You are on page 1of 87

Journal Pre-proofs

Review

Development and Application of Carbon Fiber in Batteries

Shengnan Yang, Yan Cheng, Xiao Xiao, Huan Pang

PII: S1385-8947(19)32706-8
DOI: https://doi.org/10.1016/j.cej.2019.123294
Reference: CEJ 123294

To appear in: Chemical Engineering Journal

Received Date: 26 May 2019


Revised Date: 28 September 2019
Accepted Date: 26 October 2019

Please cite this article as: S. Yang, Y. Cheng, X. Xiao, H. Pang, Development and Application of Carbon Fiber in
Batteries, Chemical Engineering Journal (2019), doi: https://doi.org/10.1016/j.cej.2019.123294

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

© 2019 Published by Elsevier B.V.


Contents

1. Introduction .......................................................................................................... 2

2. Lithium-ion battery ............................................................................................. 5

2.1 Pure carbon fiber ......................................................................................................6

2.2 Carbon fiber/silicon ................................................................................................12

2.3 Carbon fiber/Metallic oxide ....................................................................................20

2.4 Carbon fiber/Metallic sulfide ..................................................................................26

2.5 Carbon fiber/carbon material .................................................................................30

3. Lithium-sulfur battery....................................................................................... 32

3.1 Pure carbon fiber ....................................................................................................34

3.2 Carbon fiber/sulfur material ...................................................................................38

4. Other batteries ................................................................................................... 42

4.1 Sodium-ion battery ..................................................................................................42

4.2 Vanadium redox flow battery ..................................................................................47

4.3 Zinc-air battery........................................................................................................50

4.4 Aluminium-air battery .............................................................................................55

4.5 Lithium-air battery ..................................................................................................62

5. Conclusion and outlook ..................................................................................... 63


Development and Application of Carbon Fiber in Batteries

Shengnan Yang, Yan Cheng, Xiao Xiao and Huan Pang*

Shengnan Yang, Yan Cheng, Xiao Xiao and Huan Pang*


School of Chemistry and Chemical Engineering, Yangzhou University
Yangzhou, 225009, Jiangsu, P. R. China.

E-mail: huanpangchem@hotmail.com; panghuan@yzu.edu.cn


Homepage: http://huanpangchem.wix.com/advanced-material

Keywords: carbon fiber, battery

To further enhance the properties of batteries, it is important to exploit new electrode

materials. Carbon fiber has been found to play a crucial role. Various batteries, such

as Lithium-ion batteries, Lithium-sulfur batteries, Sodium-ion batteries, and

Vanadium redox flow batteries, have been investigated. Moreover, greatly improved

performance has been obtained by compositing pure carbon fibers with the metal

materials, metallic oxide materials, metallic sulfide materials, carbon materials and so

on. Through the application of carbon materials and their compounds in various types

of batteries, the battery performance has obviously been improved. This review

primarily introduces carbon fiber materials for battery applications. The relationship

between the architecture of the material and its electrochemical performance is

analyzed in detail. In addition, the development of the history, unsolved problems and

prospects with respect to the introduction of carbon fibers are also included.

1
1. Introduction

Presently, there is an energy problem facing mankind [1]. With the consumption of

petrification energy and increasing environmental deterioration, there is a growing

requirement for highly efficient, tidy, and sustainable resources of energy and new

techniques related to energy change and conservation. Reproducible energy is

undergoing unparalleled development [2]. Wind and solar resources from different

kinds of fresh power resources are the most mainstream renewable power resources to

produce electric energy [3]. However, these sources are not consecutive and cannot be

directed according to demand. For the more available use of natural resources, a

favorable power store is essential. Energy storage technologies related to

electrochemistry, for instance batteries, have been applied to portable electronics and

electric vehicles. However, with an increasing number of functions needed in the

restricted bulk of the easily portable devices, batteries with greater specific capacities

are needed [4]. Therefore, novel materials with the exceptional capacity of

Lithium-ion batteries have aroused great interest [5].

Importance of fibers as reinforcements in metal matrices has long been

recognized. It is only because of this realization that many industries like aerospace

and to some extent the automobile industry relies entirely on advanced composite

materials for their conflicting demands of high performance and lightweight materials.

However most of the work in these areas has been limited or confined to patents

created by the defense and automotive sectors. The efforts to commercialize these

materials have been restricted due to high cost of the reinforcements and the

2
manufacturing processes used. Therefore, people need to focus on the research of the

structure of the cheaper carbon fiber materials, to improve its performance, in order to

achieve its commercialization.

In addition, as excellent next generation power storage equipment, the

Lithium-sulfur battery has attracted considerable attention due to its favorable energy

density of 2600 W h kg-1 in theory, low consumption and non-toxicity [6,7]. However,

the general actual use of these batteries have been limited to increasing and

challenging difficulties including the poor electrical conductivity of sulfur and the

moving ability of soluble lithium polysulfides (Li2Sx, 4<x<8) between the anode and

cathode when cycling [8]. To solve these problems, many methods have been

researched, such as modifying the electrolyte ingredients [9,10], bonding sulfur and

the host substrates [11], covering the surface with a highly conductive polymer layer

[12] and appending an interlayer to the substrate [13]. The normally employed

approach is to utilize carbon material as a sustaining matrix for these sulfur cathodes

because of its high electrical conductivity, low cost and bulky surface proportion. As

a result, different carbonaceous materials, such as porous carbons [14], carbon fibers,

hollow carbon spheres [15] and graphene [16], have been widely regarded as sulfur

hosts. At present, many studies have offered hierarchical micro/mesopore multihole

carbon as an effective hole structure to perfect the electrochemical properties of

Lithium-sulfur batteries. Mesopores can offer a side way for the permeation of an

electrolyte to guarantee rapid conversion of the Lithium-ions during electrochemical

reactions; however, micropores can effectively decrease the spread of soluble Li2Sx

3
[17]. Guo et al. confirm that sulfur in 0.0005 µm micropores appears as small

chain-typed molecules S2-4 due to space limitations [18]. Therefore, the dissoluble

high-order Li2Sx do not appear when the redox reaction occurred. Furthermore,

micropore and mesopore carbonaceous materials often have a large surface proportion,

and it is beneficial to increase the use of sulfur, which offers further sites to the

sediment of undissolvable low-order Li2S2 and Li2S [19]. Additionally, there are other

applications of carbon fiber in other batteries, such as sodium, vanadium redox flow,

zinc, and aluminum-air batteries.

Although we have studied other nanomaterials, we found that carbon fiber in

batteries has all the advantages that no other material can match. For example,

through introduction of black phosphorous nanosheets, we get a laminated

self-standing film with enhanced capacitance and cycling stability via a one-step

electrochemical deposition method. Although it is facile to do, it is not very high in

electrical performance [20]. Carbon fiber has many excellent properties, such as high

axial strength and modulus, low density, high specific performance, no creep,

resistance to ultra-high temperature in non-oxidation environment, good fatigue

resistance, specific heat and electrical conductivity between non-metal and metal,

small coefficient of thermal expansion and anisotropy, good corrosion resistance,

good X-ray transmission. Good conductivity, electromagnetic shielding, etc. However,

the main technical problems existing in the existing technology and need to be solved

are as follows: 1. Solve the problems of dispersion technology and reinforcement

technology of carbon fiber, a special raw fiber material, in order to make a base

4
material with uniform material and excellent strength; 2. In order to ensure that the

substrate meets the required tightness and has a good surface flatness, the hot-pressing

process shall be studied to solve the problems of resin migration, thickness uniformity

and surface flatness; 3. Explore the appropriate high-temperature carbonization

process, improve the strength of the substrate and solve the problem that the substrate

is easy to break under the premise of ensuring that the substrate can reach the polar

resistivity after high-temperature carbonization.

Particularly, in recent years, different types of carbon fiber based materials have

highly influenced electrochemical energy storage; nonetheless, only a small number

of reviews have concentrated on this field. Therefore, it appears reasonable to

generate a review offering opinions on the subject. In this review, we discuss the

research progress regarding carbon fibers and their hybrid materials applied to various

batteries, such as Lithium-ion batteries, Lithium-sulfur batteries, Zinc-air batteries,

vanadium redox flow batteries, sodium-ion batteries, and aluminum-air batteries. The

synthesis procedures, the charging and discharging principles of various equipment

and the chemical properties of these materials are covered. The merits and

disadvantages regarding the structure of the carbon-fiber-based materials are also

discussed in detail. Furthermore, we present the challenges and the future of these

materials in battery applications [21].

2. Lithium-ion battery

5
Lithium-ion batteries are secondary (rechargeable) batteries that rely on lithium ions

to move between the positive and negative poles. During charging and discharging,

Li+ is embedded and de-embedded back and forth between two electrodes: during

charging, Li+ is de-embedded from the positive electrode and embedded into the

negative electrode through electrolyte, which is in the state of rich lithium. Discharge

is the opposite.

2.1 Pure carbon fiber

Because Lithium-ion batteries have great energy densities and an appropriate work

life, they have dominated the rechargeable energy storage market during the last

twenty years. In a commercial Lithium-ion battery, the cathode is often LiCoO2, and

accordingly, the anode is often graphite. Graphite is chosen because of steady

circulation, excellent coulombic efficiency, low working electric tension, and rational

capacity. Nevertheless, because graphite favors the reaction of lithium at extremely

small voltages, the graphitic materials are liable to undergo electroplating using

lithium when they charge in low-temperature environments and at rapid speeds. The

electroplating is likely to cause fatal battery breakdowns. Furthermore, top-class

Lithium-ion batteries now need major increases in the specific storage and cycling

speed to meet future energy capacity requirements in electrical cars and electrified

wire netting storage science [22,23]. Therefore, a novel anode material for

Lithium-ion batteries with enhanced invertible energy, speed capability, and battery

security is required. Table 1 summarizes the representative structure of carbon fiber

in Lithium-ion battery. Through this table, we can observe the properties of various

6
carbon fibers, including, reversible capacity, current density, capacity retention, cycle

number, and the use of electrolyte for batteries. We can clearly tell which carbon

fibers are optimal

For the sake of this goal, researchers have perfected the electrochemical

properties of carboneous materials via granule reduction to a very small extent. By

decreasing the Li+ spread lengths in carbon and adding a cover scale to all the

connected parts, a considerable enhancement in speed can be found in the carbon

anodes [24-26]. For example, Qie et al. have compounded nitrogen-doped carbon

nanofiber nebs employing the carbonation-excitation of polypyrrole nanofiber nebs

with potassium hydroxide and presented a large invertible energy of 943 mA h g-1

when the current density was 2 A g-1 [24]. Moreover, a considerable rate capability

(350 mA h g -1 at 100 mA g-1) for the carbon-nanofiber-net anode via carbonation of

electrospun polyacrylonitrile (PAN) superpolymer nanofibers was found by Kim et al

[26]. The voltage profile of the tough carbon material reveals a leaning profile rather

than a low-voltage highland, which suggests that this type of material has more secure

anodes than graphite when conducting high energy use [27]. In addition, porous

hierarchical carbon nanoribbons produced out of a penester as the anode of a Li

battery were utilized by Campbell et al., and a particular energy of ~200 mA h g-1 (on

average) at 100 mA g-1 with a considerable columbic efficiency was reported [28].

After a pyrolysis temperature of 1100 °C, the pristine anodes achieve over 260 mA h

g-1after 700 cycles and a Columbic efficiency of 101.1%, without the use of harmful

solvents or chemical activation agents. At the same time, the self-supporting

7
nanoribbon anodes demonstrate significant capacity increase as they undergo

additional charge/discharge cycles. Hence, carbon fibers have revealed enhanced

electrochemical properties, such as greater capacities, excellent cycling stabilities and

no sustained enhancement in terms of energy.

Thus, a facile compound of carbon fibers was reported from Tyromyces fissile

wild fungus with a controlled carbonation process. Scanning electron microscopy

demonstrated 40-60 μm long stereoscopic fibers as well as hollow fibers imitating the

spontaneous grain composed as-prepared carbon fibers. These carbon fibers cover

chaotic carbon construction because they have larger interlayer spacings than graphite,

as revealed by electronic testing equipment. The specific reversible energy of 340 mA

h g-1 was transferred by these carbon fibers at a speed of C/10, and an energy of 300

mA h g-1 was transferred by these fibers at a speed of C/5. The uniform decoration of

cobalt oxide particles further improved the electrochemical properties of the carbon

fibers using the solid-state thermal processing method [29].

8
Figure 1.The application of pure carbon fiber material in Lithium-ion battery
(a) Preparation of the rechargeable Lithium-ion battery using wild fungus-derived
carbon fiber anode and its utilization in a full cell to power light-emitting diodes. (b)
SEM and (c) TEM images of the carbon fibers. (d) Long cycling of TF500_3 (anode
cycled between 0 and 3 V at a C/5 rate (1C = 372 mA g-1). (e) Discharge-charge
profile (f) Rate capability comparison of TF500_3, TF800_3 and TF900_3. (g) Cyclic
voltammetry of a half-cell with TF500_3 anode. (a-g) Reprinted with permission from
Ref. [29]. Copyright 2016, American Chemical Society.

Carbon fibers were successfully fabricated by Tyromyces fissilis fungus by an

easy pyrolysis step, as shown in Figure 1a. The carbon fibers were used as anodes for

the Lithium-ion batteries after hand milling. A recharged battery including a carbon

fiber complex anode and a traditional LiCoO2 cathode with a 1 M LiPF6 bath solution

can power 2 V and 3 V lamps with different colors. In addition, various half batteries

were built to explore the electrochemical performance [30].

Further research on manual milled TF800_3 carbon fibers using scanning

electron microscopy (SEM) has shown that lengthy carbon fibers are resolved into

9
lesser debris (Figure 1b) by grinding. Every fragment included 40-60 μm long solid

fibers and 10-20 μm long hollow fiber embranchments. Figure 1c indicates that

carbon fibers consist of mainly unformed carbon with a scarce portion of onion as a

graphitic carbon structure. Because of the lower preparation temperature without a

catalyst, the brief distance ordering of a carbon slice was found [31].

A specific capacity of ~300 mA h g-1 was shown when the TF500_3 anode was

cycled at a rate of C/5 in 150 cycles (Figure 1d). The first coulombic efficiency is

merely 40% because of the large initial cycle nonreversible capacity (~400 mA h g-1).

By a nonreversible electrolyte swop with carbon functional teams, during the initial

discharge in the half battery, the initial cycle nonreversible energy is due to

considerable solid-electrolyte interphase (SEI) formation on the carbon surface [32].

Surface functional teams are generally viewed in biomass deuterogenic carbon

because of the spontaneously existence of rich oxygen and hydrogen ingredients in

this material [33]. Nevertheless, the coulombic efficiency rapidly grows to close to

100% in the terminus of the tenth cycle and maintains this level while covering the

other cycles. In general, the electrode of this material demonstrated a superb capacity

reserve of ~93% (because of the initial and 150th charging capacities). Among the

late scans, simple broad reduction peaks were observed between 1 V and 0 V but

contain many lower peak altitudes, which is likely due to the stopping of further SEI

formation by the first SEI tier [24]. Although the anodic peaks under the condition of

0.5 V slightly increased, these peaks maintained their form and strength. The

duplication of these late scans indicates the steady cycling of this anode behind the

10
first cycling. Discharging and charging outlines of this anode, as shown in Figure 1e,

suggest satisfactory agreement with the survey via the cyclic voltammetry (CV)

data(Figure 1g). During the course of the initial cycle, this abrupt lean transform at a

voltage of 0.8 V on the discharging curve shows the start of the first SEI formation on

the carbon fibers. Later cycles show overlapping discharge charge profiles as well,

revealing the invertibility of the lithium insertion course behind the initial cycle.

Moreover, the abovementioned TF500_3 anode indicates increasing capacities

under the circumstances of different C-speeds (Figure 1f) compared with the

TF800_3 and TF900_3 anodes. For instance, the specific capacities of the carbon

fiber anodes at sluggish cycling speeds of C/10 are 350, 250 and 210 mA h g-1.

TF500_3 can deliver the highest capacities that include the best class of chaotic

carbons, which have been found to transport considerable capacity in Lithium-ion

batteries [34,35].

These carbon fibers derived from Tyromyces fissilis fungus. Further performance

enhancement was achieved by decorating carbon fibers with cobalt oxide particles

through solid state mixing and subsequent thermal decomposition steps. The high

reversible capacity of 530 mAh/g at C/10 rate delivered by the CoO carbon fiber

hybrid anode with 10 mol % CoO loading is attributed to the conductive network

formed by the cross-linked carbon fibers and faster reaction kinetics promoted by

monodispersed CoO nanoclusters. This kind of pure carbon fiber shines light on the

potential of synthesizing renewable carbonaceous materials directly from fungus and

11
their utilization as anodes in Lithium-ion batteries But its performance is not very

high.

2.2 Carbon fiber/silicon

Recently, considerable attention has been focused on metallic anode materials that are

able to offer to increase energy, power density and security. A better performance

increase was obtained by compositing pure carbon fibers with silicon. The bulk

extension of silicon through honeycomb-like holes is accommodated by the carbonic

structure surrounding the core region. This type of relational carbon hull can hold

back an electrolyte from immerging into the core, and at the same time, it can increase

the electric conductivity by the touches under the fiber mesh [36].

To increase the capacity density, silicon is presented as the anode material due to

the abnormal specific capacity and poor electric potential [37,38]. Nevertheless, a

number of silicon electrodes still maintain restricted circulation ability. A volume

change of >300% will tautologically take place during the process of (de)lithiation,

which leads to the structural breakdown of the silicon electrodes; this breakage is

verified by the shattering of the material, the loss of the electric link to the conductive

agent and current collector, and the production of the erratic solid-electrolyte

interphase [39-41]. To maintain an intact electrode architecture where highly efficient

ions and electrons shift to the silicon electrodes during cycling, silicon nanoparticles

(Si NPs) in the conductive and capacious base material are considered preferable.

Carbon materials have immense advantages as anodes during use in Lithium-ion

batteries, such as electrochemical stability, mechanical strength, long lifetime,

12
electron transmission and small-scale volume expansion [42-44]. Hence, Si NPs have

been diffusely applied to the carbon matrix to manufacture silicon/carbon (Si/C)

composites with the aim of improving their cycling stability and rate performance

during the reaction process [42,45]. The outstanding electrochemical performance is

deemed to result from the stratified structure of the Si/C fibers.

Figure 2. The application of carbon fiber/silicon material in Lithium-ion battery


Schematic illustration of the preparation process and structure of the Si@IHCFs. (a)
The electrospinning process using a dual nozzle. The PAN/PMMA solution
containing Si NPs, and the PVP solution containing MWCNTs were injected into the
core and shell channels of the nozzle, respectively. (b) The hierarchical core shell
structures of the Si@IHCFs and their precursor fibers. (A color version of this figure
can be viewed online.) (c) Interconnected shell structure for the Si@IHCFs. TEM
images of (d) Si@IHCFs electrochemical performance of the Si/po-C@C electrode.
(e) Si@HCFs and Si@IHCFs with various masses measured at 0.2 A g-1 with a
voltage range of 0.005-1.5 (A color version of this figure can be viewed online.) (f)
Electrochemical performance of the Si-based electrodes. (g) Schematics of the

13
structural design of the Si/po-C@C composite fiber: Three-dimensional (3D) sketch
of the overall structure. (h) Structural formation and change in Si/po-C@C during
heat treatment and the cycling process. (i) Cycling performance of Si/po-C@C, pure
Si NPs and po-C@C measured at 0.2 A g1, and the corresponding CE for Si/po-C@C.
(j) SEM images of the cross-sectional Si/po-C@C (k) po-C@C fibers. (a-f) Reprinted
with permission from Ref. [36]. Copyright 2016, Elsevier Ltd. (g-k) Reprinted with
permission from Ref. [46]. Copyright 2015, the Royal Society of Chemistry.

During this research study, we combined Si and the interconnected hollow

carbon composite fibers (Si@IHCFs) with a layered core shell architecture by using

the double coaxial electrospinning technique [47]. The process of this particular

synthesis and structural diagram researched according to the Si@IHCFs are presented

in Figure 2a,b. Regarding the Si@IHCFs, the Si NPs were sealed into the intrinsic

porous carbon architecture by means of the interconnected carbon shell network. The

addition of the complex peroxyacetyl nitrate/polyester solution and Si NPs was

applied as the nuclear solution with the aim of manufacturing a porous carbon

structure with inbuilt Si NPs. Meanwhile, a polyvinyl pyrrolidone solution with

multiwalled carbon nanotubes was applied as the shell solution to form an interlinked

carbon shell network. The final composition of Si@IHCFs were compared with that

of Si/C with no interlinked carbon shell environment, and the results show a greater

performance in terms of the cycle stability and the rate ability. With regard to the total

fiber interwoven network, these mixed fibers are cross-connected with the mixed

carbon points and multi-walled carbon nanotubes on the surface of these tight carbon

shells, as clearly shown in Figure 2c. The construction of the Si@IHCFs was further

researched by TEM, and the resulting image is presented in Figure 2d; clearly, the Si

14
NPs are greatly dispersed into the polyporous frame core and were hermetically

sealed through an integrated shell among these fibers [48].

The cycling properties of different types of electrodes are tested in a 0.2 A g-1

environment that changed between 0.005 and 1.5 V due to the pure Si, Si/CFs,

Si@IHCFs and IHCF, as displayed in Figure 2f. The results show that poor cycling

performance is exhibited by both crude Si and the Si/CF electrodes. The capacity of

the crude Si as well as the Si/CF electrode reveals a rapid fading over the course of

the cycles. The reversible capacity decreased from 2749 mA h g-1 at the initial cycle

to 362 mA h g-1 at the final cycle of the fifty cycles for crude Si. At the same time,

after one hundred cycles for the Si/CFs, the largest invertible capacity of 1076 mA h

g-1 at the 5th cycle dropped to 416 mA h g-1. Because of this reduplicative large

volume variation in the Si NPs, the electrical link from conductive additaments on the

crude Si electrode can decrease [38,49], leading to fleet capacity fading [50].

Compared to the Si/CFs, the Si NPs are inlaid in the carbon fiber matrix, and the

modified electrical contact results in a comparatively good capacity retention among

the primal 30 cycles. Nevertheless, the capacity is visibly reduced during the course

of the following cycles. Because of the disarray of particle materials on the surface of

the fibers, and revealed firsthand, a thick SEI would cover the Si surface during the

course of the cycling process, preventing electron and ion transport [51]. Because of

the good-sized volume swell and the shrinkage of Si, the SEI could fracture and

rebuild tautologically, just at that moment the electrode surface is directly exposed to

the electrolyte [40]. Furthermore, the gathering of Si NPs along that fiber axle results

15
in an accumulated large expansion stress; hence, the Si NPs will exfoliate from the

carbon fibers, and the electric link between Si and carbon stroma would subsequently

be lost.

In addition, we also found another structure with excellent performance, and its

fabrication steps and structure are shown in Figure 2g and 2k, respectively. The

silicon/carbon (Si/C) hybrid fibers that have a straticulate nucleus-shuck architecture

is manufactured by encasing Si NPs into the mutually connected tubular carbon fibers

using the dual coaxial electrospinning method. Because of the straticulate architecture,

the Si NPs are embedded into the cellular carbon skeleton of the fiber core and then

wrapped in a web of interlocking carbon shells. Consequently, superb electrochemical

performance can be achieved due to the special structure [52]. Next, we conducted

further research to improve this result. In accordance with the general weight of the

Si/C compound, the explicit capacity of Si/po-C@C is clearly demonstrated in Figure

2i. It is essential to assess the capacity of Si in the Si/C compound. On the basis of the

thermogravimetric analyzer curve, the weight percentage of the Si constituent in

compound materials is 21.8%. According to the specific capacities of 200 mA h g-1

with when the cycle in Figure 2i and a weight ratio of 78.2% for the carbon matrix in

the Si/C compound, after removing the effect of the carbon matrix materials from the

capacity of the compound, the charging specific capacities of Si in the Si/po-C@C

electrode is recounted to be 3612 mA h g-1 during the first cycling and is able to be

maintained at 2528 mA h g-1 after one hundred and fifty cycles. An increased

procurability of Si in the Si/po-C@C compound, compared with the greatest values

16
ever reported capacity range, can be obtained [53-60]. As is clearly shown in Figure

2i, the first discharging capacity of the original Si material electrode is 3422 mA h g-1.

Nevertheless, for the Si/po-C@C composite, these Si compound materials are capable

of offering a first discharging capacity up to 4065 mA h g-1. This phenomenon

demonstrates that the porous carbon structure in this type of composite optimizes the

procurability of Si by fabricating a three-dimensional conductive network with Si NPs

in the inside of the fiber.

In addition, we also compared another sample. Freestanding membranes of the

Si@HCFs and Si@IHCFs were used as electrodes with no binder, conductive agent

and current collector to confirm the advantages of the linked framework in the fiber

net. The electrochemical properties of the independent membrane electrodes are

presented in Figure 2e. It is significant that the specific capacity of the independent

membrane anode is not as high as that for the slurry-derivative Si@HCF and

Si@IHCF electrodes and is reduced inch by inch with increasing mass and scilicet

thickness. This result indicates that the conductive structure in the film because of

these nonwoven-like stored insulates some fiber individuals. The performance of

Si@HCF membranes decreased more rapidly during the course of cycling compared

with these anodes. As shown in Figure 2e, the original invertible property of 703 mA

h g-1 reduced inch by inch to 437 mA h g-1 after 40 continuous cycles, revealing less

retention than the derived Si@HCF electrode. On the basis of the charge-discharge

curves, distinct voltage plateaus are apparent at approximately 0.7 V in the initial

discharging cycle, which is also because of the solid electrolyte interface production

17
on the carbon matrix, the cointercalation of water-soluble Li+ into the graphene sheets

[61,62], and the change between 0.2 and 0.6 V in the charge curves because of the

release of Li from these LixSi alloys [58,61]. Hence, the insulative solid electrolyte

interface layer production on the fibers is beneficial to more rapid attenuation of the

performance of the Si@HCF membrane, which reduces the structure conductivity

inch by inch in the inattentive cumulated membrane.

Figure 2e presents the Si@HCF and Si@IHCF membrane electrodes along with

the cycling performance. Notably, the capacity retentions of these two membranes

transmit adverse tendencies, also with a piecemeal recession for the Si@HCFs.

Through continuous forty cycles, the performance of the Si@IHCF materials becomes

better than that of the other material, even surpassing the original invertible capacity

for its material, even if the loading mass of Si@HCFs is less than one third of that of

the Si@IHCF materials. Therefore, we deduce that the performance and retention

properties of the Si@IHCFs is larger than those of the Si@HCFs with a similar range

of loading masses. Excellent properties can be allocated to the interlinked architecture

in the carbon fibers, leading to additional steady and valid aisles for electron shifting.

With closed and perpetual carbon-welding points, the insulative solid electrolyte

interface tier production on these fibers cannot break electric contact with the

individual carbon fibers, which is different than the Si@HCFs that are separate from

the interconnected structure. Then, the lesser performance during the incipient few

cycles may be due to the larger mass loading and terse superposition of the

Si@IHCFs. Therefore, the electron shift pathway could be longer, and the infiltration

18
of the electrolyte might be suppressed. In addition, the attenuation in the performance

with cycling can be attributed to the piecemeal activating reaction of the carbonic

material with a terse and thick shell among the processes of Li insertion/extraction

[63].

For the sake of identifying the multipurpose and stratified architecture of the

Si/po-C@C compound materials, pure porous carbon fibers (po-C@C) with no Si

were acquired from Si/po-C@C after the Si NPs were corroded via the NaOH solution.

The TEM image in Figure 2j,k characterizes the form and architecture of po-C@C.

Therefore, the original and intrinsic architecture of the carbon matrix is distinctly

proven. On the basis of this picture, fibrous carbon matrices are composed of two

cross profiles: a poriferous carbon structure in the interior core and a tight carbon tier

as an external shell. The mass of the spherical pores rooted in the process of

eliminating the Si NPs has been congruously dispersed in the three-dimensional

interlinked carbon structure, as clearly shown in Figure 2j, k. This supposition is that

the interior structure and outer shell cannot be isolated but remains an intact carbon

matrix. According to speculation, the three-dimensional interlinked network can

likely serve as a type of overly highly efficient conductive pathway during the course

of the close encounter during the carbon matrix and Si NPs [64].

In summary, for the first material, the Si@IHCFs with a novel hierarchical

coreeshell structure have been prepared based on the coaxial electrospinning

technique with a coaxial dual nozzle. For the composite fibers, Si NPs are embedded

in a honeycomb-like carbon framework in the core region, and the hybrid core is

19
further encapsulated by the cobweb-like interconnected carbon shell. The superior

electrochemical properties of Si@IHCFs are attributed to the well-designed unique

architecture. The porous carbon framework in the fiber's core could accommodate the

volume expansion of Si NPs by the void pores, as well as reinforce the electron and

ion transport by the honeycomb-like carbon skeleton and hierarchical pore structure.

The interconnected and compact carbon shell wrapping the core section could not

only block the electrolyte, but also act as continuously conductive pathways to

improve the electron transport. The second material demonstrates a high specific

capacity, stable cycling performance, excellent rate capability, as well as outstanding

accessibility of Si. The superior electrochemical properties are attributed to the

welldesigned unique architecture of the Si/po-C@C fibers. The porous carbon

framework could not only accommodate the volume expansion of Si NPs by the void

pores, but also reinforce the electron and ion transport by the honeycomb-like carbon

skeleton. The compact carbon shell wrapping the core section could block the

electrolyte, so that, a stable SEI film is formed outside the shell. It is also noted that

future work is needed to further enhance the specific capacity and cycling stability by

optimizing the structure design and preparation conditions.

2.3 Carbon fiber/Metallic oxide

Many metallic oxides have been widely used as positive pole materials for

high-performance Lithium-ion batteries due to their improved specific capacities

compared with merchant graphite [65]. Fe3O4 has been generally researched as one of

20
the possible candidates because of the supernal theoretical specific capacity of 924

mA h g-1, properly cut-price and green generated means [66]. Nonetheless, only the

Fe3O4 anode could be subjected to structural destruction, and a lack of electrical joint

of the anode material could be obtained through current collectors because of the

large volume change (93%) of Fe3O4 when the lithiation/delithiation reaction led to a

fast capacity attenuation [67,68]. The undesirable conductivity of Fe3O4 is also an

issue when attempting to acquire a large capacity and steady cycle properties at larger

current densities.

To eliminate the weakness of the pure Fe3O4 anode, carbon materials are often

used as stroma to produce Fe3O4/C composites with high electrochemical

performance due to their considerable conductivity, electron exchange rate, and

chemical and volume durability.

The Fe3O4@PCF composite shows a first invertible capacity of 1015 mA h g-1

with an 84.4% capacity conservation after 80 cycles in an environment of 0.2 A g-1.

The electrode shows a high speed capability when the current density gradually

increases to 2.0 A g-1 and the capacity conservation reaches 91% over 200 cycles at

2.0 A g-1. These singularly great properties are due to the excellent electrical

conductivity and chemical stability of these poriferous carbon fibers because of their

special architecture. On the one hand, this structure can decrease the volume variation

in Fe3O4 towards the inner space. On the other hand, the structure serves as large

carriage channels for the charged corpuscles [69]. Moreover, the pyknotic carbon

shuck is able to accelerate the generation of the steady solid electrolyte interphase on

21
the surface of the fibers [70]. The particular building-up process and structural chart

for Fe3O4@PCFs are demonstrated in Figure 3a. The SEM image of Figure 3d shows

the Fe3O4@PCFs with submicron diameters and slippery surfaces. Additionally, the

NPs are rarely observed from the outside, indicating that the Fe3O4 NPs are

adequately packaged in the fiber using a carbon cover. This architecture that the

Fe3O4 NPs are well packaged by the carbon substrate in the Fe3O4@PCFs is

ultimately confirmed via TEM observation. It is obvious that a special core–shell

architecture in Figure 3e and the Fe3O4 NPs are packaged in the poriferous carbon

cores covered by carbonic shells. This architecture that the Fe3O4 NPs are well fixed

by the carbon substrate in the Fe3O4@PCFs is ultimately confirmed using TEM

measurements. As shown in Figure 3b, the discharge and charge voltage profiles of

Fe3O4@PCFs in the context of 0.2 A g-1 from 0.01 to 3.0 V in diverse cycling

conditions are shown. This nonreversible reaction primarily appears in the 1st cycle.

Therefore, we find that the improved property of the battery is due to the conservation

of these carbon shells. Figure 3c also shows that the Fe3O4@PCFs have a greater

invertible capacity than bare Fe3O4 and PCFs.

22
Figure 3 The application of carbon fiber/metallic oxide material in Lithium-ion
battery
(a) Schematic illustration of the preparation process and structure of the Fe3O4@PCFs.
(b) Discharge/charge profiles of the Fe3O4@PCF anodes at a current density of 0.2
A-1 for the 1st, 2nd, 20th, 40th and 80th cycles. (c) Cycling performance of
Fe3O4@PCF, pure Fe3O4, and PCF anodes at 0.2 A g-1, and the corresponding CEs for
the Fe3O4@PCF anode. (d) SEM image of the Fe3O4@PCFs. (e) TEM image of
Fe3O4@PCFs. (f) Schematic illustration of the crystal structure of ZnCo 2O4. (g)
Schematic illustration of the operating principles of a rechargeable Lithium-ion
battery based on ZnCo2O4 electrodes. (h) Discharge/charge profiles of the
ZnCo2O4-urchins-on-carbon-fibers electrodes at the current densities of 0.2 °C. (i)
Cycling performance of the electrodes at 0.2 °C over 100 cycles and their
corresponding columbic efficiency (CE). (j) FESEM image. (k) TEM image of the
ZnCo2O4 urchins. (a-e) Reprinted with permission from Ref. [71]. Copyright 2015,
Elsevier Ltd. (f-k) Reprinted with permission from Ref. [72]. Copyright 2013,
Tsinghua University Press and Springer-Verlag Berlin Heidelberg.

Although the above materials have some advantages, there are still some

shortcomings. To satisfy the needs of rapid energy transformation and consumption

equipment, replaceable anodic materials with greater performance characteristics,

23
such as strong speed capability and higher continuable energy supply, have been

explored [73-75]. In particular, ZnCo2O4 is a superb choice as an anodic material

because of its great invertible capacity, long working life, and green characteristics

[76,77]. However, the low conductivity of ZnCo2O4 vivacious materials results in

superfluous property degradation when the charge and discharge cycles are tested at a

high current density. Therefore, we developed high-energy Lithium-ion batteries with

self-assembled ZnCo2O4 on these carbon fibers as the no-binder anodes that are

produced by developing ZnCo2O4 urchins on certain special carbon fibers. The

invertible capacity was also able to attain 1,180 mA h g-1 after over 100 cycles,

showing the high invertible capacity and great cycle life of the anodes. To our

surprise, the as-prepared anode exhibited an unexpectedly high rate of 20 C (18 A g-1)

with a superb capacity of 750 mA h g-1. This phenomenon exceeds that of the greatest

value that has been found for ZnCo2O4-based anodes.

Figure 3j demonstrates the SEM image of the synthesized product. Carbon

fibers with smooth surfaces were evenly covered with high-density samples on the

product. The crystal structure of spinel ZnCo2O4 is demonstrated in Figure 3f. Zn, Co,

and O atoms are represented by different color spheres that are yellow, purple, and

green, respectively. More results regarding the ZnCo2O4 urchins was achieved by

transmission electron microscopy (TEM). The magnified TEM image shown in

Figure 3k demonstrates that typical urchins with a radius of approximately 40-50 nm

have poriferous structures. Moreover, the ZnCo2O4 nanowires consist of various small

nanocrystals.

24
The operating principles of the ZnCo2O4-urchins on carbon fiber electrodes are

shown in Figure 3g. Moreover, at speeds of 0.2 C, the voltage profiles of the

ZnCo2O4-based electrodes for the first, second, tenth, and hundredth charge and

discharge cycling tests in the range of 0.01-3.0 V are demonstrated in Figure 3h. In

line with the voltage distribution, we find that every discharging curve shows obvious

plateaus within the range of 1.0 to 1.2 V. Surprisingly, the discharging capacities for

the initial cycle at rates of 0.2 and 5 C are all greater than the theoretical capacity for

ZnCo2O4 (900 mA h g-1). On the basis of previous research, this anode material would

be attributed to the nonreversible reactions that are usually seen [78-80]. Behind the

first cycle, the next charge/discharge graphs approached a steady state, explaining that

these reactions gradually went into the cycle stages [80]. Figure 3i illustrates the

cycling performance at speeds of 0.2 C for the ZnCo2O4-based electrodes. These

curves clearly show that the synthesized ZnCo2O4-based electrodes increased the

invertible discharge capacity under the condition of 0.2 C. Furthermore, its capacity

also remains at 1180 mA h g-1 for over 100 cycles, showing the great invertible

capacity and outstanding cycle lifetime [81].

In summary, people have designed a novel Fe3O4@PCF composite with 1D

hierarchical core-shell structure by embedding Fe3O4 NPs in porous carbon

framework as core, which is further wrapped by the carbon shell (Fe3O4@PCFs).

Moreover, Advanced rechargeable Lithium-ion batteries have been successfully

fabricated by using hydrothermally synthesized ZnCo2O4-urchins-on-carbon-fibers

25
structures as the binder-free anodes. A highly flexible full battery was also fabricated

with good electrical stability under fully mechanical bending.

2.4 Carbon fiber/Metallic sulfide

The FeS2@carbon fiber electrodes exhibited steady cycling properties in all the

traditional carbonated electrolytes. The discharging power density of the Al2O3-coated

FeS2@carbon fiber electrode first arrived at 1000 Wh kg-1-electrode (~1300 Wh

kg-1-FeS2 at material level) left at 840 Wh kg-1-electrode (~1110 Wh kg-1-FeS2) over

100 cycles (Figure 4b), which is greater than other reports for FeS2. Figure 4a shows

the working procedure to increase the cycle stability by covering Al2O3 at 1.0-3.0 V

(vs Li/Li+). As for the uncovered composite electrode, the repetitive volume led to

FeS2 particulates that split away from the carbon fibers, resulting in capacity damping.

Figure 4c indicates the SEM image of carbonized PAN Fe(C5H7O3)3 fibers after

conducting sulfuring treatment [82]. The TEM EDS elemental mapping in Figure 4d

shows that the nanoparticles are composed of Fe and S, proving that the nanoparticles

are FeS2.

Considering the the capacity attenuation operating principle, the cycling

stabilization of the Li-FeS2 reaction environment at 1.0-3.0 V is obviously increased

by covering the FeS2@carbon fiber electrode using a thin layer of Al2O3. The

Al2O3-covered electrode shows superb cycle properties and has a large discharge

power density.

26
Figure 4. The application of carbon fiber/metallic sulfide material in Lithium-ion
battery
(a) Schematic illustration of the mechanisms for the enhanced cycling stability by an
Al2O3 coating in the voltage range of 1.0-3.0 V (vs Li/Li+). (b) Discharge energy
density vs. the cycling number at the material level (~Wh kg-1-FeS2) and the electrode
level (~Wh kg-1-FeS2 electrode). (c) SEM image of the carbonized PAN Fe(C5H7O3)3
fibers after sulfidation. (d) Sulfur for the FeS2@carbon fiber. (e) Schematic
illustration of the synthesis process and structure of the as-designed MoS2/ACF cloth.
(f) Cycling performance of the MoS2/ACF cloth at current densities of 0.1 A g-1, 0.2
A g-1 and 0.5 A g-1. (g) Rate performance of the MoS2/ACF cloth under various
current densities ranging from 0.1 to 1.5 A g-1. (h) TEM image of the microstructure
of the MoS2/ACF cloth. (i) SEM image of the as-obtained hierarchical MoS2/ACF
cloth. (a-d) Reprinted with permission from Ref. [83]. Copyright 2016, American
Chemical Society. (e-i) Reprinted with permission from Ref. [84]. Copyright 2014,
The Royal Society of Chemistry.

Because of the hierarchical structure and desirable academic capacity, MoS 2 is

an excellent candidate as a highly competitive anode material for Li batteries.

Nonetheless, this material is subjected to fast capacity recession and an insufficient

rate capability [85]. In this research, we discovered a new layered material composed

27
of exceeding thin MoS2 nanosheets developed on the active carbon fiber (ACF) cloth

made through an easy change in the cloth shape [86]. The function of ACF cloth is a

form board and a stabilizer. The composition of the MoS2/ACF cloth possesses

multilayered porosity and a mutually connected structure. Working as an independent

and adhesive-free anode, this cloth demonstrates a great specific capacity and

excellent invertibility [87].

Figure 4e shows the two steps used to synthesize the MoS2/ACF cloth. This

MoS2/ACF cloth paper was used for future research and immediately served as an

electrode. Figure 4i shows the SEM image indicating the morphology of the

MoS2/ACF cloth. The TEM image shown in Figure 4h indicates the structure and

degree of crystallinity of the MoS2 nanosheets. The layered MoS2/ACF cloth indicates

an unusual construction that unites the merits of both one-dimensional and

two-dimensional structures. Figure 4f shows this specific discharging capacity with

excessive cycling at 0.1 A g-1, 0.2 A g-1 and 0.5 A g-1, and the capacities during the

initial cycle are 1392 mA h g-1, 1262 mA h g-1 and 1222 mA h g-1, respectively.

Moreover, the relevant charging capacities are 1125, 1036 and 908 mA h g-1, which is

superior to commercial MoS2 stives. Clearly, the nonreversible capacity damage when

the initial cycle is approximately 267 mA h g-1 is primarily due to the production of

the solid electrolyte interphase tier and the joining of Li, where MoS2 is vacant.

Nonetheless, in the next cycles, the coulombic efficiency is close to one hundred

percent. Except for the greater specific capacity in the first cycle at 0.1 A g-1, the

MoS2/ACF cloth has a high cycling property; during the next cycles, the discharge

28
capacity remains unchanged, and even damping occurs. The same situation also

occurs at 0.2 A g-1 ; that is, the MoS2/ACF cloth is steady over 200 cycles, and its

discharging capacity is maintained at 635 mA h g-1. If we continue increasing the

current density to 0.5 A g-1, then the capacity reaches 418 mA h g-1 after 200 cycles.

The MoS2/ACF cloth demonstrates a high cycling property compared with the

reported composition of MoS2/carbon [88-91] and is the same as the MoS2/carbon

fiber composite made via a hydrothermal method [92]. Figure 4g demonstrates the

speed capability in the range of 0.1 to 1.5 A g-1 for the MoS2/ACF cloth. We can also

see that the hierarchical MoS2/ACF cloth electrode has a large capacity.

In summary, we have developed a self-standing FeS2@carbon fiber electrode

with FeS2 nanoparticles either encapsulated in or attached to the interconnected

one-dimensional carbon fibers. Considering its low cost, earth abundance, relative

safety, and maturity in commercial primary batteries, it is expected that the

rechargeable Li-FeS2 system will become competitive with the Li-S system in the

future. The second material was fabricated for the first time via a simple dissolution

and sintering process, and directly used as a binder-free and freestanding anode for

Lithium-ion battery. And the preparation method developed in this work is also

applicable for other functional materials used in supercapacitors, catalysts, etc.

Table 1. Summary of the carbon fiber materials reported in the Lithium-ion battery
and Lithium-sulfur battery section.
Reversible Current Capacity Cycle
Carbon fiber Composite Electrolyte Ref.
capacity density retention number

Carbon fibers CoO 340 mA h/g - ∼93% 300 1 M LiPF6 [29]

29
Interconnected
Si 903 mA h/g 0.2 A/g 89% 100 1 M LiPF6 [36]
hollow carbon fibers

Carbon fibers Si/po-C 997 mA h/g 0.2 A/g 71% 150 1 M LiPF6 [46]

Carbon fibers Si/po-C 603 mA h/g 0.5 A/g - 300 1 M LiPF6 [46]

Carbon fibers ZnCo2O4 1180 mA h/g - - 100 1 M LiPF6 [72]

Porous carbon Fe3O4


1015 mA h/g 0.2 A/g 84.4% 80 1 M LiPF6 [71]
fibers nanoparticles

Hierarchical MoS2
Active carbon fibers 971 mA h/g 0.5 A/g 99.85% 200 1 M LiPF6 [84]
nanosheet

Carbon fibers FeS2 - 0.2 A/g 90% 100 1 M LiPF6 [83]

N-doped carbon
Li3V2(PO4)3/C 123.8 mA h/g - 98.8% 500 1 M LiPF6 [83]
fibers

Carbon fibers C/MnO2 NWs 710 mA h/g 0.2 A/g 90% 300 1 M LiPF6 [93]

Porous carbon-fiber 1M LiCF3SO3,


N/O 650 mA h/g 0.5 A/g - 100 [94]
monolith 0.3 M LiNO3

Ordered 1 M lithium bis

mesoporous carbon S 690 mA h/g 0.3 A/g - 300 imide, 0.1 M [95]

fiber LiNO3

1 M lithium bis
Porous bamboo
- 605.7 mA h/g - - 300 imide, 0.1 M [96]
carbon fibers
LiNO3

1 M lithium bis
Hierarchical carbon
S 845 mA h/g 0.25 A/g 77% 100 imide, 0.1 M [97]
fiber
LiNO3

1 M lithium bis
Hierarchical porous
S 500mA h/g - 70% 100 imide, 0.1 M [98]
carbon fibers
LiNO3

2.5 Carbon fiber/carbon material

The evolution of microelectronic devices or good-sized electric equipment requires

Lithium-ion batteries because of their superb electrical storage properties, low cost,

and most notably its stability [99]. Therefore, we seek a clever way to produce the

anodes of Lithium-ion batteries by converting dumped bamboo chopsticks into carbon

fibers. Figure 5a displays all evolution steps. Dumped bamboo chopsticks can be

30
recycled and can be briefly handled with a manageable hydrothermal reaction reacted

in alkaline solutions. In this way, the luxuriant crude fibers in the bamboo can be

parted voluntarily. Through carbonization, we can obtain a special type of carbon

fiber that has better anodic properties than voluminous bamboo carbons [100]. The

produced carbon fibers demonstrate a stronger competitiveness than merchant

graphite. SEM and TEM observations Figure 5b shows that the as-synthesized carbon

fibers are evenly distributed and exhibit a disheveled character and a large aspect ratio

(the average radius of one fiber surveyed does not exceeding 3 mm when its

maximum length actually attains several centimeters). When keeping the current

constant, we conduct charge–discharge tests with a rate of 0.37 C for 800 cycles in

Figure 5e. The long-term periodic character of carbon fibers is estimated. The

relevant information regarding the relationship between the reversible capacity and

the current density is generalized in Figure 5d. Through the research, we found that

this produced carbon fiber demonstrates excellent rate capability and capacity

conservation and provides a form of anodic substitution in Lithium-ion batteries.

Figure 5c demonstrates a typical SEM image of C/MnO2 NW/carbon fiber hybrid

products.

On the one hand, nanostructured MnO2 has a large surface-to-volume ratio to

ensure a high specific capacity. On the other hand, the internal carbon fiber

framework serves as the sturdy holder and the medium for electronic delivery. At the

same time, the carbon tiers lying on the surface of the MnO2 Nanowires (MnO2 NWs)

make great mechanical safeguards. Therefore, the properties of carbon fibers can be

31
further optimized by making nanostructured metallic oxide on the carbon fiber frame

to produce a synergetic core-shell electrode structure [101]. An excellent invertible

capacity of 710 mA h g-1 is maintained without attenuation for more than 300 cycles.

This approach shows a flexible path to utilizing bamboo chopsticks for producing

carbon fibers and provides an effective way to achieve alternative anodes for

Lithium-ion batteries and substantial multifunctional carbon-based hybrids that are

useful for electrical equipment [102].

Figure 5 The application of carbon fiber/carbon material in Lithium-ion battery


(a) Schematic diagram displaying the overall evolution of bamboo chopsticks into
uniform carbon fibers. (b) Optical and SEM images. (c) SEM observation of the
C/MnO2 NW/carbon fiber hybrid. (d) Specific capacity vs. the current density plot of
carbon fibers (green) and bulky bamboo carbons (black). (e) Long-term cyclic
performance and charge–discharge profiles. (a-e) Reprinted with permission from Ref.
[93]. Copyright 2014, The Royal Society of Chemistry.

3. Lithium-sulfur battery

32
In the Lithium-sulfur battery, sulfur is positive electrode and lithium is negative

electrode. When discharging, the negative electrode reacts with lithium to lose

electrons to become lithium ion, and the positive electrode reacts with sulfur to form

sulfide with lithium ion and electrons. The potential difference between the positive

electrode and the negative electrode is the discharge voltage provided by the lithium

sulfur battery. Under the action of applied voltage, the positive and negative reactions

of lithium sulfur battery are reversed, which is the charging process. The theoretical

discharge mass specific capacity of sulfur can be calculated as 1675 m Ah g-1

according to the electric quantity provided by the complete conversion of elemental

sulfur to S2-, and the theoretical discharge mass specific capacity of elemental lithium

can be calculated as 3860 m Ah g-1.The theoretical discharge voltage of lithium sulfur

battery is 2.287v, when sulfur and lithium completely react to form lithium

sulfide.The theoretical discharge mass specific energy of the corresponding lithium

sulfur battery is 2600 Wh kg-1.

Because of the fast decrease in fossil fuels and nonrenewable resources, such as

coal, oil and natural gas, present society needs efficient energy storage equipment by

using plentiful and cheap materials including sulfur-based and carbon-based materials

[96]. S is identified as an attractive large capacity cathode for energy storage batteries

because of its great theoretical capacity and high theoretical energy density [103-105].

Because of these merits, Li-S batteries are one of the most competitive candidates for

future power storage equipment [106]. And Table 1 also summarizes the

representative structure of carbon fiber in Lithium-sulfur battery.

33
3.1 Pure carbon fiber

Crude bamboo, as a sustainable pioneer, can produce poriferous bamboo carbon fibers

(BCFs) that can form into a BCF membrane (BCFM) as a captor interlining for the

Li2Sx intermediates between the sulfur cathode and the separator in Lithium-sulfur

batteries. As demonstrated in Figure 6a, this BCFM interlining is able to increase the

cycling property and electrochemical properties [107]. Moreover, the BCFs not only

provide an efficient conductive platform but also allow rapid mass exchange of the

electrolyte and adapt to serious volume variations in the sulfur cathode when

conducting charging/discharging. In addition, luxuriant polyporous construction of

BCFs offers vast absorbability to decrease the generation of the Li2S2/Li2S tier on the

surface of the electrode and lengthen their working time by effectively limiting sulfur

into these carbon fibers. Therefore, Lithium-sulfur batteries using the BCFM show

great electrochemical properties with a great efficiency up to 98%, poor capacity

damping, and long-playing cycle ability. The clean, cheap BCFM is able to offer a

competitive alternative to the massive commercialization of Lithium-sulfur batteries.

The micromorphology of bamboo fibers and bamboo carbon can be seen by the SEM

result in Figure 6c. The result reveals that the bamboo carbon keeps its fiber form

after the bamboo fibers are processed hardened [105]. Interestingly, the thin BCFM

interlining sample obviously has a better cycling property and rate capability than the

thick interlining sample in Figure 6b. On this occasion, the thickness of the

interlining is very significant to the mass transportation of the electrolyte and ions as

34
well as the electronic conductivity [108,109]. Distinctly, a thinner tier accelerates the

efficiency of the electrolyte and the transportation of the electrolyte and ions [110].

The result can be explained by the steady and comparatively coincident discrepancy

between the thin interlining sample and the thick sample in terms of the rate capability

[111].

A new type of carbon fiber has been produced by electrospinning and employs

resole as the source of carbon and triblock copolymer Pluronic F127 as the form

board. Later, sulfur is covered in well-organized mesoporous carbon fibers using easy

tempering [112]. The cross-linked fibrous nanostructure exhibitsan advantageously

mechanical steady and can offer a great conductive structure for sulfides during

cycling. The ordered mesopores are equipped to control the spread of long-chain

polysulfides (PS). The experimentation of the expected mesoporous structure of the

sample is shown in Figure 6d. The as-prepared ordered mesoporous carbon fiber

sulfur (OMCF-S) mixes with 63% S shows great reversal invertible capacity, high

capacity reserve and increased rate capacity when serving as the cathode in

energy-storing Lithium-sulfur batteries [113]. After measuring the capacity, we found

that even when undergoing a high cycle test, the sample electrode retains a stabilized

discharge capacity. To further observe the performance of the sample, we collected

SEM images of the special sample fibers before high-temperature processing, as

shown in Figure 6f. Apparently, the observed sample has a large fibrous morphology

with a smooth appearance and a very low mean fiber diameter. The voltage profiles of

the OMCF-S electrodes at different cycles are demonstrated in Figure 6e.

35
Figure 6 The application of pure carbon fiber material in Lithium-sulfur battery
(a) A schematic configuration of the Li–S cell with a BCFM interlayer. (b) The rate
capability of the cells with and without the interlayer. (c) SEM image of the BCFs. (d)
A schematic illustration of the experimental procedure and the formation of the ideal
mesoporous structure of the OMCF. (A color version of this figure can be viewed
online.) (e) Voltage profiles of OMCF-S. (f) SEM image of the electrospun
resol/F127/TEOS/PVB fibers. (g) Illustration of the synthesis of a carbon monolith.
(h) Rate capability of the Li–S batteries with and without a carbon monolith interlayer.
(i) SEM image of the membrane. (a-c) Reprinted with permission from Ref. [96].
Copyright 2015, The Royal Society of Chemistry. (d-f) Reprinted with permission
from Ref. [95]. Copyright 2015, Elsevier Ltd. (g-i) Reprinted with permission from
Ref. [94]. Copyright 2014, The Royal Society of Chemistry.

In our impression, carbon nanotubes, microporous carbon and activated carbon

are usually employed, but a nonconducting adhesive needs to be demanded to gather

these components into a single-layer film [105,114,115]. Graphene is a great

component because the relevant product can be simply manufactured based on π-π

and hydrophobic reciprocities [116]. Nonetheless, the fabrication of reduced graphene

oxide cannot meet the purpose of being environmentally friendly and takes

36
considerable time. For this reason, a new approach can be developed to attain the

efficient synthesis of porous monolithic carbon. At present, a comprehensive strategy

using renewable resources has always been the focus of people's attention [117-120].

Microtherm hydrothermal synthesis is the most promising method for changing

biomass into a highly porous uniwafer carbon [121-123]. Nonetheless, in an

experienced hydrothermal reaction, a greater extent of pyrolysis is needed to increase

the conductivity of carbon for the electrode [124]. Currently, projecting and

producing uniwafer carbon materials becaues of the integration of different aspects

has attracted considerable attention. For example, cellulose monoliths, as great

carbonization raw materials, can be effectively fabricated by the metabolism of

bacteria [125]. Apart from this method, microorganisms themselves are sustainable

carbonization raw materials and can reproduce by using biomass as nutrients

[126,127]. Nonetheless, based on what we know, only a minority of samples have

favorably produced porous monolithic carbon depending on the microorganisms.

Therefore, we found a novel approach that can synthesize this type of carbon fiber by

producing filamentous fungi as the raw material. The method is described as follows.

The fungi were preproccessed in solution, this preprocessing continues for three days

with biomass as a nutrient. Notably, the concentration of the sample reached

approximately 11 mg mL-1. Because of the reasonable control of fungi filtration and

drying, fungi membrane or aerogel was achieved, as shown in Figure 6g. SEM is a

great approach for characterizing the product. As shown in Figure 6i, the carbon

monolith (membrane and aerogel) is composed of even and mutually connected

37
one-dimensional (1D) structures. Pyrolysis occurred under an inert atmospheric

condition, the whole carbon fiber was produced, and its electric conductivity was over

1 S cm-1. The rate capability of the Li–S batteries is revealed in Figure 6h. The results

indicate that the Lithium-sulfur batteries without embedded carbon can only be cycled

at a rate of 0.5 C, when maintaining a capacity of ~230 mA h g-1. When the rate is 1 C,

the capacity faded to approximately 0 mA h g-1. However, if the carbon interlining is

inserted, then the capacity can increase to ~620 mA h g-1, even if the cycle is 2 C.

Therefore, we found that the carbon fiber produced at 800 °C was composited by

N (~2.4 at%) and O (~1.3 at%) and showed a BET surface area of ~305 and ~20 m2

g-1, respectively. From this type of carbon material, we observe different pore sizes,

such as mesopores and macropores. The carbon-fiber monolith demonstrates a

competitive ability to develop the cycle and capacity of Lithium-sulfur batteries and

has the potential to serve as a multifunctional electrode in energy storage equipment.

3.2 Carbon fiber/sulfur material

A new layered structural C/S composite is found because of carbon fiber matrices.

This composite is prepared using electrospinning. The carbon fibers are composed of

air-corn graphited carbon globes that employ catalytic Ni nanoparticles as firm

stencils. Sulfur material is fixed to the carbon substrates by heat evaporation [128].

The layered carbon fiber/S composite and components are characterized using

different approaches. The electrochemical property is estimated by cyclic

voltammetry and galvanostatic charge–discharge. Previous reports show an initial

38
discharging capacity of 845 mA h g-1 at 0.25 C. At the same time, the retained rate

reached 77% after100 cycles, which is also a considerable result. The discharging

capacity is able to reach 533 mA h g-1 at 1.0 C. The high cycle property and rate

capability support the even dispersal of sulfur materials and the high electrical

conductivity of the carbon fibers and air-corn graphitized carbon globes. Figure 7d

and Figure 7e show SEM images of various fibers used in this experiment. This

diagram helps us to understand their microscopic morphology. The PAN/Ni precursor

fibers shown in Figure 7d have a mean radius of 75 nm and exhibit different sharp

curves with pure PAN fibers (Figure 7d, illustration) because of the existence of Ni

salt. After a period of stability and then carbonization treatment, many granules in the

size range of 100 to 250 nm are seen lying on these carbon fibers (Figure 7e).

According to speculation, these granules are likely to be elemental Ni. The charging

and discharging performance of the HCF/S composite at a rate of 0.1 C is

demonstrated in Figure 7b. As shown in the graph, a charge potential terrace at a

voltage of 2.35 V and two discharging potential terraces at a voltage of 2.3 V and 2.1

V are observed in agreement with the report. The nonreversible capacity is under 80

mA h g-1, signifying the remission of the shuttle conditions. The first discharge

capacity reached 978 mA h g-1. When over 30 cycles, the cathode also has a 783 mA

h g-1 discharging capacity. To understand the rate capability, the initial charging and

discharging states of the HCF/S composite cathode as well as the property of the cycle

at different rates is indicated in Figure 7c. The discharging capacity and potential

platforms decline as the rate increases. Nonetheless, a first discharging capacity of

39
533 mA h g-1 is also available, while the rate reaches 1.0 C. When conducting the

different rates of circulation, the battery shows an invertible capacity of 438 mA h g-1

at a rate of 2.0 C and increases to 620 mA h g-1 at a rate of 0.2 C. The high rate

property manifests that the HCF/S cathode has both great conductivity between the

electrons and ions and the stabilized architecture [129].

Figure 7 The application of carbon fiber/sulfur material in Lithium-sulfur


battery
(a) Schematic illustration of the synthetic route of the HCF/S composite. (A color
version of this figure can be viewed online.) (b) The potential profiles at 0.1 C. (c)
The rate capability. (d) SEM image of PAN/Ni precursor fibers. (e) Pure PAN fibers
(a inset), HCF/Ni. (f) Scheme of S/CHPCF preparation. (g) Schematic illustration of
the visual–electrochemical study. and Schematic of a typical voltage profile. (h)
Electrochemical properties of the S/HKC and S/CHPCF electrodes. (i) S/HKC and (d)
S/CHPCF at various C-rates from 0.5C to 15C. (j) Scanning electron microscopy
(SEM) image. (k) TEM image of CHPCF at a different magnification.(a-e) Reprinted
with permission from Ref. [97]. Copyright 2015, Elsevier Ltd. (f-k) Reprinted with
permission from Ref. [98]. Copyright 2016, The Royal Society of Chemistry.

40
One-dimensional (1D) directed cross-linking hierarchical porous carbon fibers

(CHPCFs) were projected and planned as the sulfur immobilizer using Lithium-sulfur

batteries [130]. The entire manufacturing process is shown in Figure 7f. The CHPCFs

show a large length/diameter rate, an interwoven construction and a rational

hierarchic poriferous state, which offers “great aisles” for electron and Lithium-ion

transfer. In addition, the CHPCFs possess large poriferous surfaces to limit the PS

spread. Consequently, the S/CHPCF composite cathodes concurrently obtain high

C-rate properties and a steady cycling of 535 mA h g-1 at 15 C and a 0.076% capacity

damping for one cycle at 5 C and 500 cycles. The architecture and property relation of

the carbon fibers and Lithium-sulfur batteries need to be researched in particular.

Therefore, research studies are needed to reveal Lithium-sulfur batteries with

excellent C-rate property. To further illustrate the high cycling property of S/CHPCFs,

an electrochemical experiment was conducted to determine the relationship between

PS and the host by surveying the color variation in the electrolyte in clear containers

under relevant discharge conditions (Figure 7g). The C-rate properties of both the

S/HKC and S/CHPCF electrodes are shown in Figure 7h and Figure 7i, respectively.

The phenomenon that the discharge capacities decreased but the rate of cycle

increased is due to a raised overpotential [131].

In the first material, Ni nanoparticles play an important role in the formation of

hierarchical structure. They catalyze the graphitization of carbon fibers at lower

temperature and promote the formation of nano-size hollow carbon spheres as hard

templates. For the second material, the crosslinking structure between the carbon

41
fibers enables fast (direct) electron transfer “through fiber joint”, instead of the

traditional “through point contact” between the contacted carbon units.

4. Other batteries

In addition to Lithium-ion battery and Lithium-sulfur mentioned above, we also

summarize the application of carbon fiber materials in the other batteries. The

research results of carbon fiber and their derivatives for other batteries in the past few

years were shown in Table 2.

4.1 Sodium-ion battery

Carbon fiber is an excellent electrode material and has been widely used. Therefore,

the sources of carbon fiber are cheap and green, which has drawn considerable

attention with regard to the electrode material. MoS2/cotton-derived carbon fibers

(MoS2/CDCFs) were produced by a hydrothermal method and were later carbonized.

The production procedure of this composite allowed the MoS2 nanosheets to grow on

the surface of the carbon fibers. The composite provides rapid routes to shift electrons

and simultaneously serves as a firm substrate to decrease the volume transformation

of the MoS2 nanosheets when conducting circulation. The architecture and

characterization of different samples were conducted using SEM and TEM, as shown

in Figure 8b and Figure 8c, respectively. The results indicate that Na batteries have

high rate properties and increased cycling stability because of the superiority of this

composite. The MoS2/CDCFs-1 demonstrates a first invertible capacity of 504.9 mA

42
h g-1 at 0.1 A g-1. At the same time, the sample maintains 444.5 mA h g-1 after 50

cycles. Although the current density increases continuously, MoS2/CDCFs-1 still

maintains a considerable value, which is greater than that of the pure MoS2

nanosheets. Therefore, we further learned about the importance of carbon fibers. The

increased electrochemical property indicates the excellent method of using carbon

fibers to make layered composites for Na batteries.

The development is closely connected to the increased stability in the

performance because CDCFs have great electrical conductivity. As shown in Figure

8a, the CDCFs, where MoS2 nanosheets exist, provide a rapid method for the transfer

of electrons and ions, which greatly help electrode reactions. In addition, the CDCFs

prevent the stacking of MoS2. The holes are able to decrease the volume variations in

the MoS2 nanoslices when the repetitive sodium storage cycle can guarantee steady

circulation.

The electrochemical properties of this composite were assessed as anode

materials by using the half batteries. The images of as-prepared samples were

measured when the rate of scanning was 0.2 mV s-1 during the initial five cycles in

Figure 8e. In the initial scan, the peak represents a reduction reaction at

approximately 0.72 V because the sodium inserts into MoS2 along with a phase

change, that is, MoS2 converts into NaxMoS2 [132]. The peak lying 0.6 V is related to

the production of a solid electrolyte interface arising from pole falling into a decline.

The later reduction peak at 0.25 V is consistent with the transformation reaction of

NaxMoS2 and sodium, generating Na2S and Mo [133-137]. The peak at the site of 1.8

43
V indicates the variation in Na2S and a part of Mo to reproduce MoS2. At the same

time, the peak at the site of 2.2 V is due to the disintegration of remaining Na2S into S

through the scan of the anode [138]. Through the initial cycling of CV, MoS2 mixed

with Mo and S consists of the electrode instead of the original pure MoS2.[90] Hence,

in the next cycles, the cathodic peaks move in comparison to the peaks in the initial

cycle. In addition, we can deduce other electrode reactions [139]. The CV curves

remain largely unchanged, showing great stability of the sample when cycling. Figure

8d shows the rate properties of various samples. We can find from this diagram, in

spite of the fact that the CDCFs demonstrate capacities less than 150 mA h g-1,

MoS2/CDCFs-1 and MoS2 NSs have similar capacities in the inchoate cycles. The

reason is that the proportion of CDCFs existing in this sample is not high, and the

CDCFs with great conductivity favor a charge shift.

Figure 8 The application of carbon fiber materials in Sodium-ion battery

44
(a) Schematic illustration of Na+ storage in the MoS2/CDCFs. (b) and (c) SEM
images of MoS2/CDCFs-1. (d) CV profiles of MoS2/CDCFs-1 at a scan rate of 0.2
mV s-1 during the initial five cycles. (e) Rate performance of MoS2/CDCFs-1, the
MoS2 NSs and the CDCFs. (f) Schematic illustration of preparing the CFC/S
composite. (g) Cross-sectional SEM image of a single carbon fiber in the CFC/S-2
composite and the corresponding elemetalmappings for C and S. (h) Rate
performance of Na-S batteries based on CFC/S-2 composite and conventional C/S
composite. (i) First discharge and charge curves of Na-S batteries based on CFC/S
composites and conventional C/S composite at 0.05 C. (a-e) Reprinted with
permission from Ref. [140]. Copyright 2017, Elsevier B.V. (f-i) Reprinted with
permission from Ref. [141]. Copyright 2017, Elsevier B.V.

Moreover, another carbon fiber material can be used in Na batteries that also

have great electrochemical performance. The steps of this carbon fiber synthesis are

shown in Figure 8f. These composites were produced by covering S in the carbon

fiber cloth (CFC) that is achieved from cloth that has experienced carbonization.

Through observation, the surface of the CFC/S is slippery and without massive sulfur;

however, the even distribution of S is shown in this carbon fiber. The elementary

mappings of the composites further illustrate that the elements of carbon and sulfur

were symmetrically distributed upon these carbon fibers, as shown in Figure 8g. This

CFC/S material is a superb cathode for a normal-temperature Na-S battery because of

its all-around interlocked network and great electrolyte absorbance. In addition,

because the independent CFC/S material shows outstanding elasticity and

electroconductibility, it can be used as the electrode materials of sodium-ion batteries.

To prove the feasibility of this theory, sodium-ion batteries were produced using these

composites. Surprisingly, the properties of the batteries remain steady even when the

batteries are bent.

45
Besides, due to the great flexibility and electroconductibility, these materials are

equipped to immediately serve as no-binder and no current-collector electrodes for the

sodium-sulfur battery. Then, we further studied the performance of the as-prepared

battery. The initial discharge-charge curve of CFC/S at the rate of 0.05 C is

demonstrated in Figure 8i. In this image, we can also use the reversible hydrogen

electrode basis of the site of peaks to deduce the reactions. For example, the peaks at

approximately 2.2 and 1.6 V are observed, representing S changing into Na2Sx during

the process of discharging. Moreover, the peaks at approximately 1.9 and 2.4 V

represent short-chain sodium PS translated into long-chain sodium PS as well as S.

The CFC/S-1 sample shows a first discharging capacity of 549 mA h g-1 at 0.05 C and

attains an invertible specific charging capacity of 511 mA h g-1 during the initial

charge. When the addition of S reaches 3 mg cm-2, the peak of this curve is not very

apparent, illustrating that this battery experiences a large polarization. Comparing the

three samples, we find that CFC/S-1 has a low area capacity, whereas CFC/S-3

exhibits a larger polarization. Therefore, the CFC/S-2 composite is eclectic. In

contrast with the CFC/S-2 composite, the traditional C/S material with equal radio of

S should be produced by covering the super P and sulfur nanoparticles and then

placing them on the surface of the Al slice. The result of the comparison is that the

novel material have a better discharge capacity than the traditional C/S material. This

result can be verified in Figure 8h.

Recently, SnO2 has garnered much attention because of its high abundance,

large theoretical capacity and nontoxicity. However, the large volume expansion and

46
low intrinsic conductivity greatly plague its practical applications. The enhanced

electrochemical performance can usually be achieved by nanosizing SnO2 and

binding with the conductive matrix. However, despite these strategies that can

effectively suppress the volume expansion and accelerate the ionic and electronic

transportation to some extent, they cannot retain the long-term stability and rate

capability. There is a robust strategy for crafting oxygen vacancy-containing SnO2-x

nanoparticle-encapsulated carbon nanofibers with outstanding electrochemical

performance for advanced Sodium-ion battery. But compared with using carbon fiber,

this method is more complex and difficult to operate. [142]

4.2 Vanadium redox flow battery

The vanadium redox flow battery is one of the most prospective power storage

devices and has many advantages, such as safety, abundant energy capacity, flexible

design and an extended lifetime. The characteristics of vanadium redox flow batteries

reduce the risk of power outages by saving enough power during off-peak hours and

then using it during peak hours. A representative vanadium redox flow battery system

is composed of several batteries, pumps and vessels for electrolytes. An independent

cell is composed of a counter electrode, flow frames and a diaphragm. The electrode

is usually produced by great conductive carbon felt, which offers a place for the

reaction and the paths of the electrons. Carbon fiber-based fabrics, including felts,

cloths or papers, are all great poriferous electrode materials.

47
A carbon fiber/polyethylene (BP) composite (BPCFE) can be researched to

reduce the touch resistance that exists in the BP and CFE (carbon felt electrode). The

synthesis steps of this sample are as follows. The main raw materials of the sample are

plain-weave carbon fabric and PE powder. In addition, the mixture of conductive

particles and the PE powder can decrease the large resistance of the BPs. Because of

the dominant size and price of material, carbon black is elected as the conductive

particle. First, we need to achieve a mixture of the carbon fabric and PE

powder/carbon black. Then, the mixture is washed and dried and evenly dispersed.

Second, the compound is precisely acted on the as-prepared carbon fiber by a filter to

ensure unification. Third, the unwound sample should be cured. A closed die

containing an admixture of carbon fiber, PE powder and carbon black is placed into a

thermocompressor to give the mixture high pressure and a quantity of heat. At the

same time, the hot press is kept at 160 °C for 30 minutes so that the carbon fiber can

be entirely covered with PE powder. Then, the temperature of this thermocompressor

is decreased to normal atmospheric temperature, which lasts for two hours in a of

MPa environment. Finally, the CFEs are placed on the surface of the as-prepared

carbon/PE composite B and then sited into a thermocompressor to mix with the CFEs.

This bonding method is referred to as part thermoplastic welding. The concepts of this

method are to raise the surface temperature of the composite material. In this process,

the CFE fibers can thread the surface resin-rich area in the clear to directly touch the

BP fibers; at the same time, the mechanical performance of the BP is reserved.

Therefore, we undergo the experiments in a 1.2 Mpa environment to test its

48
mechanical performance. The compression tests were also carried out, and the quality

variation in the CFE was measured in an environment of the compaction pressure.

The result is that mass reduction occurs because of the fiber breakage of the CFE

while the compaction pressure surpasses 1.2 MPa. Hence, the pressure for local

heating is confirmed to be 1.2 Mpa that is a maximal pressure to ensure not destroying

the CFE during fabrication.

As is known, under the circumstances of an equivalent circuit, the magnitude of

the resistance is impacted by the poorest resistance element to a great extent. Hence,

the bypass electric conduction offered by these directly touching fibers obviously

reduces the nominal resistance of the BP-CFE module compared to the interfacing

touch resistance about BP and CFE. In addition, because of those influences, the

BP-CFE module exhibited relatively little sensitivity towards pressure so that the

VRFB can achieve steady state.

Furthermore, a gas permeability experiment shows that the material has poor air

permeability that is nearly zero if the time of local heating is just right or under 6 s.

Nonetheless, the BP-CFE module is produced under the same heating condition of

over 6 s, which had great permeability that illustrates the degeneration of the PE

matrix of the BP. Hence, the carbon/polyethylene composite material BP-CFE

component has good durability under the circumstances of vanadium sulfuric acid

[143,144]. By means of a sole battery charging and discharging test, the exploited

BP-CFE module has a slightly higher power efficiency than that of the traditional BP.

This finding is because compared with traditional graphite BP, the battery resistance

49
decreases due to a reduction in the total resistance, thus reducing the ohmic loss. At

the same time, it was discovered that the exploited BP-CFE components did not

degrade after 100 cycles of testing. Hence, we have developed this new type of

assembly material that is appropriate for vanadium redox flow battery application and

is also able to displace the traditional BP [145].

4.3 Zinc-air battery

Because of the rapid development of portable electronic devices, the need for soft

high-performance power storage equipment has become increasingly urgent. To date,

large efforts have been made to improve soft storage [146]. Nonetheless, the

performance characteristics, such as the low energy density and short cycle life, of

Lithium-ion batteries are not very great [147,148]. When metal batteries were

considered future candidates for storage equipment that possesses outstanding power

capacities [149], researchers have begun to study them vigorously. In particular, the

Zinc-air battery is of wide concern because of its great energy density, low cost,

safety and reliability [150]. However, at present, most air cathodes that are used in

Zinc-air batteries are bulky and can hardly satisfy the special requirements of portable

Zinc-air batteries. In addition, it is still a great problem to find efficient bifunctional

electrocatalysts for electrode reactions using in portable chargeable Zinc-air batteries

[151], despite significant developments in conventional Zinc-air batteries. An

excellent catalyst is needed for chargeable Zinc-air batteries, where the kinetics is

primarily controlled by the main discharge and the charge reactions that appear on the

50
cathode [152,153]. For conventional metal chargeable batteries, platinum and

ruthenium are typically used as the element of the catalyst. However, due to large

costs, rareness, and short lifetime, the commercialization of chargeable Zinc-air

batteries has been suffocated [154,155]. Recently, many studies have shown that

materials without metals and other heteroatoms are one of the most competitive

catalysts for rechargeable metal-air batteries; of course, Zinc-air batteries are included

[156]. Because of adequate carbon resources and superb electrochemical performance

as well as handling flexibility [157], carbon-based catalysts can serve as satisfactory

cheap, nimble and very active bifunctional catalysts for flexible Zinc-air batteries.

Nonetheless, to our knowledge, the possibility of heteroatom-doped carbon

nanomaterials as super bifunctional catalysts in flexible Zinc-air batteries has not been

realized. Maintaining the flexibility of the air cathode and good bifunctional

electrocatalytic activity is a challenge when building flexible Zinc-air batteries. After

several investigations, researchers have realized that newly produced nanometer

porous carbon nanofiber films (NCNFs) with large superficial areas are agile, while

they serve as cathodes in Zinc-air batteries working in surrounding air, have large

voltages, high power densities and large energy densities. This relevant chargeable

fluid Zinc-air battery shows little charging/discharging voltage difference and large

invertibility as well as outstanding stability. Moreover, flexible all-solid-state

chargeable Zinc-air batteries and air cathodes reveal great mechanical and cycling

stability under the condition of a small overpotential and a long cycling life after

repetitive bending. For example, if five of the newly produced flexible Zinc-air

51
batteries relying on this NCNF air cathode are developed, then they can serve as

powerful commercial light-emitting diodes that can be conveniently carried as flexible

energy resources. Therefore, we should undergo further research.

The NCNFs were produced by the hyperthermal pyrolysis of electrospun

polyimide (PI) films in an argon atmosphere, as shown in Figure 9a. Figure 9b shows

the polarization and energy density curves of these fluid elementary Zinc-air batteries

using these as-prepared films. The open-circuit voltage (OCV), maximal energy

density of the Zinc-air battery that uses the NCNF-1000 catalyst was found and

reached values that surpass these kinds of samples on the basis of the Pt/C composites.

As is clearly presented in Figure 9c, the catalyst offers the LSV curves that are tested

with a rotating disk electrode (RDE) at a 1600 rpm rate in a 0.1 mol/L potassium

hydroxide solution, with a similar onset potential for each sample. Additionally, a

larger ultimate current density was demonstrated for NCNF-1000 than that of other

samples. The consequences show that NCNF-1000 is the greatest electrocatalyst,

which is indeed superior to the Pt/C catalyst.

52
Figure 9 The application of carbon fiber materials in Zinc-air battery
(a) Schematic representation of the fabrication procedure towards the NCNF. (b)
Polarization and power density curves of the primary Zn-air batteries with different
catalysts. (c) LSV curves of different catalysts for ORR in O2-saturated at 1600 rpm.
(d) NCNF-1000. (e) Schematic illustration of the preparation of the Zn-Co-S
nanosheets, nanoparticles, and nanoneedles on CF. (f) OER polarization curves of
commercial RuO2, bare CFP, and Zn-Co-S NS, NP, NN/CFP at a scan rate of 5 mV
s-1 in 1.0 M KOH. (g) Discharge/recharge curves of rechargeable Zn-air batteries
catalyzed by Zn-Co-S NN/CFP and Pt/C + RuO2 at 10 mA cm-2 with a duration of
400 s per cycle. (a-d) Reprinted with permission from Ref. [158]. Copyright 2016,
Wiley-VCH. (e-g) Reprinted with permission from Ref. [159]. Copyright 2017,
American Chemical Society.

This result means a lasting challenge to exploit low cost, high-efficiency and

steady electrocatalysts to accelerate the slow electrode reactions that occur in the

53
chargeable metal batteries and fluid electrolyte system [160]. Through observation

and research, investigators have found a synthesis method that is easy to control. This

method obtains samples by growing a mixture of zinc, cobalt and sulfide

nanostructure on the carbon fiber paper CFP. At the same time, this method makes

full use of the electrochemical catalytic effect of Zn-air batteries and the electrolysis

of water from hydrogen and oxygen [161]. The synthesis of various forms of

complexes is due to the cooperative impact of the decomposing urea product as well

as NH4F corrosion. In the prepared complex, the nanoneedle arrays sustained on CFP

show an excellent trifunctional property for these redox reactions at the electrode

compared to its nanosheet and relevant counterpart by means of a half-reaction test.

At the same time, the catalytic endurance of this complex is better than that of Pt/C

and RuO2. In addition, this newly prepared electrode realizes a chargeable Zinc-air

battery with a poor overpotential, outstanding efficiency, extended cycling time and

extensive properties for water splitting. These properties give rise to the complex

nanoneedle/CFP nanoarchitecture that not only offers increased active areas but also

promotes the shift of ions and gases between the catalyzing surface and electrolyte,

thereby maintaining an effective interface among the solid, liquid and gas states for

electrocatalysis. The consequences show that the zinc-cobalt-S nanoneedle and CFP

assembly is a cheap, high efficiency and enduring rechargeable Zinc-air battery

electrode, which is suitable for the electrolysis of water in an alkaline solution [162].

Three different types of CFP-supported Zn-cobalt-S nanoarchitectures are

produced by a simple hydrothermal method to obtain the relevant carbonate

54
hydroxide precursor, which is then managed through thermal vulcanization, as shown

in Figure 9e. To understand the electrocatalytic performance of the catalyzing oxygen

evolution reaction (OER), this three-element-composition complex and its

counterparts, pure CFP and RuO2, are measured in a solution of potassium hydroxide

employing the three-electrode system. As described in Figure 9f, the sample has a

current density of 10 mA cm-2 at 320 mV. The performance of this sample is better

than that of the other two samples under different conditions with higher

overpotentials. However, the performance of this complex is similar to RuO2. In

particular, the first charging/discharging voltages of the as-prepared electrode are

~2.03 and 1.18 V at 10 mA cm-2. These values demonstrate a small overpotential and

a high cycling efficiency, resembling the Pt/C+RuO2 electrode, indicating the great

bifunctional catalytic ability of the ORR and the inverse OER. Interestingly, the

zinc-cobalt-S NN/CFP-catalyzed cells showed excellent rechargeability after 200

cycles (Figure 9g), in contrast with t/C+RuO2 (under 55 cycles), manifesting better

cycle durability. The reversibility of rechargeable Zinc-air batteries shown here can be

compared with the results of recently published studies.

4.4 Aluminium-air battery

Alkaline metal air batteries are considered a promising candidate with regard to

security and clean and massive power storage technologies. Additionally, these

batteries include the use of electrode catalysts without Pt, have smaller corrosive

electrolytes than acidic media, have a very large energy density and does not release

55
greenhouse gases [149,165]. Towards alkaline metal-air batteries, the active metal,

such as Li, Mg, Al, Fe and Zn, undergoes oxidizing reactions at the anode, giving off

an electron that transits between the outer circuit and the cathode. At the same time,

the oxygen attains these electrons, after which a reducing reaction occurs on the

surface of the cathode, referred to as an oxygen reduction reaction [166-169]. As the

largest proportion of metal in the earth's crust and an extremely low molecular weight,

aluminium is one of the best candidates for metal-air batteries in alkaline solution. In

fact, aluminium air batteries have a large theoretical specific energy capacity, second

only to lithium batteries, which can output a great current density [170]. In order to

further confirm the property of this prepared electrode in the surrounding of an actual

working battery, the Ag/CFP electrode directly served as the air cathode in the

primary aluminum air battery (Figure 10b). The cathode is split by a type of

macromolecule membrane with an Al anode. The use of hydrophobic CFP as a

collector is suitable for offering a well-developed solid-liquid-gas interface for the

oxygen reduction reaction. Obviously, the solution of 4 M NaOH is used in this cell

due to the smaller corrosion potential than the same concentration of the potassium

hydroxide solution to the aluminium alloy [171]. Nonetheless, aluminium air batteries

have become prosperous massive commercial power reserve equipment, but is still a

challenge due to the stagnant kinetics of the reaction of the cathode, which obviously

adds overpotential during the discharge process and results in a decline in the battery

properties. The important approach is to have steady and highly efficient air

electrodes with a low over-potential in the environment of large current density. It is

56
important to fabricate this type of electrode, including the development of high

efficiency, cheap and long lifetime electrocatalysts for oxygen reduction reactions

[172,173]. The widespread and employed valuable material of Pt and its composites

are considered the most effective catalysts for the oxygen reduction reaction;

nonetheless, they are expensive, have poor stability and can rapidly deactivate

[174,175]. Hence, wider and cheaper metal catalysts, including Ni, Cu, Co and Mn,

that are more stable in an alkaline solution have been used to replace this Pt-based

catalyst when developing the air electrode [176,177]. For example, a research and

development team has investigated an air-breathe gas spread electrode by employing

a nanoscale Mn catalyst for the application of a metal air battery. In various metal

catalysts, Ag catalysts have great promise to displace Pt material because of their

fairly high activity, great durability and inexpensive properties, which are 0.02 the

price of Pt. Although some researchers have explored the oxygen reduction reaction

activity of Ag electrocatalysts in an alkaline solution by means of density functional

theoretical calculations, there have only been a few reports on the preparation and

evaluation of silver-containing air electrodes that can be used in real air batteries. In

industrial applications, it is universally believed that the air electrode should not only

reduce the load on the catalyst but also increase the catalytic property and stability of

the catalyst under actual operational conditions. These obstacles can be overcome by

properly structured support materials [178]. Moreover, the supporting base structure

must promote the diffusion between the oxygen and electrolytes to the active site in

the air cathodes [179]. In recent years, triaxial carbon fiber materials, an appropriate

57
base to support electrocatalysts because of their three-dimensional poriferous

structure, great electronic conductivity and formidable stability, are widely used in

electrochemical devices for electrode substrate manufacturing. As shown in Figure

10c, CFP has a three-dimensional structure, including well-linked carbon fibers with

diameters of approximately 10 µm. Therefore, we can determine how the Ag/CFP

electrode is fabricated in Figure 10a. First, this commercial CFP base is initially

oxidized in a gentle manner by H2SO4/HNO3 at a high concentration to achieve vast

negative functional groups, such as -COO- on the surface. Later, the prepared surface

oxidizing CFP is immersed in an aqueous solution including a 2 mm Ag precursor.

Due to surface functionalization, the CFP and Ag precursors are self-assembled by

electrostatic interactions and are located via the in situ nucleation and growth of Ag

metal in combination with Ag+ [180]. Via the intensive selection of the electrode

positional potential, an Ag/CFP unbonded air electrode is immediately acquired, and

the granule size and distribution of Ag can be controlled. Then, the configuration and

morphology of Ag/CFP are measured and analyzed in particular. On the side of a

three-dimensional carbon felt, an air electrode based on CFP, because of its good

porosity and low density, offers an unimpeded gas diffusion channel and continuously

provides a sufficient oxygen reactant to the reaction catalytic site of this cathode,

which has made greater progress [181]. In preceding studies, CFPs combined with

electrocatalytic functional material appear to have enhanced oxygen reduction

reaction properties. The combination of the conductive carbon fibers and the

high-activity catalysts form strong three-dimensional structures, which is conducive to

58
the mass shift of the electrolyte to the electrode, as well as the electron to the active

center and later adsorbent oxygen molecule [182]. Nonetheless, the composite of most

air electrodes requires various stringent technologies, including the high degree of the

vacuum environment, complicated and inexpensive multistep reactions, the use of

vast, expensive and poisonous reagents to obtain the command of the morphology,

and the use of enduring polymer adhesives [151,183], hindering extensive commercial

use. Therefore, the combination of Ag-based catalysts with a CFP base is an efficient

tactic to improve the properties of the air electrode. More importantly, we should

exploit a simple, nonpolluting, and moderate method to synthesize no-binder air

electrodes. On this basis, we can research a cheap, ordinary and viable in situ

synthesis approach to build a highly integrated Ag catalyst layer and CFP that does

not need to use a dangerous polymer binder. Ag precursors self-assemble on the

surface of functional CFP (high concentration acid pretreatment) via electrostatic

interactions in both Ag+ and negative ions and then slowly are reduced to form Ag

particles. This newly researched no binder Ag and CFP air electrode shows superb

and stable electrochemical properties for the oxygen reduction reaction and singularly

excellent properties for primary aluminum air batteries. Moreover, the

electrochemical analysis test shows that the granule size and dispersity of the silver

catalyst on this matrix surface are very important to the activity of catalysis; at the

same time, the degree of coverage to the electroactive silver depends on the catalytic

property.

59
Figure 10 The application of carbon fiber materials in Aluminium-air battery
(a) Schematic diagram of the synthesis process of the Ag/CFP electrode. (b) A
schematic of the primary Al-air battery. (c) SEM image of the pristine CFP. Inset in
panel (c): Schematic illustrating the interpenetrating 3D network structure of CFP. (d)
Linear sweep voltammograms. (e) Polarization curve (red) and corresponding power
density plot (green) of the batters using Ag/CFP-3 (solid line) and QSI-Nano
electrode (dotted line) as the cathodes. (a-e) Reprinted with permission from Ref.
[164]. Copyright 2017, Te Author(s).

Table 2. Summary of the carbon fiber materials reported in the other battery section.
Cycle
Carbon fiber composite Reversible capacity Current density Electrolyte Ref.
number

Sodium-ion battery

cotton-derived carbon
MoS2 444.5 mA h/g 0.1 A/g 50 1 M NaClO4 [140]
fibers

cotton-derived carbon
MoS2 323.1 mA h/g 0.5 A/g 150 1 M NaClO4 [140]
fibers

1.5 M NaClO4 0.2 M


carbon fiber cloth S 390 mA h/g - 300 [141]
NaNO3

Vanadium redox flow

battery

carbon fiber polyethylene 320 mA h/g - 100 vanadium sulfuric [163]

Zinc-air battery

Nanoporous carbon 6 M KOH, 0.2 M


graphite 378 mA h/g 0.2 A/g 48 [158]
fiber films Zn(CH3COO)2

6.0 M KOH, 0.2 M


Carbon fiber paper Zn-Co-S NS 453 mA h/g 0.3 A/g 200 [159]
ZnCl2

6.0 M KOH, 0.2 M


Carbon fiber paper Zn-Co-S NP 467 mA h/g 0.3 A/g 200 [159]
ZnCl2

6.0 M KOH, 0.2 M


Carbon fiber paper Zn-Co-S NN 484.7 mA h/g 0.3 A/g 200 [159]
ZnCl2

60
Aluminium-air battery

Carbon fiber paper Ag 783.5 mA h/g 0.3 A/g 100 4.0 M NaOH [164]

The linear sweep voltammograms for the oxygen reduction reaction activity of

these composites are shown in Figure 10d. The initial CFP shows slight activity, but

the Ag/CFP-3 offers a -0.07 V start potential that is more positive than the other two

samples. The property of Ag/CFP-3 is enhanced, as demonstrated by the lower Tafel

slope. In addition, the more positive starting potential shows an easier oxygen

reduction reaction process started on that electrode [182]. Therefore, we have

determined that the greatest cathode is Ag/CFP-3 through this test. The efficiency of

electrocatalysis depends on the degree of electroactive silver covered on the surface of

these samples. The result shows that the more the coverage area of these silver

materials is, the faster the catalyzing oxygen reduction reaction rate of the identically

sized air cathode is. Therefore, the great oxygen reduction reaction catalyzing the

activity of Ag/CFP-3 shows that the granule size and dispersal of silver catalysts on

these carbon fibers are vital to improve the activity of the oxygen reduction reaction.

Figure 10e displays the discharge polarization curves of the sample, and the

results demonstrate that the Ag/CFP cathodes have greater properties than commercial

QSI-nanobased cathodes. The peak energy density of Ag/CFP-3 is 109.5 mW cm-2 at

0.98 V. However, the QSI-nanobased battery is 90.8 mW cm-2 at 0.81 V. Both the

discharge voltage and peak energy density of the as-prepared samples with a low

mass load exceeded the values of the batteries produced by QSI-nanobased cathodes

due to the existence of a forceful coupling between the transcendental

61
oxygen-reduction-reaction activity of these equipped silver particles and the

conductive CFP substrate [184]. If the discharge continues, then the aluminum sheet

becomes thinner little by little, and at the same time, the electrolytes have an

increasing content of soluble interaluminate.

4.5 Lithium-air battery

Lithium-air batteries are considered one of the most promising energy conversion

systems for delivering large specific energy. However, practical applications still face

major challenges, including poor rate capability, short cycle life, and low round-trip

efficiencies. Reactions of cathode are including gas, liquid and solid phase reactions

in Lithium-air batteries system. Thus, the requirement of the air electrode is extremely

high. Considerable efforts have been focused on increasing specific surface area of

electrodes, preparation of highly active catalyst and optimizing the structure of air

electrode.

In theory, Lithium-air batteries are more energy dense than Lithium-ion batteries

because they can replace bulky solid-state electrodes with a porous carbon electrode

that stores energy by capturing oxygen from the air floating over it, which combines

with lithium ions to form lithium oxide. The latest research is a step forward. The

carbon electrode has more pores than other carbon electrodes, so that when the battery

discharges, it has more pores to store solid lithium oxide. We can use the chemical

vapor deposition process to grow arrays of vertically aligned carbon nanofibers,

which act as blankets as energy storage 'scaffolds' with high electrical conductivity

62
and low density. During the discharge process, lithium peroxide particles appear on

the carbon fiber, and the carbon will increase the weight of the battery. Therefore, it is

very important to minimize the amount of carbon and leave enough space for lithium

peroxide, which is the active chemical substance formed in the discharge process of

lithium air and air power pool. Because the arrangement of the carbon fiber electrode

carbon particles is very orderly, and carbon particles in the other electrode is very

chaotic, therefore, are more likely to use scanning electron microscope to observe the

behavior of the electrode in the middle of the charging state, which would help them

to improve the efficiency of the battery, also helps explain why the existing system

after many times charge discharge cycle, performance will decline.

5. Conclusion and outlook

In summary, the review chiefly presents applications of carbon fibers and their

ramifications in terms of batteries. In addition, carbon fibers can also be used in

extensive electrochemical applications, including Lithium-ion, Lithium-sulfur, Sodion,

Znic-air, and Aluminium-Air batteries. The synthetic approaches, morphology and

electrochemical performance of these batteries have been specifically recommended.

Carbon fibers exhibited promising accomplishments during previous research

studies, and their characteristic superiorities, stable electrochemical performance,

excellent mechanical strength, high electrical conductivity, great electron transmission

and small variation of volume are included in this study. Moreover, carbon fibers

have acquired good invertibility, low charge-transfer impedance and highly stable

63
performance during charging and discharging. In addition, carbon fibers are also

ordinary, inexpensive and efficient, and extraordinarily promising for low cost

technical manufacturing. The carbon fiber can be applied as the anode of

multipurpose structural Lithium-ion batteries. Carbon fibers are more efficient and

bring about less reaction than other materials that are applied as the anode in

multipurpose structural batteries and in other areas wanting simultaneous

high-intensity and outstanding electrochemical performance. Therefore, the

combination of carbon fibers and batteries produce superior electrochemical

performance.

However, the performance of pure carbon fiber is still not good enough. Under

these circumstances, a growing number of investigators have been devoted to

overcoming difficulties, with the aim of acquiring further promising applied materials.

The idea of manufacturing one type of complex combining pure carbon fiber and

other materials appears to be the most prospective and effective choice not only to

eliminate shortcomings of the pure carbon fiber but also the advantages of other

materials could be brought into the compounds. Fortunately, these kinds of

compounds, particularly carbon fiber/metallic complex materials, have undergone

tremendous progress over the past several years. As has been reported in previous

studies, metallic materials and their relevant materials have exhibited outstanding

performance, and with a combination of these materials, many imperfections of pure

carbon fibers can be eliminated. For instance, the stability, selectivity, adsorption and

electroconductivity were extensive to a great extent, guaranteeing that the compound

64
demonstrates a modified property in many applications. As emphasized during the

report, synthesis approaches of carbon fiber-based materials compound have been

highly enhanced, offering potential for taking advantage of their best performance in

wider applications.

In view of the reported literature, there are a few difficulties to be addressed in

future developments:

(1) The conductivity of pure carbon fibers is not very high and can be enhanced by

uniting these materials with great electrical conductivity. Composite materials not

only have the merits of all components but also eliminate the deficiency of the

independent portions to further improve the battery performance.

(2) That is, it is essential to research the effects of the structure and performance of a

compound. In the meantime, changes in the structure often exhibit enhancements

in terms of performance. Therefore, further research still needs to be focused on

their relationship, which is also advantageous to better understand and employ

compound materials.

(3) Transition metals are increasingly used by people due to their excellent properties.

We find that we can effectively take advantage of the large superficial area of

transition metal-based carbon fibers to further improve the battery performance.

Nevertheless, some of these are not able to be effectively utilized because the

electrolyte cannot permeate into all small holes. This is also a problem that needs

to be urgently solved. The hole size is a key issue for transition-metal-based

carbon fibers.

65
(4) Conductive materials, for instance, nanoarchitectured carbon materials and

conductive superpolymers, which are able to immensely improve the

electroconductibility of whole composite materials, work out the inherent low

conductivity of batteries.

(5) Hierarchical hollow structures currently present better properties than nonhollow

hierarchical structures because this special construction is able to buffer the effect

of the variation in the volume during charging/discharging. Hence, it is a great

choice to hollow out the material to a create a proper electrode.

(6) Binder-free electrodes are a significant portion of Lithium-ion battery research. In

terms of traditional electrodes, synthetic processes currently use macromolecule

binders or other additives to paste active materials to electrodes. These binders

block the flow of electrons to a great extent, and the electrical capacity is greatly

reduced. On this point, we urgently need to find a material to reduce the use of

binders.

As a result, it is significant to employ the presented method to improve properties.

In particular, great power density means a large risk of safety, and it is also necessary

to satisfy all requirements regarding security and the development efficient,

continuable and green power sources. Despite many difficulties that need to be

overcome, composites of carbon fiber materials offer great prospects for the

expansion of applications of carbon fiber-based batteries. In the meantime, the

technology deserves further research, which will result in promising accomplishments

in the future.

66
Acknowledgements

This work was supported by the National Natural Science Foundation of China

(NSFC- 21671170, 21673203, and 21201010), the Top-notch Academic Programs

Project of Jiangsu Higher Education Institutions (TAPP), Program for New Century

Excellent Talents of the University in China (NCET-13-0645), Postgraduate Research

& Practice Innovation Program of Jiangsu Province (XKYCX17-038), the Six Talent

Plan (2015-XCL-030), and Qinglan Project. We also acknowledge the Priority

Academic Program Development of Jiangsu Higher Education Institutions.

67
Reference

[1] H. Tang, M. Zheng, Q. Hu, Y. Chi, B. Xu, S. Zhang, H. Xue, H. Pang, J. Mater.

Chem. A. 6 (2018) 13999–14024.

[2] N. Zhang, X. Xiao, H. Pang, Nanoscale Horizons. 4 (2019) 99–116.

[3] P. Wan, X. Wen, C. Sun, B. K. Chandran, H. Zhang, X. Sun, X. Chen, Small, 11

(2015), 5409–5415.

[4] W. Lu, X. Guo, Y. Luo, Q. Li, R. Zhu, H. Pang, Chem. Eng. J. 355 (2019)

208–237.

[5] T. Wang, Y. Guo, P. Wan, H. Zhang, X. Chen and X. Sun, Small, 12 (2016),

3748–3756.

[6] A. Manthiram, Y. Fu, S.-H. Chung, C. Zu, Y.-S. Su, Chem. Rev. 114 (2014)

11751–11787.

[7] C. Xing, G. Jing, X. Liang, M. Qiu, Z. Li, R. Cao, X. Li, D. Fan and H. Zhang,

Nanoscale, 9 (2017), 8096–8101.

[8] R. Xu, J. Lu, K. Amine, Adv. Energy Mater. 5 (2015) 1500408.

[9] L. Suo, Y.-S. Hu, H. Li, M. Armand, L. Chen, Nat. Commun. 4 (2013) 1481.

[10] N. Mahmood, C. Zhang, H. Yin, Y. Hou, J. Mater. Chem. A. 2 (2014) 15–32.

[11] X. Qian, X. Yang, L. Jin, D. Rao, S. Yao, X. Shen, K. Xiao, S. Qin, J. Xiang,

Mater. Res. Bull. 95 (2017) 402–408.

[12] W. Li, Q. Zhang, G. Zheng, Z. W. Seh, H. Yao, Y. Cui, Nano Lett. 13 (2013)

5534–5540.

68
[13] S.-H. Chung, P. Han, R. Singhal, V. Kalra, A. Manthiram, Adv. Energy Mater. 5

(2015) 1500738.

[14] Z. Huang, Z. Zhang, X. Qi, X. Ren, G. Xu, P. Wan, X. Sun and H. Zhang,

Nanoscale, 8 (2016), 13273–13279.

[15] W. Zhou, C. Wang, Q. Zhang, H. D. Abruña, Y. He, J. Wang, S. X. Mao, X.

Xiao, Adv. Energy Mater. 5 (2015) 1401752.

[16] Y. Qiu, W. Li, W. Zhao, G. Li, Y. Hou, M. Liu, L. Zhou, F. Ye, H. Li, Z. Wei, S.

Yang, W. Duan, Y. Ye, J. Guo, Y. Zhang, Nano Lett. 14 (2014) 4821–4827.

[17] S. Zhang, M. Zheng, Z. Lin, N. Li, Y. Liu, B. Zhao, H. Pang, J. Cao, P. He, Y.

Shi, J. Mater. Chem. A. 2 (2014) 15889–15896.

[18] S. Bai, C. Sun, H. Yan, X. Sun, H. Zhang, L. Luo, X. Lei, P. Wan and X. Chen,

Small, 11 (2015), 5807–5813.

[19] Y. Wu, M. Gao, X. Li, Y. Liu, H. Pan, J. Alloys Compd. (2014), 608, 220–228.

[20] S. Luo, J. Zhao, J. Zou, Z. He, C. Xu, F. Liu, Y. Huang, L. Dong, L. Wang and H.

Zhang, ACS Appl. Mater. Interfaces, 10 (2018), , 3538–3548.

[21] R. Zhu, J. Ding, Y. Xu, J. Yang, Q. Xu, H. Pang, Small. 14 (2018) 1803576.

[22] Z. Yang, J. Zhang, M. C. W. Kintner-Meyer, X. Lu, D. Choi, J. P. Lemmon, J.

Liu, Chem. Rev. 111 (2011) 3577–3613.

[23] V. Etacheri, R. Marom, R. Elazari, G. Salitra, D. Aurbach, Energy Environ. Sci.

4 (2011) 3243.

[24] L. Qie, W.-M. Chen, Z.-H. Wang, Q.-G. Shao, X. Li, L.-X. Yuan, X.-L. Hu,

W.-X. Zhang, Y.-H. Huang, Adv. Mater. 24 (2012) 2047–2050.

69
[25] Z.-J. Fan, J. Yan, T. Wei, G.-Q. Ning, L.-J. Zhi, J.-C. Liu, D.-X. Cao, G.-L.

Wang, F. Wei, ACS Nano. 5 (2011) 2787–2794.

[26] C. Kim, K. S. Yang, M. Kojima, K. Yoshida, Y. J. Kim, Y. A. Kim, M. Endo,

Adv. Funct. Mater. 16 (2006) 2393–2397.

[27] V. G. Pol, M. M. Thackeray, Energy Environ. Sci. 4 (2011) 1904–1912.

[28] B. Campbell, R. Ionescu, Z. Favors, C. S. Ozkan, M. Ozkan, Sci. Rep. 5 (2015)

14575.

[29] J. Tang, V. Etacheri, V. G. Pol, ACS Sustain. Chem. Eng. 4 (2016) 2624–2631.

[30] S. Ifuku, R. Nomura, M. Morimoto, H. Saimoto, Materials (Basel). 4 (2011)

1417–1425.

[31] V. G. Pol, J. Wen, K. C. Lau, S. Callear, D. T. Bowron, C.-K. Lin, S. A.

Deshmukh, S. Sankaranarayanan, L. A. Curtiss, W. I. F. David, D. J. Miller, M.

M. Thackeray, Carbon N. Y. 68 (2014) 104–111.

[32] J. Collins, G. Gourdin, M. Foster, D. Qu, Carbon N. Y. 92 (2015) 193–244.

[33] J. Liu, Z. A. Wang, X. W. Wu, X. H. Yuan, J. P. Hu, Q. M. Zhou, Z. H. Liu, Y.

P. Wu, J. Power Sources. 299 (2015) 301–308.

[34] H. Fujimoto, K. Tokumitsu, A. Mabuchi, N. Chinnasamy, T. Kasuh, J. Power

Sources. 195 (2010) 7452–7456.

[35] W. Xing, J. Electrochem. Soc. 144 (1997) 1195.

[36] J. Wu, X. Qin, C. Miao, Y. B. He, G. Liang, D. Zhou, M. Liu, C. Han, B. Li, F.

Kang, Carbon N. Y. 98 (2016) 582–591.

[37] M. N. Obrovac, V. L. Chevrier, Chem. Rev. 114 (2014) 11444–11502.

70
[38] I. Kovalenko, B. Zdyrko, A. Magasinski, B. Hertzberg, Z. Milicev, R. Burtovyy,

I. Luzinov, G. Yushin, Science (80-. ). 334 (2011) 75–79.

[39] H. Wu, Y. Cui, Nano Today. 7 (2012) 414–429.

[40] H. Wu, G. Chan, J. W. Choi, I. Ryu, Y. Yao, M. T. McDowell, S. W. Lee, A.

Jackson, Y. Yang, L. Hu, Y. Cui, Nat. Nanotechnol. 7 (2012) 310–315.

[41] Y. Han, P. Qi, X. Feng, S. Li, X. Fu, H. Li, Y. Chen, J. Zhou, X. Li, B. Wang,

ACS Appl. Mater. Interfaces. 7 (2015) 2178–2182.

[42] A. Magasinski, P. Dixon, B. Hertzberg, A. Kvit, J. Ayala, G. Yushin, Nat. Mater.

9 (2010) 353–358.

[43] N. Liu, Z. Lu, J. Zhao, M. T. McDowell, H.-W. Lee, W. Zhao, Y. Cui, Nat.

Nanotechnol. 9 (2014) 187–192.

[44] Y. Fu, A. Manthiram, Nano Energy. 2 (2013) 1107–1112.

[45] J. Wu, X. Qin, H. Zhang, Y.-B. He, B. Li, L. Ke, W. Lv, H. Du, Q.-H. Yang, F.

Kang, Carbon N. Y. 84 (2015) 434–443.

[46] H. Zhang, X. Qin, J. Wu, Y.-B. He, H. Du, B. Li, F. Kang, J. Mater. Chem. A. 3

(2015) 7112–7120.

[47] Y. Wang, M. Zhao, Q. Zhao, Q. Li, H. Pang, Nanoscale. 10 (2018)

15755–15762.

[48] Y. Xu, Q. Li, H. Xue, H. Pang, Coord. Chem. Rev. 376 (2018) 292–318.

[49] H. Kim, M. Seo, M.-H. Park, J. Cho, Angew. Chemie Int. Ed. 49 (2010)

2146–2149.

71
[50] Y. Xu, B. Li, S. Zheng, P. Wu, J. Zhan, H. Xue, Q. Xu, H. Pang, J. Mater. Chem.

A. 6 (2018) 22070–22076.

[51] Y. Zheng, S. Zheng, H. Xue, H. Pang, Adv. Funct. Mater. 28 (2018) 1804950.

[52] P. Wu, Y. Xu, J. Zhan, Y. Li, H. Xue, H. Pang, Small. 14 (2018) 1801479.

[53] A. Magasinski, P. Dixon, B. Hertzberg, A. Kvit, J. Ayala, G. Yushin, Nat. Mater.

9 (2010) 461–461.

[54] T. H. Hwang, Y. M. Lee, B.-S. Kong, J.-S. Seo, J. W. Choi, Nano Lett. 12 (2012)

802–807.

[55] X. Zhou, L.-J. Wan, Y.-G. Guo, Small. 9 (2013) 2684–2688.

[56] H. Wu, G. Zheng, N. Liu, T. J. Carney, Y. Yang, Y. Cui, Nano Lett. 12 (2012)

904–909.

[57] Y. Li, B. Guo, L. Ji, Z. Lin, G. Xu, Y. Liang, S. Zhang, O. Toprakci, Y. Hu, M.

Alcoutlabi, X. Zhang, Carbon N. Y. 51 (2013) 185–194.

[58] Z.-L. Xu, B. Zhang, J.-K. Kim, Nano Energy. 6 (2014) 27–35.

[59] B.-S. Lee, S.-B. Son, K.-M. Park, J.-H. Seo, S.-H. Lee, I.-S. Choi, K.-H. Oh,

W.-R. Yu, J. Power Sources. 206 (2012) 267–273.

[60] X. Zhou, Y.-G. Guo, J. Mater. Chem. A. 1 (2013) 9019.

[61] Y. Xu, Y. Zhu, F. Han, C. Luo, C. Wang, Adv. Energy Mater. 5 (2015) 1400753.

[62] Y. Chen, X. Li, K. Park, J. Song, J. Hong, L. Zhou, Y.-W. Mai, H. Huang, J. B.

Goodenough, J. Am. Chem. Soc. 135 (2013) 16280–16283.

[63] X. Li, S. Zheng, L. Jin, Y. Li, P. Geng, H. Xue, H. Pang, Q. Xu, Adv. Energy

Mater. 8 (2018) 1800716.

72
[64] B.-S. Lee, H.-S. Yang, H. Jung, S.-Y. Jeon, C. Jung, S.-W. Kim, J. Bae, C.-L.

Choong, J. Im, U.-I. Chung, J.-J. Park, W.-R. Yu, Nanoscale. 6 (2014) 5989.

[65] Y. Wang, X. Xiao, Q. Li, H. Pang, Small. 14 (2018) 1802193.

[66] H. Bin Wu, J. S. Chen, H. H. Hng, X. Wen (David) Lou, Nanoscale. 4 (2012)

2526.

[67] W. Wei, S. Yang, H. Zhou, I. Lieberwirth, X. Feng, K. Müllen, Adv. Mater. 25

(2013) 2909–2914.

[68] C. He, S. Wu, N. Zhao, C. Shi, E. Liu, J. Li, ACS Nano. 7 (2013) 4459–4469.

[69] Y. Yan, Y. Luo, J. Ma, B. Li, H. Xue, H. Pang, Small. 14 (2018) 1801815.

[70] S. Zheng, B. Li, Y. Tang, Q. Li, H. Xue, H. Pang, Nanoscale. 10 (2018)

13270–13276.

[71] X. Qin, H. Zhang, J. Wu, X. Chu, Y.-B. He, C. Han, C. Miao, S. Wang, B. Li, F.

Kang, Carbon N. Y. 87 (2015) 347–356.

[72] B. Liu, X. Wang, B. Liu, Q. Wang, D. Tan, W. Song, X. Hou, D. Chen, G. Shen,

Nano Res. 6 (2013) 525–534.

[73] G. Wang, H. Liu, J. Liu, S. Qiao, G. M. Lu, P. Munroe, H. Ahn, Adv. Mater. 22

(2010) 4944–4948.

[74] C. M. Doherty, R. A. Caruso, B. M. Smarsly, P. Adelhelm, C. J. Drummond,

Chem. Mater. 21 (2009) 5300–5306.

[75] X.-L. Wu, L.-Y. Jiang, F.-F. Cao, Y.-G. Guo, L.-J. Wan, Adv. Mater. 21 (2009)

2710–2714.

[76] D. Deng, J. Y. Lee, Nanotechnology. 22 (2011) 355401.

73
[77] L. Hu, B. Qu, C. Li, Y. Chen, L. Mei, D. Lei, L. Chen, Q. Li, T. Wang, J. Mater.

Chem. A. 1 (2013) 5596.

[78] N. Du, Y. Xu, H. Zhang, J. Yu, C. Zhai, D. Yang, Inorg. Chem. 50 (2011)

3320–3324.

[79] B. Liu, J. Zhang, X. Wang, G. Chen, D. Chen, C. Zhou, G. Shen, Nano Lett. 12

(2012) 3005–3011.

[80] Y. Wang, H. J. Zhang, L. Lu, L. P. Stubbs, C. C. Wong, J. Lin, ACS Nano. 4

(2010) 4753–4761.

[81] L. Chen, X. Guo, W. Lu, M. Chen, Q. Li, H. Xue, H. Pang, Coord. Chem. Rev.

368 (2018) 13–34.

[82] M. Zheng, H. Tang, Q. Hu, S. Zheng, L. Li, J. Xu, H. Pang, Adv. Funct. Mater.

28 (2018) 1707500.

[83] Y. Zhu, X. Fan, L. Suo, C. Luo, T. Gao, C. Wang, ACS Nano. 10 (2016)

1529–1538.

[84] C. Wang, W. Wan, Y. Huang, J. Chen, H. H. Zhou, X. X. Zhang, Nanoscale. 6

(2014) 5351–5358.

[85] Q. Li, Y. Xu, S. Zheng, X. Guo, H. Xue, H. Pang, Small. 14 (2018) 1800426.

[86] Y. Li, Y. Xu, W. Yang, W. Shen, H. Xue, H. Pang, Small. 14 (2018) 1704435.

[87] X. Yang, J.-K. Sun, M. Kitta, H. Pang, Q. Xu, Nat. Catal. 1 (2018) 214–220.

[88] S. Ding, J. S. Chen, X. W. D. Lou, Chem. - A Eur. J. 17 (2011) 13142–13145.

[89] K. Bindumadhavan, S. K. Srivastava, S. Mahanty, Chem. Commun. 49 (2013)

1823.

74
[90] L. Yang, S. Wang, J. Mao, J. Deng, Q. Gao, Y. Tang, O. G. Schmidt, Adv. Mater.

25 (2013) 1180–1184.

[91] S. Liang, J. Zhou, J. Liu, A. Pan, Y. Tang, T. Chen, G. Fang, CrystEngComm.

15 (2013) 4998.

[92] H. Yu, C. Zhu, K. Zhang, Y. Chen, C. Li, P. Gao, P. Yang, Q. Ouyang, J. Mater.

Chem. A. 2 (2014) 4551–4557.

[93] J. Jiang, J. Zhu, W. Ai, Z. Fan, X. Shen, C. Zou, J. Liu, H. Zhang, T. Yu, Energy

Environ. Sci. 7 (2014) 2670–2679.

[94] L. Zhang, Y. Wang, B. Peng, W. Yu, H. Wang, T. Wang, B. Deng, L. Chai, K.

Zhang, J. Wang, Green Chem. 16 (2014) 3926.

[95] H. Wang, C. Zhang, Z. Chen, H. K. Liu, Z. Guo, Carbon N. Y. 81 (2015)

782–787.

[96] X. Gu, C. Lai, F. Liu, W. Yang, Y. Hou, S. Zhang, J. Mater. Chem. A. 3 (2015)

9502–9509.

[97] F. Wu, L. Shi, D. Mu, H. Xu, B. Wu, Carbon N. Y. 86 (2015) 146–155.

[98]X. Yang, Y. Yu, N. Yan, H. Zhang, X. Li, H. Zhang, J. Mater. Chem. A. 4 (2016)

5965–5972.

[99] Y. Shi, X. Pan, B. Li, M. Zhao, H. Pang, Chem. Eng. J. 343 (2018) 427–446.

[100] Y. Luo, Y. Tang, S. Zheng, Y. Yan, H. Xue, H. Pang, J. Mater. Chem. A. 6

(2018) 4236–4259.

[101] M. Zheng, H. Tang, L. Li, Q. Hu, L. Zhang, H. Xue, H. Pang, Adv. Sci. 5

(2018) 1700592.

75
[102] B. Li, Y. Shi, K. Huang, M. Zhao, J. Qiu, H. Xue, H. Pang, Small. 14 (2018)

1703811.

[103] C. Zu, Y.-S. Su, Y. Fu, A. Manthiram, Phys. Chem. Chem. Phys. 15 (2013)

2291.

[104] X. Gu, Y. Wang, C. Lai, J. Qiu, S. Li, Y. Hou, W. Martens, N. Mahmood, S.

Zhang, Nano Res. 8 (2015) 129–139.

[105] Y.-S. Su, A. Manthiram, Chem. Commun. 48 (2012) 8817.

[106] B. Li, P. Gu, G. Zhang, Y. Lu, K. Huang, H. Xue, H. Pang, Small. 14 (2018)

1702184.

[107] S.-H. Chung, A. Manthiram, Adv. Mater. 26 (2014) 1360–1365.

[108] S.-H. Chung, A. Manthiram, ChemSusChem. 7 (2014) 1655–1661.

[109] W. G. Wang, X. Wang, L. Y. Tian, Y. L. Wang, S. H. Ye, J. Mater. Chem. A.

2 (2014) 4316–4323.

[110] Y. Shu, B. Li, J. Chen, Q. Xu, H. Pang, X. Hu, ACS Appl. Mater. Interfaces.

10 (2018) 2360–2367.

[111] L.-L. Xing, K.-J. Huang, S.-X. Cao, H. Pang, Chem. Eng. J. 332 (2018)

253–259.

[112] L. Zhang, Q. Li, H. Xue, H. Pang, ChemSusChem. 11 (2018) 1581–1599.

[113] P. Geng, S. Zheng, H. Tang, R. Zhu, L. Zhang, S. Cao, H. Xue, H. Pang, Adv.

Energy Mater. 8 (2018) 1703259.

[114] X. Han, Y. Xu, X. Chen, Y.-C. Chen, N. Weadock, J. Wan, H. Zhu, Y. Liu,

H. Li, G. Rubloff, C. Wang, L. Hu, Nano Energy. 2 (2013) 1197–1206.

76
[115] Y.-S. Su, A. Manthiram, Nat. Commun. 3 (2012) 1166.

[116] X. Wang, Z. Wang, L. Chen, J. Power Sources. 242 (2013) 65–69.

[117] S. Du, J. A. Valla, G. M. Bollas, Green Chem. 15 (2013) 3214.

[118] M. Biswal, A. Banerjee, M. Deo, S. Ogale, Energy Environ. Sci. 6 (2013)

1249.

[119] L. Wang, G. Mu, C. Tian, L. Sun, W. Zhou, P. Yu, J. Yin, H. Fu,

ChemSusChem. 6 (2013) 880–889.

[120] H. Bi, Z. Yin, X. Cao, X. Xie, C. Tan, X. Huang, B. Chen, F. Chen, Q. Yang,

X. Bu, X. Lu, L. Sun, H. Zhang, Adv. Mater. 25 (2013) 5916–5921.

[121] C. Zhang, W. W. Tjiu, T. Liu, RSC Adv. 3 (2013) 14938.

[122] N. Brun, C. A. García-González, I. Smirnova, M. M. Titirici, RSC Adv. 3

(2013) 17088.

[123] R. J. White, N. Brun, V. L. Budarin, J. H. Clark, M.-M. Titirici,

ChemSusChem. 7 (2014) 670–689.

[124] N. Brun, S. A. Wohlgemuth, P. Osiceanu, M. M. Titirici, Green Chem. 15

(2013) 2514.

[125] L.-F. Chen, Z.-H. Huang, H.-W. Liang, W.-T. Yao, Z.-Y. Yu, S.-H. Yu,

Energy Environ. Sci. 6 (2013) 3331.

[126] J. Huh, S. K. Saikin, J. C. Brookes, S. Valleau, T. Fujita, A. Aspuru-Guzik, J.

Am. Chem. Soc. 136 (2014) 2048–2057.

[127] H. Sun, W. He, C. Zong, L. Lu, ACS Appl. Mater. Interfaces. 5 (2013)

2261–2268.

77
[128] Y. Lu, B. Li, S. Zheng, Y. Xu, H. Xue, H. Pang, Adv. Funct. Mater. 27 (2017)

1703949.

[129] X. Xiao, S. Zheng, X. Li, G. Zhang, X. Guo, H. Xue, H. Pang, J. Mater.

Chem. B. 5 (2017) 5234–5239.

[130] X. Li, S. Ding, X. Xiao, J. Shao, J. Wei, H. Pang, Y. Yu, J. Mater. Chem. A.

5 (2017) 12774–12781.

[131] L. Zhang, S. Zheng, L. Wang, H. Tang, H. Xue, G. Wang, H. Pang, Small. 13

(2017) 1700917.

[132] X. Zuo, K. Chang, J. Zhao, Z. Xie, H. Tang, B. Li, Z. Chang, J. Mater. Chem.

A. 4 (2016) 51–58.

[133] R. Wang, C. Xu, J. Sun, Y. Liu, L. Gao, H. Yao, C. Lin, Nano Energy. 8

(2014) 183–195.

[134] Z. Wang, T. Chen, W. Chen, K. Chang, L. Ma, G. Huang, D. Chen, J. Y. Lee,

J. Mater. Chem. A. 1 (2013) 2202–2210.

[135] J. Yang, N. Huo, Y. Li, X.-W. Jiang, T. Li, R. Li, F. Lu, C. Fan, B. Li, K.

Nørgaard, B. W. Laursen, Z. Wei, J. Li, S.-S. Li, Adv. Electron. Mater. 1 (2015)

1500267.

[136] H. Fang, C. Battaglia, C. Carraro, S. Nemsak, B. Ozdol, J. S. Kang, H. A.

Bechtel, S. B. Desai, F. Kronast, A. A. Unal, G. Conti, C. Conlon, G. K. Palsson,

M. C. Martin, A. M. Minor, C. S. Fadley, E. Yablonovitch, R. Maboudian, A.

Javey, Proc. Natl. Acad. Sci. 111 (2014) 6198–6202.

78
[137] L. Wang, J. Sun, R. Song, S. Yang, H. Song, Adv. Energy Mater. 6 (2016)

1502067.

[138] F. Zhou, S. Xin, H.-W. Liang, L.-T. Song, S.-H. Yu, Angew. Chemie Int. Ed.

53 (2014) 11552–11556.

[139] Y. Xia, B. Wang, X. Zhao, G. Wang, H. Wang, Electrochim. Acta. 187 (2016)

55–64.

[140] X. Li, Y. Yang, J. Liu, L. Ouyang, J. Liu, R. Hu, L. Yang, M. Zhu, Appl.

Surf. Sci. 413 (2017) 169–174.

[141] Q. Lu, X. Wang, J. Cao, C. Chen, K. Chen, Z. Zhao, Z. Niu, J. Chen, Energy

Storage Mater. 8 (2017) 77–84.

[142] D. Ma, Y. Li, H. Mi, S. Luo, P. Zhang, Z. Lin, J. Li and H. Zhang, Angew.

Chemie, 130 (2018), 9039–9043.

[143] G. Zhang, X. Xiao, B. Li, P. Gu, H. Xue, H. Pang, J. Mater. Chem. A. 5

(2017) 8155–8186.

[144] H. Pang, X. Li, Q. Zhao, H. Xue, W.-Y. Lai, Z. Hu, W. Huang, Nano Energy.

35 (2017) 138–145.

[145] J. Yu, C. Mu, B. Yan, X. Qin, C. Shen, H. Xue, H. Pang, Mater. Horizons. 4

(2017) 557–569.

[146] P. Gu, M. Zheng, Q. Zhao, X. Xiao, H. Xue, H. Pang, J. Mater. Chem. A. 5

(2017) 7651–7666.

[147] Y. J. Nam, S.-J. Cho, D. Y. Oh, J.-M. Lim, S. Y. Kim, J. H. Song, Y.-G. Lee,

S.-Y. Lee, Y. S. Jung, Nano Lett. 15 (2015) 3317–3323.

79
[148] L. Li, S. Peng, H. Bin Wu, L. Yu, S. Madhavi, X. W. D. Lou, Adv. Energy

Mater. 5 (2015) 1500753.

[149] Y. Li, H. Dai, Chem. Soc. Rev. 43 (2014) 5257–5275.

[150] G. Nam, J. Park, M. Choi, P. Oh, S. Park, M. G. Kim, N. Park, J. Cho, J.-S.

Lee, ACS Nano. 9 (2015) 6493–6501.

[151] X. Liu, M. Park, M. G. Kim, S. Gupta, X. Wang, G. Wu, J. Cho, Nano

Energy. 20 (2016) 315–325.

[152] D. Geng, N. Ding, T. S. Andy Hor, Z. Liu, X. Sun, Y. Zong, J. Mater. Chem.

A. 3 (2015) 1795–1810.

[153] L. Dai, Y. Xue, L. Qu, H.-J. Choi, J.-B. Baek, Chem. Rev. 115 (2015)

4823–4892.

[154] G. S. Park, J.-S. Lee, S. T. Kim, S. Park, J. Cho, J. Power Sources. 243 (2013)

267–273.

[155] V. M. Dhavale, S. Kurungot, ACS Catal. 5 (2015) 1445–1452.

[156] R. Li, Z. Wei, X. Gou, ACS Catal. 5 (2015) 4133–4142.

[157] T. Chen, L. Dai, J. Mater. Chem. A. 2 (2014) 10756.

[158] Q. Liu, Y. Wang, L. Dai, J. Yao, Adv. Mater. 28 (2016) 3000–3006.

[159] X. Wu, X. Han, X. Ma, W. Zhang, Y. Deng, C. Zhong, W. Hu, ACS Appl.

Mater. Interfaces. 9 (2017) 12574–12583.

[160] X.-C. Li, Y. Zhang, C.-Y. Wang, Y. Wan, W.-Y. Lai, H. Pang, W. Huang,

Chem. Sci. 8 (2017) 2959–2965.

[161] X. Li, H. Xue, H. Pang, Nanoscale. 9 (2017) 216–222.

80
[162] T. Wang, S. Chen, H. Pang, H. Xue, Y. Yu, Adv. Sci. 4 (2017) 1600289.

[163] J. W. Lim, D. G. Lee, Compos. Struct. 134 (2015) 483–492.

[164] Q. Hong, H. Lu, Sci. Rep. 7 (2017) 3378.

[165] X. Ge, A. Sumboja, D. Wuu, T. An, B. Li, F. W. T. Goh, T. S. A. Hor, Y.

Zong, Z. Liu, ACS Catal. 5 (2015) 4643–4667.

[166] D. Wittmaier, N. A. Cañas, I. Biswas, K. A. Friedrich, Adv. Energy Mater. 5

(2015) 1500763.

[167] Y. Xue, H. Miao, S. Sun, Q. Wang, S. Li, Z. Liu, J. Power Sources. 297

(2015) 202–207.

[168] H. Yadegari, Q. Sun, X. Sun, Adv. Mater. 28 (2016) 7065–7093.

[169] J. Y. Cheon, K. Kim, Y. J. Sa, S. H. Sahgong, Y. Hong, J. Woo, S.-D. Yim,

H. Y. Jeong, Y. Kim, S. H. Joo, Adv. Energy Mater. 6(2016) 1501794.

[170] K.-N. Jung, J. Kim, Y. Yamauchi, M.-S. Park, J.-W. Lee, J. H. Kim, J. Mater.

Chem. A. 4 (2016) 14050–14068.

[171] J. Ma, J. Wen, J. Gao, Q. Li, Electrochim. Acta. 129 (2014) 69–75.

[172] X. Liu, H. Zheng, Z. Sun, A. Han, P. Du, ACS Catal. 5 (2015) 1530–1538.

[173] X. Han, F. Cheng, T. Zhang, J. Yang, Y. Hu, J. Chen, Adv. Mater. 26 (2014)

2047–2051.

[174] W. Xia, A. Mahmood, Z. Liang, R. Zou, S. Guo, Angew. Chemie Int. Ed. 55

(2016) 2650–2676.

[175] M.-S. Park, J. Kim, K. J. Kim, J.-W. Lee, J. H. Kim, Y. Yamauchi, Phys.

Chem. Chem. Phys. 17 (2015) 30963–30977.

81
[176] X. Liu, M. Park, M. G. Kim, S. Gupta, G. Wu, J. Cho, Angew. Chemie Int.

Ed. 54 (2015) 9654–9658.

[177] X. Wu, F. Chen, Y. Jin, N. Zhang, R. L. Johnston, ACS Appl. Mater.

Interfaces. 7 (2015) 17782–17791.

[178] J. Kim, J. Lee, J. You, M.-S. Park, M. S. Al Hossain, Y. Yamauchi, J. H.

Kim, Mater. Horizons. 3 (2016) 517–535.

[179] Z. Lu, W. Xu, J. Ma, Y. Li, X. Sun, L. Jiang, Adv. Mater. 28 (2016)

7155–7161.

[180 L. Pu, K. Li, Z. Chen, P. Zhang, X. Zhang, Z. Fu, J. Power Sources. 268 (2014)

476–481.

[181] K. J. Kim, M.-S. Park, Y.-J. Kim, J. H. Kim, S. X. Dou, M. Skyllas-Kazacos,

J. Mater. Chem. A. 3 (2015) 16913–16933.

[182] T. Y. Ma, J. Ran, S. Dai, M. Jaroniec, S. Z. Qiao, Angew. Chemie Int. Ed. 54

(2015) 4646–4650.

[183] X. Wu, F. Chen, N. Zhang, A. Qaseem, R. L. Johnston, J. Mater. Chem. A. 4

(2016) 3527–3537.

[184] Y. Yan, P. Gu, S. Zheng, M. Zheng, H. Pang, H. Xue, J. Mater. Chem. A. 4

(2016) 19078–19085.

82
Huan Pang received his Ph. D. degree from Nanjing University
in 2011. He is now a university distinguished professor in Yangzhou University. In
the past 10 years, his group has been engaged in the design and synthesis of functional
nanomaterials, especially for MOF-based materials. He is on the editorial board of
FlatChem and managing editor of EnergyChem, Elsevier. He has published more than
200 papers in peer-reviewing journals including Nature Catalysis, Chemical Society
Reviews, Energy Environ. Sci., Advanced Materials, with 8000 citations
(H-index=53). His research interests include the development of inorganic
nanostructures and their applications in nanoelectrochemistry with a focus on energy
devices.

83
1. A broad overview of carbon fiber materials for batteries.

2. Synthetic strategy, morphology, structure, and property have been researched.

3. Carbon fiber composites can improve the conductivity of electrode material.

4. Challenges in future development of carbon fiber materials are addressed.

84
85

You might also like