and significantly improve the electrochemical property.

The as-synthesized NiCo2O4@MnMoO4 and NiCo2O4@NiMoO4 core–shell

nanoflowers are uniformly attached on the substrate, the MnMoO4 and NiMoO4

nanoflowers are very thin, which can be observed in the magnified images of Fig.

10(c–e). NiCo2O4 nanowire is fully wrapped by a great amount of MnMoO4

34
ACCEPTED MANUSCRIPT

nanosheets. The electrochemical test reveals that the obtained hybrid

NiCo2O4@MnMoO4 electrode exhibits superior pseudocapacitor performances with

high capacitance of 1705.3 F g-1 at 5 mA cm−2, and retaining 92.6% after 5000 cycles.

However, the bare NiCo2O4 electrode only shows a capacitance of 839.1 F g−1 and a

retention of 90.9% at the corresponding current density after 5000 cycles. In addition,

T
the charge-transfer resistance (Rct) of NiCo2O4 and NiCo2O4@MnMoO4 electrodes

IP
before and after cycles is investigated. It can be observed that the Rct increased with

CR
the increase of cycling number with 0.12 Ω at 50 cycles, 0.17 Ω at 2500 cycles, and

US
0.56 Ω at 5000 cycles, respectively. There is a conductance deterioration during

cycling. However, the Rct is still satisfactory, indicating that there is no great change
AN
of Rct for both electrodes even after 5000 cycles. Notably, the NiCo2O4@MnMoO4
M

shows higher ASC (area specific capacitance) than pristine NiCo2O4. As shown in Fig.

10(f), the ASCs of NiCo2O4@MnMoO4 are 1.91, 1.78, 1.62, 1.39, and 1.19 F cm−2 at
ED

current densities of 1, 2, 5, 10, and 20 mA cm−2, respectively. However, the


PT

corresponding ASCs of pristine NiCo2O4 are only 0.93, 0.84, 0.73, 0.62, and 0.52 F

cm−2, respectively. Obviously, NiCo2O4@MnMoO4 has much higher ASCs than the
CE

pristine NiCo2O4 at each current density. In addition, the cycling stability of the
AC

NiCo2O4@MnMoO4 hybrid electrodes was more excellent than NiCo2O4 electrodes,

as shown in Fig. 10(g).

35
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Fig. 10. (a) Schematic fabrication process of NiCo2O4@MnMoO4 hybrid NCAs on Ni

foam [136], (b) schematic mechanism model of NiCo2O4@MnMoO4 core−shell


PT

nanoarrays [138], (c) typical SEM images of the NiCo2O4@NiMoO4 nanoflowers


CE

(NFRs) on Ni foam [137], (d) TEM image of NiCo2O4@MnMoO4, the inset image

shows the SAED of NiCo2O4@MnMoO4 corresponding to the area where both core
AC

and shell are characterized as the red circle assign [138], (e) TEM image of

NiCo2O4@MnMoO4 (the inset image shows the SAED of NiCo2O4@MnMoO4

corresponding to the area where only shell is characterized as the red circle assign

[138]), (f) ASC and (g) cycling performance of NiCo2O4 and NiCo2O4@MnMoO4

electrodes at various current densities [136]. Reproduced from Refs. [136,138] with
36
ACCEPTED MANUSCRIPT

permission from American Chemical Society. Reproduced from Refs. [137] with

permission from The Royal Society of Chemistry.

The conductivity of NiCo2O4 is two orders of magnitude larger than NiO or

Co3O4, but it is necessary to explore the effects of the combination between NiCo2O4

T
and other oxides on the electrochemical performance. In fact, the NiCo2O4 composites

IP
are also potential candidates for high-performance supercapacitor. NiCo2O4, NiO,

CR
ZnO and Co3O4 were P-type semiconductors with the energy band gap of 2.1 [139],

US
3.6 [140], 3.37 [141] and 2.2 eV [142], respectively. To forecast the electrochemical

performance of these transition metal oxides electrode materials, it is necessary to


AN
consider essential influences, such as pore structure [143–146], the electrical
M

conductivity [147,148] and the degree of crystallization [149,150]. The electrical

conductivity of transition metal oxides can be improved by changing band effects, and
ED

the improved electronic conductivity can usually the improve charge-transfer


PT

processes and reduce internal resistance of the electrodes [151,152]. The

electrochemical activity of metal oxides usually depends on the crystallinity. The


CE

appropriate microstructure will contribute high specific surface area and the
AC

electrolyte permeation [153−155]. Interestingly, smarter integrated architectures

combined with different genres of materials, such as MnO2, NiO, and Co3O4, possess

many competitive advantages, including rich approachable electroactive sites, high

electron aggregation efficiency, short ion transfer pathways, and even excellent

synergetic effects or multifunctional properties of the nanostructure components [156].

37
ACCEPTED MANUSCRIPT

For these reasons, it is vital to coordinate microstructure, crystallinity, and electrical

conductivity of these metal oxides-based materials for their potential application in

ECs. In summary, researchers manifest that the fabrication of oxides composite with

unique structure is an effective way to improve electrochemical behavior of the

electrodes.

T
Recently, Zhou et al. [157] successfully anchored two types hierarchical

IP
core–shell arrays, namely, homogeneous NiCo2O4 nanosheet@NiCo2O4 nanorod and

CR
heterogeneous NiCo2O4 nanosheet@NiO nanoflake (NiCo2O4 sheets regard as the

US
core skeleton for bearing shell materials) via facile solution hydrothermal methods

and chemical bath deposition process in combination with a simple thermal treatment.
AN
NiCo2O4 nanorods disperse on the NiCo2O4 nanosheets in many directions. NiCo2O4
M

nanorods and NiCo2O4 nanosheets are interconnected. Similarly, for the

NiCo2O4@NiO hybrid nanoarrays, the whole nanosheets obviously become thicker


ED

and the surfaces are fully covered by plentiful leaf-like NiO nanoflakes, forming a
PT

highly porous core–shell structure with convenient ion/electron transport paths,

enhanced reactivity, and improved structural stability (Fig. 11a). When the two
CE

as-synthetized hierarchical nanoarrays were evaluated as supercapacitor electrodes,


AC

they exhibited high specific capacitance and enhanced rate capability, and improved

cycling stability. It is observed from Fig. 11(h), that the specific capacitances

retentions of NiCo2O4@NiCo2O4 and pristine NiCo2O4 nanosheets are 76.4% and

73% when the current densities fall from 0.5 to 20 A g-1, and both electrodes are

higher than the NiCo2O4@NiO (43.2%). If excessive amounts of NiO nanoflakes

38
ACCEPTED MANUSCRIPT

cover the NiCo2O4 nanosheets, the formed NiO shell will impede electrolyte diffusion

during rapid charge/discharge processes, resulting in the low electrochemical

performance of NiCo2O4@NiO. Even so, the NiCo2O4@NiO nanoarrays also exhibit

higher specific capacitances (955 F g-1) than the bare NiCo2O4 nanosheets (685 F g-1)

at high current density of 20 A g-1. Hierarchical NiCo2O4@NiCo2O4 core/shell

T
nanoflake arrays (Fig. 11b) were also synthesized by Liu et al. [158] via the facile

IP
hydrothermal and chemical bath deposition for high-performance supercapacitor

CR
materials. The TEM images of NiCo2O4 nanosheets arrays and NiCo2O4@MnO2

US
core–shell nanosheet arrays were displayed in Fig. 11(f and g). Notably, NiCo2O4

nanosheets with wrinkles are beneficial to the composite between NiCo2O4 nanosheets
AN
and MnO2 nanosheets because of the enhanced specific surface area. In addition, it
M

seems that the NiCo2O4 nanosheets were covered with ca. 40–70 nm MnO2

nanosheets. As shown in Fig. 11(i), the effective electron transfer and faster ion
ED

diffusion in the core/shell nanoflake arrays are further verified by the EIS
PT

measurements. The semicircle of the Nyquist plot corresponds to the Faradic reactions

and its diameter stands for the interfacial charge-transfer impendence, and the result
CE

indicates that the interfacial charge-transfer impendence of the NiCo2O4@NiCo2O4


AC

core/shell nanoflake arrays is smaller than that of the bare one. It is noteworthy that

the framework of the 3D NiCo2O4@NiO structure with huge amounts of mesopores

offers a shortened diffusion distance for electrons/ions, which improves the full use of

the high pseudocapacitance of NiCo2O4@NiO and relieves the tension caused by the

volume change and phase transformation [159]. Sennu group [160] also reported the

39
ACCEPTED MANUSCRIPT

preparation of high energy asymmetric supercapacitor (ASC) adopting

pseudocapacitive 3D microstructured composite NiCo2O4@Co3O4 and double layer

forming activated carbon (AC), obtained from the bio-waste. The NiCo2O4@Co3O4

consisted of rod-like NiCo2O4 microstructure was grown on full surfaces of Co3O4

nanosheets (Fig. 11d). The ASC could deliver a high energy density of 42.5 Wh kg-1

T
and power density of 80 W kg-1. Additionally, the ASC provides excellent cycle ability,

IP
the capacitance retention maintains ~97% of initial capacitance after 7000 cycles. A

CR
novel hierarchical NiCo2O4@MnO2 core–shell nanosheet arrays composite directly

US
grown on nickel foam formed integrated electrode for supercapacitors was designed

and fabricated through simple two-step hydrothermal method [161]. Remarkably, it is


AN
discovered that the density of MnO2 nanosheets has been increased as the extended
M

reaction time. When the reaction time of MnO2 nanosheets deposited on bare NiCo2O4

nanosheet lasts for 9 h, the particle size of a core–shell nanostructure is about 500 nm
ED

(Fig. 11c). The unique nanostructure has high electrochemical performance,


PT

displaying a capacitance of 2.39 F cm-2 as well as excellent cycling stability, which is

superior to the pure NiCo2O4 nanosheet arrays. In these above electrode design, not
CE

only both the “core” and “shell” active materials can be fully utilized but also a strong
AC

synergistic effect can come true.

40
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M

Fig. 11. (a) Schematic illustration of the controlled synthesis of homogeneous


ED

NiCo2O4 nanosheet@NiCo2O4 nanorod and heterogeneous NiCo2O4 nanosheet@NiO


PT

nanoflake core–shell nanoarrays [157], (b) SEM image of hierarchical

NiCo2O4@NiCo2O4 core–shell heterostructured nanowire arrays on Ni foam substrate


CE

[158], (c) SEM images of NiCo2O4@MnO2 core–shell nanosheet arrays [161], (d)
AC

SEM images of the 1D NiCo2O4 nanorods grown over and below the Co3O4

nanosheets [160], (e) SEM images of hierarchical NiCo2O4@NiO core–shell

heterostructured nanowire arrays on CC substrate [159], (f) HRTEM images of

as-synthesized NiCo2O4 nanosheet [161], (g) HRTEM images of as-synthesized

NiCo2O4@MnO2 nanosheet [161], (h) SC of the electrodes as a function of current

41
ACCEPTED MANUSCRIPT

density [157] and (i) Nyquist plots of bare NiCo2O4 and NiCo2O4@NiCo2O4 electrode

[158]. Reproduced from Refs. [157,159] with permission from The Royal Society of

Chemistry. Reproduced from Ref. [158] with permission from American Chemical

Society. Reproduced from Refs. [160,161] with permission from Elsevier.

3.2.2 NiCo2O4- conducting polymer composites

T
In 1976, conducting polymers with good electric conductivity, large capacitance,

IP
low cost was discovered and then applied in capacitors. The conducting polymers is

CR
an important genre of electrode material for pseudocapacitors [162]. In the past few

US
years, great effort has been devoted to conducting polymers owing to their huge latent

applications in supercapacitors. Commonly, the most commonly used conducting


AN
polymers often include polypyrrole (PPy), polyaniline (PANI), and poly [3,
M

4-ethylenedioxythiophene] (PEDOT) and their derivatives [163]. Conducting

polymers can be compounded via chemical and electrochemical methods, which are
ED

operated by polymerization of the monomer. In general, these polymers have several


PT

redox states from insulator to metal, which could be used in supercapacitors with a

pseudocapacitance mechanism [164]. Commonly, conducting polymer possess a very


CE

relative large specific capacitance compared with the clear majority carbon materials,
AC

e.g., 1284 F/g for PANI, 480 F g-1 for PPy, and 210 F g-1 for PEDOT [165]. Moreover,

works would pay more attention to their advantages, such as relatively low cost,

environment-friendly, high charge density, and high electrical conductivity in doped

states.

Recently, great efforts have been taken to increasing the capacitance of NiCo2O4

42
ACCEPTED MANUSCRIPT

through exploring miscellaneous nanostructures to provide huge specific surface area

and shorten ion diffusion path [166,167]. However, the poor conductivity problem of

NiCo2O4 is still limited its further applications. The conducting polymer, like

polyaniline (PANI), due to the high theoretical capacitance (up to 3407 F g−1),

controllable electrical conductivity, low cost and facile synthesis has drawn intensive

T
attention among various pseudocapacitive materials [168]. Ppy is also a desired

IP
electrode material with high charge density and superior electrical conductivity in

CR
doped states (10−100 S cm−1), which has also been widely adopted as electrode for

US
SCs. For example, Ppy and PEDOT were wrapped up mesoporous NiCo2O4

nanosheet arrays via a facial and controllable electrodeposition process by Xu et al.


AN
[169]. Results indicate that the NiCo2O4@PPy electrode has the highest areal

capacitance of 4.1 F cm− 2 at 2 mA cm−2, which is much higher than the values
M

obtained for the NiCo2O4@PEDOT electrode (0.86 F cm−2) and NiCo2O4 electrode
ED

(0.65 F cm−2). What’s more, even at high current density of 30 mA cm−2, the areal
PT

capacitance of 2.7 F cm−2 can be achieved for NiCo2O4@PPy electrode. Moreover,

the NiCo2O4@PPy electrode shows much smaller equivalent series resistance (ESR)
CE

than NiCo2O4@PEDOT and NiCo2O4 electrodes. By wrapping the highly conductive


AC

and porous polymer support, the distinct core-shell heterostructure is available for the

best use of the pseudocapacitive behavior of PANI layer, offering a feasible measure

to develop high-performance electrodes for supercapacitors [170]. Additionally,

conducting polymers have reversible electrochemical redox properties, good

environmental stability and strong charge storage capacity, especially, high electrical

43
ACCEPTED MANUSCRIPT

conductivity. Recently, conducting polymer-metal nanocomposites are widely

researched as electrode materials because of their excellent stability under universal

environmental conditions, good electrical conductivity and compatibility with high

charge density, which has also been widely adopted as electrode material for high

performance SCs by solving the poor electron transport and ion diffusion issues [171].

T
Typically, in the synthesis process, self-supported NiCo2O4 NWAs on carbon

IP
textiles were prepared by a modified hydrothermal growth method, and then the PPy

CR
or other conducting polymers was fabricated by the polymerization of pyrrole

US
monomer and related reagent via carried out at 0−5 °C. For example, the synthesis of

3D hierarchical NiCo2O4@PPy coaxial nanowire arrays on carbon textiles as


AN
electrode for high-performance SCs was first reported by Kong group [172]. As

shown in Fig. 12(a), the inlayer mesoporous NiCo2O4 nanowires work as the “core”
M

with high capacitance and the coated PPy nanospheres serve as the “shell” with
ED

excellent conductivity. The 3D PPy nanospheres coupled with 3D NiCo2O4 nanowire


PT

arrays (NiCo2O4@PPy NWAs) electrode exhibit a high specific capacitance (∼2244 F

g−1) at the current density of 1 A g−1 in 3 M KOH aqueous solution, excellent rate
CE

capability of ∼60.5% and good cycling stability with the capacity retention of 89.5%
AC

after 5000 cycling as well as 82.9% after 10000 cycling (Fig. 12d and e). The

NiCo2O4@PPy NWAs obtained after 1.5 h polymerization exhibit the highest specific

capacitance and lowest internal resistance (IR) among all samples. The (IR) of the

NiCo2O4@PPy NWAs hybrid electrode was ∼0.016 V, which is much smaller than

that of the NiCo2O4 NWAs electrode (0.03 V). The result shows that the PPy shell can

44
ACCEPTED MANUSCRIPT

effectively reduce the total resistance and thus improve the charge transport and

electron collection efficiency. However, the thick and dense PPy layer could prevent

the ion penetration to the inner core of NiCo2O4 NWAs. As a result, the PPy

polymerization time is a key factor in the design of high-performance NiCo2O4@PPy

electrode.

T
Compared with PPy, PANI has a much higher theoretical capacitance.

IP
Furthermore, ordered PANI nanowire arrays are supposed to have a good capacitance

CR
performance owing to their special structure [173]. A reasonable design with an easy

US
synthesis of PANI-based composite electrode has been realized by constructing novel

core-shell architecture and adopting porous NiCo2O4 nanorod arrays as substrate.


AN
Typical structures of the NiCo2O4@PANI NRAs (Fig. 12b and c) were synthesized by
M

Jabeen’s group [174]. It can be found that the surface of NiCo2O4 NRAs is thoroughly

covered with a thin PANI film (about 40 nm), indicating a representative core−shell
ED

heterostructure. As displayed in Fig. 12(f), the mean capacitance retention achieved as


PT

high as 91±4% at 10 A g−1 after 3000 continuous cycles in 1 M H2SO4 aqueous

solution. The super cycling stability can be ascribed to the optimized synergetic
CE

effects and logical design of the patent core−shell NiCo2O4@PANI NRAs. The inset
AC

in Fig. 12(f) suggests typical charge−discharge plots from the 3000 successive cycles.

The curves were symmetric corresponding to charge−discharge process,

demonstrating no significant structural collapse of the NiCo2O4@PANI NRAs

electrode. Generally, the metal/conductive polymer composite electrode materials

with unique core-shell structure are easy to be prepared. In the meantime, the high

45
ACCEPTED MANUSCRIPT

capacitance of metal oxides can be fully achieved. The microscale supercapacitors can

meet the energy demand of different electronic devices on chips. The combination of

new functions between metal oxides and conductive polymer composites could

further widen the application field of supercapacitors.

T
IP
CR
US
AN
M
ED
PT
CE
AC

Fig. 12. (a) Schematic illustration of the fabrication process of hierarchical

mesoporous NiCo2O4@PPy hybrid NWAs on carbon textiles (left: NiCo2O4@PPy

NWAs top view; right: schematic representation of rechargeable supercapacitor based

46
ACCEPTED MANUSCRIPT

on NiCo2O4@PPy NWAs on carbon textiles) [172], (b) TEM images of the

NiCo2O4@PANI NRAs, (c) an individual core−shell nanorod [174], (d)

current−density dependence of the specific of the NiCo2O4 NWAs and NiCo2O4@PPy

NWAs electrode [172] and (e) cycling performance of the NiCo2O4@PPy NWAs

electrodes (10000 cycles), compared to NiCo2O4 NWAs electrodes (5000 cycles)

T
[172]. (f) Cycle performance of the NiCo2O4@PANI NRAs electrode for 3000

IP
successive charge−discharge cycles at a large current density of 10 A g−1 and

CR
corresponding Coulombic efficiency (inset is the typical charge−discharge curves)

US
[174]. Reproduced from Refs. [172,174] with permission from American Chemical

Society.
AN
3.2.3 NiCo2O4-MOFs (metal-organic frameworks) composites
M

The discovery of metal–organic frameworks (MOFs), as a new class of

crystalline porous materials (CPMs) with high surface areas, remarkable porosity,
ED

enormous structure diversity, and multifunctional tunability with accuracy at the


PT

molecular level has unlocked the potential opportunities for scientists to solve some

pressing problems related to sustainable energy, adsorption, sensors, supercapacitors,


CE

and catalysts [175]. MOFs are assembled by joining inorganic metal nodes (secondary
AC

building units, SBUs) with organic linkers (Fig. 13a), which are consisted of metal

ions and organic binding ligands and have increasingly drawn urgent attention

[176–178]. The pseudocapacitance behavior can be found in MOFs, and

pseudocapacitive MOFs are mainly from 3d transition metals, where the metal nodes

supply the redox activity [179]. Compared with familiar inorganic porous materials,

47
ACCEPTED MANUSCRIPT

their crystalline structures with high-porosity, larger specific surface areas, diversity

of structures and controllable functionalities [180–186]. Recently, owing to their

porous structures with high surface areas, MOFs have been used as desired sacrificial

templates to support various porous active materials, such as carbonaceous materials,

metal oxides, and metal phosphides [187–191]. In addition, its structure is novel and

T
strange. As shown in Fig. 13(b), the as-prepared products of metal oxides/hydroxides

IP
with various structures, such as hollow nanocages, nanoframes, multishelled

CR
microboxes, can be obtained [192–196].

US
AN
M
ED
PT
CE
AC

Figure 13. (a) Building blocks and crystal structures of MOF-5 [179], (b) examples of

some MOF-based nanostructures [191] and (c) schematic representation of electron

and electrolyte conduction in MOF and MOF interwoven by PANI [175]. Reproduced

from Ref. [175,191] with permission from The Royal Society of Chemistry.

Reproduced from Ref. [179] with permission from American Association for the

48
ACCEPTED MANUSCRIPT

Advancement of Science.

MOFs are promising porous precursors for the construction of various functional

materials for high-performance electrochemical energy storage and conversion system

because they can offer abundant active sites and shorten ion diffusion distance as well

as relax the tension during the electrochemical reactions, thus keeping the wholeness

T
of the electrodes [197–205]. For making use of the advantage of MOF materials

IP
possibly and well, it is necessary to further study situation of all sides in

CR
supercapacitors. In recent years, Wang et al. covered the recent research advancement

US
in utilizing MOFs and their composites or derivatives by focusing on the relationship

of architecture, composition, and function for energy storage and conversion systems
AN
[206]. Choi et al. [207] reported the electrochemical performance of M7M-MOF-74
M

containing seven different metal elements, such as Co, Mg, Ni, Zn, Mn, Fe and Cd.

For example, a novel hollowed cubic NiS nanoframe electrode was synthesized by
ED

anion exchange of a Ni-Co MOF with S2- ions, which exhibited improved
PT

electrochemical performance in supercapacitors [208]. According to early research, it

is noteworthy to explore the effect of surface properties of MOFs on their capacity of


CE

charge storage. For this purpose, Lee et al. [209] studied Co MOFs with different pore
AC

dimensions and morphology and reported the influence of pore dimensions and

morphology on the supercapacitive performance, which is under the guidance of

coordination copolymerization of topologically different linkers with identical

coordinating functionality. Wang et al. [210] reported a novel tactic for fabricating

high performance SCs and solved the insulating problems of MOFs by introducing

49
ACCEPTED MANUSCRIPT

polyaniline (PANI) chains into Co-MOF crystals (ZIF-67) (Fig. 13c) with an

extraordinary areal capacitance of 2146 mF cm–2 at 10 mV s–1. Recently, Hu et al.

developed a MOFs template-engaged method to prepare multicompositional

Fe2O3-SnO2 microboxes, and it exhibited more excellent lithium storage performance

than the single-compositional micro-boxes [211]. Notably, a kind of amphoteric metal

T
oxide (such as ZnO) can be dissolved in acid or alkali, which impedes its extensive

IP
applications as an excellent electrode material in SCs [212–214]. To solve this

CR
problem, Zeng and co-workers [215] introduce a metal-organic framework (MOF)

US
structure to effectively protect the ZnO nanorods from corrosion in the alkali

electrolyte such as KOH or NaOH solution. Ultimately, the carbon coating ZnO
AN
nanorods array derived from zeolitic imidazole framework-8 (ZIF-8) was obtained
M

(Fig. 14d and e). Cao et al. [216] reported a facile two-step solution method to design

a novel hollow NiCo2O4 nanowall arrays electrode on flexible carbon cloth substrate
ED

(Fig.14a and b). As exhibited in Fig. 14(c), it is obvious that the hollow characteristic
PT

is presented clearly by the top cross section, and the surface of the nanowall arrays

has been wrapped with cross-linked ultrathin nanoflakes. These results show that
CE

hollow and porous structure is successfully prepared on the carbon cloth. The
AC

as-obtained NiCo2O4 nanostructure arrays can provide rich reaction sites and shorten

ion diffusion path. When evaluated as a flexible electrode material for supercapacitor,

the as-fabricated NiCo2O4 nanowall electrode shows remarkable electrochemical

performance with excellent rate capability and long cycle life. In addition, the hollow

NiCo2O4 nanowall electrode exhibits promising electrocatalytic activity for oxygen

50
ACCEPTED MANUSCRIPT

evolution reaction. The CC@ NiCo2O4 exhibits a high capacitance of 1055.3 F g−1 at a

small current of 2.5 mA cm−2 and a high capacitance of 483.3 F g−1 at the current

density increased to 60 mA cm-2. The capacitance retention can reach 45.8% even the

current increased 24 times. Additionally, hollow structures with high complexity in

shell architecture and composition have attracted tremendous interest because of their

T
great importance for both fundamental studies and practical applications. Recently,

IP
Hu et al. [217] reported the synthesis of novel box-in-box nanocages (NCs) with

CR
different shell compositions, namely, Co3O4/NiCo2O4 double-shelled nanocages

US
(DSNCs). The particles still retain a polyhedral shape but exhibit a rougher surface

constructed by small nanosheets, as shown in Fig. 14(f). The Co3O4/NiCo2O4 DSNCs


AN
exhibits a high capacitance of 972 F g−1 at a current density of 5 A g−1 and an

impressive capacitance of 615 F g−1 even at a current density of 50 A g−1. This means
M

that the capacitance retention can retain 63.2% even the current density increases
ED

from 5 to 50 A g−1, indicating superior rate capability. Compared with Co3O4/NiCo2O4


PT

DSNCs, the Co3O4 NCs delivers inferior performance in terms of much lower specific

capacitance and poor rate capability. Furthermore, the electrochemical stability over
CE

cycling is evaluated at a current density of 10 A g−1 by repeated charging and


AC

discharging for 12000 cycles. As shown in Fig. 14(g), the specific capacitance of 870

F g−1 for the first cycle gradually decreases to a value of 805 F g−1 after 12000 cycles

with a capacitance retention of 92.5%. The superior cycling stability is likely related

to the excellent structural stability of the Co3O4/ NiCo2O4 DSN.

51
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED
PT
CE
AC

Figure 14. (a) Schematic illustration of the formation of hollow NiCo2O4

nanostructure from 2D Co-MOF solid nanowalls [216], (b, c) NiCo2O4 hollow

nanostructure. The arrows in (c) indicate the hollow feature of the materials [216], (d)

high magnifications SEM of ZnO@ZIF-8 core-shell NRs grown on flexible cloth

52
ACCEPTED MANUSCRIPT

[215], (e) high resolution TEM images of a ZnO@ZIF-8 core-shell nanorod [215], (f)

TEM image of an individual Co3O4/NiCo2O4 DSNC [217] and (g) cycling

performance at a current density of 10 A g–1 [217]. Reproduced from Refs. [215,217]

with permission from The Royal Society of Chemistry. Reproduced from Ref. [216]

with permission from Wiley-VCH.

T
In general, the applications of transition-metal (Fe, Co, Ni) MOFs and their

IP
derivatives in supercapacitors and batteries are attracted by many researchers. The

CR
conductivity of most transition-metal (Fe, Co, Ni) based MOFs electrodes is poor.

US
However, it could be enhanced by combining with materials of high conductivity such

as conducting polymer, carbon nanotube, reduced graphene oxide. Additionally,


AN
MOFs material is a new type of multi-functional compound material which has many
M

excellent properties the other materials do not have [218]. Hence, the composite

between MOFs and NiCo2O4 electrode usually shows unique characteristic.


ED

3.2.4 NiCo2O4-metal sulfide composites


PT

Transition metal oxides and their compounds are widely applied to

supercapacitors due to their high abundance, high theoretical capacitances (>1000 F


CE

g−1) and environmentally friendly. However, the low intrinsic conductivity [219],
AC

inferior rate capability and poor cycling stability hindered the further practical

applications in the field of energy storage [220]. In addition, the rate capability and

efficiency of NiCo2O4 electrode material are still unacceptable despite its remarkable

high theoretical capacitance. The rate capability of the electroactive material, as

well-known, is principally determined by the kinetics of ion diffusion and electron

53
ACCEPTED MANUSCRIPT

transport, electronic conductivity [221]. Whereas, transition metal sulfides (TMS),

especially nickel sulfide (Ni2S3, etc.), cobalt sulfide (Co9S8, Co3S4, etc.) and AB2S4 (A

and B=Fe, Co, Ni, etc.) spinel structures bimetallic sulfides were generally served as

the promising pseudocapacitive electrode materials owing to their superior electrical

conductivity and high electrochemical activity [222,223]. Moreover, the synthesis

T
methods of sulphide are usually simple. Such metallic sulfides with unique and

IP
variable morphologies can be easily transformed from their corresponding metal

CR
oxide/hydroxide precursors by anion exchange reactions and Kirkendall effects.

US
Therefore, a combination between transition metal oxides and sulphides has drawn

much attention as excellent pseudo-capacitors materials owing to their excellent


AN
electrical conductivity and more available metal oxidation states to further redox
M

reactions. As a result, various metal sulfides have been introduced into the NiCo2O4

material to form composite materials, and the rate capability performance and cycling
ED

stability of NiCo2O4 can be obviously improved due to the synergistic effect. For
PT

example, as shown in Fig. 15(a), Chu et al. [224] have reported that electrodeposition

of nickel sulfide (Ni-S) on NiCo2O4 (NiCo2O4@Ni-S) nanoplate arrays forms a novel


CE

hierarchical NiCo2O4@NiS functional-material, which delivers high specific


AC

capacitances of 1.85 F cm−2 at 8 mA cm−2. As can be seen, the NiCo2O4 nanoplate has

a width of about 30 nm, and a thickness ranging from 0.5 to 1.5 nm. Ni-S nanoplates

were uniformly covered by a thin layer nanostructure, forming thicker core-shell

NiCo2O4@Ni-S nanoplate arrays. Three-dimensional hierarchical core/shell

nanostructure is favourable for faster ion and electron transfer owing to large specific

54
ACCEPTED MANUSCRIPT

areas and small bandgaps for anion exchange. Consequently, NiCo2O4@Ni3S2

core/shell arrays [225,226], can be synthesized by stepwise synthesis approach, which

includes hydrothermal synthesis and electrodeposition or hydrothermal sulfurization

process. Moreover, Fig. 15(d and g) confirms that the NiCo2O4@Ni3S2 has a

core/shell nanostructure, and it makes the NiCo2O4 as the core and Ni3S2 as the shell.

T
In addition, 3D hierarchical NiCo2O4 nanosheets/Co3S4 nanoneedle hybrid

IP
nanostructure were synthesized firstly by Shi’s group [227]. The optimized

CR
NiCo2O4/Co3S4 electrode shows excellent pseudocapacitor behaviors with high

US
specific capacity of 1468 F g−1 (1 A g−1). Co3S4 nanoneedles can be observed attached

tightly on the surface of NiCo2O4 (Fig. 15e), which can increase the specific surface
AN
area and shorten the diffusion path of ions and electrons during electrochemical
M

reaction process. TEM image in Fig. 15(f) further confirms that the detailed

microstructure of composite electrode material is composed of Co3S4 nanoneedles and


ED

NiCo2O4 nanosheets. It's noteworthy that MoS2 exhibits greatly enhanced electrical
PT

conductivity due to the intrinsic small bandgaps for the anion exchange [228]. MoS2

nanowires/NiCo2O4 nanosheets hybrid nanostructure grown on Ni foam network as


CE

binder-free composite electrode was prepared through a facile hydrothermal method


AC

[229]. As displayed in Fig. 15(b and c), the NiCo2O4 nanosheets stick to the Ni foam

surface, which can offer a mass of sites to grow MoS2 nanowires. The nanowires and

nanosheets are twined together to form composites, and then restrain the

agglomeration of nanosheets. Hence, the specific surface area is increased, and the

transport path of ions/electrons is shortened. Notably, compared with corresponding

55
ACCEPTED MANUSCRIPT

single-component nickel sulphides and cobalt sulphides, NiCo2S4 possesses more

redox reactions with mixed valence state [230]. It also shows lower optical band gap

energy than NiCo2O4 [231]. Importantly, it shows high electrical conductivity of ~1×

103 S cm−1 [232], which is about 100 times as much as NiCo2O4. Interestingly,

three-dimensional NiCo2O4/NiCo2S4 hybrid nanostructure on Ni-foam as a conductive

T
substrate by a facile ammonia evaporation technique, exhibited a striking specific

IP
capacitance of 3671 F g−1 at a current density of 1.8 A g−1 and 2767 F g−1 at 9 A g−1

CR
with a capacity retention of 84% at 10 A g−1 after 2000 cycles [233]. These

US
capacitance performances were much more superior to bare NiCo2O4 nanosheets (see

Fig. 15h and i). It indicates that the synergistic effect between NiCo2O4 and NiCo2S4
AN
improves the capacitance at the high current density due to the high electrical
M

conductivity of NiCo2S4.

In general, both the NiCo2O4 with high theoretical capacity and metal sulfide
ED

with relatively good conductivity are good pseudo-capacitive materials. The unique
PT

hierarchical core/shell nanostructures with enhanced mechanical stability would

provide a short and fast transport pathway, thus accelerate the efficient
CE

electrochemical reaction process. It provides a feasible way to improve the


AC

electrochemical performance of other pseudo-capacitive materials in the future

research development.

56
ACCEPTED MANUSCRIPT

(a)

(b) (c) (d)

T
IP
CR
(e) (f) (g)

US
AN
(h) (i)
M
ED
PT

Figure 15. (a) Schematic illustration of the two-step synthesis of hierarchical

NiCo2O4@NiS core shell nanoplate arrays [224], (b) typical SEM images of the
CE

sample MoS2/NiCo2O4 supported on Ni foam at different magnifications [229], (c)


AC

typical TEM image of the MoS2/NiCo2O4 supported on Ni foam [229], (d) TEM of

NiCo2O4 and NiCo2O4@Ni3S2 core/shell mesoporous nanothorn arrays [226], (e)

typical SEM image of NC-6 [227], (f) typical TEM image of NC-9 [227], (g) the

NiCo2O4@Ni3S2 hierarchical nanostructure [225], (h) summary of specific

capacitance as a function of current densities [233] and (i) cycling stability of

57
ACCEPTED MANUSCRIPT

electrodes [233]. Reproduced from Refs. [224,227,229] with permission from Elsevier.

Reproduced from Refs. [225,226,233] with permission from The Royal Society of

Chemistry.

3.2.5 NiCo2O4-carbon based composites

Carbon materials such as templated porous carbons, carbon aerogels, carbon

T
nanotubes, graphene [234–236], are the most commonly used supercapacitor

IP
materials. They possess large activated surface areas, excellent electrical conductivity,

CR
strong mechanical strength and high chemical stability however. However, the

US
specific capacitance of carbon is much lower than that of transition metal oxides,

which limits their practical applications to a certain degree. Fortunately, carbon


AN
modification can improve the electrical conductivity of NiCo2O4-based electrode
M

materials. In fact, various carbon additives such as carbon nanotubes (CNTs),

graphene, have been introduced into the transition metal oxides material to form
ED

composite materials, and the rate capability performance and cycling stability of
PT

NiCo2O4 can be obviously improved due to the improved conductivity [237–239]. In

conclusion, a combination between NiCo2O4 and carbon materials may take full
CE

advantage of the superior electrical conductivity and large specific surface areas of
AC

carbon materials as well as the high specific capacitance of NiCo2O4. So far, great

efforts have been made in this field and the following section will describe briefly.

Typically, graphene as a type of two-dimensional carbon with large specific area

and high conductivity, has been investigated as electrode material incorporated with

metal oxides [240,241]. Especially, NiCo2O4/graphene composite exhibited

58
ACCEPTED MANUSCRIPT

outstanding electrochemical performance owing to the synergetic effect between

conductive graphene and high theoretical capacity of NiCo2O4. For example, Zhang et

al. reported few-crystalline, porous NiCo2O4 nanosheets (FCP-NiCo2O4/RGO/CNTs)

grown on graphene and carbon nanotubes synthesized in reflux at 100 ℃ for 12 h (Fig.

16a). The FCP-NiCo2O4/RGO/CNTs not only provide an effective charge transport

T
path due to high conductive graphene and CNTs, but also avoid the aggregation of

IP
individual components and facilitate the diffusion of charge/ions within the electrode.

CR
As can be seen in Fig. 16(b), there are lots of the NiCo2O4 nanosheets attached on the

storage capability of NiCo2O4


US
RGO surface to form porous structure, which is in favor of enhancing the charge

nanosheets [242]. The hollow NiCo2O4


AN
nanoparticle/graphene composite was firstly prepared through a template-free method
M

[243]. As shown in Fig. 16(c and d), the hollow NiCo2O4 nanoparticle/graphene

composite presents high specific capacitance (1238 F g−1 at 1 A g−1) and excellent
ED

cycle stability (604 F g−1 at 10A g−1 for 1000 consecutive cycles). In a subsequent
PT

study, Li et al. [244] prepared C@NiCo2O4 hollow microspheres (HSs) composite by

a two-step strategy of hard-template method. Fig. 16(e and f) displays the SEM image
CE

of the C@NiCo2O4 material with and without templates, respectively. Obviously, an


AC

obvious morphology difference of both samples can be found, and the C@NiCo2O4

sample synthesized by template shows obvious core-shell structure. As shown in Fig.

16(g), NiCo2O4 nanoparticles were tightly grown on the sidewall of CNTs, forming a

3D core-shell network structure with open space size ranging from several hundred

nanometers to several micrometers, resulting in much reduced specific surface area

59
ACCEPTED MANUSCRIPT

[245]. Carbon nanofibers (CNFs) are cylindric nanostructures with graphene layers,

which also possess large specific surface area and high conductivity. Therefore,

porous NiCo2O4/C nanofibers are synthesized with cotton as template via mild

hydrothermal route [246]. As exhibited in Fig. 16(h), the porous NiCo2O4/C

nanofibers deliver a specific capacitance of 1029 F g−1 at a current density of 1 A g−1,

T
which is much higher than that of bare NiCo2O4 nanoparticles. More notably,

IP
NiCo2O4/C nanofibers material shows remarkable rate capability. Even though the

CR
current density is up to 60 A g−1, its specific capacitance still retains at 771 F g−1.

US
Obviously, the charge-transfer resistance (Rct) of NiCo2O4/C nanofibers electrode is

very low, indicating an efficient ion diffusion and rapid redox reaction process (see
AN
Fig. 16i).
M

At the end, taking full use of the good conductivity and high surface area of

carbon materials are the key points in preparing NiCo2O4@C materials composite
ED

electrodes. The synergistic effects of NiCo2O4 and carbon nanomaterials can


PT

maximize the properties of different components, for example, to further improve the

conductivity of the NiCo2O4 materials. The carbon coated NiCo2O4 nanocomposites


CE

have large surface areas, resulting in better electrolyte wettability and high
AC

conductivity, which contribute to cycling stability. Therefore, the excellent

electrochemical performance and reasonable architectural design of NiCo2O4@C

materials exhibits promising potential for practical application in energy storage.

60
ACCEPTED MANUSCRIPT

(a)

T
(b) (c) (d)

IP
CR
(e) (f)
US (g)
AN
M

(h) (i)
ED
PT
CE
AC

Figure 16. (a) Scheme showing the synthesis of NCO NRs and RG/NCO NCs [242],

(b) TEM image of the FCP-NiCo2O4/RGO/CNTs composite [242], (c) TEM image of

hollow NiCo2O4/graphene composite [243], (d) cycle performance of the hollow

NiCo2O4 nanoparicle/graphene composite [243], (e) SEM image of C@NiCo2O4 HSs

[244], (f) SEM image of NiCo2O4 NSs [244], (g) FESEM images of as-prepared 3D

61
ACCEPTED MANUSCRIPT

NiCo2O4/CNT core–shell network structure [245], (h) comparison of specific

capacitance between NiCo2O4 nanoparticles and porous NiCo2O4/C nanofibers at

various scan rates (the inset is typical FE-SEM image for porous NiCo2O4/C

nanofibers [246]) and (i) EIS curves of porous NiCo2O4/C nanofibers before and after

2000 cycles [246]. Reproduced from Refs. [243,244,246] with permission from

T
Elsevier. Reproduced from Refs. [242,245] with permission from Royal Society of

IP
Chemistry.

CR
4. Conclusions and perspectives

US
Generally, supercapacitors possess a very wide range in energy storage field

applications prospect. Recent advances in frequently-used synthesis method and


AN
structures of NiCo2O4-based electrochemical capacitors, including pristine NiCo2O4
M

and their composites with diverse morphology, have been summarized in this review.

The review emphasizes the advantages of NiCo2O4 as the materials of supercapacitors


ED

with super rate capability, high energy density, good power density, and excellent
PT

cycling stability. Meanwhile, we have advanced reasonable improvements in the

electrochemical performance of NiCo2O4 electrode materials and anticipate that this


CE

review can facilitate further studies and advancements in this field:


AC

(1) Many methods have been used to prepare NiCo2O4-based electrode materials,

including sol-gel, bio-template, electrodeposition, electrospinning and hydrothermal

methods. By comparison, the hydrothermal process is a direct and effective way to

synthesize 1D, 2D and 3D nanostructured NiCo2O4-based materials with unique

architectures and superior electrochemical performances by adjusting temperature

62
ACCEPTED MANUSCRIPT

parameters, the kinds and concentration of surfactants. Compared with 1D and 2D

nanostructures, 3D structural morphology not only forms the desired hierarchical

porous continuous channels for electrolyte ions by serving as an ion buffering

reservoir, but also has higher electrical conductivity and excellent structural

mechanical stability, which can alleviate the volume expansion effect during

T
charge/discharge process.

IP
(2) The electrochemical performance of NiCo2O4-based materials can be

CR
improved by controlling the pore properties and morphology. The novel

US
nanostructures such as hollow, core-shell and hollow-core-shell are explored to

further improve the electrochemical performances with feasible approaches.


AN
NiCo2O4-based materials with 3D hierarchical core-shell structure have attracted
M

much attention because of the high surface/volume ratios, larg surface areas, excellent

permeabilities and more surface-active sites, which is vital to achieve high


ED

energy/power densities in electrochemical supercapacitors.


PT

(3) The electrochemical performance of NiCo2O4-based materials can be

improved by taking full advantage of the synergistic effects of NiCo2O4 and other
CE

metal oxides, sulfides, conducting polymers or carbon nanomaterials. The synergistic


AC

effects can maximize the properties of different components, for example, to further

improve the conductivity or capacity of the NiCo2O4 materials.

(4) The spinel NiCo2O4 shows high performance, but there still are many

challenges to be conquered before commercialization. In concept, the NiCo2O4

material as a hybrid supercapacitor or battery-type supercapacitor is being in

63
ACCEPTED MANUSCRIPT

controversy. There is no doubt that high theoretical capacitance of spinel NiCo2O4

was known very well. In addition, the working voltage window of NiCo2O4 in alkali

electrolyte is only approximately 0.45–0.5 V, which shows a relatively low energy and

power density. Therefore, it is necessary to assemble with a carbon material into an

asymmetric capacitor to build a stable and higher operating voltage window.

T
In general, NiCo2O4-based supercapacitor devices may play a critical role in

IP
renewable energy, and it may be expected to solve environmental pollution problems

CR
and reduce the use of hydrocarbon fuels. With rapid development of NiCo2O4-based

US
electrodes, the high-performance supercapacitors with safe operation and friendly

environment are suitable as practical energy storage devices. Therefore, the excellent
AN
electrochemical performance and reasonable architectural design of NiCo2O4-based
M

materials exhibits promising potential for practical application in energy storage.

Acknowledgments
ED

This work was financially supported by the National Natural Science Foundation
PT

of China (nos. 51774002 and 51672156), and Anhui Provincial Science Fund for

Excellent Young Scholars (no. gxyqZD2016066), the National Key Basic Research
CE

Program of China (2014CB932400), Guangdong special support program


AC

(2015TQ01N401), Shenzhen Technical Plan Project (KQJSCX20160226191136).

References

[1] S. Chu, A. Majumdar, Nature 488 (2012) 294−303.

[2] J. Lv, T. Liang, M. Yang, S. Ken, M. Hideo, J. Energy Chem. 26(2017) 330-335

[3] B. D. Boruah, A. Misra, J. Mater. Chem. A 4 (2016) 17552−17559.

64
ACCEPTED MANUSCRIPT

[4] Y. Tu, D. Deng, X. Bao, J. Energy Chem. 25(2016) 957-966

[5] T. Brezesinski, J. Wang, S. H. Tolbert, B. Dunn, Nat. Mater. 9 (2010) 146-151.

[6] C. Liu, F. Li, L.P. Ma, H. M. Cheng, Adv. Mater. 22 (2010) E28−E62.

[7] Q. Zhang, B. Zhao, J. Wang, C. Qu, H. Sun, K. Zhang, M. Liu, Nano Energy 28

(2016) 475−485.

T
[8] M. Armand, J.M. Tarascon, Nature 451 (2008) 652-657.

IP
[9] M.S. Whittingham, Chem. Rev. 104 (2004) 4271−4302.

CR
[10] B.E. Conway, J. Electrochem. Soc. 138 (1991) 1539−1548.

US
[11] E. Frackowiak, Phys. Chem. Chem. Phys. 9 (2007) 1774−1785.

[12] H. Jiang, P. S. Lee, C. Li, Energy Environ. Sci. 6 (2013) 41−53.


AN
[13] Q. Meng, K. Cai, Y. Chen, Y. Chen, L. Chen, Nano Energy, 36 (2017) 268−285.
M

[14] C. Largeot, C. Portet, J. Chmiola, P.L. Taberna, Y. Gogotsi, P. Simon, J. Am.

Chem. Soc. 130 (2008) 2730–2731.


ED

[15] J. Yan, Q. Wang, T. Wei, Z. Fan, Adv. Energy Mater. 4 (2014) 157−164.
PT

[16] Z. Yu, L. Tetard, L. Zhai, J.Y. Thomas, Energy Environ. Sci. 8 (2015) 702−730.

[17] G. Yu, X. Xie, L. Pan, Z. Bao, Y. Cui, Nano Energy 2 (2013) 213−234.
CE

[18] L. Hu, W. Chen, X. Xie, N. Liu, Y. Yang, H. Wu, Y. Yao, M. Pasta, Y. Cui, H.N.
AC

Alshareef, ACS Nano 5 (2011) 8904−8913.

[19] L. Nyholm, G. Nyström, A. Mihranyan, M. Strømme, Adv. Mater. 23 (2011)

3751−3769.

[20] A. Huang, J. Yan, H. Zhang, X. Li, H. Zhang, J. Energy Chem. 26(2017)

121-128.

65
ACCEPTED MANUSCRIPT

[21] L. Wei, H. E. Karahan, S. Zhai, Y. Yuan, Q. Qian, K. Goh, A. K. Ng, Y. Chen, J.

Energy Chem. 25(2016) 191-198

[22] X. Wang, M. Wang, X. Zhang, H. Li, X. Guo, J. Energy Chem. 25 (2016) 26-34

[23] Y. Zhai, Y. Dou, D. Zhao, P.F. Fulvio, R.T. Mayes, S. Dai, Adv. Mater. 23 (2011)

4828−4850.

T
[24] X. Zhang, H. Zhang, Z. Lin, M. Yu, X. Lu, Y. Tong, Sci. China Mater. 59 (2016)

IP
475–494.

CR
[25] V. Augustyn, P. Simon, B. Dunn, Energy Environ. Sci. 7 (2014) 1597−1614.

US
[26] D.P. Dubal, O. Ayyad, V. Ruiz, P. Gomez-Romero, Chem. Soc. Rev. 44 (2015)

1777−1790.
AN
[27] S. Zheng, Z.S. Wu, S. Wang, H. Xiao, F. Zhou, C. Sun, X. Bao, H.M. Chen,

Energy Storage Materials 6 (2017) 70−97.


M

[28] B. Dunn, H. Kamath, J.M. Tarascon, Science 334 (2011) 928−935.


ED

[29] Y. Gogotsi, Nature 509 (2014) 568−570.


PT

[30] J. Chmiola, G. Yushin, Y. Gogotsi, C. Portet, P. Simon, P.L. Taberna, Science 313

(2006) 1760−1763.
CE

[31] G.L. Soloveichik, Nature 505 (2014) 163−165.


AC

[32] H. Wang, H. Dai, Chem. Soc. Rev. 42 (2013) 3088−3113.

[33] C. Zhong, Y. Deng, W. Hu, J. Qiao, L. Zhang, J. Zhang, Chem. Soc. Rev. 44

(2015) 7484−7539.

[34] Y.-Z. Zhang, Y. Wang, T. Chen, W.-Y. Lai, H. Pang, W. Huang, Chem. Soc. Rev.

44 (2015) 5181−5199.

66
ACCEPTED MANUSCRIPT

[35] N.S. Malvankar, M. Vargas, K.P. Nevin, A.E. Franks, C. Leang, B.C. Kim, K.

Inoue, T. Mester, S.F. Covalla, J.P. Johnson, V.M. Rotello, M.T. Tuominen, D.R.

Lovley, Nat. Nanotechnol. 6 (2011) 573−579.

[36] J. Chmiola, C. Largeot, P.L. Taberna, P. Simon, Y. Gogotsi, Science 328 (2010)

480−483.

T
[37] D.W. Wang, F. Li, M. Liu, G.Q. Lu, H.M. Cheng, Angew. Chem. Int. Ed. 47

IP
(2008) 373−376.

CR
[38] Y. Wang, C.Y. Foo, T.K. Hoo, M. Ng, J.Y. Lin, Chem.–Eur. J. 16 (2010)

3598−3603.
US
[39] Y. Zhang, B. Wang, F. Liu, J. Cheng, X. Zhang, L. Zhang, Nano Energy 27 (2016)
AN
627−637.
M

[40] G.S. Gund, D.P. Dubal, S. Shinde, C.D. Lokhande, ACS Appl. Mater. Interfaces 6

(2014) 3176−3188.
ED

[41] X. Wu, Z. Han, X. Zheng, S. Yao, X. Yang, T. Zhai, Nano Energy 31 (2017)
PT

410−417.

[42] G. Ye, Y. Gong, K. Keyshar, E. A. M. Husain, G. Brunetto, S. Yang, R. Vajtai, P.


CE

M. Ajayan, Part. Part. Syst. Charact. 32 (2015) 817−821.


AC

[43] J. Chang, M. Jin, F. Yao, T.H. Kim, V. T. Le, H. Yue, F. Gunes, B. Li, A. Ghosh, S.

Xie, Y.H. Lee, Adv. Funct. Mater. 23 (2013) 5074−5083.

[44] Y. Zhang, L. Li, H. Su, W. Huang, X. Dong, J. Mater. Chem. A 3 (2015) 43−59.

[45] D.P. Dubal, P. Gomez-Romero, B.R. Sankapal, R. Holze, Nano Energy 11 (2015)

377−399.

67
ACCEPTED MANUSCRIPT

[46] G. Zhang, X.W. Lou, Adv. Mater. 25 (2013) 976−979.

[47] Q. Wang, B. Liu, X. Wang, S. Ran, L. Wang, D. Chen, G. Shen, J. Mater. Chem.

22 (2012) 21647−21653.

[48] Q.W. Ye, Y.C. Xiang, T.J. Ping, Q.Q. Zhou, Electrochim. Acta 56 (2011)

7517−7522.

T
[49] H. Wang, Q. Gao, L. Jiang, Small 7 (2011) 2454−2459.

IP
[50] R. Zou, K. Xu, T. Wang, G. He, Q. Liu, X. Liu, Z. Zhang, J. Hu, J. Mater. Chem.

CR
A 1 (2013) 8560−8566.

US
[51] H. Jiang, J. Ma, C. Li, Chem. Commun. 48 (2012) 4465−4467.

[52] B. Cui, H. Lin, Y. Liu, J. Li, P. Sun, X. Zhao, C. Liu, J. Phys. Chem. C 113 (2009)
AN
14083−14087.
M

[53] J. Zhang, Q. Mei, Y. Ding, K. Guo, X. Yang, J. Zhao, ACS Appl. Mater. Inter. 9

(2017) 29771−29781.
ED

[54] J.F. Marco, J.R. Gancedo, M. Gracia, J.L. Gautier, E. Ríos, F.J. Berry, J. Solid
PT

State Chem. 153 (2000) 74−81.

[55] C. Yuan, J. Li, L. Hou, X. Zhang, L. Shen, X.W. Lou, Adv. Funct. Mater. 22
CE

(2012) 4592−4597.
AC

[56] M. Lenglet, R. Guillamet, J. Dürr, D. Gryffory, R.E. Vandenberghe, Solid State

Commun. 74 (1990) 1035−1039.

[57] Z Wu, Y Zhu, X Ji, J. Mater. Chem. A 2 (2014) 14759−14772.

[58] X.Y. Liu, S.J. Shi, Q.Q. Xiong, L. Li, Y.J. Zhang, H. Tang, C.D. Du, X.L. Wang,

J.P. Tu, ACS Appl. Mater. Inter. 5 (2013) 8790−8795.

68
ACCEPTED MANUSCRIPT

[59] Y. Chen, B. Qu, L. Hu, Z. Xu, Q. Li, T. Wang, Nanoscale 5 (2013) 9812−9820.

[60] H. Jiang, J. Ma, C. Li, Chem. Commun. 48 (2012) 4465−4467.

[61] Y. Zhang, M. Ma, J. Yang, H. Su, W. Huang, X. Dong, Nanoscale 6 (2014)

4303−4308.

[62] J. Mei, T. Liao, L. Kou, Z. Sun, Adv. Mater. 29 (2017) 1700176.

T
[63] L. Wen, F. Li, H.-M. Cheng, Adv. Mater. 28 (2016) 4306-4337.

IP
[64] D. Chen, Q. Wang, R. Wang, G. Shen, J. Mater. Chem. A 3 (2015) 10158-10173.

CR
[65] L.L. Hench, J.K. West. Chem. Rev. 90 (1990) 33-72.

US
[66] Y.Q. Wu, X.Y. Chen, P.T. Ji, Q.Q. Zhou, Electrochim. Acta 56 (2011)

7517-7522.
AN
[67] T.Y. Wei, C.H. Chen, H.C. Chen, S.Y. Lu, Adv. Mater. 22 (2010) 347-351.
M

[68] M.-C. Liu, L.-B. Kong, C. Lu, X.-M. Li, Y.-C. Luo, L. Kang, X. Li, F.C. Walsh, J.

Electrochem. Soc. 159 (2012) A1262-A1266.


ED

[69] M.-C. Liu, L.-B. Kong, C. Lu, X.-M. Li, Y.-C. Luo, L. Kang, ACS Appl. Mater.
PT

Interfaces 4 (2012) 4631−4636.

[70] Taguchi, F. Schüth, Microporous Mesoporous Mater. 77 (2005) 1-45.


CE

[71] Y. Zhang, L. Li, H. Su, W. Huang, X. Dong, J. Mater. Chem. A 3 (2015) 43-59.
AC

[72] C. Yuan, J. Li, L. Hou, J. Lin, G. Pang, L. Zhang, L. Lian, X. Zhang, J. Mater.

Chem. A 3 (2013) 18573-18578.

[73] S. Liu, S. Sun, X.-Z You, Nanoscale 6 (2014) 2037-2045.

[74] G. Lota, K. Fic, E. Frackowiak, Energy Environ. Sci. 4 (2011) 1592-1605.

[75] H.-W. Wang, Z.-A. Hu, Y.-Q. Chang, Y.-L. Chen, H.-Y. Wu, Z.-Y. Zhang, Y.-Y.

69
ACCEPTED MANUSCRIPT

Yang, J. Mater. Chem. 21 (2011) 10504−10511.

[76] M.F. El-Kady, V. Strong, S. Dubin, R.B. Kaner, Science 335 (2012) 1326−1330.

[77] L.L. Li, S.J. Peng, Y.L. Cheah, P.F. Teh, J. Wang, G. Wee, Y.W. Ko, C.L. Wong,

M. Srinivasan, Chem.-Eur. J. 19 (2013) 5892–5898.

[78] W. Xiong, Y. Gao, X. Wu, X. Hu, D. Lan, Y. Chen, X. Pu, Y. Zeng, J. Su, Z. Zhu,

T
ACS Appl. Mater. Interfaces 6 (2014) 19416-19423.

IP
[79] L. Shen, L. Yu, X.Y. Yu, X. Zhang, X.W. Lou, Angew. Chem. Int. Edit. 54 (2015)

CR
1868-1872.

US
[80] X.-L. Wu, T. Wen, H.-L. Guo, S. Yang, X. Wang, A.-W. Xu, ACS Nano 7 (2013)

3589−3597.
AN
[81] L. Li, S. Peng, Y. Cheah, P. Teh, J. Wang, G. Wee, Y. Ko, C. Wong, M.
M

Srinivasan, CHEM-EUR. J. 19 (2013) 5892-5898.

[82] J. Yan, Y.P. Huang, Y.E. Miao, W.W. Tjiu, T.X. Liu, J. Hazard. Mater. 283 (2015)
ED

730−739.
PT

[83] S. Cavaliere, S. Subianto, I. Savych, D.J. Jones, J. Roziere, Energy Environ. Sci.

4 (2011) 4761−4785.
CE

[84] P. Zhang, C. Shao, Z. Zhang, M. Zhang, J. Mu, Z. Guo, Y. Liu, Nanoscale 3 (2011)
AC

3357−3363.

[85] C. Kim, K.S. Yang, Appl. Phys. Lett. 83 (2003) 1216−1218.

[86] Y.P. Huang, Y.E. Miao, L.S. Zhang, W.W. Tjiu, J.S. Pan, T.X. Liu, Nanoscale 6

(2014) 10673−10679.

[87] J. Wang, Y. Yang, Z. Huang, F. Kang, Electrochim. Acta 56 (2011) 9240−9247.

70
ACCEPTED MANUSCRIPT

[88] M.J. Zhi, A. Manivannan, F.K. Meng, N.Q. Wu, J. Power Sources 208 (2012)

345−353.

[89] Y.P. Huang, Y.E. Miao, W.W. Tjiu, T.X. Liu, RSC Adv. 5 (2015) 18952−18959.

[90] F. Lai, Y.E. Miao, Y Huang, T.-S. Chung, T. Liu, J. Phys. Chem. C 119 (2015)

13442−13450.

T
[91] M. Musiani, Electrochim. Acta 45 (2000) 3397−3402.

IP
[92] J. Du, G. Zhou, H. Zhang, C. Cheng, J. Ma, W. Wei, L. Chen, T. Wang, ACS Appl.

CR
Mater. Interfaces 5 (2013) 7405−7409.

US
[93] Y. An, Z. Hu, B. Guo, N. An, Y. Zhang, Z. Li, Y. Yang, H. Wu, RSC Adv. 6 (2016)

37562−37573.
AN
[94] Y. Liu, Y. Jiao, Z. Zhang, F. Qu, A. Umar, X. Wu, ACS Appl. Mater. Interfaces,
M

6 (2014) 2174−2184.

[95] J. Wu, P. Guo, R. Mi, X. Liu, H. Zhang, J. Mei, H. Liu, W.-M. Lau, L.-M. Liu, J.
ED

Mater. Chem. A 3 (2015) 15331−15338.


PT

[96] L. Huang, D. Chen, Y. Ding, Z.L. Wang, Z. Zeng, M. Liu, ACS Appl. Mater.

Interfaces 5 (2013) 11159–11162.


CE

[97] G. Zhang, T. Wang, X. Yu, H. Zhang, H. Duan, B. Lu, Nano Energy 2 (2013)
AC

586–594.

[98] S. Feng, R. Xu, Accounts Chem. Res. 34 (2001) 239−247.

[99] X. Zhang, X. Cheng, Q. Zhang, J. Energy Chem. 25(2016) 967-984.

[100] J. Pu, J. Wang, X. Jin, F. Cui, E. Sheng, Z. Wang, Electrochim. Acta 106

(2013) 226−234.

71
ACCEPTED MANUSCRIPT

[101] X.Y. Yu, X.Z. Yao, T. Luo, Y. Jia, J.H. Liu, X.J. Huang, ACS Appl. Mater.

Interfaces 6 (2014) 3689−3695.

[102] R. Ding, L. Qi, M. Jia, H. Wang, J. Appl. Electrochem. 43 (2013) 903−910.

[103] X. Lai, J.E. Halpert, D. Wang, EnergyEnviron. Sci. 5 (2012) 5604−5618.

[104] Y. Zhang, M. Ma, J. Yang, H. Su, W. Huang, X. Dong, Nanoscale 6 (2014)

T
4303−4308.

IP
[105] J. Wang, Y. Zhang, J. Ye, H. Wei, J. Hao, J. Mu, S. Zhao, S. Hussain, RSC

CR
Adv. 6 (2016) 70077−70084.

[106]
US
C. Yuan, J. Li, L. Hou, J. Lin, X. Zhang, S. Xiong, J. Mater. Chem. A 1 (2013)

11145−11151.
AN
[107] C.T. Hsu, C.C. Hu, J. Power Sources 242 (2013) 662−671.
M

[108] C. An, Y. Wang, Y. Huang, Y. Xu, C. Xu, L. Jiao, H. Yuan, CrystEngComm

16 (2014) 385−392.
ED

[109] M. Rui, X. Li, L. Gan, T. Zhai, H. Zeng, Adv. Funct. Mater. 26 (2016)
PT

5051−5060.

[110] S. Buller, J. Strun, J. Energy Chem. 25(2016) 171-190


CE

[111] G. Zhang, X. Xiao, B. Li, P. Gu, H. Xue, H. Pang, J. Mater. Chem. A 5 (2017)
AC

8155–8186.

[112] Y. Liu, X. Peng, Appl. Mater. Today 8 (2017) 104–115.

[113] B.X. Li, Y.F. Wang, J. Phys. Chem. C 114 (2010) 890-896.

[114] G. Zhang, L. Yu, H.E. Hoster, X.W. Lou, Nanoscale 5 (2013) 877−881.

[115] C. Tang, M.-M. Titirici, Q. Zhang, J. Energy Chem., 26(2017) 1077-1093.

72
ACCEPTED MANUSCRIPT

[116] Y. Li, L. Kong, M. Liu, W. Zhang, L. Kang, J. Energy Chem., 26(2017)

494-500

[117] X. Wang, X.L. Wu, Y.G. Guo, Y. Zhong, X. Cao, Y. Ma, J. Yao, Adv. Funct.

Mater. 20 (2010) 1680−1686.

[118] D. Ge, L. Yang, Z. Tong, Y. Ding, W. Xin, J. Zhao, Y. Li, Electrochim. Acta

T
104 (2013) 191−197.

IP
[119] J. Liang, Z. Fan, S. Chen, S. Ding, G. Yang, Chem. Mater. 26 (2014)

CR
4354-4360.

[120]
US
X.F. Lu, D.J. Wu, R.Z. Li, Q. Li, S.H. Ye, Y.X. Tong, G.R. Li, J. Mater. Chem.

A 2 (2014) 4706-4713.
AN
[121] C. Ji, F. Liu, L. Xu, S. Yang, J. Mater. Chem. A 5 (2017) 5568−5576.
M

[122] A.L.M. Reddy, S.R. Gowda, M.M. Shaijumon, P.M. Ajayan, Adv. Mater. 24

(2012) 5045−5064
ED

[123] H. Zhang, D. Xiao, Q. Li, Y. Ma, S. Yuan, L. Xie, C. Chen, C. Lu, J. Energy
PT

Chem. 27 (2018) 195−202.

[124] H.Y. Wu, H.W. Wang, Acta Phys.-Chim. Sin. 29 (2013) 1501–1506.
CE

[125] H. Zhao, L. Liu, R. Vellacheri, Y. Lei, Adv. Sci. 4 (2017) 1700188.


AC

[126] J. Du, G. Zhou, H.M. Zhang, C. Cheng, J.M. Ma, W.F. Wei, L.B. Chen, T.H.

Wang, ACS Appl. Mater. Interfaces 5 (2013) 7405–7409.

[127] R.J. Zou, K.B. Xu, T. Wang, G.J. He, Q. Liu, X.J. Liu, Z.Y. Zhang, J.Q. Hu,

J. Mater. Chem. A 1 (2013) 8560–8566.

[128] C, Wu, J. Cai, Y. Zhu, K. Zhang, RSC Adv. 6 (2016) 63905–63914.

73
ACCEPTED MANUSCRIPT

[129] X. Liu, S. Shi, Q. Xiong, L. Li, Y. Zhang, H. Tang, C. Gu, X. Wang, J. Tu,

ACS Appl. Mater. Interfaces 5 (2013) 8790−8795.

[130] H. Jiang, J. Ma, C.Z. Li, Chem. Commun. 48 (2012) 4465–4467.

[131] Q.F. Wang, X.F. Wang, B. Liu, G. Yu, X.J. Hou, D. Chen, G.Z. Shen, J. Mater.

Chem. A 1 (2013) 2468–2473.

T
[132] C. Guan, X.H. Xia, N. Meng, Z.Y. Zeng, X.H. Cao, C. Soci, H. Zhang, H.J.

IP
Fan, Energy Environ. Sci. 5 (2012) 9085–9090.

CR
[133] J.P. Liu, J. Jiang, C.W. Cheng, H.X. Li, J.X. Zhang, H. Gong, H.J. Fan, Adv.

Mater. 23 (2011) 2076–2081.

[134]
US
S. Chen, G. Yang, Y. Jia, H. Zheng, J. Mater. Chem. A 5 (2017) 1028–1034.
AN
[135] D.P. Cai, B. Liu, D.D. Wang, L.L. Wang, Y. Liu, H. Li, Y.R. Wang, Q.H. Li,

T.H. Wang, J. Mater. Chem. A 2 (2014) 4954−4960.


M

[136] C. Cui, Xu, L. Wang, D. Guo, M. Mao, J. Ma, T. Wang, ACS Appl. Mater.
ED

Interfaces 8 (2016) 8568–8575.


PT

[137] Z. Gu, X. Zhang, J. Mater. Chem. A 4 (2016) 8249–8254.

[138] Y. Yuan, W. Wang, J. Yang, H. Tang, Z. Ye, Y. Zeng, J. Lu, Langmuir 33


CE

(2017) 10446–10454
AC

[139] L. Hu, L. Wu, M. Liao, X. Fang, Adv. Mater. 23 (2011) 1988–1992.

[140] J.M. Choi, S. Im, Applied Surf. Sci. 244 (2005) 435-438.

[141] B. Li, Y. Wang, J. Phys. Chem. C 114 (2010) 890–896.

[142] R.V. Kumar, Y. Diamant, A. Gedanken, Chem. Mater. 12 (2000) 2301−2305.

[143] Z.R. Tian, W. Tong, J.Y. Wang, N.G. Duan, V.V. Krishnan, S.L. Suib, Science

74
ACCEPTED MANUSCRIPT

276 (1997) 926−930.

[144] Y. Rao, D.M. Antonelli, J. Mater. Chem. 19 (2009) 1937−1944.

[145] M. Cabo, E. Pellicer, E. Rossinyol, M. Estrader, A. Lopez-Ortega, J. Nogues,

O.E. Castell, S. Surinach, M.D. Baro, J. Mater. Chem. 20 (2010) 7021−7028.

[146] M. Dai, L. Song, J.T. LaBelle, B.D. Vogt, Chem. Mater. 23 (2011)

T
2869-2878.

IP
[147] J.W. Lee, T. Ahn, D. Soundararajan, J.M. Ko, J.D. Kim, Chem. Commun. 47

CR
(2011) 6305–6307.

US
[148] C. Sassoye, C. Laberty, H. Le Khanh, C. Sophie, B. Cédric, A. Markus, S.

Clément, Adv. Funct. Mater. 19 (2009) 1922–1929.


AN
[149] L.Q. Mai, F. Yang, Y.L. Zhao, X. Xu, Y.Z. Zhao, Nat. Commun. 2 (2011)
M

381.

[150] J. Liu, Y. Li, X. Huang, G. Li, Z. Li, Adv. Funct. Mater. 18 (2008) 1448-1458.
ED

[151] D. Zhao, W. Zhou, H. Li, Chem. Mater. 19 (2007) 3882-3891.


PT

[152] G.W. Yang, C.L. Xu, H.L. Li, Chem. Commun. 48 (2008) 6537−6539.

[153] N. Fertig, M. George, M. George, R.H. Blick, J.C. Behrends, Appl. Phys.
CE

Lett. 81 (2002) 4865−4867.


AC

[154] G.R. Fu, Z.A. Hu, L.J. Xie, X.Q. Jin, Y.L. Xie, Y.X. Wang, Z.Y. Zhang, Y.Y.

Yang, H.Y. Wu, Int. J. Electrochem. Sci. 4 (2009) 1052−1062.

[155] J.H. Kim, S.H. Kang K. Zhu, J.Y. Kim, N.R. Neale, A. Frank, J. Chem.

Commun. 47 (2011) 5214−5216.

[156] L. Yu, G. Zhang, C. Yuan, X.W. Lou, Chem. Commun. 49 (2013) 137–139

75
ACCEPTED MANUSCRIPT

[157] W. Zhou, D. Kong, X. Jia, C. Ding, C. Cheng, G. Wen, J. Mater. Chem. A 2

(2014) 6310–6315.

[158] X. Liu, S. Shi, Q. Xiong, L. Li, Y. Zhang, H. Tang, C. Gu, X. Wang, J. Tu,

ACS Appl. Mater. Interfaces 5 (2013) 8790–8795.

[159] W. Yang, Z. Gao, J. Ma, X. Zhang, J. Liu, J. Materials Chem. A 2 (2014)

T
1448–1457.

IP
[160] P. Sennu, V. Aravindan, Y.S. Lee, J. Power Sources 306 (2016) 248–257.

CR
[161] F. Bao, Z. Zhang, W. Guo, X. Liu, Electrochim. Acta 157 (2015) 31–40.

[162]

[163]
US
B.E. Conway, V. Birss, J. Wojtowicz, J. Power Sources 66 (1997) 1–14.

E. Frackowiak, V. Khomenko, K. Jurewicz, K. Lota, F. Béguin, J. Power


AN
Sources 153 (2006) 413–418.
M

[164] A.J. Heeger, Rev. Mod. Phys. 73 (2001) 681–700.

[165] K. Wang, H. Wu, Y. Meng, Z. Wei, Small 10 (2014) 14–31.


ED

[166] Q. Wang, X. Wang, B. Liu, G. Yu, X. Hou, D. Chen, G. Shen, J. Mater. Chem.
PT

A 1 (2013) 2468–2473.

[167] G.Q. Zhang, H.B. Wu, H.E. Hoster, M.B. Chan-Park, X.W. Lou, Energy
CE

Environ. Sci. 5 (2012) 9453–9456.


AC

[168] X. Peng, K.F. Huo, J.J. Fu, X.M. Zhang, B. Gao, P.K. Chu, Chem. Commun.

49 (2013) 10172−10174.

[169] K. Xu, X. Huang, Q. Liu, R. Zou, W. Li, X. Liu, S. Li, J. Yang, J. Hu, J.

Materials Chem. A 2 (2014) 16731–16739.

[170] J. Pu, F.L. Cui, S.B. Chu, T.T. Wang, E.H. Sheng, Z.H. Wang, ACS

76
ACCEPTED MANUSCRIPT

Sustainable Chem. Eng. 2 (2013) 809–815.

[171] Y. Yan, T. Wang, X. Li, H. Pang, H. Xue, Inorg. Chem. Front. 4 (2017)

33–51.

[172] D. Kong, W. Ren, C. Cheng, Y. Wang, Z. Huang, H.Y. Yang, ACS Appl.

Mater. Interfaces 7 (2015) 21334–21346.

T
[173] K. Wang, J. Huang, Z. Wei, J. Phys. Chem. C 114 (2010) 8062–8067.

IP
[174] N. Jabeen, Q. Xia, M. Yang, H. Xia, ACS Appl. Mater. Interfaces 8 (2016)

CR
6093–6100.

[175]
US
W. Zhang, H. Pang W. Sun, L.-Ping, Y. Wang, Electrochem. Commun. 87

(2017) 80–85.
AN
[176] D.Y. Lee, S.J. Yoon, N.K. Shrestha, S.-H. Lee, H. Ahn, S.-H. Han,
M

Microporous and Mesoporous Mater. 153 (2012) 163–165.

[177] J. Zhou, B. Wang, Chem. Soc. Rev. 46 (2017) 6927–6945.


ED

[178] H. Furukawa, K.E. Cordova, M. O’Keeffe, O.M. Yaghi, Science 341 (2013)
PT

1230444.

[179] A.P. Cote, A.I. Benin, N.W. Ockwig, M. O’Keeffe, A.J. Matzger, O.M. Yaghi,
CE

Science 310 (2005) 1166-1170.


AC

[180] C.V. McGuire, R.S. Forgan, Chem. Commun. 51 (2015) 5199-5217.

[181] S.L. James, Chem. Soc. Rev. 32 (2003) 276-288.

[182] I. Stassen, M. Styles, G. Grenci, H.V. Gorp, W. Vanderlinden, S.D. Feyter, P.

Falcaro, D.D. Vos, P. Vereecken, R. Ameloot, Nat. Mater. 15 (2015) 304-310.

[183] W.J. Li, M. Tu, R. Cao, R.A. Fischer, J. Mater. Chem. A 4 (2016)

77
ACCEPTED MANUSCRIPT

12356-12369.

[184] Z.-H. Yan, D. Li, X.-B. Yin, Sci. Bull. 62(2017) 1344-1354.

[185] Y. Liu, H. Dong, F. Hu, Y. S. Zhao, Sci. Bull. 62(2017) 3-4.

[186] L. Wang, X. Feng, L. Ren, Q. Piao, J. Zhong, Y. Wang, H. Li, Y. Chen, B.

Wang, J. Am. Chem. Soc. 137 (2015) 4920-4923.

T
[187] K.M. Choi, H.M. Jeong, J.H. Park, Y.B. Zhang, J.K. Kang, O.M. Yaghi, ACS

IP
Nano 8 (2014) 7451-7457.

CR
[188] J. Liu, C. Wu, D. Xiao, P. Kopold, L. Gu, P.A. van Aken, J. Maier, Y. Yu,

Small 12 (2016) 2354-2364.

[189]
US
H. Furukawa, K.E. Cordova, M. O'Keeffe, O.M. Yaghi, Science 341 (2013)
AN
1230444.
M

[190] C.V. McGuire, R.S. Forgan, Chem. Commun. 51 (2015) 5199-5217.

[191] H.-C. Zhou, S. Kitagawa, Chem. Soc. Rev. 43 (2014) 5415-5418.


ED

[192] W. Xia, A. Mahmood, R. Zou, Q. Xu, Energy Environ. Sci. 8 (2015)


PT

1837-1866.

[193] Z. Jiang, Z. Li, Z. Qin, H. Sun, X. Jiao, D. Chen, Nanoscale 5 (2013)


CE

11770-11775.
AC

[194] C. Sun, J. Yang, X. Rui, W. Zhang, Q. Yan, P. Chen, F. Huo, W. Huang, X.

Dong, J. Mater. Chem. A 3 (2015) 8483-8488.

[195] S. Chen, M. Xue, Y. Li, S. Qiu, J. Mater. Chem. A 3 (2015) 20145-20152.

[196] J.K. Sun, Q. Xu, Energy Environ. Sci. 7 (2014) 2071-2100.

[197] Y. Zhao, X. Li, J. Liu, C. Wang, Y. Zhao, G. Yue, ACS Appl. Mater. Interfaces

78
ACCEPTED MANUSCRIPT

8 (2016) 6472-6480.

[198] A. Morozan, F. Jaouen, Energy Environ. Sci. 5 (2012) 9269-9290.

[199] J. Qi, X. Lai, J. Wang, H. Tang, H. Ren, Y. Yang, Q. Jin, L. Zhang, R. Yu, G.

Ma, Z. Su, H. Zhao, D. Wang, Chem. Soc. Rev. 44 (2015) 6749-6773.

[200] A. Mahmood, W. Guo, H. Tabassum, R. Zou, Adv. Energy Mater. 6 (2016)

T
1600423.

IP
[201] S. Wang, J. Wang, M. Zhu, X. Bao, B. Xiao, D. Su, H. Li, Y. Wang, J. Am.

CR
Chem. Soc. 137 (2015) 15753-15759.

[202]
US
L. Shang, H. Yu, X. Huang, T. Bian, R. Shi, Y. Zhao, G.I. Waterhouse, L.Z.

Wu, C.H. Tung, T. Zhang, Adv. Mater. 28 (2015) 1668-1674.


AN
[203] S. Mao, Z. Wen, T. Huang, Y. Hou, J. Chen, Energy Environ. Sci. 7 (2014)
M

609-616.

[204] C. Sun, J. Yang, X. Rui, W. Zhang, Q. Yan, P. Chen, F. Huo, W. Huang, X.


ED

Dong, J. Mater. Chem. A 3 (2015) 8483-8488.


PT

[205] R. Díaz, M.G. Orcajo, J.A. Botas, G. Calleja, J. Palma, Mater. Lett. 68 (2012)

126-128.
CE

[206] H. Wang, Q. L. Zhu, R. Zou, Q. Xu, Chem. 2 (2017) 52-80.


AC

[207] K.M. Choi, H.M. Jeong, J.H. Park, Y.-B. Zhang, J.K. Kang, O.M. Yaghi, 8

(2014) 7451-7457.

[208] X.Y. Yu, L. Yu, H.B. Wu, X.W. Lou, Angew. Chem. 127 (2015) 5421-5425.

[209] D.Y. Lee, D.V. Shinde, E.-K. Kim, W. Lee, I.-W. Oh, N.K. Shrestha, J.K. Lee,

S.-H. Han, Microporous Mesoporous Mater. 171 (2013) 53-57.

79
ACCEPTED MANUSCRIPT

[210] L. Wang, X. Feng, L. Ren, Q. Piao, J. Zhong, Y. Wang, H. Li, Y. Chen, B.

Wang, J. Am. Chem. Soc. 137 (2015) 4920-4923.

[211] L. Zhang, H.B. Wu, X.W. Lou, J. Am. Chem. Soc. 135 (2013) 10664–10672.

[212] Z. Liu, C. Liu, J. Ya, E. Lei, Renew. Energy 36 (2011) 1177-1181.

[213] J. Yang, Y. Lin, Y. Meng, Y. Liu, Ceram. Int. 38 (2012) 4555-4559.

T
[214] X. Gan, X. Li, X. Gao, W. Yu, J. Alloy. Compd. 481 (2009) 397-401.

IP
[215] W. Zeng, L. Wang, H. Shi, G. Zhang, K. Zhang, H. Zhang, F. Gong, T. Wang,

CR
H. Duan, J. Mater. Chem. A 4 (2016) 8233-8241.

[216]
US
C. Guan, X. Liu, W. Ren, X. Li, C. Cheng, J. Wang. Adv. Energy Mater. 7

(2017) 1602391.
AN
[217] H. Hu, B. Guan, B. Xia, X.W. Lou, J. Am. Chem. Soc. 137 (2015)

5590−5595.
M

[218] H. Li, M. Eddaoudi, M. O'Keeffe, O.M. Yaghi, Nature 402 (1999) 276-279.
ED

[219] J. Jiang, Y. Li, J. Liu, X. Huang, C. Yuan, X. W. (David) Lou, Adv. Mater. 43
PT

(2015) 5166-5180.

[220] Y. Yang, D. Cheng, S. Chen, Y. Guan, J. Xiong, Electrochim. Acta 193 (2016)
CE

116–127
AC

[221] Y.P. Gao, K.J. Huang, Chem. Asian J. 12 (2017) 1969-1984.

[222] W. Zhou, X. Cao, Z. Zeng, W. Shi, Y. Zhu, Q. Yan, H. Liu, J. Wang, H.

Zhang, Energy Environ. Sci. 6 (2013) 2216-2221.

[223] X. Xia, C. Zhu, J. Luo, Z. Zeng, C. Guan, C. F. Ng, Small 10 (2014)

766–773.

80
ACCEPTED MANUSCRIPT

[224] Q. Chu, W. Wang, X. Wang, B. Yang, X. Liu, J. Chen, J Power Sources 276

(2015) 19-25.

[225] B. Liu, D. Kong, Z.X. Huang, R. Mo, Y. Wang, Z. Han, C. Cheng, H.Y. Yang,

Nanoscale 8 (2016) 10686–10694.

[226] J. Wang, S. Wang, Z. Huang, Y. Yu, J. Mater. Chem. A 2 (2014)

T
17595-17601.

IP
[227] Y. Liu, S. Wen, W. Shi, Mater. Lett. 214 (2018) 194–197.

CR
[228] L.N. Wang, Y. Ma, M. Yang, Y.X. Qi, Electrochim. Acta, 186 (2015)

391-396.

[229]
US
S. Wen, Y. Liu, F. Zhu, R. Shao, W. Xu, Appl. Surf. Sci. 428 (2018) 616–622.
AN
[230] L. B. Ma, Y. Hu, R. P. Chen, G. Y. Zhu, T. Chen, H. L. Lv, Y. R. Wang, J.
M

Liang, H. X. Liu, C. Z. Yan, H. F. Zhu, Z. X. Tie, Z. Jin, J. Liu, Nano Energy 24

(2016) 139−147.
ED

[231] H. Chen, X.L. Liu, J.M. Zhang, F. Dong, Y.X. Zhang, Ceram. Int. 42 (2016)
PT

8909-8914.

[232] C. Xia, P. Li, A.N. Gandi, U. Schwingenschlogl, H.N. Alshareef, Chem.


CE

Mater. 27 (2015) 6482-6485.


AC

[233] S. Raj, S. K. Srivastava, P. Kar, P. Roy, RSC Adv. 6 (2016) 95760–95767.

[234] X. Li, B. Wei, Nano Energy 2 (2013) 159-173.

[235] T. Chen, L. Dai, Mater. Today 16 (2013) 272-280.

[236] X. Shi, S. Zheng, Z.-S. Wu, X. Bao, J. Energy. Chem. 27 (2018) 25-42.

[237] S. Abouali, M.A. Garakani, Z.L. Xu, J.K. Kim, Carbon 102 (2016) 262-272.

81
ACCEPTED MANUSCRIPT

[238] J. Yang, C. Yu, S. Liang, S. Li, H. Huang, X. Han, C. Zhao, X. Song, C. Hao,

J. Qiu, Chem. Mater. 28 (2016) 5855–5863.

[239] S. Sun, S. Wang, S. Li, Y. Li, Y. Zhang, J. Chen, Z. Zhang, S. Fang, P. Wang,

J. Mater. Chem. A 4 (2016) 18646-18653.

[240] G. A.M. Ali, S. A. A. Manaf, A. Divyashree, K. F. Chong, G. Hegde, J.

T
Energy Chem. 25(2016) 734-739.

IP
[241] Y. Xu, L. Wang, P. Cao, C. Cai, Y. Fu, X. Ma, J. Power Sources 306 (2016)

CR
742–752.

[242]
US
L. Zhang, L. Dong, M. Li, P. Wang, J. Zhang, H. Lu, J. Mater. Chem. A 6

(2018) 1412-1422.
AN
[243] S. Sun, S. Li, S. Wang, Y. Li, L. Han, H. Kong, P. Wang, Mater. Lett. 182
M

(2016) 23-26.

[244] W. Li, F. Yang, Z. Hu, Y. Liu, J. Alloys Compounds 749 (2018) 305-312.
ED

[245] X. Li, W. Sun, L. Wang, Y. Qi, T. Guo, X. Zhao, X. Yan, RSC Adv. 5 (2015)
PT

7976–7985.

[246] L. Huang, W. Zhang, J. Xiang, Y. Huang, J. Materiomics 2 (2016) 248-255.


CE
AC

82
ACCEPTED MANUSCRIPT

Graphic Abstract

T
IP
CR
US
AN

Strategies in the synthesis of NiCo2O4-based materials are discussed. Progress in


M

understanding and modifying NiCo2O4-based materials from various aspects is


ED

summarized. Key strategies for improving the cycling stability of

NiCo2O4-based material are discussed.


PT
CE
AC

83
ACCEPTED MANUSCRIPT

Yan-Mei Li received her Bachelor’s degree from Anhui University of Technology in

T
2016. At present, she is pursuing her M. S. degree in Prof. Ting-Feng Yi's group at

IP
Anhui University of Technology. Her current research focuses on the application of

CR
functional materials in electrochemical supercapacitors.

US
AN
M
ED
PT

Xiao Han received his Bachelor’s degree from Anhui University of Technology in

2016. He is now a second year Master student in Prof. Ting-Feng Yi's group at Anhui
CE

University of Technology. His current research focuses on the application of transition


AC

metal oxides in lithium-ion batteries.

84
ACCEPTED MANUSCRIPT

Ting-Feng Yi received his Ph.D. degree from Harbin Institute of Technology in 2007.

T
He joined Anhui University of Technology as an assistant professor in 2007 and a full

IP
professor in 2011. He has now joined Northeastern University at Qinhuangdao as a

CR
full professor. He has published more than 95 papers in peer-reviewed journals as the

first author or corresponding author, with 2600 citations (H-index=28). His research

US
interests include synthesis of electrochemical functional materials and their
AN
application in lithium-ion battery, supercapacitor and lead-acid battery. For detail

please see his Research ID: http://www.researcherid.com/rid/F-4594-2012.


M
ED
PT
CE

Yan-Bing He received his Ph.D. degree from the Department of Applied Chemistry,

Tianjin University in 2010. He worked as a post-doctoral fellow at Graduate School at


AC

Shenzhen, Tsinghua University from 2010 to 2012 and a visiting scholar at Hong

Kong University of Science and Technology from 2012 to 2013. He is currently an

associate professor of Graduate School at Shenzhen, Tsinghua University. His

research interests mainly focus on lithium ion power batteries and materials, solid

state electrolyte and lithium metal anode.


85
ACCEPTED MANUSCRIPT

T
IP
Xifei Li is currently a full professor at Xi’an University of Technology and Tianjin

CR
Normal University, and he is an executive dean of Institute of Advanced

Electrochemical Energy at Xi’an University of Technology. Dr. Li earned his bachelor

US
degree in Electrochemistry Engineering from Harbin Institute of Technology, and he
AN
then completed his Master Degree and Ph.D. from General Research Institute for

Nonferrous Metals of China and Xi’an Jiaotong University, respectively. He was


M

awarded Ontario PDF and Mitacs Elevate Fellowship from 2009~2013 in Canada. Dr.
ED

Li’s research group is currently working on design, synthesis, as well as performance

improvement of the anodes and the cathodes with various structures for high
PT

performance lithium-ion batteries, lithium-sulfur batteries, sodium-ion batteries, and


CE

supercapacitors. Dr. Li has authored/co-authored more than 160 peer-reviewed

research articles with more than 6000 citations and has fled over 15 patents and patent
AC

applications.

86

You might also like