Вы находитесь на странице: 1из 88

Authors Accepted Manuscript

Electrospun hollow nanofibers for advanced


secondary batteries

Linlin Li, Shengjie Peng, Jeremy Kong Yoong Lee,


Dongxiao Ji, Madhavi Srinivasan, Seeram
Ramakrishna
www.elsevier.com/locate/nanoenergy

PII: S2211-2855(17)30405-6
DOI: http://dx.doi.org/10.1016/j.nanoen.2017.06.050
Reference: NANOEN2056
To appear in: Nano Energy
Received date: 24 May 2017
Revised date: 26 June 2017
Accepted date: 30 June 2017
Cite this article as: Linlin Li, Shengjie Peng, Jeremy Kong Yoong Lee, Dongxiao
Ji, Madhavi Srinivasan and Seeram Ramakrishna, Electrospun hollow nanofibers
for advanced secondary batteries, Nano Energy,
http://dx.doi.org/10.1016/j.nanoen.2017.06.050
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Electrospun hollow nanofibers for advanced secondary batteries
Linlin Lia,c, Shengjie Penga,b*, Jeremy Kong Yoong Leeb, Dongxiao Jib, Madhavi Srinivasanc,
Seeram Ramakrishnab
a
Jiangsu Key Laboratory of Materials and Technology for Energy Conversion, College of
Materials Science and Technology, Nanjing University of Aeronautics and Astronautics,
Nanjing 210016, China
b
Department of Mechanical Engineering, National University of Singapore, 117574,
Singapore
c
School of Materials Science and Engineering, Nanyang Technological University, 639798,
Singapore

pengshengjie@nuaa.edu.cn
mpepeng@nus.edu.sg

Abstract

Electrospinning is a versatile and efficient technique to fabricate hollow structured nanofibers

of metal oxides, carbon, metals, composites, and others, which present great potential in

energy conversion and storage devices. In this review, we provide a comprehensive overview

of the preparation of electrospun inorganic hollow nanofibers and their applications in

advanced secondary batteries. We rst describe different synthetic methodologies for

fabricating hollow nanofibers as well as the compositional and geometric manipulation of

these nanofibers based on electrospinning. We next go on to discuss the recent research

activities involving electrospun hollow nanofibers in advanced secondary batteries, including

lithium-ion batteries, sodium-batteries, Li-S batteries, and metal-air batteries. In each

application, some typical examples and a critical commentary are provided to demonstrate the

superior properties of the hollow structures. The correlation between geometric properties of

hollow structures and their performance are also highlighted. These results demonstrate that

electrospun hollow nanofibers with unique properties such as large surface area, engineered

high porosity, and free-standing characteristics have a direct impact on the final performance

in their applications. Finally, in conclusion, the remaining challenges of adopting electrospun

hollow nanofiber materials in their future applications are further discussed.

1
Graphic abstract

The review summarizes the recent progress of the preparation of electrospun inorganic hollow

nanofibers and their applications in advanced secondary batteries, including lithium-ion

batteries, sodium-batteries, Li-S batteries, and metal-air batteries.

Keywords: Eletrospinning; Hollow nanofibers; One-dimensional structures; Advanced


secondary batteries; Lithium batteries

1. Introduction
The over exploitation of non-renewable natural resources such as fossil fuels to drive the

engines of economic growth have unambiguously had severe and detrimental environmental

impacts. In order to sustainably meet the ever-increasing global energy demand, it is

imperative that renewable, clean energy technologies are developed. Of the existing energy

conversion and storage technologies, rechargeable batteries, such as lithium ion batteries

(LIBs),[1, 2] with high energy density, excellent rate capability, long cycle lives, and low self-

discharging rates, represent a promising class of energy storage devices which can be used as

a high-energy, mobile electrochemical power source, providing clean electrical energy from

stored materials time after time.[3] The unique characteristics LIBs possess stand them in

good stead for time-dependent power system applications such as in portable electronic

devices, an industry estimated to be worth over ten billion dollars. In addition, a large number

of future industries, such as the automotive industry with full electric vehicles (EVs), hybrid

2
electric vehicles (HEVs), and plug-in hybrid electric vehicles (PHEVs) all gaining popularity,

will depend on battery technologies, paving the way for potentially staggering market growth

for LIBs.[4] Therefore, the design and materials necessary for the advanced secondary

batteries with superior electrochemical performance has attracted substantial research

attention.

Although the LIBs can trace its heritage back over 20 years when it first emerged as the

dominant portable energy source in electronic devices, it still falls a long way short of

meeting the energy density requirements for long range EVs and power density needs for

load-levelling applications, not to mention the requirement for long term stability.[5] There is

therefore considerable interest in developing batteries with superior performance in terms of

energy density, power density, and cycle stability. Furthermore, the scarcity of lithium

resources and its high cost may hinder the further development of LIB technologies. This has

led to a trend in the exploration of exploring alternative materials that exploit the more

abundant elements, enabling the industrial fabrication of lower cost batteries. Sodium for

instance, which has the second lowest mass and is the smallest alkali metal, is one such cheap,

readily available alternative. Sodium-ion batteries (NIBs) also share much of the same energy

storage mechanisms and configurations with conventional LIBs, making them promising

candidates to replace LIBs.[6] Apart from NIBs, much activity has been devoted to the study

of metal-air and Li-S batteries as these technologies have demonstrated much greater

volumetric and gravimetric energy density then their LIB counterparts, useful in EVs and grid

scale energy storage applications, as shown in Figure 1.[7, 8] The performance of these

alternative rechargeable batteries has accelerated in recent years, and are anticipated to

approach levels signaling commercial viability in the near future.

3
Figure 1. The gravimetric energy densities (Wh kg-1) for various types of rechargeable
batteries compared to gasoline. The theoretical density is based strictly on thermodynamics
and is shown as the blue bars while the practical achievable density is indicated by the red
bars and numerical values. For Li-air, the practical value is just an estimate. For gasoline, the
practical value includes the average tank-to-wheel efficiency of cars.

One critical aspect in the design of batteries with higher energy and power densities is the

use of functional electrodes with nanoscale engineering since these nanostructured electrodes

play a deterministic role in the overall performance of the battery.[9-12] Of the

nanoarchitectures that have been studied, one-dimensional (1D) nanofibers possess a less

agglomerated configuration which has been proven to yield improved battery electrochemical

performance.[13, 14] Recent developments in materials science related to hollow

nanostructures have generated significant buzz concerning their potential to be applied in

domains as diverse as medicine, from drug delivery,[15] molecular sensors,[16] and

biomedical engineering,[17] to energy, including energy storage[18-21] and catalysis[22].

Electrospinning has proven to be an excellent technique in processing polymer melts or

viscous polymer solutions into continuous 1D nanofibers with diameters ranging from the

order of ten nanometers to several micrometers. Best known for its merits of being simple,

efficient, cheap, and highly reproducible,[23, 24] the technique has received much academic

scrutiny, and found a diverse range of applications including fabricating filtration membranes,

bioluminescent materials, tissue engineering scaffolds, electrode materials, and nanoscale

4
cables.[25-28] Current research efforts are focused on electrospinning more complex

architectures such as hollow, core-shell, and porous nanofibers by modifying the

electrospinning setup and optimizing the processing parameters. Coaxial electrospinning is a

popular route to obtain hollow nanofibers, wherein a coaxial spinneret replaces the single

spinneret, allowing two solutions to be fed and spun into nanofibers in a coaxial configuration

before appropriate post treatments such as calcination or core-solvent extraction can then be

employed.[29, 30] 1D nanofibers with mesoporous walls fabricated from by electrospinning

tend to have characteristics suitable for energy storage and conversion applications, including

greater specific surface areas, higher aspect ratios, and improved pore interconnectivity.[27]

Compared with the solid nanofibers, the advantages for the hollow nanofibers include a

greater surface area and high aspect ratio which provides more active sites for ion adsorption;

higher pore volume and surface permeability which facilitates improved kinetics vis--vis

intercalation/de-intercalation of active species, shortens the diffusion pathways for ions and

electrons, and also increases the number of energy storage sites.[31-34] Besides these, the

voids of the hollow nanostructure allow for a degree of buffering against volume changes

which accompany the Li insertion/desertion, staving off pulverization and agglomeration of

the electrode, thereby improving electrochemical performance and cycling life for LIBs and

NIBs.[35] With regard to the Li-S batteries, the large internal void space of the hollow

structures not only allows relatively high loading of sulfur, whereas solid fibers can only load

sulfur on their exposed surfaces, but also provides enough space in the hollow interior to

accommodate large volumetric expansions during the lithiation process.[36] Furthermore, the

integrated shells can promise more ecient confinement of soluble LiPSs, which is different

from the simple adsorption of LiPSs on the surfaces of solid-type hosts. When applied in

metal-air batteries, the hollow structure with a high surface area can offer a large reaction area

for nucleation or the accommodation of Li2O2.[37] Meanwhile, the high surface are can

provide more electrocatalytic sites, which result in higher specic capacity. In addition, the
5
porous hollow structure could offer more abundant oxygen and electrolyte transportation

paths in the electrode, facilitating the formation and decomposition of the discharge product

and thus improving the reversibility of the O2 electrode. These improvements go some way in

addressing the bottlenecks in the development of high performance battery technologies. In

particular, compared to conventional powder based nanomaterials, self-supporting carbon

nanofiber electrodes eschew the use of polymeric binders which block the active sites, inhibit

diffusion, and increase series resistance.[38, 39] At the time of writing, there already exist

numerous reports in the literature on the superior electrochemical performance of batteries

that employ fibers of organic polymers or inorganic materials in core-shell and hollow

configurations.[40-42]

Compared to the hollow nanofibers obtained by other methods, electrospun hollow

nanofibers have several advantages. Firstly, it is applicable to all condensable materials (to the

best of the authors experience) to obtain hollow nanofibers, especially by the well-

established coaxial electrospinning or co-electrospinning technique, requiring only the careful

selection of inorganic precursors (salts or alkoxides) and the corresponding solvents.

Secondly, from an industrial/commercial point of view, electrospinning is a cost-effective

method that can be easily scaled-up, making it very economically attractive and compatible

with industrial requirements.[43] Thirdly, compared with hollow structured materials obtained

by other fabrication methods, the dimensions, in particular, the diameter of electrospun

hollow nanofibers, can be easily tuned from the order of 10 nm to a few microns by simply

altering the electrospinning parameters such as polymer concentration, applied electric field,

and flow rate of the feed solution.[44] Furthermore, the length and orientation of the

nanofibers can also be varied by controlling these parameters as well as the geometry and size

of the collector mandrel.[45] Thus, the electrospinning process should be optimized based on

the aforementioned parameters to obtain uniform, long, continuous nanofibers of a desired

orientation and diameter. Besides hollow structures, electrospinning also enables the
6
fabrication of complex, multilevel interior hollow nanofibers which are not easily fabricated

using other methods. Fourthly, besides randomly oriented webs, electrospun nanofibers can

be collected as aligned mats if a rotating collector or two parallel electrodes are used.

Furthermore, individual nanofibers can be collected via this approach, which magnifies the

possible uses of electrospinning to prepare materials for microscale devices. Finally,

electrospun hollow nanofibers prepared via coaxial electrospinning or emulsion

electrospinning have been investigated by theoretical and in-situ optical analysis, which can

help provide insight to understanding the formation mechanisms and guide future

synthesis.[46, 47].

As is well documented, electrospinning has proven to be an excellent technique in

processing polymer melts or viscous polymer solutions into continuous 1D inorganic

nanofibers with diameters ranging from the order of ten nanometers to several micrometers.

With regard to energy applications, electrospun materials have demonstrated superior

electrochemical performance in batteries, supercapacitors, solar cells, etc, and have

correspondingly attracted immense research interest leading to the publication of several

review articles to summarize the work done in this flourishing field.[10, 14, 15, 25, 48-50]

Although there are several excellent review articles which have summarized the developments

in the field of electrospun nanofibers, we believe that our review manuscript with an in-depth

analysis specifically on the hollow nanostructure, from their preparation to their application in

advanced secondary batteries is significant, focused, and timely. The authors believe it will

attract interest from the scientific community, both from researchers familiar with

electrospinning and are looking to design materials with more complex nanoarchitectures in

energy devices, as well as from energy researchers looking for alternative fabrication methods

to synthesize nanomaterials. This comprehensive review will suit experienced researchers and

aspiring scholars alike with its holistic content incorporating both fabrication and application

7
of this class of material, coupled with the authors distinct conclusions on the challenges

remaining and informed perspectives on future research directions.

In this review, we concisely address these aspects, underscoring the importance of tuning

the nanostructural evolution to obtain an optimal combination of chemical and physical

properties for the rational design of electrode materials. A summary of the recent results in the

downstream implementations of these materials in advanced secondary batteries is also

provided. Given the breadth of the materials being studied, for brevity we shall be focusing

only on the materials with relevance to electrospinning, namely inorganic materials such as

carbon, metal oxides, and silicon composites. We first give a brief description of the

electrospinning process, followed by a critical review of the fabrication and design of

electrospun hollow nanofibers, as shown in Figure 2. To establish the unique attributes arising

from the hollow nanostructure, a range of examples will be discussed, illustrating how this

nanostructure broadly contributes to the morphological variation and electrochemical

performance in recent studies of advanced batteries such as LIBs, NIBs Li-O2, and Li-S

batteries. In addition, a discussion on how particular characteristics influence particular

properties of the electrodes is provided. Lastly, we propose a challenge and postulate on the

future research directions related to the structure and composition of electrode materials. It is

hoped that through this review paper, a new perspective and better understanding of how to

design and develop hollow electrospun nanofibers for electrode materials can be gleaned.

8
Figure 2. A schematic diagram of electrospun hollow nanofiberes for advanced secondary
batteries, including lithium-ion battery, sodium-ion battery, Li-S battery and metal-air battery.

2. Hollow nanofibers based on electrospinning

Electrospinning is an electrostatically driven process used to fabricate nanofibers. Advances

in the methodologies and strategies for synthesizing hollow nanofibers over the past few years

has encouraged their adoption and application in many fields.[51-53] Generally, modifying a

number of electrospinning parameters can result in nanofibers with hollow structures, for

instance, changing the spinneret to a coaxial spinneret, using an emulsion based spinning dope,

and the control over the heating conditions, as shown in Figure 3. These synthesis methods

have been well-researched to manipulate the resulting structures and improve the materials

properties such as increased surface area, enhanced mechanical properties, and tunable

anisotropic charge carrier transport characteristics. In this section, the protocols developed for

the synthesis of electrospun hollow nanofibers will be discussed in detail.

9
Figure 3. Schematic of an electrospinning setup to obtain hollow nanofibers with various
structures, which are influenced by the properties of precursor polymers, electrospinning
parameters, and post-treatment.

2.1 Electrospinning

2.1.1 Electrospinning process

As illustrated in Figure 1A, the electrospinning setup is relatively straightforward and

facilitates easy modification. The rudimentary setup consists of only three major components,

namely a high voltage power source, a spinneret, and a conductive collector (typically in a flat

plate or rotating drum configuration).[49, 54] A general electrospinning experiment involves

applying a strong electric field to a droplet of polymer solution to effect the drawing of the

nanofiber. A syringe loaded with the polymer solution is coupled to a syringe pump which

feeds the solution at a controlled rate to a spinneret with a thin nozzle (with an inner diameter

approximately between 100 m to 1 mm). In most laboratory scale setups, the air-gap

distance between collector and spinneret is approximately between 5 to 25 cm. The nozzle

functions as an electrode, while the collector acts as its counter electrode when a high voltage
10
(100-3000 kV m-1) is applied between them. The electric field electrifies the pendant drop of

polymer solution on the nozzle tip, inducing charges that are distributed uniformly over the

surface of the droplet. When the electric field strength reaches a critical level, the surface

tension of the polymer solution is overcome by the electrostatic forces, resulting in the

evolution of a fine liquid jet that is drawn toward the counter electrode. It is during this

movement that the solvent evaporates or the melt solidifies, precipitating a long, thin, solids

fiber that collects on the collector as a non-woven, randomly oriented mat.

2.1.2 The effect of electrospinning parameters

The eventual characteristics of these nanofibers are heavily influenced by a number of

parameters such as the solution characteristics (including concentration, viscosity, surface

tension, and conductivity), electrospinning conditions (including tip-to-collector distance,

solution feed rate, and voltage), and ambient environment (including humidity, and

temperature). Among these process parameters, both voltage and viscosity could play

important parts in the morphology of the final nanofibers. Generally, high voltage and

viscosity are beneficial for the formation of smooth and uniform nanofibers. While, low

voltage and viscosity usually lead to the generation of beads or even no any fibrous structures

formed. With the development of materials science and technology, inorganic and ceramic

nanofibers can also be fabricated by electrospinning when combined with a subsequent

carbonization or calcination step.[50, 55] Starting with a suitable composite solution of a

polymer loaded with the precursor of interest, electrospinning under appropriate conditions

has been found to yield polymer nanofibers with precursor salts uniformly distributed

throughout the nanofiber. Thermal treatment or treatment in the presence of reducing agents

(e.g. hydrogen) serves to reduce the precursor salts to the target metal or metal oxides whilst

maintaining the 1D nanostructure, forming composite nanofibers. The organic component of

the composite nanofibers can also be removed by the application of suitable thermal

treatments in air, resulting in nanofibers of pure metal or metal oxide. Control over the
11
processing conditions allows the manipulation of the diameters of the resulting inorganic

nanofiber in the order of 10 to 100 nm with narrow size distributions. Carbonaceous

nanofibers can also be fabricated by using polymers with high carbon yield such as

polyaniline (PAN), by subjecting the electrospun polymer nanofibers to stabilization and

carbonization treatments.[50] Subsequent activation steps may be necessary to modify the

morphology and surface chemistry should an improved capacity for ion adsorption or

enhanced electrochemical performance be desired. Many examples for the successful

fabrication of metal oxide nanofibers like ZnO, SnO2, and TiO2 have been reported by

calcining the composite nanofibers in air.[30, 56, 57] Other results reported in the literature

suggest that in addition to PAN and Poly(vinylpyrolidone) (PVP) polymers in various

solvents such as N, N-dimethylformamide (DMF) and ethanol, Poly(ethylene oxide) (PEO) in

water can also be used as polymer carriers hosting inorganic precursors to produce

electrospun nanofibers, hinting at the many possible combinations of polymer, solvent, and

precursor to synthesize a wide variety of inorganic nanofiber materials.[58]

2.2 Strategies

A number of review articles in the last decade have detailed the broad array of strategies used

to influence the electrospinning of nanofibers and their subsequent properties.[14, 50, 59]

These articles considered how these strategies can be handled to control the nanofiber

diameter, porosity, branching, morphology, beading, as well as the deposition patterns

resulting from the angle of the conical envelope of the bending path. As shown in Figure 4,

coaxial elctrospinning, emulsion electrospinning, heating process and hard templates are

employed to synthesize the hollow nanofibers.

12
Figure 4. Scheme Schematic of an electrospinning setup to obtain hollow nanofibers with

various structures,

2.2.1 Coaxial electrospinning

The advent of coaxial electrospinning in 2003 has enabled a wide range of nanofiber

architectures such as the novel core-shell structure to be developed.[60] At present, coaxial

electrospinning is generally used to fabricate nanotubes and core shell nanofibers composed

of various polymers, semiconductors, and ceramics of varying compositions.[61-64] In brief,

core-shell electrospinning utilizes two precursor solutions, fed simultaneously into different

chambers of a specially designed spinneret arranged in an inner die and outer die

configuration with concentric nozzles. The coaxial spinneret allows for the formation of a

compound jet to be ejected from the tip of the droplet upon the application of an electric field.

The compound jet undergoes the typical bending instabilities, stretching and thinning while

the solvent evaporates, eventually depositing on the collector as a mat comprising thin core-

shell fibers. Several studies on the behavior of the compound jet have detailed the conditions

13
necessary for the formation of this hollow structures as well. The rapid precipitation of the

shell material and good wettability of the shell by the core material have been shown to favor

the formation of tube like structures.

The choice of the core and shell solution to be electrospun can affect the eventual

morphology of the nanofibers. For instance, PVP with a titanium precursor when selected as

the shell material with mineral oil as the core has been reported to yield hollow titania

nanofibers upon the removal of the mineral oil core and thermal decomposition of the PVP

followed by calcination.[65, 66] Carbonaceous nanotubes have also been electrospun from a

poly(methyl methacrylate) (PMMA) based precursor core solution and a PAN based shell

solution by subjecting the spun polymer fibers to a two stage heating treatment which served

to decompose the PMMA at a lower temperature and carbonize PAN at a higher temperature.

Besides these approaches, a multi-fluidic coaxial electrospinning technique has also been used

to produce ultra-thin, core-shell nanofibers with a unique nanowire-in-microtube design, as

shown in Figure 5a-c. In this approach, a spinneret comprising three coaxial capillaries was

used, in which a chemically inert fluid acts as a buffer between the outer and inner polymer

solutions. The buffer middle fluid is then removed, creating a void between the inner solid

nanofiber, and the outer solid microtube. This approach was successfully demonstrated in

fabricating a TiO2 nanofiber encased in a TiO2 microtube.[67, 68]

In addition to its usefulness in fabricating hollow structures, the triple-coaxial

electrospinning technique can be adapted to produce other novel fiber architectures.[69] The

basic coaxial electrospinning is illustrated in Figure 5d-f, and consists of the hierarchically

assembled compound nozzle comprising two or more metal capillaries separately leading to a

single, blunted metallic needle. Careful selection of the outer shell polymer solution, middle

fluid, and inner core polymer solutions in terms of miscibility is critical in the stable ejection

of the multi-fluidic compound jet to produce multi-channel microtubes wherein the middle

14
wall can be further divided into two to four channels, proving the versatility and simplicity of

this technique.

Alterations to the electrospinning system has enabled the production of coaxially

electrospun nanotubes from sol-gels and composite materials having immiscible cores.

Modifications such as incorporating active compounds into the core dope have been shown to

be effective in altering the inner surface of the nanofibers. This modified electrospinning

technique has also been employed in tandem with post processing steps like carbonization,

selective dissolution, and spinning into liquid nitrogen for the fabrication of coagulating

solutions and blends to achieve novel architectures such as porous structures and

microchannels.

Figure 5. a) Schematic illustration of the multifluidic coaxial electrospinning experimental


setup used to generate hollow fibers with the nanowire@nanotube structure; b) SEM and c)
TEM images of TiO2 with nanowire@nanotube structure; Reproduced with permission. [68]
Copyright 2010, American Chemical Society. d) Side-view SEM image of sample after the
organics have been removed; e-h) SEM images of multichannel tubes with variable diameter
and channel number from two to five. Reproduced with permission.[69] Copyright 2007,
American Chemical Society.

2.2.2 Emulsion electrospinning

While promising as a technique to produce hollow oxide structures, coaxial electrospinning

suffers from complications arising from the complex interplay between factors such as the
15
miscibility between the solutions used, the relative viscosity between them, the interfacial

tension, etc., rendering it difficult to achieve consistent concentricity of the spun nanofibers.

To address these limitations, much effort has been devoted to developing single capillary

electrospinning. The single nozzle co-electrospinning technique employs the use of two

different polymer solutions to induce phase separation.[61] The use of such a one-pot, single

spinneret method to prepare hollow nanofibers overcomes the abovementioned limitations,

and has been demonstrated for hollow nanofibers of many different compositions including

Fe2O3, SnO2, TiO2, NiO, CoFe2O4, ZnO-SnO2, etc.

A classic example of this one-pot, single spinneret approach uses a blend of PMMA and

PAN in DMF. The metastable polymer blend has a tendency to form an emulsion upon

mixing for one day, and further separate into distinct layers with time, as observed in the

optical images.[70, 71] The emulsion consists of PMMA/DMF solution organized in 100-200

m-diameter droplets floating in the PAN/DMF solution. During electrospinning, a core-shell

jet is ejected from the short-lived coreshell Taylor cone where the PMMA/DMF droplet is

trapped in the tip of the needle. Core-shell nanofibers with wall thicknesses ranging from 200

nm to 1 m, and exterior diameters between 0.5 and 5 m were obtained, and subject to

pyrolysis to decompose the PMMA component, thereby forming hollow nanofibers, and even

allowing the fabrication of multi-channeled, hollow carbon microtubes with diameters

approximately 200 nm. It was also found that tuning the PMMA molecular weight had a

profound effect on the eventual structure. Serveral other notable works have deployed this

emulsion single nozzle co-spinning strategy for blends of PMMA/CA, PAN/PA, etc. Of

particular interest is a study by Lou et al. on the 3D interconnected microchannel carbon, spun

from a blend of PAN/PS, resembling the structure of lotus roots.[72] The channel structure

was easily tuned by varying the ratio of PS and PAN used, from 1:10 to 1:1. Increasing the PS

content served to increase the number and diameter of the channels inside the nanofibers as

show in the Figure 6a-h.


16
Figure 6i and j. show the phase separation of a PAN/PVP/tetrabutyl orthotitanate

(TBOT)/DMF emulsion as the compound emulsion is jetted from the nozzle tip, resulting in

the elongation PAN/DMF droplets floating in the PVP/TBOT/DMF solution while the DMF

is evaporated during the stretching and thinning, to form the multi-channel nanofibers.[73]

The relatively small TBOT can also migrate without impediment to the exterior regions of the

nanofiber during this process, enabling the development of carbon free, hollow, multi-channel

structured TiO2 nanofibers upon calcination in air to remove the polymer component.

In another example, a PVP/Sn solution system was prepared with appropriate co-solvents to

produce PVP/Sn precursor nanofibers with a core-shell structure. The PVP was then rapidly

decomposed to retain the fibrous structure, establishing a concentration gradient between Sn

precursor and SnO2 as a result of surface diffusion and the Kirkendall effect during

calcination, as shown in Figure 6k and l.[74] The resulting SnO2 nanofibers possessed hollow

structures attributed to the differential evaporation rate of the solvents, where the interior

solvent evaporates faster than the diffusion through the surface, leading to a buildup of

pressure inside the fiber and formation of voids. Exploiting this phenomenon, Sanjay and co-

workers were able to produce hollow LiFeO2 nanofibers.[75]

Mai and co-workers have of late designed a gradient electrospinning cum controlled

pyrolysis protocol to prepare various 1D nanostructured materials such as continuous

nanowires, pea-like nanotubes, and mesoporous nanotubes, as show in Figure 6m-o.[51] At

the core of their method lies the use of different molecular weight polymers to during

electrospinning, which under strong electrostatic forces, favors the separation according to

their weights. They electrospun low, medium, and high molecular weight poly(vinyl alcohol)

which separated into three layers according to the mass gradient upon application of the

electric field. Subjecting the resulting nanofibers to sintering in air, inorganic mesoporous

nanotubes composed of small interconnected inorganic nanoparticles due to the PVA mass

loss. Additionally, they found that annealing the produced fibers under an argon atmosphere
17
at high temperatures results in carbonization of the PVA at the surface while in the interior,

the inorganic materials coalesce into larger nanoparticles, giving rise to the pea-pod like

nanotube structure.

Figure 6. a-d) Schematic diagrams, eh, m, n) FESEM and il, o, p) TEM images of lotus
root-like multichannel carbon nanobers based on various PAN/PS weight ratio: a, e) 1:0.1, b,
f) 1:0.2, (c, g) 1:0.5, d, h) 1:1; Reproduced with permission.[72] Copyright 2015, Nature
Publishing Group. SEM and TEM images of TiO2 nanofiber with fiber-in-tube nanostructure
heat-treated at 450 C under an oxygen atmosphere; Reproduced with permission.[76]
Copyright 2015, Wiley. k) SEM and l) TEM images of SnO2 nanotubes with a fiber-in- tube
structure obtained by heating to 500 C over 2 h at a rate of 5 C min-1; Reproduced with
permission.[77] Copyright 2015, Wiley.m, n) SEM and o) TEM images of Li3V2(PO4)3 (m)
mesoporous nanotubes and n, o) pea-like nanotubes , respectively. Reproduced with
permission.[51] Copyright 2015, Nature Publishing Group.

2.2.3 Heating process

18
The polymer based composite nanofibers containing inorganic precursors are often subject to

one or many heat treatment processes under controlled atmospheres of air, nitrogen, argon,

hydrogen, organic gases like ethylene, or mixtures of these gases. The heat treatment is

employed to effect one or more of the following chemical changes to the prepared materials,

(i) carbonize the carrying polymer, (ii) decompose the sacrificial polymer, (iii) convert the

metallic precursor to their corresponding metal, oxide, or hybrid, and (iv) functionalize the

material. The strong influence calcination treatments have on the eventual crystallization and

structural development of the hollow nanofibers is critical, thus to a large extent determines

their eventual functional electrochemical properties. The versatility of the one-pot

electrospinning technique is demonstrated as it allows the preparation of a wide range of

functional materials such as TiO2, SnO2, MnO2, etc. Generally, molecular chains of the

polymer in the electrospinning dope solution carry induced positive charges in the presence of

the applied voltage, causing them to migrate to the ejected jet surface. This results in the

formation of a gel-like shell on the precursor fibers associated with rapid phase separation and

solvent evaporation. Upon subjecting these fibers to heat treatment, the gel-like shell stiffens,

allowing the material to retain the 1D fibrous morphology. The polymer content in the gel

layer is eliminated during the calcination step while the formation of the inorganic particles

occurs simultaneously. Inorganic salts present at the surface of the fiber exposed to air during

the treatment form an oxide layer as a shell, whereas in the interior bulk of the fiber, a lack of

oxygen prevents the decomposition of the salt to its corresponding oxide. As the heat

treatment temperature is ramped up, a difference in the concentration gradient drives the

remaining salts to migrate from the fiber bulk to the surface, eventually forming hollow

inorganic nanofibers. This approach has been adopted with success for a variety of oxides. In

a separate work, Kang and co-workers also reported the fabrication of fiber-in-tube structures

of SnO2 and TiO2 by initially heating the electrospun nanofibers to form an intermediate

19
product of metal oxide-carbon nanofiber composites, then subsequently applying a further

heat treatment to derive the desired product.[77, 78]

Apart from the control of the calcination temperature, the concentration of the precursor

solution and the heating rates applied have also been shown to dramatically affect the

development of the hollow structure. Reports have suggested that such hollow nanostructures

evolve along a prescribed pathway during heat treatment, the so-called heterogenous

contraction, in which the heating ramp rate dictates the final geometry, forming either hollow

structures or more complex architectures like double or triple walled structures. Based on this

finding, our group has devised a generic facile protocol to fabricate tube-in-tube structures by

rational adjusting the heating rate applied to various mixed metal oxides such as CoFe2O4,

NiCo2O4, and CoMn2O4, taking advantage of the contractile and adhesive forces induced

during the heat treatment.[53, 79, 80]

With regard to manipulating the precursor concentrations, Xu and co-workers made use of

an intermediate concentration (between what was necessary to fabricate hollow fibers and

solid fibers) of TBOT in the precursor solution to fabricate tube-in-tube structured TiO2.[78]

Controlling the concentration of TBOT causes differences in the melting and gasification

behavior between the interior and the surface of the fiber. Initially, pressure differences

arising from the melting of the outer shell drive TiO2 nanoparticles to move to the surface,

forming the exterior tube. Subsequently, the melting of the inner region induces differential

pressures, bringing about the migration of the TiO2 nanoparticles to coalesce at the surface of

the melted inner region, forming the interior tube structure of the tube-in-tube hollow fiber.

Their methodology can be applied in the preparation of other multi-tubular metal oxide

hollow fibers such as tube-in-tube structures of In2O3, SnO2, SiO2, and ZrO2.

2.2.4 Hard templates

Among the various strategies for fabricating hollow structures, template-based methods are

considered one of the more efficient ways of creating nanostructured pores in various
20
materials. This approach offers certain advantages including a greater degree of control over

the shape, size, and uniformity of the final product since these properties are dictated by the

template used, but comes at the expense of more complex protocols. There are two general

template based approaches to elecrospinning hollow nanofibers, the first utilizes a spinning

solution containing a hard template, and the second electrospins the nanofibers as the

templates directly. Both approaches require the eventual removal of the templates to obtain

the hollow structures.

For hard template based strategies, SiO2 spheres, with their straightforward fabrication,

high yield, and their uniform diameter distributions, have been widely used as hard templates

introduced into the electrospinning solution.[81, 82] Hollow nanofibers are then obtained by

removing the core of SiO2 in the electrospun nanofibers through a chemical etching process

with NaOH, HF, or NH4HF2. Drawing inspiration from the spawning process of golden toads

and the unique structure of the eggs, Zhang et al. utilized this method and reported the

fabrication of lightweight, self-standing, hierarchical macroporous active carbon ber

(MACF) by calibrating the viscosity of the solution and adjusting the ratio of the SiO2

template to carbon precursor used. Pre-oxidizing and carbonizing the electrospun material

gives rise to the integration of macroporous spheres of carbon into the macroporous fiber,

with the resulting structure consisting of hollow macroscopic carbon spheres, approximately

180 nm in diameter, containing numerous surface cavities and interconnected pores, ranging

from 20 to 50 nm in diameter, as shown in Figure 7a and b. The pores originate from the

removal of SiO2, evaporation of volatile compounds during pyrolysis, and the regions

between neighboring template spheres. Another result reported by Cui and co-workers used

Tetraethyl orthosilicate (TEOS) directly as the source of the SiO2 template to obtain graphitic

carbon nanofibers resembling bamboo. These fibers showed well balanced porosity on the

micro-, meso-, and macro- scales.

21
The second commonly employed template based approach uses electrospun nanofibers as

the template, benefiting from the low cost, simple manipulation of their dimensions, and high

yield of nanofiber production. The negative surface charge on various polymer and carbon

nanofibers facilitates the interaction between the precursor and core materials modulated by

electrostatic forces, allowing them to act as scaffolds. Modification of the electrospun fibers

imbues them with the ability to draw inorganic precursor alkoxides and salts toward the

surface of the template and interacts with them via various physical and chemical interactions.

Once the nanofiber scaffolds are decorated with the inorganic shell, the templates are

removed leaving the compacted inorganic hollow shell intact. The versatility of this approach

enables the fabrication of various types of carbonaceous and polymeric hollow

nanostructures.[83, 84] An example of this nanofiber template synthesis found in the literature

involves the study of the hydrophilic surface of poly(styrene sulfonate) sodium nanofibers

(PSSNa) making it favorable for vapor deposition polymerization and subsequent

carbonization of the sheaths to occur. In another report, the abundance of the -CN groups

present in PAN was found to allow a dense coating of Co(OH)2 nanosheets formed from

strong coordination bonding via a solvothermal process in the presence of an alcoholic

Co(NO3)2 solution. This coating can then be transformed into low crystallinity C-doped

Co3O4 hollow nanofibers by subjecting the material to precarbonization to eliminate the

sacrificial PAN nanofibers (Figure 7c).[85]

Of the materials which hold promise in replacing the mature commercial graphite

technology in LIB applications, silicon is one of the most extensively studied. Kang and co-

workers published research on the synthesis of Si nanotubes, adopting a surface sol-gel

reaction using an electrospun organic nanowire scaffold - in this case, pyridine-like nanowires

obtained from the pyrolysis of PAN in air.[86] TEOS undergoes the sol-gel reaction on the

template surface resulting in well-defined hybrid nanowires. With the removal of the

pyridine-like scaffolds by air pyrolysis, reduction of the remaining material yields large scale
22
silicon nanotubes, as shown in Figure 7d-g. Besides the surface sol-gel technique, atomic

layer deposition (ALD), a self-limiting technique possessing the advantages of superior step

coverage, coverage homogeneity, and atomic scale control of the deposited layer thickness,

has been successfully used to prepare SnO2 nanotubes using electrospun polymer nanofiber

templates.[87]

Figure 7. a and b) SEM images hierarchical macroporous active carbon bers inset,
photograph; Reproduced with permission.[81] Copyright 2016, Wiley. c) SEM and the
magnified part of the C-doped Co3O4; Reproduced with permission.[85] Copyright 2016, Wiley.
d, e and f) SEM and g) TEM images of silica nanotubes after removing pyridine nanowires;
Reproduced with permission.[86] Copyright 2012, Wiley.

2.2.5 Pros and cons of each synthetic strategy

Although various approaches are widely employed to synthesize hollow nanofibers, each

currently used method has its pros and cons, as shown in Table 1. The coaxial electrospinning

technique is widely reported as the most versatile method for the fabrication of hollow

nanofibers and offers unprecedented exibility and control for tailoring the architecture of the

bers generated. This technique stands out from other methods, owing to its ability to easily

prepare one-dimensional core-shell nanoscale heterostructures on a large scale via a one-step

synthesis process. However, a coaxial spinneret is a complex piece of equipment that is not

commonly available. Furthermore, the selection of a solvent for the inner solution and the

23
control over the spinning parameters are problematic to obtain core-sheath bers of high

quality, especially when a water-soluble polymer is to be spun as the inner core. Compared

with coaxial electrospinning, emulsion electrospinning, which uses a common stainless steel

needle without any complicated coaxial assembly, is an attractive alternative because it is a

very simple, effective, top-down method to prepare hollow fibers of lengths of up to several

centimeters on a large scale after calcination. However, emulsion electrospinning is affected

by many factors, such as the miscibility/immiscibility of the solution pairs, viscosity ratio,

interfacial tension, and the fact that two different appropriate polymer systems with metal

alkoxides and metal salts able to induce phase separation are needed, renders severe

limitations to the choices of materials available for this technique. Furthermore, there is a

major concern over the pot-life of the prepared solutions, where the emulsion prepared is a

metastable system that could demulsify when stored for an extended time or even during the

electrospinning process. Hard-template methods are a universal and straightforward approach

to obtain hollow nanofibers with various morphologies. However, removing the templates by

thermal (sintering) or chemical (etching) means can be a very complicated and energy-

consuming procedure, not only complicating the process significantly, but also detrimentally

affects the quality of the final products to some extent. Moreover, the sacricial template

method is limited due to the insufcient availability of suitable raw materials to act as

compatible sacrificial templates. The calcination process is often a necessary step for

inorganic materials for two purposes, the removal of the polymeric phase by decomposition at

high temperature, and the conversion of the precursor component to produce the desired

inorganic materials by high temperature nucleation and growth. This technique is simple and

conducive for preparing hollow nanofibers with controlled diameters and shells by adjusting

the polymer and inorganic precursor concentrations. However, the mechanism of forming

hollow or complex hollow nanofibers induced by calcination are not well understood.

Although such a fabrication method for hollow nanostructures has been reported, it is quite
24
uncommon and not suitable to produce nanostructures of numerous metal oxides, thus needs

greater theoretical and experimental basis to be understood and utilized. In summary, based

on the pros and cons of each synthesis strategy, choosing a facile, efcient, and controlled

route to hollow nanofibers is desirable.

Table 1 Summaries of the main synthetic strategies for hollow nanofibers based on
electrospinning.

Strategy Advantages Disadvantages


Coaxial electrospinning Simple and versatile approach, coaxial spinneret, difcult to
uniform thickness obtain thick lms, complex
procedure to remove inner
material
Emulsion electrospinning Simple, single spinneret Limited precursor solvents,
hard to control
Heating process Easy process, complex tube hard to control, precise heat
structure process
Hard templates Controlled hollow or pore size, Limited templates, selected
high yield, functionalization solubility in precursor solution,
Limitation on ber dimensions

3. Lithium ion batteries

As one of the most popular rechargeable batteries, lithium-ion batteries (LIBs) have

undoubtedly become the dominant power source for portable electronic devices over the last

two decades, and expectedly have the potential for large-scale electric grids and to power

electric vehicle in the near future, mainly due to their unique advantages of no memory effect,

high energy density, environmental friendly and long lifetime. Nevertheless, the ever-growing

energy consumption is now pushing technological and scientific efforts to find better LIBs in

terms of enhanced safety, smaller size, lower cost, lighter weight, and longer lifespan.

Typically, a conventional LIBs stores and generates energy through reversible shuttling of Li +

ions between the positive and negative electrode across an electrolyte-filled insulating

separator. It would come as no surprise that the electrochemical performance of LIBs

ultimately depends on the development of electrode materials. So far, the most commonly

used cathode materials are lithium metal oxides with high positive redox potentials, such as

LiMn1-xMxO2, LiMn2-xMxO4, LiCo1-xMxO2, and polyanion-based compounds (LiMPO4, M =

Fe, Co, Ni, etc). Carbonaceous materials (normally graphite) with low cost and long lifespan

25
are used as anode materials to pair with the cathode materials. In the past decade, it is well

known that the exploration of high capacity lithium storage materials has fostered the

development of high-energy batteries.

Recently, considerable attention has devoted to various novel electrode materials with

unique nanostructures such as transition metal oxides and Sn/Si alloys, which possess a

theoretical specific capacity several times higher than that of conventional used carbon-based

anode materials. Unfortunately, their practical applications are far below the expectations

mainly owing to the rapid capacity degradation even at low current rates, the poor cycling

stability, and intrinsic low conductivity, which greatly hinder the employment of these active

materials in LIBs. Currently, one representative strategy to mitigate these problems is to

construct one-dimensional (1D) nanostructures with hollow or porous features.[88-90]

Generally, the enhanced performance mainly comes from the enlarged electrode-electrolyte

contact area and reduced pathways for charge transport, which will remarkably promote the

electrochemical reactions, better accommodate volume changes and withstand mechanical

strains. Thus, it is of great significance to develop effective approaches to fabricate high-

performance electrode materials with 1D hollow nanostructures. Compared with other

strategies, such as hydrothermal/solvothermal, template-directed growth, vapour-liquid-solid

method, electrospinning is one favourable and cost-effective technique that can be applied to

design 1D hollow nanofibers for advanced LIBs. Comprehensive summaries of developments

in LIBs can be seen in a numerous of review articles. In this contribution, instead of giving a

thorough survey, we will give an overview of the recent process and development of

electrospun hollow nanofibers and highlight their promising applications for LIBs.

3.1 Cathode materials

3.1.1 Electrospun lithium metal oxides hollow nanofibers cathode

As is known to all, the most representative lithium-insertion cathode materials is lithium

transition metal oxides, which mainly include LiMO2 (M often refers to Co, Mn, Ni) with
26
layered structure and LiM2O4 with spinel structure. In addition, recent advances in

electrospinning technique has witnessed a continuously development of electrospun hollow

nanofibers as cathodes for LIBs. Some typical examples are shown in Table 2.

Table 2 Summarize of some of representative electrospun hollow nanofibers as cathodes in


LIBs.
Electrospinning
Materials Techniques Structures Performance Ref.
precursors
-1
0.4Li2MnO30.6Li PAN/Li(Ac)2+Ni(Ac)2+Co Electrospinning and 199 mAh g after 50
Nanotubes [91]
Ni1/3Co1/3Mn1/3O2 (Ac)2+Mn(Ac)2/DMF thermal treatment cycles at 0.1 C
Mesoporou -1
Li1.2Mn0.54Ni0.13Co0. PAN/Li(Ac)2+Ni(Ac)2+Co Electrospinning and two- ~ 140 mAh g after
s [92]
13O2 (Ac)2+Mn(Ac)2/DMF steps thermal treatment 300 cycles at 3 C
nanotubes
PVP/Li(Ac)2+Ni(Ac)2+Co(
LiNi0.8Co0.1Mn0.1O2 Coaxial electrospinning Hollow 174 mAh g-1 after 50
Ac)2+Mn(Ac)2/citric [93]
-MgO and thermal treatment nanofibers cycles at 0.02 A g-1
acid/H2O
0.5Li2MnO3-
PVA/Li(Ac)2+Ni(Ac)2+Co( Electrospinning and Hollow ~ 200 mAh g-1 after 20
0.5LiNi1/3Co1/3Mn1/ [94]
Ac)2+Mn(Ac)2/H2O thermal treatment nanowires cycles at 0.01 A g-1
3O2
103 mAh g-1 after
PVP/Li(NO3)2+Mn(Ac)2/et Electrospinning and Hollow
LiMn2O4 1250 cycles at 0.15 A [95]
hanol thermal treatment nanofibers
g-1
PVA/Li(Ac)2+Ni(NO3)2+M Electrospinning and Porous 109 mAh g-1 after 50
LiNi0.5Mn1.5O4 [96]
n(Ac)2/ethanol thermal treatment nanofibers cycles at 0.15 A g-1
PMMA/LiH2PO4+Fe(NO3) Coaxial electrospinning
Hollow 135 mAh g-1 after 20
C-coated LiFePO4 3+Fe(SO3)2/DMF- and two-steps thermal [97]
nanofibers cycles at 2 C
PAN/DMF treatment
-1
PVP/LiOH+Mn(SO3)2+Fe Electrospinning and Hollow ~ 100 mAh g after 50
LiFe1-yMnyPO4/C [98]
(SO3)2/H2O thermal treatment nanofibers cycles at 4 C
Coaxial electrospinning -1
PVP/LiOH+Fe(NO3)3+H3 Hollow 138 mAh g after 0.2
LiFePO4/C/Ag and two-steps thermal [99]
PO4+AgNO3/DMF nanofibers C
treatment
PVP/vanadium Electrospinning and Porous 105 mAh g-1 after 250 [100
V2O5
acetylacetone/DMF thermal treatment nanotubes cycles at 2 A g-1 ]
C-encapsulated Electrospinning and Porous 241 mAh g-1 after 100 [101
PAN/VO(acac)2/DMF
V2O5 thermal treatment nanotubes cycles at 0.1 A g-1 ]

The classic LiCoO2 is the most successfully commercialized cathode materials for LIBs

due to its low self-discharge, high energy density, and excellent cycling stability. Despite a

high theoretical capacity (274 mAh g-1 assuming complete Li extraction), only half of this

value can be practically used because of the severe chemical and structural instability. Besides,

the slow solid state diffusion of Li+ ions also limits the performance of LIBs. One of the most

investigated strategies to address these drawbacks is to fabricate electrospun LiCoO2 hollow

nanofibers with large specific surface area, good mechanical stability, as well as short

electronic and ionic diffusion pathways.[102] Although LiCoO2 shows good electrochemical

properties, the recognized shortcomings, such as the limited abundance, toxicity, and high-

cost of cobalt, lead to the exploration for alternative cathode materials. Recently, an appealing

27
way is to fully or partially replace Co by using more environmentally friendly, cheaper, more

abundant elements (e.g. Mn or Ni). For example, electrospun LiNiO2 and LiNixM1-xO2 (M =

metal) hollow nanofibers have been reported.[91, 103, 104] Abdelkader and co-workers

reported that the hollow Li1.2Mn0.54Ni0.13Co0.13O2 electrode exhibited good cycling

performance and achieved a capacity retention of ~ 89% even after 300 cycles under a high

current density of 3 C, which could be related to the well-guided charge transfer kinetics,

short ionic diffusion pathways, and large effective contact area with electrolyte during the

cycling process.[38] Moreover, it is well accepted that coating with the metal oxides could

effectively improve the electrode performance. By utilizing coaxial electrospinning, hollow

LiNi0.8Co0.1Mn0.1O2-MgO coaxial nanofibers was fabricated by Jian and co-workers, and

presented a higher discharge capacity than those of corresponding solid nanofibers.[93] After

50 cycles, the discharge capacity of ~ 179 mAh g-1 could be delivered, corresponding to

89.2% of their initial capacities, as shown in Figure 8ac. These results indicated that the

electrochemical properties of the hollow LiNi0.8Co0.1Mn0.1O2-MgO coaxial nanofibers were

remarkably influenced by their specific morphology.

The spinel LiMn2O4 and its doped variants (LiNi0.5Mn1.5O4) have been considered as a one

of the promising substitute and are being established as cathode materials for high-power

electric device applications because of its high operating voltage (above 4.5 V vs. Li), low-

cost and nontoxicity.[94, 105] Its major drawback is the unsatisfactory long-term cycling

performance due to Mn dissolution in the electrolyte and Jahn-Teller distortion of Mn3+.[9]

One effective approach to increase the capacity retention is to focus on the nanostructured

spinels such as LiMn2O4 hollow nanofibers. Madhavi and co-workers reported that the porous

LiMn2O4 hollow nanofibers synthesized by electrospinning technique exhibited extraordinary

lithium storage performance with capacity retention of ~87% after 1250 cycles at the 1 C rate

(Figure 8d).[95] Very recently, the porous electrospun LiNi0.5Mn1.5O4 hollow nanofiber was

prepared and demonstrated exceptional performance.[96] Taking the advantages of the porous
28
hollow 1D architectures, the full cell based on LiNi0.5Mn1.5O4 cathode and TiO2 anode gave a

relatively high operating potential of ~ 2.8 V and achieved excellent cycling stability with a

capacity retention of ~ 86% even after 400 cycles (Figure 8e and f). The outstanding

performance elucidates that the nanostructuring of the electroactive materials in a 1D hollow

architecture is an efficient way to realize a high performance Li-ion configuration.

3.1.2 Electrospun polyanion-based compounds hollow nanofibers cathode

Polyanion-based compounds, LixMy(XO4)z (M = metal; X = P, S, Mo, etc.), is of interest

cathode materials for advanced LIBs. Among them, the class of phosphor-olivines with the

chemical formula LiMPO4 (M = Fe, Co, Ni, Mn) has become the focus of research efforts due

to their high discharge potential, relatively high specific capacity (170 mAh g-1), clean and

safety nature of the ion phosphate, as well as the favourable cycling stability.[97, 99, 106,

107] By virtue of the unique phosphate framework, such materials display remarkable

electrochemical and thermal stability. However, the intrinsic low electrical conductivity

makes a coating with an electronically conductive component (e.g., carbon) necessary.

Compared with LiFePO4, LiMnPO4 have similar theoretical capacities but higher operating

potential because of the larger redox potential of Mn3+/Mn2+ (~ 4.1 V) than Fe3+/Fe2+ (~ 3.7

V), enabling higher energy densities. In addition, both Fe and Mn are abundant and non-toxic,

which makes them particularly attractive for large-scale use in industry. However, the poor

capacity retention and sluggish kinetics of LiMnPO4 limits their application in LIBs.

Replacing some of the Mn with Fe to form mixed metal olivine compounds provides the

possibility to enhance the kinetics of the Li-ion transfer while maintaining higher voltages. In

view of the above, a representative example is LiFe1-yMnyPO4/C composite electrode, which

was prepared by innovative electrospinning method.[98] The electrospun LiFe1-yMnyPO4/C

hollow nanofiber composite electrode with an optimal Mn : Fe ratio manifests an impressive

electrochemical performance as cathode for LIBs. Moreover, it is found that a substitution of

up to half of Fe with Mn led to a higher energy density and reversibility. While increasing the
29
content of Mn more than half, a signicant lattice distortion in the Mn3+/Mn2+ transition

would occur, thereby resulting in poor electrochemical performance.

3.1.3 Electrospun non-lithium metal oxides hollow nanofibers cathode

Vanadium pentoxide (V2O5) is well-accepted as one of the most encouraging cathode

substitutions. Benefiting from the layered structure, large quantities of electrons can be

captured during the lithiation/delithiation process, thus resulting in high theoretical capacity

(~ 400 mAh g-1). Nevertheless, the low electronic conductivity, structural instability, as well

as slow lithium diffusion rate upon cycling process strongly restrain the realistic use of V 2O5

as cathode in LIBs. To resolve these problems, the use of electrospinning to fabricate V2O5

hollow nanofibers offers a new avenue.[108] For instance, zhang and co-workers have

controllable synthesized V2O5 nanotubes, hierarchical V2O5 nanofibers, and single-crystalline

V2O5 nanobelts by simply varying the annealing temperature.[100] More importantly, as

cathode materials for LIBs, the V2O5 nanotubes achieve a high energy density 201 W h kg-1,

whilst the power density maintained as high as 40.2 kW kg-1 (Figure 8g). Moreover, carbon

coating is recognized as an efficient way to enhance the electrical conductivity of active

materials. With this in mind, the carbon-encapsulated V2O5 hollow structure was fabricated

through electrospinning and postcalination.[101] By optimizing the postcalcination conditions,

the unique V2O5 architecture which included interior void spaces, well-defined pores, and a

uniform carbon layer, can be obtained. Compared with commercial V2O5, electrospun carbon-

encapsulated V2O5 hollow nanofibers demonstrated excellent cycling stability (241 mAh g-1

after 100 cycles) and improved high-rate performance (155 mAh g-1 at 1 A g-1) (Figure 8h and

i). The excellent electrochemical performance can be clarified by unique 1D nanostructure,

which facilitated Li-ion diffusion and improved electrolyte infiltration in the electrode. All of

these results suggested that further optimization or improvement of the electrochemical

performance of electrode materials could be achieved by the rational design of 1D

architecture with hollow structures.


30
Figure 8. a) The initial charge discharge curves and b) capacity retention versus cycle number
plots of three ber electrodes at a current density of 20.0 mA g-1; c) SEM picture of the
hollow coaxial bers after 50 cyclings; Reproduced with permission.[93] Copyright 2008,
American Chemical Society. d) long-term cyclability of LiMn2O4 hollow nanofibers recorded
between 3.54.3 V at a current density of 150 mA g-1; inset: the SEM image of LiMn2O4
hollow nanofibers; Reproduced with permission.[95] Copyright 2013, The Royal Society of
Chemistry. e) plots of discharge capacity vs. cycle number and f) the cycling performance for
LiFePO4/TiO2 cells cycled between 0.9 and 2.5 V at various current densities from 0.051 A
g-1 in room temperature; Reproduced with permission.[96] Copyright 2014, The Royal Society
of Chemistry. g) rate performance of the porous V2O5 nanotubes at different current densities
in the range 2.54.0 V; inset: Ragone plots; Reproduced with permission.[100] Copyright 2012,
Wiley. h) Cycling durability of the chargedischarge capacities of commercial
V2O5,C/HPV2O5-10, C/HPV2O5-30, and C/HPV2O5-60 up to 100 cycles; (i) The high-rate
performance at current densities of 50, 100, 300, 500, 700, 1000, and 50 mA g-1; Reproduced
with permission.[101] Copyright 2016, American Chemical Society.

3.2 Anode materials

31
Electrospun hollow nanofibers have been widely employed as functional nanomaterials in

many fields due to their unique chemical and physical properties. This section will summarize

the recent advances in the area of electrospun hollow nanofibers as anode materials for LIBs,

which will greatly help to understand the primary advantages of the use of electrospun hollow

nanofibers for energy storage devices. The silicon or tin-based alloys as well as the TMOs-

based hollow nanofibers are introduced in this section. Table 3 shows some typical examples

of such electrospun nanomaterials as anode materials for LIBs.

Table 3 Summarize of some of representative electrospun hollow nanofibers as anodes in


LIBs.
Electrospinning
Materials
precursors
Techniques Structures Performance Ref.
Electrospinnig and sol-
Carbon coated ~ 800 mAh g-1 after 90
PAN/DMF-TEOS gel reaction, Mg nanotubes [86]
Si cycles at 0.4 A g-1
reduction
PMMA/Si/acetone+DM Core/shell 1% capacity loss after [109
Si@C Coaxial electrospinnig
F/PAN/DMF nanofibers 300 cycles at 3 C ]
Electrospinning and Particles
872 mAh g-1 after 200 [110
Si@CT PVP/Si-TEOS/ethanol thermal treatment (HF encapsulated in
cycles at 1 A g-1 ]
etching) nanotubes
Particles
Electrospinning and
embedded in 1104 mAh g-1 after 100 [111
Si@PCNF PAN/Si-SiO2/DMF thermal treatment (HF
porous cycles at 0.5 A g-1 ]
etching)
nanofibers
Electrospinning and
PMMA-PAN/Tin Multichannel 774 mAh g-1 after 140
Sn@C two-steps thermal [71]
octoate/DMF nanotubes cycles at 0.5 C
treatment
Bamboo-like
Mineral oil/TBT- Coaxial 737 mAh g-1 after 200
Sn@C hollow [66]
PAN/DMF electrospinning cycles at 0.5 C
nanofibers
Electrospinning and Particles filled
PMMA- 410 mAh g-1 after 200 [112
SnSb/C two-steps thermal porous
PAN/SnO2+Sb2O5/DMF cycles at 0.5 A g-1 ]
treatment nanofibers
Electrospinning and
PVP/ TBT/acetic acid/ Fiber-in-tube 177 mAh g-1 after 1000
TiO2 multi-steps thermal [76]
ethanol structures cycles at 0.2 A g-1
treatment
Multichannel
PVP-PAN/ Electrospinning and
TiO2 hollow 156 mAh g-1 at 20 C [73]
Ti(OiPr)4/DMF thermal treatment
nanofibers
Coaxial
PVP/Ti(OiPr)4/ethanol/ electrospinning and Hollow 156 mAh g-1 after 100 [113
N-doped TiO2
mineral oil thermal treatment with nanofibers cycles at 0.2 C ]
NH3
PVP/Zn(NO3)2+SnCl2/D Electrospinning and Porous 584 mAh g-1 after 45 [114
SnO2/ZnO
MF+ethanol thermal treatment nanotubes cycles at 0.168 A g-1 ]
C&N co-doped Electrospinning with 1648 mAh g-1 after 100 [115
PVP/SnCl2/DMF Nanotubes
SnO2 Ppy coating cycles at 0.2 A g-1 ]
Nanofibers
PVP/Sn(oct)2/acetic Electrospinning with 663 mAh g-1 after 250 [116
SnOx comprising yolk-
acid/ethanol kirkendall reaction cycles at 2 A g-1 ]
shell SnOx
Electrospinning and
PVP/Sn(oct)2/acetic thermal treatment Fiber-in tube 640 mAh g-1 after 300
SnO2-C [77]
acid/ethanol based on phase structures cycles at 1 A g-1
seperation
Bubble-
Electrospinning with 812 mAh g-1 after 300 [117
Fe2O3-C PAN/Fe(acac)3/DMF Nanorod-
kirkendall reaction cycles at 1 A g-1 ]
structure
Electrospinning with
PVP/Fe(acetylacetonat Porous 987 mAh g-1 after 250 [118
Fe2O3 varied metal precursor
e)3/DMF nanotubes cycles at 0.2 A g-1 ]
concentration
PAN/mineral Electrospinning and 600 mAh g-1 after 100 [119
Fe3O4/C nanotubes
oil/Fe(acetylacetonate)3 thermal treatment cycles at 0.15 C ]

32
/DMF
Electrospinning and Hollow 978 mAh g-1 after 150 [120
C-coated Fe3O4 PVP/Fe(NO3)3/DMF
hydrothermal nanofibers cycles at 1 A g-1 ]
Coaxial Porous 1826 mAh g-1 after 100 [121
Co3O4 PVP/Co(Ac)2/DMF
electrospinning nanotubes cycles at 0.3 A g-1 ]
Electrospinning and Hierarchical
PAN/Co(Ac)2/DMF, 1281 mAh g-1 after 200 [122
Co3O4/CNT multi-step thermal tubular
Co(Ac)2-PAN@ZIF-67 cycles at 0.1 A g-1 ]
treatment structures
Electrospinning and
PAN/DMF with
hydrothermal with Hollow 1121 mAh g-1 after 100
C-doped Co3O4 Co(NO3)2 for [85]
multi-step thermal nanofibers cycles at 0.2 A g-1
hydrothermal
treatment
Electrospinning and
PAN/DMF with
C&N co-doped hydrothermal with Hollow 1501 mAh g-1 after 60 [123
Co(NO3)2 for
Co3O4 ammonia thermal nanofibers cycles at 0.2 A g-1 ]
solvothermal
treatment
PAN/DMF with KMnO4 Electrospinning and Nanotubes with 875 mAh g-1 after 100 [124
C-coated Mn2O3
solution treatment microwave treatment porous wall cycles at 0.1 A g-1 ]
PVP/ethanol- Electrospinning and Porous 565 mAh g-1 after 50 [125
CaSnO3
Ca(NO3)2+SnCl2/DMF thermal treatment nanotubes cycles at 0.06 A g-1 ]
Electrospinning with
PAN/Ca(NO3)2+SnCl2/D Eggroll-like 648 mAh g-1 after 50 [126
CaSnO3 varied metal precursor
MF nanotubes cycles at 0.06 A g-1 ]
concentration
Nanofibers
Electrospinning and
CoFe2O4@onion PAN/Co(Ac)2+ composed of 930 mAh g-1 after 1000 [127
multi-step thermal
-like C Fe(acac)3/DMF hollow cycles at 1 A g-1 ]
treatment
nanoparticels
PVP/
Electrospinning and 1228 mAh g-1 after 160 [128
CoFe2O4 Co(NO3)2+Fe(acetylace Nanotubes
thermal treatment cycles at 0.05 A g-1 ]
tonate)3/DMF
PVP/PAN/
Electrospinning and 816 mAh g-1 after 50 [129
CuFe2O4 Cu(NO3)2+Fe(NO3)2/DM Nanotubes
thermal treatment cycles at 0.2 A g-1 ]
F
PVP/Zn(NO3)2+Co(NO3 Electrospinning and Porous 1454 mAh g-1 after 30 [130
ZnCo2O4
)2/ethanol thermal treatment nanotubes cycles at 0.1 A g-1 ]
PVP/Mn(Ac)2+Co(Ac)2/ Electrospinning and Hollow ~ 500 mAh g-1 after 100
CoMn2O4 [53]
methanol thermal treatment nanofibers cycles at 0.4 A g-1
PVA with different
Electrospinning and
molecular Tube-in-tube 565 mAh g-1 after 500 [131
CoMn2O4 multi-steps thermal
weight/Mn(Ac)2+Co(Ac) nanotubes cycles at 2 A g-1 ]
treatment
2/H2O
Electrospinning and Double wall
PVP/Mn(Ac)2+Co(Ac)2/ 997 mAh g-1 after 50 [132
MnCo2O4 thermal treatment with hollow
DMF cycles at 0.2 C ]
varied heating rate nanofibers
PVP/Ni(NO3)2+Co(NO3) Electrospinning and 732 mAh g-1 after 200 [133
Au@NiCo2O4 Nanotubes
2+HAuCl4/DMF thermal treatment cycles at 0.1 C ]

3.2.1 Electrospun carbon-based materials hollow nanofibers anode

In terms of their unique structure features, including high specic surface area and extremely

long ber length, 1D carbon nanomaterials have greatly attracted tremendous attention.[134]

Particularly, the development of continuous 1D carbon nanomaterials by a combination of

electrospinning and subsequent thermal treatment has various advantages such as low cost,

ease of processing, and environmental benignity. Moreover, they show improved

electrochemical performance as electrode for LIBs.[135] However, the lithium storage

capability and cycling stability of these carbon materials are still limited by volume expansion

and less additional sites for Li+ storage during the repeated lithium lithiation/delithiation

process. Recently, hollow electrospun nanofibers may prove to be especially beneficial for
33
application in energy storage devices, since the hollow structure could provide more

additional sites for Li+ storage and act as a buffer to accommodate the volume expansion.

To fabricate hollow nanofibers, coaxial electrospinning of core/shell precursor nanofibers

followed by subsequent post treatment like calcination or washing were employed. Different

polymers have been introduced to prepare core/shell nanofibers such as PAN/PMMA,

PVP/PVDF, PAN/PSSNa, and PAN/SAN.[136-140] Among carbon precursors, PAN and

SAN have been widely used to create hollow structures. This is because SAN is a very

suitable materials as the sacrificial core due to its immiscibility with PAN and good thermal

sustainability to prevent PAN shell from shrinking during thermal treatment procedure. For

instance, Yu and co-workers synthesized various hollow nanofibers using coaxial

electrospinning of SAN core and PAN shell combined with subsequent thermal

treatments.[141] The effect of the carbonization temperature on their microstructure and

electrochemical properties was systematically studied. With the increases of carbonization

temperature, both crystallite length and thickness showed a signicantly increases while the

initial irreversible capacity decreases. It is found that the initial capacities and initial

irreversible capacity strongly associated with the microstructures. A quite stable cycling

performance with coulombic efciencies exceed 99.3% (99.51% for 1600 C) could be

observed. These predictable configuration and electrochemical properties of electrospun

hollow carbon nanofibers gave us important insight for the design of anode materials with

novel architectures such as Sn or Si encapsulated electrospun hollow carbon nanofibers.

Moreover, it has been greatly investigated that the electrochemical performance of carbon

materials with desirable pores could be improved. In general, the pores could increase the

accessibility between the electrode and electrolyte, relieve the stress induced by volume

variation during lithiation/delithiation process, and facilitate charge transfer at the interface of

electrode and electrolyte. As for that, Yu and co-workers developed a facile preparation

technique to create porous electrospun hollow carbon nanofibers and investigated the
34
influence of pore on their electrochemical performance.[142] Porous electrospun hollow

carbon nanofibers were fabricated by the coaxial electrospinning of a sacrificial SAN core

solution and an emulsified PAN shell solution containing sacrificial SAN islands for pore

formation. According to the HRTEM images, the relationship between the pore size in the

porous hollow carbon nanofibers and the size of the SAN/DMF islands in the emulsion was

elucidated, proving the possibility of manufacturing various porous hollow carbon nanofibers

with controlled pore sizes and volumes. By virtue of the introduced pores in hollow carbon

nanofibers, their initial capacity and capacity retention were enhanced signicantly to 1003

mAh g-1 and 61.8%, respectively, compared to those (653 mAh g-1 and 53.9%) of nonporous

hollow carbon nanofibers. The increased pore size and expanded graphene layers are believed

to facilitate lithium insertion/extraction behaviour, leading to the exceptional electrochemical

performance. Furthermore, to control the size of micropore and reduce the initial irreversible

capacity, a chemical vapor depositon (CVD) process was introduced between the stabilization

and carbonization procedure, thus inducing the deposition of carbon atoms in the pores.[143]

The initial coulombic efciencies of the hollow carbon nanofibers were signicantly

increased by using CVD process. The mechanism behind such enhancement was revealed by

morphological, microstructural, surface, and electrochemical analyses, suggesting that control

of the pores in hollow carbon nanofibers is an effective way to improve the electrochemical

performance of hollow carbon nanofibers.

Although the porous electrospun hollow nanofibers displayed promising advantages as

electrode materials for LIBs, the graphitization degree of these structures is poor, thereby

resulting in low electrical conductivities. To overcome these problems, a novel 1D hybrid

material of amorphous carbon nanostructures combined with graphitized hollow carbon

spheres (ACNHGCNs) have been developed by using electrospinning technique followed by

subsequent calcination and acid treatment.[144] The TEM images in Figure 9a and the isnet

show that hollow graphitic carbon nano-spheres with a diameter of ~15 nm were formed on
35
the hollow carbon fibers. The incorporation of graphitized hollow carbon spheres in the

electrospun carbon 1D architectures could significantly increase the conductivity of the

electrode materials, which facilitate the transport and collection of electrons during cycling

process. As anode materials for LIBs, such unique structures exhibited reversible capacities of

960, 748, 573, 456, 400 and 330 mAh g-1 at current densities of 0.05, 0.1, 0.25, 0.5, 1.85 and

3.7 A g-1, which are higher than those of CNF, CNT, raphene sheets (GNS), GNS/CNF,

natural graphite and hollow carbon nanospheres (Figure 9b). Furthermore, ACNHGCNs

exhibited good cyclic performance and maintained a reversible capacity of ~965 mAh g-1 after

100 cycles (Figure 9c). The lower resistance shown in Electrochemical impedance

spectroscopy (EIS) measurements inidicated that the ACNHGCN electrodes possess a higher

electrical conductivity and a more rapid charge-transfer reaction for lithium ion insertion and

extraction (Figure 9d). These show that ACNHGCNs have a good cycling stability at both

low and high current density and outstanding rate performance. In addition, zhou and co-

workers successfully employed a triple-coaxial electrospinning technique to fabricate a

unique nanostructures containing of amorphous carbon nanotubes decorated with hollow

graphitic carbon nanospheres.[145] Due to the highly graphinized hollow carbon nanospheres

that have been well-dispersed throughout the whole carbon nanotubes, the special

nanoarchitectures possess high electronic conductivity and facilitate fast Li+ diffusion inside

the electrode so that very good rate capability can be achieved, which render them good

candidates as anode for high energy and high power LIBs.

Besides, chemical doping of carbon materials with heteroatoms, such as phosphorus, sulfur,

boron, and nitrogen,[146-148] is considered beneficial to modify their electronic and chemical

properties. Among the various dopants, nitrogen has been regarded as the most fascinating

element. Specifically, nitrogen not only possess a comparable atomic size with carbon but

also can easily bond with neighboring carbon atoms due to its electronegativity (3.04)

compared with that of carbon (2.55). Moreover, nitrogen doping can create defect sites that is
36
favourable for lithium storage, thereby rendering large capacity. Furthermore, the electronic

conductivity of carbon materials could be greatly enhanced because of the induced n-type

conducting behaviors by N doping. Therefore, it is obvious that electrospun hollow carbon

nanofibers with nitrogen incorporation are highly desirable for achieving excellent LIB

performance, including high lithium storage capacity, excellent rate capability and long-term

cycling performance. Based on this concept, a few previous reports have shown that nitrogen-

doped electrospun hollow carbon nanofibers could act as excellent anode materials for LIBs.

As one representative example, kang and co-workers reported the successful fabrication of N-

enriched electrospun carbon nanofibers by using low cost melamine and polyacrylonitrile as

precursors through electrospinning combined with subsequent carbonization and NH3

treatments.[149] Such unique configuration presented ultrahigh lithium storage capacity,

excellent cycling stability with a reversible capacity of as high as 1323 mAh g-1 at a current

density of 50 mA g-1. These attractive features make it very promising anode materials for

advanced LIBs.

37
Figure 9. (a) TEM and HRTEM (inset) images of the composite nanotubes after the
dissolution of nickel; (b) cycling performance of amorphous carbon nanotubes decorated with
hollow graphitic carbon nanospheres electrodes at different current densities; (c) their rate
performance and (d) Nyquist plots of ACNHGCNs and amorphous carbon nanober
electrodes after 10 cycles at a rate of 0.05 A g-1. Reproduced with permission.[145] Copyright
2012, The Royal Society of Chemistry.

3.2.2 Electrospun lithium alloys hollow nanofibers anode

Some main-group elements such as Si, Sb, Sn, Al, Ge, and Zn that can alloy with lithium at a

low potential, have been investigated for ages. Typically, Si and Sn have been widely

explored as electrode for LIBs due to its high theoretical capacity (4200 and 993 mAh g-1,

respectively). Unfortunately, huge volume changes (up to 400% for Si anode) occur during

the electrochemical lithiation/delithiation process, leading to cracking and eventual

pulverization of active materials from the unstable solid electrolyte interface (SEI) layer and

current collector. Another limitation associated with the Si anode is the poor electronic

conductivity, which impedes rapid lithiation. To overcome these hurdles, substantial effort

has been focused on nanostructuring of Si in the form of nanofibers, nanotubes, and nanorods,

which are more exible to facilitate electron transportation and endure the strain induced by

volume variations because of their axially directed structure and empty space. As expected,

such structures have demonstrated superior cycling stability and rate capability compared to

bulk Si. For example, silicon nanotubes have been synthesized by a facile surface sol-gel

reaction on electrospun organic nanowires followed by a simple magnesium reduction.[86]

Electrospun PAN nanowires were used as template for a sol-gel reaction of tetraethyl

orthosilicate (TEOS) on the surface. The elimination of PAN scaffolds leaded to the

formation of silica nanotubes. Then, silicon nanotubes could be obtained through a

magnesium reduction route without destruction of whole tube-like architectures. Finally,

graphitic carbon was uniformly coated on silicon nanotubes to further improve its

conductivity. By virtue of the hollowness of the nanotubes, the cycling stability of fabricated
38
Si nanostructures outperform that of commercial silicon nanoparticles. Unlike the

magnesiothermic reduction, a unique Si-based multicomponent was also synthesized by using

coaxial electrospinning and a subsequent aluminothermic reduction reaction, which enabled

successfully transformation of silica to Si as well as the formation of multicomponents (Si,

alumina, and titanium silicide).[150] The Si-based multicomponent nanotubular structures

exhibited excellent electrochemical properties such as enhanced cycling performance (~ 75%

after 280 cycles), high rate capability (650 mAh g-1 at a rate of 5C), and extremely suppressed

volume changes (~ 14% after 100 cycles) compared with bare Si nanotubes. Nevertheless, the

direct preparation of Si with special 1D nanostructures usually suffers from low yield.

Meanwhile, the procedures involved normally costly and complex because of the lack of

precursors which could be easily converted to Si. Therefore, the use of commercially

available Si nanoparticles would be a more straightforward route. A representative approach

is to incorporate the Si nanoparticles into carbon nanofibers through electrospinning technique

followed by subsequent thermal treatment.

Numerous research has revealed that the Si/carbon composite is a promising candidate for

advanced LIBs. In this regard, Choi and co-workers reported that coreshell Si NPs/CNFs

could be easily prepared by employing a dual nozzle electrospinning technique.[109] Two

electrospinning solutions was used. Si nanoparticles and PMMA dissolved in the co-solvent

of DMF and acetone were loaded into the core channel of the nozzle, whereas PAN dissolved

in DMF was used for the carbon shell generation. The formation of void space in the core

plays key role to buffer the volume variation of Si during cycling performance. Besides, the

miscibility between adjacent solutions is one critical factor to the multiple coaxial

electrospinning. By using acetone, they successfully prevent the solution mixture between the

core and shell. When evaluated as anode for LIBs, the unique structure with Si nanoparticles

encapsulated in the robust carbon shell exhibited outstanding cycling stability (~ 800 mA h g-1

after 300 cycles at 3C) and an excellent rate capability (~ 721 mA h g-1 at 12 C). The
39
electrospun coreshell Si@C nanofibers demonstrated a new design concept for robust and

scalable Si anode. Similarly, the Si core/C shell nanofibers were prepared through coaxial

electrospinning of Si nanoparticles in SAN core solution and PAN shell solution followed by

calcination.[151] The in situ observation of the contact-lithiation found that the C shell and Si

core reacted with Li+ in an independent and stable manner during cycling process. The

incorporation of of Si (9.9 wt.%) into hollow carbon nanofibers ensured the enhancement of

reversible capacity with capacity retention of ~ 92% after 50 cycles. Also, yu and co-workers

demonstrated the synthesis of Si (core)-hollow carbon nanofiber (sheath) nanocomposite

through a coaxial electrospinning technique.[152] In this work, Si and mineral oil was

selected as the core precursor, while PAN was used as sheath precursor. After thermal

treatment, Si-hollow carbon nanofibers can be developed. In terms of electrochemical

performance, this composite electrode delivered a high reversible capacity of 1300 mA h g-1

even after 80 cycles at 0.5 C. Even at relatively high rate of 3 C, a reversible discharge

capacity as high as 700 mA h g-1 still can be achieved, suggesting the promising potential of

SiC composite as anode for LIBs. Beyond that, coaxial electrospinning has proved itself as a

powerful technique to prepare nanomaterials with hollow coreshell architectures.

Apart from the core-shell structures for Si/C nanocomposites, tremendous efforts have been

devoted to create additional void space between the neighbouring Si nanoparticles, which was

expected to further accommodate the volume variation of Si during lithiation/delithiation

process. Cui and co-workers successfully confined Si nanoparticles in hollow carbon fibres,

which engineered hollow space between the nanoparticles and the carbon fibres wall to

alleviate the volume expansion, as shown in Figure 10a and b.[110] Firstly, the continuous

silicon dioxide nanofibers with embedded Si nanoparticles could be generated by

electrospinning. After conformal conducting thin carbon coating, HF aqueous solution was

used to remove SiO2 and leave Si to produce void space around Si nanoparticles. Such unique

structures could efficiently overcome these problems for Si electrode, including unstable SEI
40
film, mechanical instability and current collector contact issues. The fabricated Si electrode

illustrated a high specific capacity (~ 1000 mAh g-1 based on the total mass), which is 3 times

higher than that of traditional graphitic anodes (370 mAh g-1). The SiNP@CT electrode

showed superior cycling stability and the discharge capacity retention after 50, 100, and 200

cycles was 95, 95, and 90%, respectively (Figure 10c). The superior electrochemical

performance of SiNP@CT electrode is due to the stable coating layer and the stable SEI. It

can be seen that the carbon tubes remained continuous and unbroken after cycles and a stable

solid electrolyte interface (SEI) layer formed outside carbon tubes (inset of Figure 10c).

Besides, Guo and co-workers designed an elegant structure with encapsulated Si nanoparticles

in porous carbon nanofibers (Si@PCNF).[111] The coating of SiOx on the surface of Si

nanoparticles was introduced to develop a homogeneous suspension of Si@SiOx

nanoparticles in PAN/DMF solution for electrospinning, and function as a solid source for

hydrogen uoride (HF) etching to generate empty space between Si nanoparticles and carbon

shell. Thus, such favourable structures not only promote the electronic transportation from

carbon shell to Si nanoparticles but also provided sufcient empty space to cushion the

volume changes of Si during alloying/dealloying process. As a consequence, this novel

Si@PCNF demonstrate enhanced cycling stability (~ 1104 mA g h-1 after 100 cycles at

current density of 0.5 A g-1) as well as encouraging rate capabilities (~ 485 mAh g-1 at current

density of 10 A g-1), showing great potential as anode material for LIBs.

Thus far, many studies mainly focused on a simple incorporation of crystalline Si

nanoparticles into electrospun carbon nanofibers. However, poor electrical conductivity

caused by the high content of insulating Si remains a major challenge. To address the problem,

a novel Si/C hybrid system with Si encapsulated in hollow graphitized carbon nanofibers

derived from polydopamine (PDA) has been reported.[153] Benefiting from its multi-layered

structure, N-doping and nanometer-scale thickness, C-PDA coating could significantly

contribute to the improvement of electrochemical performance. As expected, the C-PDA-Si


41
nanofibers showed much better cycling performance and rate capability (500 mAh g-1 at

current density of 5 A g-1) compared to conventional C/Si nanofibers, which could be related

to excellent electrical conductivity of the carbonized PDA (C-PDA) shell and the high

electrochemical activity of C-PDA, the hollow nature of the nanobers, as well as the

buering eect of the remaining PAN-derived carbon around the Si nanoparticles. Moreover,

moon and co-workers reported the salami-like core-sheath composites consisting of Si

nanoparticles with indium tin oxide (ITO) coating.[154] Conductive ITO shell could offer

conduction pathways for electrons, which is important to enhance the electrochemical

performance. These engineered composites ensured the unique anode with a good capacity

retention of 95.5%, and outstanding rate capability with a retention of 75.3% from 1 to 12 C.

It is well accepted that not only the intrinsic conductivity for the carbon-based

nanocomposite but also the conductive pathway among active materials have signicant effect

on the electrochemical properties of the electrode in LIBs. Thus, the conductivity among

carbon nanofibers is also critical for the electron transfer in the electrode. In this respect,

concentrically tri-layered (C-core/Si-medium/C-shell) nanofibers, were fabricated by triple

coaxial electrospinning.[155] The tri-layered nanoarchitectures exhibited outstanding specific

capacity, excellent rate capability, and stable cycling behaviour, which could be associated to

the continuous C-core. These enhancements indicated that the additional conductive C-core

significantly facilitated the participation of numerous Si nanoparticles in the

alloying/dealloying reaction. It has been demonstrated that the cross-linked structure could

provide more access sites for electron transfer in the bre network, resulting in significantly

enhanced rate capability. Li and co-workers prepared Si@IHCFs composite nanofibers with

Si nanoparticles embedded in a honeycomb-like carbon framework.[156] The well-dened

Si@IHCFs showed a reversible capacity of 903 mAh g-1 and a capacity retention of 89% after

100 cycles at a rate of 0.2 A g-1. With the current density gradually increasing to 2.0 A g-1, the

electrode shows a specic capacity of 743 mAh g-1, exhibiting superior rate capability
42
compared to the Si/C bers without connected structures. The hierarchical core-shell structure

and cross-linked network of the Si/C composite rendered it the remarkable electrochemical

properties. Inspired by the development of electrospinning and electrospray techniques, a

flexible 3D Si/C fiber paper electrode have been fabricated through simultaneously

electrospraying nano-Si-PAN clusters and electrospinning PAN fibers followed by

carbonization.[157] The flexible electrode manifested a high capacity and exceptional rate

capability. In addition to the energy related application, the combined technology also can be

considered to use in more research fields.

Similar with Si-based electrode, the huge volume changes and aggregation of Sn

nanoparitcles also greatly hampered their implementation in LIBs. To tackle these limitations,

an interesting approach is to confine Sn nanoparticles in conductive carbon to avoid the

aggregation during alloying/dealloying process. In particular, Sn nanoparticles encapsulated

in porous multichannel carbon microtubes was fabricated by single-nozzle electrospinning

method using co-polymers (PMMA and PAN) as the template.[71] The Sn nanoparticles were

uniformly dispersed in the carbon framework with sufficient space to buffer the volume

changes during cycling process. This material displayed good electrochemical performance in

terms of reversible capacities, rate capability, and cycling stability, which can be attributed to

its unique carbon scaffold. Moreover, the evaporation of Sn at high temperature should be

considered in the synthesis routes. Guo and co-workers reported a novel 3D Sn/C core/shell

nanostructures with cross-linked channels via electrospinning followed by glucose-

hydrothermal process and subsequent thermal treatment.[158] The evaporation loss of Sn at

high temperature could be effectively reduced due to the protective function of the carbon

coating. In addition, many pores and hollow carbon protuberances can be observed in the

composite, and especially, channel-like structures formed by the conglutinated and cross-

stacked Sn/C hybrid nanobers, which could greatly improve lithium storage properties,

facilitate lithium ion diffusion, and enlarge the contact area between active materials and
43
electrolyte. Another approach to enhance electrochemical performance is to confine Sn

nanoparticles in hollow carbon matrix, which provides more buffer zone to alleviate the

volume variation of reactants. Recently, Sn@carbon nanoparticles in bamboo-like hollow

carbon nanofibers has been synthesized via coaxial electrospinning, which employ PAN/DMF

for the shell and tributyltin (TBT)/mineral oil for the core, as shown in Figure 10d and e.[66]

The particular Sn@carbon composite not only has high Sn content (around 70 wt% Sn) but

also offer enough void space to compensate large volumes fluctuation, thus limiting

aggregation of the Sn nanoparticles. This composite deliver excellent cyclability of 737 mAh

g-1 after 200 cycles at 0.5 C (Figure 10f). The rate capability of the Sn/C composite electrode

in Figure 10g delivered a rate capacity of about 650 mAh g-1, when first cycled at 1 C, 550

mAh g-1 at 3 C, 480 mAh g-1 at 5 C, and finally back to 610 mAh g-1 at 1 C again. Also, Sn

embedded porous multichannel carbon nanotubes (Sn-PMCMT) were prepared via co-

electrospun tributylphenyl-tin/polyacrylonitrile-polymethymethacrylate (TBPT/PAN-PMMA)

and carbonization.[159] The overall features endow the Sn-PMCMT composite with

exceptional cycling stability (340 mAh g-1 after 400 cycles at 1 A g-1). Considering that the

size of the Sn nanoparticles is also critical for the stability of LIBs. Dun and co-workers

reported the synthesis of Sn quantum dots (QDs) uniformly dispersed in N-doped carbon

nanofibers.[160] This newly designed Sn QD@CNF highlighted the importance of the

combination of ultrasmall Sn QDs and CNFs matrices to enhance the electrochemical

performance.

In addition to Si and Sn nanocomposites, Zhang and co-workers reported the preparation of

SnSb nanoparticles in porous/hollow carbon nanofiber.[112, 161] They suggested that the

choice of active particle also plays important role to create small pores around the active

particles. Meanwhile, the unique structures guaranteed SnSb/C composite with good cycling

stability and rate capability. Compared with the above mentioned alloy anodes, Ag is also an

attractive choice due to its good electrical conductivity and high diffusivity for lithium,
44
thereby leading to high rate performance. To achieve good cycling stability, a free standing

hybrid electrode consisted of Ag nanoparticles dispersed in hollow carbon nanofibers have

been synthesized by coaxial electrospinning followed by thermal treatment.[162] The

encapsulating of Ag nanoparticles in hollow carbon nanofibers acts as a dual role, which not

only improves electronic conductivity leading to faster charge\discharge kinetics, but also

reacts with lithium to contribute additional lithium storage sites by forming AgLix.

Figure 10. (a) Schematic outlining the material fabrication process. Si nanoparticles in SiO2
nanofibers were first prepared by electrospinning. After carbon coating and removal of SiO 2
core, the SiNP@CT structure was obtained. (b) TEM images of synthesized samples. Lower
inset shows TEM image with higher magnification; upper inset shows SAED pattern of the
sample. (c) Chargedischarge cycling test of SiNP@CT electrodes at a current density of 1A
g-1, showing 10% loss after 200 cycles. SEM image (left) of SiNP@CT electrode after 200
electrochemical samples. A thin layer of SEI can be observed outside tubes. SEM image
(right) of the same cycled SiNP@CT electrode after dissolving the SEI layer by HCl washing.
Reproduced with permission.[110] Copyright 2012, American Chemical Society. (d)
Preparation of Sn@carbon nanoparticles encapsulated in hollow carbon nanofibers; (e) TEM
image of Sn@carbon nanoparticle encapsulated in ahollow carbon nanofiber; (f) Capacity
cycle number curves of a Sn/C composite electrode and a commercial Sn nanopowder
(diameter: 100 nm) electrode at a cycling rate of 0.5 C. (g) Discharge capacity of a Sn/C
composite electrode as a function of discharge rate (15 C). Reproduced with permission.[66]
Copyright 2009, Wiley.

3.2.3 Electrospun transition metal oxides hollow nanofibers anode

45
It is anticipated that transition metal oxides (TMOs) are now one of the most promising anode

materials for LIBs. Compared to silicon or carbon materials, TMOs exhibit substantial merits,

including high capacity, enhanced safety, and widespread availability. Numerous TMOs, such

as binary, ternary, and complex oxides, have been widely investigated. Unfortunately, the

employment of TMOs in commercial LIBs is still hindered by their intrinsic low conductivity

and poor cycling stability. One effective way to guarantee a superior electrochemical

performance is to fabricate 1D nanostructured TMOs (especially with hollow interior).

It is considered that the research on TiO2 is very likely to make a breakthrough as electrode

for novel LIBs due to the relatively high working voltage, low cost, and structure stability.

Although there are various polymorphs of TiO2, including anatase, rutile, brookite, or TiO2

(B) (bronze), only anatase and TiO2 (B) phase shows satisfactory results owing to their unique

crystal structure. With the development of nanotechnology, it might be anticipated that

superior lithium storage properties can be achieved by engineering TiO2 with favorable

architecture.[163, 164] To achieve these aims, TiO nanofibers with multichannel hollow

structures were synthesized through single-nozzle electrospinning method without

template.[73] The generation of hollow multichannel architectures is based on the

evaporation-induced phase separation during electrospinning. Meanwhile, the unique hollow

nanobers enable easy electrolyte access to the bulk active material and provides effect

pathways for fast lithium ion transport, thus resulting in high rate performance. More recently,

a variety of configurations also have been proposed as good anode hosts with improved

electrochemical performance. For instance, TiO2 with fiber-in-tube structures has been

introduced using a traditional electrospinning process.[76] It demonstrated that the crystal

structures and morphologies of electrospun nanofibers were strongly affected by the post-

treatment conditions. The key requirement to obtain the TiO2 fiber-in-tube nanostructure is

the burning of TiO2-C nanofibers at oxygen atmosphere with a short time. Besides, the

repeated combustion and contraction of the electrospun nanofibers rendered the formation of
46
novel TiO2 fiber-in-tube configuration. The TiO2 fiber-in-tube nanostructures manifested low

charge transfer resistance and improved structural stability during repeated charge/discharge

process and excellent cycling and rate performances (177 mAh g-1 after 1000 cycles at 0.2 A

g-1) (Figure 11a and b).

Notably, the intrinsic poor electronic conductivity and low lithium diffusion rate

associated with TiO2 anodes limit its practical use. Considerable efforts have been devoted to

surmount these issues. As mentioned previously elsewhere in this review paper, metal or

carbon coating and N doping have been widely applied to enhance the conductivity of

electrode materials. Considering the attractive design, nitridated TiO2 hollow nanofibers has

been prepared via co-electrospun PVP/Titanium isopropoxide and mineral oil combined with

subsequent nitridization treatment.[113] Through NH3 treatment, the functionalized

conductive layers could be uniformly coated on both outer and inner surface of TiO2 hollow

nanofibers. As a result, the nitridated TiO2 hollow nanofibers exhibited excellent rate

capability (two times higher than that of pristine TiO2 nanofibers at 5 C), which could be

regarded as the conducting shell layer and hollow geometry. To realize homogeneous coating

or doping, kang and co-workers reported the fabrication of TiO2Nx porous 1D nanostructures

with high conductivity and fast Li+ diffusion.[165] Various questions, such as where does

nitrogen, what is formed and why electrochemical performances are enhanced, have been

clearly clarified, which further pave the way to improve the electrochemical performance of

LIBs via N doping. Recently, inspired by the development of film electrode, non-woven films

consisted of TiO2/Ag composite hollow fibers have been prepared by coaxial electrospinning

technique.[166] The incorporation of Ag in the composite structures significantly enhanced

the rate capability compared with TiO2 film without Ag. However, TiO2/Ag composite anode

still showed insufficient Li+ storage. From the point of the capacity level, multichannel hollow

nanofibers consisted of TiO2/C as binder-free anode has been produced via forcespinning of a

PVP/Titanium butoxide/ethanol/acetic acid/mineral oil and subsequent thermal


47
treatment.[167] The TiO2/C hollow fibers electrode demonstrated superior cycling

performance with a high specific capacity of 228.9 mAh g-1 after 100 cycles at 100 mA g-1.

As a promising anode materials for LIBs, SnO2 shows great advantages such as high

theoretical capacity (782 mAh g-1), relatively low cost, and abundance. It is found that the

lithium storage reaction between SnO2 and lithium can be ascribed as follow:

(1)

(2)

the alloying/de-alloying reaction (eq. 2) accounts for the high lithium storage capacity but the

huge volume variation during the formation of LixSn alloy lead to the poor capacity retention,

which significantly hinder the real application of SnO2 anodes. Recently, Sn with various 1D

hollow nanostructures have been created and illustrated enhanced electrochemical

performances. Porous SnO2 nanotubes have been reported, which exhibit superior

performance with a high discharge capacity of 807 mAh g-1 after 50 cycles.[168] The porous

hollow architecture not only accommodated the large volume changes during cycling process

but also facilitate the electrolyte diffusion into the active materials, thus accounting for the

remarkable electrochemical performance. Moreover, Kang et al. prepared fiber-in-tube SnO2

nanotubes based on the new proposed synthetic route.[77] It revealed the formation procedure

of SnO2@void@SnO2 fiber-in-tube nanostructures, which evolved from C-SnO2 composite

intermediate product to C-SnO2@void@SnO2 and finally produced the SnO2 fiber-in-tube

configurations during the heat treatment process at 500 C in air. The unique

fiber@void@tube configuration is beneficial to enable their structure stability during

electrochemical cycling, thereby resulting in excellent capacity retention (640 mAh g-1 after

300 cycles at 1 A g-1) and rate performance. Meanwhile, they also developed a unique

configuration with yolk-shell Sn@void@SnO/SnO2 structures.[116] During the two-step

post-treatment process, the electrospun precursor nanofibers were first reduced to carbon

nanofiber with embedded Sn nanospheres, then the oxidation of the Sn-C composite
48
nanofibers produced the yolk-shell Sn@void@SnO/SnO2 structures at 400 C in air. The

nanoscale Kirkendall diffusion process during the oxidation process caused the transformation

the Sn nanospheres into hollow SnO2 nanostructures. The formation of Sn/SnO core-shell

outer layer depended on the oxidation of Sn nanoparticles on the surface. As anode for LIBs,

the Sn@void@SnO/SnO2 nanostructures delivered a high discharge capacity of 663 mAh g-1

after 250 cycles at a high current density of 2 A g-1, corresponding to capacity retention of

77% (calculated from the second cycle) (Figure 10b). To increase the electrical conductivity,

a lot of design strategy to incorporate carbon or graphene layer in SnO2 anode to create higher

electrical networks have also been proved as an efficient way to improve the electrochemical

performance of LIBs.[115, 169-171] Doping one or more electrochemically active counter

atoms/ions have also been implemented. As one of the representative structures, SnO2-ZnO

hybrid nanotubes manifested excellent electrochemical performance, which can be related to

the hollow tubular nanostructures and the synergistic effect from the heterostructures of SnO2

and ZnO.[114, 172]

Different from intercalation and alloying mechanisms, another representative TMOs

(MxOy, M = Fe, Co, Ni, Mn etc.) react with lithium through a reversible conversion reaction,

which can be described as follows.

(3)

Generally, these kinds of TMOs could incorporate more than one Li ions per metal, thus

giving higher lithium storage capacities compared with those of graphite anodes.[173, 174]

However, the conversion reaction also leads to large volume changes upon electrochemical

cycling, which is like the alloying/dealloying reactions. Since the reversible lithium storage

occurs more easily with 1D hollow nanostructured electrode. Therefore, numerous TMOs and

TMOs/carbon composites with 1D hollow architectures have been fabricated through

electrospinning technique and a subsequent thermal treatment procedures.

49
1D nanostructured iron oxides,[175] Fe2O3 and Fe3O4, have been widely investigated as

anode materials for LIBs because of their high theoretical capacity, low cost, and

environmentally friendly nature. Unfortunately, the rapid capacity fading, which mainly

caused by the low intrinsic electric and huge volume changes, strongly hampered the

commercial application of iron oxides in current LIBs. To overcome these, the interstation of

void space into the structure has been suggested as an efficient approach to not only alleviate

the mechanical stress during electrochemical cycling but also enlarge the contact area

between the electrode and electrolyte, therefore significantly enhancing the lithium storage

properties. Based on the above analysis, bubble-nanorod-structured Fe2O3-C composite with

hollow Fe2O3 nanospheres embedded in 1D carbon matrix, has been fabricated by controlling

thermal treatment condition.[117] By virtue of the synergetic effect of hollow Fe2O3

nanospheres and conductive 1D carbon matrix, a superior cycling stability (812 mAh g-1 after

300 cycles at current density of 1 A g-1) and excellent rate performance can be achieved

(Figure 11c). Also, wang et al. reported the preparation of porous Fe2O3 nanotube by precisely

adjusting the concentration of iron precursor in the electrospun precursor solution.[118] With

the increase of concentration, the morphology of electrospun products changed from

nanobelts into nanotubes. Meanwhile, it suggested that the porous, hollow, and continuous 1D

structure of Fe2O3 plays a key role to obtain high specific capacity and good cycling stability

(987 mAh g-1 after 250 cycles at a rate of 200 mA g-1). Likewise, the spinel Fe3O4 is also

suitable for high-performance LIBs. Particularly, Fe3O4/C nanotubes were fabricated via

typical electrospinning route based on the phase separation process, followed by stabilization

and carbonization.[119] The composited nanotubes presented an unexpected cycling stability

with capacity of 600 mAh g-1 after 100 cycles at 0.15 C. Also, carbon-coated Fe3O4 hollow

nanofibers have demonstrated remarkable electrochemical properties with high reversible

capacity and excellent cycling performance as anode for advanced LIBs.[120]

50
In the case of cobalt oxides, it is essential to optimize their lithium storage properties by

controlling their nanostructures. Cao et al. fabricated porous Co3O4 nanotubes (P-Co3O4-NTs)

through coaxial electrospinning technique, which displayed tubular structure with porous wall

and hollow interior.[121] By simply adjusting the mass ratio of reactants, Co3O4 with various

nanostructure could be obtained. The unique porous tubular structure of P-Co3O4-NTs

endowed outstanding cycling stability (1826.2 mAh g-1 after 100 cycles at current density of

0.3 A g-1). Electrospun PAN-cobalt acetate nanofiber were selected as the bi-functional

template. Next, tubular-like structure of ZIF-67 nanocrystals can be achieved via a chemical

transformation process. In order to obtain hierarchical tubular CNT/Co3O4 nanoarchitecture, a

two-step thermal treatment was applied. As anode for LIBs, the CNT/Co3O4 hybrid structure

delivered a high reversible capacity of 1281 mAh g-1 at 100 mA g-1 with long cycle life over

200 cycles and exceptional rate capability (Figure 11d and e). Impressively, low et al.

reported the synthesis of hierarchical tubular structures consisted of Co3O4 hollow

nanoparticles and carbon nanotubes (CNT) by an efficient multi-step route (Figure 11f).[122]

Despite the success in the enhancement electrochemical performance of LIBs by designing

tubular structures, the conductivity of pure Co3O4 is still a huge challenge, which can be

improved by incorporating non-metallic conductive species to form hybrid structure. As

carbon doping nanostructures has been certified to be an effective way to enhance their

intrinsic conductivity by modifying the band structure. Recently, carbon-doped (C-doped)

Co3O4 hollow hybrid were synthesized by combining electrospinning and hydrothermal

method.[85] It is found that the C doping and 1D hollow structure guarantee a larger

delocalization of band structure, more Co2+ ions for faster Co2+/Co0 redox reaction, shorter

electron and Li+ ions diffusion pathway, as well as sufficient void space for alleviating

volume changes. All of these features are significant for the improvement of battery

performance. Notably, sun et al. designed novel C&N co-doped Co3O4 hollow structures and

also give evidence of the experimentally enhanced conductivity in unique C&N co-doped
51
Co3O4 hollow structures.[123] Benefiting from the uniqueness feature, such hybrid structures

exhibit ultrahigh specific capacity (1501 mAh g-1 after 60 cycles) and excellent cycling

stability.

Compared to Fe and Co, manganese oxides (MnO, MnO2, Mn2O3, Mn3O4) are another

important electrode materials for LIBs due to their high theoretical capacity, naturally

abundance, and environmental friendly. Thus, the use of manganese oxides or their

composites to enhance the battery performance is a prospect that should not be neglected.

Recently, Ahn et al. reported the synthesis of hierarchical C-encapsulated Mn2O3 architecture

with hollow tubular structures and porous walls based on a combination of a microwave

process and hydrothermal method.[124, 176] The uniqueness of this strategy is to offer the

beneficial synergistic effects in terms of electrochemical contact areas, structural stability, as

well as shorten electrical and ionic pathways. The optimized novel Mn2O3 electrode

demonstrated good electrochemical performances including high capacity and long cycling

life (875 mAh g-1 capacity retention up to 100 cycles) and superior rate capability (729 mAh

g-1 at 700 mA g-1) compared to previously reported Mn2O3-based anode materials.

Additionally, many other TMOs with hollow structures, such as nickel oxides[177, 178] and

indium oxides,[179] have also been prepared by electrospinning technique and employed as

anode for LIBs.

52
Figure 11. a) Cycling performances, and b) rate performances of the TiO2 nanofibers with
fiber-in-tube structure and TiO2 nanofibers with a filled structure and the commercial SnO2
nanopowders; Reproduced with permission.[76] Copyright 2015, Wiley. c) cycle performances
at a constant current density of 2.0 A g-1 of tin oxide porous nanobers; Reproduced with
permission.[116] Copyright 2015, Wiley. (d) cycling performances, and (e) rate performance of
Fe2O3 hollow nanobers and bubble-nanorod-structured Fe2O3-C composite nanobers;
Reproduced with permission.[117] Copyright 2015, American Chemical Society. f) rate
performance of the hierarchical CNT/Co3O4 microtubes. Reproduced with permission.[122]
Copyright 2016, Wiley.

As a promising family, mixed transition metal oxides (MTMOs), which incorporate one

or more metal oxides, have attracted wide interest due to the variety of physical and chemical

characteristics, as well as rich crystallographic arrangements, which render them applicable as

electrode for LIBs. Normally, MTMOs are immediately converted to nanocomposites of

single-metal oxides and Li2O after the initial conversion reaction. These single-metal oxides

could act as buffering matrices to alleviate the volume changes during electrochemical

cycling process.[180, 181] Moreover, it has been suggested that the ideal way to combine

these metal oxides is to start with ternary oxides (denoted as ABxOy) rather than mechanical

mixing of single-metal oxides. The complex chemical compositions and their synergetic

effects help to achieve desirable electrochemical performance.[182] Take the spinel ternary

oxides as an example, the presence of two different cations provide more flexibility to tune

the working voltages and control the energy densities.

Keeping the concept of MTMOs in mind, 1D CaSnO3 with various morphology

including nanotubes or eggroll-like nanotubes, have been introduced.[125, 126] It suggested

that the morphology of final products strongly depended on various preparation parameters

such as the concentration of metal precursor, thermal treatment condition, as well as some

other electrospinning parameters. For example, by rationally adjusting the metal precursor

concentration, the morphology of 1D CaSnO3 nanostructures could be tuned from eggroll-like

53
nanotubes to nanorods (Figure 12a). Also, by simply adjusting the annealing temperature,

CaSnO3 nanofibers, nanotubes, and nanobelts can be obtained. As the electrochemical

performance of electrode materials is sensitive to their diverse morphologies. It provide us

more opportunity to achieve high-performance electrodes. When evaluated as anode for LIBs,

the Ca elements do not take part in the electrochemical reactions but act as inactive buffering

matrices, which is critical for the enhancement of cycling stability of Sn-based anodes. It can

be observed that CaSnO3 nanotubes delivered a higher reversible capacity of 648 mAh g1

after 50 cycles, which is about 1.6 times of that for CaSnO3 nanorods (410 mA h g-1) (Figure

12b). Besides, both the unique interior hollow structures and channel-like surface also account

for the excellent electrochemical performance of eggroll-like CaSnO3 nanotubes. Different

from inactive buffering matrices, electrochemical active elements matrices could also

introduced to form ternary isostructures without sacrificing the electrochemical

performance.[53, 127-129, 132, 133] Considering that both Co and Zn are electrochemically

active with respect to lithium. Porous ZnCo2O4 nanotubes were fabricated by electrospinning

followed by calcination (Figure 12c).[130] The different rate of mass transfer during the

thermal treatment procedure might promote the generation of tubular structures. Such unique

configuration favours fast Li+ transportation and accommodates the volume changes during

cycling process. Besides, the Zn and Co could act as self-matrices, which not only

contribute to the large specific capacity but also perfectly prevent the aggregation of active

materials, therefore resulting in high capcacity, good cyclability, and superior rate

performance. As seen in Figure 12d, the specic discharge capacities of 1011, 841 and 794

mAh g-1 were retained even after 30 dischargecharge cycles at the current densities as high

as 500,1000 and 2000 mA g-1. Similarly, Mai et al. demonstrated the facile interface-

modulated approach towards the synthesis of various MTMOs multilevel nanotubes with

tunable interior structures via electrospinning combined with controlled heat treatment.[131]

The electrospun precursor nanofibers can be transformed into shrinkable wire-in-tube and
54
tube-in-tube nanotubes by rationally controlling the polymer/metal oxides interface.

Remarkably, this versatile strategy can be extended to prepare various metal oxides such as

CoMn2O4, MnCo2O4, as well as NiCo2O4. These multilevel nanotubes enables a high surface

area, shorten diffusion pathway, and good strain accommodation, which are beneficial for

LIBs.

Figure 12. a) SEM image of egg-like CaSnO3 hollow nanofibers, b) Capacity and Coulombic
eciency versus cycle number of the electrospun CaSnO3 nanomaterials (current density = 60
mA g-1); Reproduced with permission.[126] Copyright 2013, The Royal Society of Chemistry.
c) SEM image of ZnCo2O4 hollow nanofibers, and d) Capacity vs. cycle numberof the porous
ZnCo2O4 nanotubes at current densities of 100, 500, 1000, and 2000 mA g-1 in the voltage
range of 0.013 V. Reproduced with permission.[130] Copyright 2012, The Royal Society of
Chemistry.
3.3 Electrospun polymer electrolyte

Conventional LIBs consist of liquid electrolytes and solid electrodes, which can have

potential security risks, such as flammability, explosion and volatilization, due to the presence

of organic solvents. Demand for safe and high energy LIBs is increasing. Polymer electrolytes

have been actively studied as a possible solution to eliminate most of the safety concerns

encountered with liquid electrolytes, mainly because of high safety, low cost, wide

electrochemical window, good compatibility with lithium salts, and high thermal stability. It

is known that polymer electrolytes act as dual role, an electrolyte and a separator,

55
simultaneously, in a solid-state configuration. Thus, many essential electrochemical

characteristics must be concerned, including ionic and electronic conductivity, chemical,

thermal and mechanical stabilities. Up to now, preparation of polymeric electrolyte using

various host polymers such as poly(ethylene oxide) (PEO), poly(acrylonitrile) (PAN),

poly(vinylidene fluoride) (PVDF), and poly(methyl methacrylate) (PMMA) have been

investigated. In this part, we mainly focus on systematically summarizing the recent

developments and issues concerning electrospun polymer electrolytes for LIBs.

As a key characteristic, ionic conductivity is a basic requirement of a suitable electrolyte

for LIBs. Nevertheless, the pursuit of polymer electrolytes remains active out of the

enthusiasm for a neat polymeric media that allows for ion/electron transport while

simultaneously providing strong mechanical support. Researchers have devoted to find

different approaches to improve the ionic conductivity of polymer electrolytes, such as,

modifying, blending, making polymer derivatives and so on. PEO are the earliest reported and

most popular polymer hosts of polymer electrolytes for LIBs by virtue of its complexing

ability with Li+ ions and highly flexible backbone. However, the high crystallinity of PEO at

room temperature limits the conductivity, and its mechanical strength is too low to be used

alone. Thus, di- and tri-block copolymers of PEO as polymer matrices is drawing much

interesting in the recent reports. For example, Li et al. reported the fabrication of PEO/PVDF

blending polymers by electrospinning.[183] In the composite nanofibrous membranes, PEO

benefited to good electrochemical property, while PVDF acted as mechanical support. The

PEO/PVDF exhibited improved ionic conductivity, stable electrochemical window (4.8 V),

and good cycling stability in LIBs. Similarly, shan et al. used electrospun PVDF membrane as

scaffold and then build up PEO-LiClO4-TiO2 electrolyte by infiltration to be the ionic

conducting phase.[184] The effect of TiO2 nanoparticles on the enhancement of Li+ ion

conductivity has been studied. And it is found that the real dispersion state of TiO2

nanoparticles in the nanocomposite electrolytes is in the form of large aggregates instead of


56
the individual primary nanoparticles, which would tremendously depress the surface effect of

the individual nanoparticles and greatly impact on the enhancement of the ionic conductivity.

The electrospun PVDF-HFP composite nanofibers have been widely adopted as host

matrices to prepare polymer electrolyte for LIBs, and it is well-accepted that the incorporation

of nano-sized ceramic fillers is an efficient way to enhance the ionic ionic conductivity of

polymer electrolytes. Ahn et al. reported that the nano-sized ceramic fillers were incorporated

into PVDF-HEP membranes during the electrospinning process.[185, 186] The existence of

the ceramic nanoparticles is beneficial for the improvement of ionic conductivity and the

electrochemical stability window of the electrospun PVDF-HFP-based polymer electrolyte. It

is found that the cell Li/LiFePO4 based the on the electrospun PVDF-HFP-BaTiO3 polymer

electrolyte delivers a discharge capacity of 164 mAh g-1, which corresponds to 96.5%

utilization of the active material. The enhanced performance of the PVDF-HFP-BaTiO3

polymer electrolyte is attributed to its better interaction with the host polymer and

compatibility with lithium metal. Moreover, the investigation on a new electrospun polymer

electrolyte containing of thermoplastic polyurethane (TPU) and PVDF has been made.[187,

188] Cao et al. fabricated nanofibrous membranes based on PVDF doped with TPU and

polystyrene (PS) by electrospinning. the electrospun TPU/PVDF/PS polymer electrolyte

presented a high tensile strength and elongation.[189] With the help of excellent

polyelectrolyte complexes, the prepared nanofibrous membranes present superior

electrochemical and mechanical performances, which are very promising for application in

LIBs. Recently, novel organic ionic plastic crystal (OIPC) modified PVDF composite

nanofiber has been successfully prepared by electrospinning.[190] Compared with pure

electrospun PVDF, which have an electroactive phase phase and a small amount of non-

polar phase, the ion-dipole interaction between OIPC and the polymer in the co-electrospun

composite system can reduce the non-polar phase PVDF, resulting in almost entirely

electroactive phase PVDF, which is desirable for optimum electrochemical activity.


57
Consequently, the incorporation of the OIPC in electrospun fibers can greatly improve the

conductivity of OIPC/PVDF composites. The increased ionic conductivity of these

composites is very beneficial for their use as solid state electrolyte for batteries.

Apart from the above mentioned electrospun polymer electrolyte, electrospun PAN and

its composite,[191-193] PMMA,[194] as well as polyimide (PI),[195] has also been

investigated as polymer electrolyte for LIBs. However, it is still difficult to obtain an

electrolyte that displays equivalent ionic conductivity to that of a liquid electrolyte, superior

safety, and suitable mechanical properties and contact with electrodes. The study of

electrospun polymer electrolyte is still faced with a lot of unsolved problems. Therefore,

numerous scientific and technical challenges must be overcome to realize commercial use of

solid electrolytes.

4. Sodium ion batteries

Sodium ion batteries (NIBs) have shown immense potential as alternatives to LIBs in energy

storage applications, particularly for large scale grid storage, owing to the relative abundance

of Na, and the low cost of the raw materials needed for their fabrication.[6, 181, 196-198]

However, performance wise, the larger ionic radius of Na compared to Li (1.02 vs 0.76 ),

and their tendency to preferentially coordinate in prismatic and octahedral sites, poses serious

challenges in designing a suitable crystalline host that allows for sufficient electrochemical

capacity and repeated cycling of the sodium ion insertion/desertion reactions during charge

and discharge cycles.[199] Thus, is has become a priority to develop NIBs electrode materials

that give high capacity and cyclability to advance the adoption of NIBs technology.

Numerous oxide based cathode materials have been fabricated and tested as cathodes for

NIBs.[200] Among them, layered materials having the chemical composition NaMO2 (M = V,

Cr, Mn, Fe), appear to hold much promise. Specifically, layered oxides of Fe-Mn has been

extensively studied largely due to their high redox potentials, high energy densities, and low

58
cost of the Fe and Mn precursor materials.[201] The considerable difficulty in dissolving the

raw precursors to prepare an electrospinning solution render it challenging to electrospin these

mixed oxides as NIBs cathodes directly. Mai and co-workers prepared mesoporous nanotubes

with chemical composition Na0.7Fe0.7Mn0.3O2 by an electrospinning process and evaluated

their sodium storage performance.[51] These nanotubes are composed of ultra-thin carbon

nanotubes of approximately 200 nm diameter, decorated with Na0.7Fe0.7Mn0.3O2 nanoparticles

of diameters typically about 10 nm (Figure 13a). The research group reported stable voltage

plateaus when measurements were taken at current densities from 100, 200, 300, and 500 mA

g-1, corresponding to a pair of distinct peaks, with the anodic peak at 3.9 V and the cathodic

peak at 3.6 V as shown in Figure 13b. These hollow nanostructured fibers demonstrated good

performance in cyclability tests, retaining 90% of its initial capacity after 1000 cycles at

current density 100 mA g-1. Even when a higher current density of 500 mA g-1 was applied,

these nanofibers showed a capacity retention of 70% after 5,000 cycles (Figure 13c).

Comparing the performance of the mesoporous Na0.7Fe0.7Mn0.3O2 nanotubes against their

nanoparticle counterparts, a much improved capacity and cycling behavior was observed. This

was attributed to the large surface area, high conductivity, and unique mesoporosity of the

ultra-thin continuous carbon nanotubes, which also act to effectively accommodate the

stresses induced during the sodium insertion/desertion cycles.

During early development of SIBs, hard carbon was the material most commonly used as

the anode, but presented many practical limitations and were eventually phased out in favor of

the TMO originally developed for LIBs.[202, 203] Benefitting from a more efficient

electrochemical conversion mechanism, these TMO based anodes allowed higher theoretical

specific capacities to be achieved.[204-206] Compared to related metals and oxides, metal

chalcogenides are expected to display improved mechanical stability using ether-based

electrolytes because of their smaller volume change and higher first cycle efficiency conferred

by their greater reversibility, which inspire researchers to develop metal chalcogenides anodes
59
materials in SIB.[207-210] Metal selenide materials designated for use in rechargeable

batteries have, however, not been prepared via the electrospinning process. As such, hollow

(Co1/3Fe2/3)Se2 composite materials have been fabricated by electrospinning and subsequent

selenization process.[198, 211] The selenization process can transform smooth CoFe2O4 tube-

in-tube nanostructures into hierarchically (Co1/3Fe2/3)Se2 fiber-in-tube nanostructures. The

SEM image shows ultrafine (Co1/3Fe2/3)Se2 nanorods are uniformly decorated on the

nanofibers (Figure 13d). When applied in SIB, (Co1/3Fe2/3)Se2 exhibits a discharge capacity of

594 mA h g-1 after 60 cycles (Figure 13e). The rate performance shows that the reversible

discharge capacity decreased slightly from 585 to 497 mA h g-1 as the current density was

increased from 0.1 to 5.0 A g-1 (Figure 13f). Compared to the (Co1/3Fe2/3)Se2-Se-C composite,

the better electrochemical performance of the unique (Co1/3Fe2/3)Se2 with the fiber-in-tube

structure is attributed to the synergetic effects of the hierarchical structures and ultrafiber

nanorods with high electrical conductivity, which is indicated by the lower charge transfer

resistance.

Figure 13. (a) TEM image of the Na0.7Fe0.7Mn0.3O2 mesoporous nanotubes; (b) Charge
discharge curves of Na0.7Fe0.7Mn0.3O2 measured at 100, 200, 300 and 500 mAg-1, respectively.
The inset is the CV collected at a scan rate of 5 mVs-1 in the potential range 3.04.5V; (c)
Cycling performance of Na0.7Fe0.7Mn0.3O2 mesoporous nanotubes tested for 1,000 cycles at
100 mAg-1; Reproduced with permission.[51] Copyright 2015, Nature Publishing Group. (d)
SEM image of the (Co1/3Fe2/3)Se2 nanobers with ber-in-tube structures; (e) cycling
60
performances, and (f) rate performances of the hierarchically structured (Co1/3Fe2/3)Se2
nanobers and (Co1/3Fe2/3)Se2SeC composite. Reproduced with permission.[211] Copyright
2016, The Royal Society of Chemistry.

5. Li-S batteries

An unambiguous trend facing the development of advanced energy storage devices is the ever

increasing energy density requirements, laying the gauntlet for researchers to design materials

which can deliver the necessary performance. Unconventional cathode materials are a

burgeoning research area, where reports indicate materials like sulfur and oxygen offer much

superior capacities over their conventional insertion cathode counterparts.[212, 213] Sulfur

possesses the dual advantages of both being an abundant material, and being

electrochemically active, able to accept two electrons per atom at 2.1 V against Li/Li +. This

theoretically gives it an exceptionally high capacity of 1675 mA h g-1, and when applied as

Li-S batteries, a theoretical energy density approximately 2600 W h kg-1. Additionally, the

low toxicity of sulfur as opposed to expensive, highly hazardous transition metals such as

cobalt, make it a viable low cost, environmentally benign, and safer alternative.[214]

The path to commercialization for Li-S batteries however face several significant obstacles,

with its lack of electric and ionic conductivity proving the most challenging.[215, 216] One

approach to mitigate this shortcoming is to incorporate the sulfur into a conducting matrix

such as a metal, carbon, or conducting polymer, this coming at the cost of reducing the energy

density. A second limitation is the poor mechanical stability and diminishing capacity

resulting from the complex chemical and structural changes accompanying the formation of

soluble polysulfide intermediates during the operation of the battery. Thirdly, the dissolution

of the polysulfide results in a loss of active sulfur and decreases the batterys capacity during

charge/discharge cycles. The fourth issue to be addressed is the formation of a solid-

electrolyte-interphase (SEI) as the reactive lithium metal reacts with the electrolyte, which

61
hampers the reversibility of the reaction and deteriorating coulombic efficiency over the

course of the plating/stripping cycles.

Much of the research effort centered on sulfur based electrode materials concerns

preserving the electrodes structural integrity, fully utilizing the active material, and

delivering adequate cycle life with overall systemic efficiency. One potential route to

achieving these objectives is to engineer these materials at the nanoscale, exploiting the

nanoscale architectures which give them enhanced resistance to mechanical degradation.

These nanostructured sulfur electrodes can be designed with varying morphologies including

porous carbon-sulfur composites, as well as nanotubes containing sulfur, which besides

improving rate capability, also serve to provide voids which buffer volume changes during

charge/discharge cycling.[217] The free volume minimizes the strains experienced by the

electrode, staving off electrode pulverization, and improving cycling stability. These

improvements brought about by tuning the electrodes nanostructure add up to significantly

alleviate the problems plaguing the Li-S battery system, and are thus seen as a promising

avenue for further exploration.

High performance sulfur cathodes for Li-S batteries that have both excellent capacity and

long-lived cycling stability should ideally possess certain characteristics such as (1) an

adequate sulfur content, (2) high electrical and ionic conductivity (which can be achieved

through the incorporation of conducting materials within the cathode), (3) a mechanically

robust structure able to accommodate the strains resulting from large volume changes, (4)

mechanisms which enable the trapping of polysulfides at the cathode.[218] In other words, the

desirable cathode should have high content of sulfur with no compromise in superior ionic

and electronic conductivities and with resistance to pulverization while retaining the

polysulde intermediates within the electrode.

With those requirements in mind, three-dimensional nanoarchitectures are a promising

platform to obtain high sulfur content since they do not require the addition of binders,
62
conductive fillers (like carbon black), nor current collectors, thereby attaining higher mass

and volumetric fractions of the eventual electrode compared to the conventional electrodes

prepared from slurry coatings. For instance, Lou and co-workers used electrospinning to

obtain pie-like paper electrodes that were free standing, and constituted from numerous

multi-channel carbon nanofibers resembling lotus roots (LRC), shown in the Figure 14.[72]

High sulfur loadings of 3.6 mg cm-2 was achieved from this method, with hydrophobic

interactions between the pyrolytic carbon of the supporting nanofiber and the impregnated

sulfur confining the sulfur to within these 3D interconnected channels. Elemental mapping

performed during TEM analysis of a segment of a single free-standing nanofiber further

evidenced the successful incorporation of sulfur in the channels of the LRC, forming a lotus

root-like carbon fiber/sulfur composite electrode (LRC/S) (Figure 14a). Upon the application

of a thin coating of functionalized reduced graphene oxide (EFG) on the LRC/S surface, the

initial discharge capacity of the electrodes was measured to be as high as 1,215 mAh g-1,

approaching 72% of sulfurs theoretical capacity at 1,675 mAh g-1. This LRC/S@EFG

performed well in the following charge/discharge cycling tests, showing much improved

stability with a mere 0.1% decay in the reversible capacity per cycle, measured at 950 mAh g-
1
at 200 cycles (Figure 14b). The rate performance in Figure 13b shows that reversible

discharge capacity is found to stabilize at ~300 mAh g-1 (4.7 mAh cm-2) at initially 0.1 C, and

gradually decreases to 1,113 (4.0), 801 (2.9), 688 (2.5) and 363 mAh g-1 (1.3 mAh cm-2) as

the current rate is increased to 0.2, 0.5, 1 and 2 C, respectively. When the current rate is

reduced abruptly back to 0.1 C again, the LRC/S@EFG electrode is able to recover most of

the original capacity, indicating outstanding stability and robustness of the pie-structured

electrode (Figure 14c and d). At a current density of 1.2 mA cm-2, the cells with different

(three, two and one) layers of LRC/S@EFG electrodes deliver initial discharge capacities of

10.7 (993), 7.2 (995) and 3.8 mAh cm-2 (1,054 mAh g-1), respectively, and show good

capacity retention (Figure 14e). These improvements can be traced to several aspects of the
63
engineered 3D nanostructure of LRC/S@EFG, firstly, the large porosity enables high loading

of the active sulfur material in the electrode. Secondly, the highly aligned channel walls

running parallel within the nanofiber gives good contact between the sulfur/lithium sulfide

and the supporting carbon framework. The unique design of the electrode comprising an

interconnected framework of LRC nanofibers loaded with active sulfur material acts to reduce

the charge transport resistance necessary for the electrochemical redox processes to occur,

thereby resulting in the high discharge capacity observed.

Figure 14. (a) TEM image and corresponding elemental mappings of carbon and sulfur of the
LRC/S composite; (b) Cycle performance of LRC/S@EFG in comparison with LRC/S at a
current density of 0.2 C; (c) Voltage proles and (d) discharge capacities at various current
densities from 0.1 to 2 C; (e) Areal capacities of layer-by-layer structured LRC/S@EFG
electrodes during cycling at a current density of 1.2 mA cm-2. Reproduced with permission.[72]
Copyright 2015, Nature Publishing Group.

6. Metal-air batteries

The tremendous theoretical energy density of metal-air batteries, far outstripping that of its

peers in energy storage technologies, has attracted much interest over the past few years.[219-

221] This superlative energy density of lithium-oxygen batteries is largely due to the pure

64
metal anode, and the vast external availability of the cathode oxidant (oxygen in air). The

lightweight Li-air battery, employing a metallic Li anode possesses the highest theoretical

specific energy in the order of 13,300 Wh kg-1 with respect to the anode, making them front

running candidates for rechargeable batteries in all manner of applications, especially those

requiring large capacity energy storage like in electric vehicles having a single-charge range

of 500 miles, rivalling that of combustion engine vehicles.[222]

The standard assembly of an Li-O2 battery comprises a metal lithium anode, porous air

cathode in contact with oxygen, and an aqueous or non-aqueous organic electrolyte.[223] The

standard assembly of a Li-O2 battery comprises a metal lithium anode, porous air cathode in

contact with oxygen, and an aqueous or non-aqueous organic electrolyte. Since non-aqueous

electrolytes tend to give higher theoretical energy densities than their aqueous counterparts,

they have been the more intensively researched variety. Of these components, the most

challenging aspect to enhancing the performance of the Li-O2 battery system lies in improving

the air electrode, which will dictate the power density, energy density and efficiency of the

non-aqueous rechargeable system. The air electrode is responsible for the electrocatalysis of

the oxygen reduction and oxygen evolution reactions as described by the following equations.

ORR O2 + 2Li++2e Li2O2 (4)

OER Li2O2 O2 + 2Li++2e (5)

These catalyst reactions play a pivotal role in the operation of the battery, improving the

charge reactions by lowering the potential necessary to trigger the dissociation of the reaction

products like LiO2 into lithium and oxygen[224] Because these factors greatly dictate the

performance of the battery system, investigations into improving the sluggish electrochemical

reaction kinetics and stability of the compounds produced have drawn much attention.

Employed directly by itself, as conductive filler, or as an electrocatalyst support, carbon is

the most widely extensively used material for air electrodes in Li-air batteries.[225, 226] The

65
use of bare porous carbon nanofibers is limited by the relatively smaller pore sizes which

encumbers their ability to accommodate the large amounts of discharge products consisting of

non-conducting and insoluble Li2O2.[227] Moreover, the small surface openings, low surface

permeability, and the absence of interconnected channels across the macroporous carbon

spheres, contribute to the clogging of the susceptibility to clogging by Li2O2 and poor mass

transport kinetics. These characteristics result in a poor utilization ratio, impairing the

performance of the electrode in terms of reducing specific capacity in addition to lowering the

rate capability, once more emphasizing the importance of rational design of the carbon

composites. Work by Zhang and co-workers on lightweight, electrospun, hierarchical

macroporous active carbon fiber (MACF) decorated with RuO2 (R-MACF) represents a

classic electrocatalyst for Li-O2 batteries.[81] Nanoparticles of TiO2 measuring 1.5 to 4 nm in

diameter were uniformly distributed throughout the hollow carbon sphere surfaces. Employed

as self-standing, binder free cathodes in Li-O2 battery cells, under high current densities of

1000 and 2000 mA g-1, R-MACF managed to achieve high specific capacities of 13,290 and

9,222 mA h g-1 respectively. The cyclic stability performance of R-MACF cathodes in Li-O2

batteries is also noteworthy, with the terminal discharge voltage maintained above 2.0 V after

being subject to 154 cycles. This observed excellence in cyclic stability and rate capability is

derived from the hierarchically porous nanostructure, conductive materials, mechanical

robustness, and outstanding thermal characteristics of the R-MACF, allowing the mass

transport of Li+, O2, and discharge products while maintaining the mechanical stability of the

binder-free electrode during charge/discharge cycling.

Oxides are another class of commonly used oxygen catalysts for aqueous Li-air batteries

apart from carbon based noble metals. Metal oxides, perovskites in particular, are being

touted as prime candidate electrocatalyst materials for Li-O2 batteries owing to their

exceptional ionic and electronic conductivity.[41, 228-230] In a reported example,[41]

nanotubes of La0.75Sr0.25MnO3 (PNT-LSM) were electrospun, taking advantage firstly of the


66
materials 1D tubular structure which promotes fast electron transport, secondly of the high

porosity facilitating reduced diffusion lengths and homogenous distribution of the electrolyte

and O2, and thirdly of the hollow structure offering increased exposed surface area for

electrolyte contact thereby increasing the accessibility to the catalytic sites available. These

design features allowed a much reduced charge voltage for the assembled Li-O2 battery, with

PNT-LSM/KB being 200 mV lower compared to that of neat KB. Furthermore, good cycling

stability was observed, with PNT-LSM/KB able to undergo more than 124 cycles at high

current densities of 1000 mAh g-1.

In a recently published work, Kim and co-workers fabricated a tube-in-nanotube

configuration of RuO2/Mn2O3 (RM-TIT/KB) for highly efficient Li-O2 batteries.[42] Figure

15a shows that Mn2O3 coexisted in both the inner and outer tubes. The rationale for

incorporating RuO2 lies in improving the electrical conductivity and sluggish OER kinetics

faced by MnOx. RM-TIT/KB demonstrated an outstanding capacity of 2000 mAh g-1, (Figure

15b) and excellent cycling stability in excess of 100 cycles at a discharge capacity of 1000

mAh g-1 and voltage window between 2.0 and 4.8 V (Figure 15c). This excellent

electrochemical performance was a result of the novel double walled structure of the fibers

which facilitated the efficient formation and decomposition of Li2O2. The most plausible

models for catalytic reaction mechanism of the RuO2/Mn2O3 ber-in-tube (RM-FIT) and RM-

TIT based air electrodes are proposed, and corresponding results of ex-situ observation were

exhibited in Figure 15d-m. After discharging, Li2O2 were mainly accumulated on the surface

of the outer tubes of RM-FIT due to the outstanding ORR properties of Mn2O3 (Figure 15e, f),

while some of the Li2O2 on the outer surface of the RuO2 shells in RM-FIT are insuciently

decomposed due to the relatively low OER catalytic activity of Mn2O3. Due to the firmly

distribution of Mn2O3 and RuO2 nanocomposite in the double-walled bers, the

predominantly formed Li2O2 at near the Mn2O3 components in inner and outer tubes during

discharge process (Figure 15j, k) can be easily catalyzed (decomposed) by adjacent RuO2
67
components in recharge process (Figure 15l, m). The ecient formation and decomposition

of Li2O2 induced by the evenly distributed RuO2 and Mn2O3 catalysts lead to stable cycling

and rate performance in Li-O2 cells. The rational design of the hierarchical structure unlocked

significant improvements in efficiency for OER and ORR catalytic activity, resulting in better

electrochemical performance and long term cycling stability of the batteries.In a separate

work, the same group fixated Co3O4 on both surfaces of 2D non-oxidized graphene

nanoflakes and applied the material in Li-O2 batteries.[231]

Figure 15. (a) SEM image of RM-TIT; (b) First discharge-charge curves with a capacity limit
of 2000 mAh g-1 at a current density of 100 mA g-1 with voltage window between 2.35 and
4.35 V; (c) cycle performance, of KB, RM-FIT/KB, and RM-TIT/KB with a capacity limit of
1000 mAh g-1 at a current density of 400 mA g-1 with voltage window between 2.0 and 4.8 V;
Proposed reaction mechanism of (d) RM-FIT and (i) RM-TIT used electrodes during the
operation of LiO2 cells, which are represented by four states as (i) on discharging, (ii)
discharged state, (iii) on recharging, and (iv) recharged state. The ex situ SEM and TEM

68
images of (e, f) discharged and (g, h) recharged states in RM-FIT used electrode, and ex situ
SEM and TEM images of (i, k) discharged and (l, m) recharged states in RM-TIT used
electrode. The dashed circles in TEM images (f, h, k) indicate the discharge products, Li2O2.
Reproduced with permission [42] Copyright 2016, American Chemical Society.

Zinc-air batteries generally consist of a Zn based anode (which can take the form of the

metal itself, Zn paste, or Zn powder mixed with other additives in the presence of an alkaline

electrolyte), an alkaline electrolyte, a battery separator, and an air cathode (commonly carbon

based with catalyst particles). These batteries differ from non-aqueous Li-air batteries in that

the 9 M, 37 wt% KOH alkaline solution electrolyte is often used in Zn-air batteries since it is

highly conductive and does not result in significant corrosion gassing. Where Zn-air batteries

find similarities with their Li-air contemporaries is that both are limited by poor ORR kinetics,

throwing into focus the urgent need to develop highly efficient bi-functional catalysts for both

ORR and OER reactions in order to achieve commercialization of these Zn-air batteries.[232]

To this end, electrospun carbon nanofibers doped with nitrogen have shown encouraging

performance in Zn-air batteries.[230] It was found that the heat treatment carried out at 1100
o
C yielded the best performing material (N-CNF-1100), with the highest positive onset

potential, and a limiting current density on par with expensive commercial gold standard Pt/C

catalysts. Evaluating the N-CNF-1100s performance in a Zn-air cell, peak power densities of

194 mW cm-2 was recorded, again comparable to that of Pt/C catalysts. Another conclusion

drawn from this reported work is how dependent the eventual electrochemical performance of

the battery is on contributory factors such as carbonization temperature, number of exposed

active sites, and extent of nitrogen doping.

7. Conclusions, Challenges, and Prospects

With high performance batteries poised to play a growing role as the primary or backup

power supplies for the electric vehicles and portable electronic devices, battery technology

research, especially in the field of materials science is currently in vogue, has drawn much

69
attention in developing ever more efficient, high performance electrode materials. Over the

past several years, the sheer variety of materials being investigated as advanced electrode

materials has been overwhelming, including carbon, metal oxides, nitrides, carbides, and their

many derivatives. While the materials used may be different, there is an undeniable consensus

that novel, low-cost, active materials with engineered nanostructures are the key in unlocking

substantial gains in performance. The versatile, efficient, simple electrospinning method of

preparing nanofibers of a vast array of materials has thus gained popularity, and is taking on

greater significance in the fabrication energy storage and conversion devices.

Electrospun hollow nanofibers (EHNF) are fast becoming one of the most effective active

electrode materials in advanced battery devices because of their superior electrical

conductivity, greater reactivity, and robust mechanical properties. In this review, we endeavor

to condense the recent developments of EHNF in terms of the approaches to synthesize them,

their morphologies, properties, characterization, and applications in some of the more

prominent battery technologies such as the LIBs, NIBs, Li-S, and metal-air batteries. It was

discovered that the synthesis of EHNF with finely-tuned crystal structures, particle sizes,

structural defects, morphologies, compositions, and dimensions played a critical role in

maximizing the active site exposure, and in controlling the properties of the synthesized

material. Given its importance, we therefore must first embark on an exploration of the

methodologies adopted for electrospinng hollow nanofibers. Covering the control of the

parameters involved in electrospinning and heat treatments, we introduce modified

electrospinning protocols such as coaxial electrospinning, and emulsion electrospinning,

along with the various post processing treatments available, and discussed how they can be

manipulated to exert an influence over the properties such as the morphology, graphitization

degree, and functional chemistry of the spun nanofibers. Of the techniques introduced, coaxial

electrospinning is an efficient approach having the distinction of being the most widely used,

affording the versatility to fabricate multi-shelled, hollow micro/nanostructures of various


70
metal oxides. However, the complexity involved in employing concentric needles and

controlling solution characteristics reduces its viability at an industrial level. The development

of novel alternative one-pot, single spinneret approaches such as emulsion electrospinning

and precisely controlled heat treatments to obtain these hollow nanofibers provided an avenue

to spin a broad range of materials in a more consistent and scalable manner, but more

significantly, enabled the fabrication of even more complicated, hollow, multi-shelled

nanoarchitectures.

Bearing in mind the desirable characteristics possessed by these hollow nanofibers such as

large surface areas, high porosity, and membranous morphology, EHNF have exhibited

outstanding performances when applied in advanced energy devices like advanced batteries

(Li ion, Li-S, and metal-air batteries) and supercapacitors. The superlative performance along

the metrics of large capacity, long term cycling stability, and high electrocatalytic reactivity

are derived from the following remarkable features of the EHNF: (1) hollow nanostructures

give a large electrode/electrolyte contact area and allow the storage of Li ions on the surface,

in the pores beside the grain lattice, and at the interfaces, all of which result in higher

capacity; (2) mesoporous structures can buffer the large volume fluctuations accompanying

charging/discharging, allowing the retention of mechanical integrity, better stability and

cyclability. Furthermore, the porous structures act as traps to effectively confine the dissolved

polysuldes of the sulfur cathode, preventing active material loss. (3), the hollow structure

improves electrolyte-electrode contact area, increasing the number of accessible

electrocatalytic active sites for ORR/OER during LiO2 batteries operation. Additionally,

these nanostructures provide more free volume to accumulate the discharged products which

improves both capacity and cyclability. (4) these flexible electrodes do not contain binders or

additives, but offer short ion and electron diffusion pathways for intrinsically high charge

carrier conductivity. Exploiting these features can thus be seen as an opportunity central to the

future development of flexible batteries.


71
The accelerated rate of development in the electrospinning of hollow nanofibers is likely to

see progress in the energy density, power density, and stability of the advanced battery

electrodes. However, this promising technology as applied to hollow nano/microspheres is

still in its infancy, and certainly comes with many technical challenges to be addressed (to be

detailed in the following paragraphs), providing opportunities for groundbreaking work to

take the next leap forward in academic research, and possible commercial adoption.

Firstly, although there exist innumerable reports of the successful synthesis of hollow

nanofibers, the inherent difficulty in fabricating these micro/nanostructures with

compositional and geometric accuracy makes it hard to derive large quantities of consistently

high quality materials suitable for specific real-world applications outside the laboratory. The

way forward could be to focus fine-tuning the synthesis and structural control of these hollow

nanofibers by firstly developing a mechanistic understanding of the nanoporosity evolution at

these confined nanoscale dimensions. Insights from this could allow for better defined

protocols in controlling the eventual structure and properties of the nanofiber, and tune them

for various applications. In another regard, it has been demonstrated that hollow nanofibers

with multi-shells or complex interiors possess some obvious advantages over the same sized

materials with simple structures, due to the high specic areas, abundant inner voids,

multiphase heterogeneous interfaces, making them larger volumetric capacity, better

structural and electrochemical stabilities and higher energy and power densities. However,

compared with the reported hard-template methods, especially carbon spheres for multi-

hollow spheres, the examples of multi-shelled hollow nanofibers in batteries are limited, due

to the difficulty in synthesizing them, which needs researchers to take more efforts to

investigate it. Apart from this, surface modification strategies may prove to be an effective

post-treatment such as by manipulating crystal grain boundaries, introducing dangling bonds,

and attaching functional groups etc. to the prepared materials. These alterations can affect the

resulting fiber performance characteristics like the ionic storage capacity, thereby invoking
72
improvements to their application specific performances. Additionally, investigations into the

mechanisms behind the improved capacity of multi-shelled electrode materials for LIBs and

SCs can yield important insights beyond the currently poorly understood model, where multi-

level surfaces provide a higher density of reactive sites for charge storage. These preliminary

steps could go a long way in understanding how to effectively tailor and ne tune the

nanofibers physical, chemical and mechanical properties through the carefully design and

precise engineering of their structure and chemical composition on the nanometer scale.

The second challenge lies in achieving scalability in production, as the current rate of

production is rather limited, and insufficient to meet both the scale and the cost effectiveness

targets necessary for industrial applications. Existing up-scaled electrospinning setups

employing multi-needle arrays or needleless rotating drums are able to attain high throughputs

of solid nanofibers, but unable to be adapted to produce complex geometries necessary for the

hollow variety. Emulsion co-electrospinning may have had some success in creating

nanofibers of more complex geometries, but issues concerning the compatibility of the

constituent polymer, metal salt/alkoxide, and solvent to form the precursor emulsion limit the

variety of materials that can be synthesized by this sol-gel method. Optimized heat treatments

on the other hand are an integral part of the fabrication process, allowing the control over

product structures, which makes scaling the process up to the kilogram scale interesting to the

industry and vital to the push for adoption of these materials.

Thirdly, although LIB are the state-of-the-art in rechargeable battery technology, there is

still a substantial room for improvement in order to meet practical demands. One of the more

critical limitations affecting the use of hollow nanofibers is their lack of cycling stability,

thereby causing diminishing capacity over time. One potential solution to address this is

hybridization with carbon materials. Electrospun hollow carbon nanofibers are blessed with

many attractive features such as high electrical conductivity, a porous structure, and a large

surface area, which serve to prevent active material loss, and also to ensure good contact
73
between the active material and carbon during charge/discharge, thus achieving excellent long

term stability of the battery. One can also observe that much of the research work has been

focused on the anode component of these batteries and not the cathode, leaving much room

for the exploration of electrospun cathode materials in future research. Attaining the

maximum charge transport rates in LIBs can be potentially be achieved through vertical

patterning of electrospun hollow nanofibers, and could be explored by research groups.

Another potential research topic could be how to increase batterys volumetric capacity by

improving the density of the prepared materials via the growth of secondary structures on the

surfaces of the hollow nanofiber materials, wherein the findings could have implications in

the development of more compact, high energy density devices like flexible batteries for

wearable electronics. Lastly, the thermodynamic and kinetic basis behind the electrochemical

performance of these hollow nanostructures is relatively less well understood compared to

some of the other nanostructures, and should be explored through first principle theoretical

calculations and simulations, coupled with more experimental rigor through in-situ techniques

such as synchrotron X-ray, in-situ XRD, and in-situ Raman spectroscopy.

Fourthly, although electrospun hollow carbon nanofibers have been applied in Li-S

batteries, they have been significantly outperformed by many other metal compounds

prepared by other methods. Research should therefore be focused in that direction,

specifically on how to improve the performance of electrospun hollow micro/nanostructured

metal compounds and their derivatives as host for sulfur. To be effective hosts, the hollow

nanofiber should be composed of carbonaceous supporting scaffolds to provide the necessary

electrical conductivity and electrochemical reactivity. Also, since sulfur does not actively

contribute to the energy storage capacity, strategies should be developed to design the host

material in such a way as to minimize the sulfur content without impacting the performance.

Another key technical issue to resolve is to develop strategies allowing for the infusion of

74
sulfur at specific locations, necessitating the tailoring of the interior structure and surfaces of

the hollow host material.

Lastly, from the metal-air battery development perspective, the emergence of highly active

EHNF having controlled pore sizes as bifunctional catalysts for both ORR and OER is

promising, but still warrants a more in-depth investigation. The pores of these EHNF promote

the transport of all reactant species (electrons, Li+, and oxygen) to the active catalyst surface

and provide adequate space for the storage of the lithium oxide by products, thereby reducing

the considerable cathode overpotentials during charge/discharge appreciably. The catalysts

developed should also ideally have little to no effect in promoting the electrolyte

decomposition. To this end, carbon supported composite catalysts with rationally designed

nanostructures could prove a versatile material. Research could also look into how the

membrane-like properties can be utilized to develop high throughput, mechanically robust,

air-breathing membranes to function as sieves, separating oxygen from the ambient air to feed

the reaction. This prevents the accumulation of H2O, CO2 and other environmental

contaminants from adversely affecting the lifetime of the catalysts. Finally, much more work

needs to be done to deduce the fundamental reaction mechanisms governing Liair batteries

which could serve as a guide, empowering researchers to design and synthesize ever more

efficient cathode catalysts.

Although the development of electrospinning for hollow nanobers in advanced energy

device has reached an unprecedented level of success. However, it should be pointed out that

certain essential studies are still required and a number of technical issues remain to be

addressed. (i) We should design the hollow nanofibers more controllably. Uniform nanofibers

with diameters below 50 nm should be developed to provide short diffusion pathways and fast

ion kinetics. Modification of the hollow nanofiberes into hierarchical nanostructures or hybrid

structures (core-shell, 0D, 1D, 2D or 3D) with support materials can further enhance their

electrochemical performance. Furthermore, uniaxially aligned nanofibers with controlled size


75
should be produced to achieve maximum electron transport in energy and electronic devices.

(ii) Compared to polymer counterparts, electrospinning of inorganic nanobers is more

challenging due to the involvement of hydrolysis, condensation, and gelation reactions. Some

strategies for polymer nanofibers may be not useful for the synthesis of inorganic nanobers,

and it still needs great attention in the elaborate design before mass production. At the present

state, the synthesis of hollow nanofibers is mainly based on experiment. Furthermore,

understanding the mechanism of electrospinning hollow nanobers is of critical importance to

guide the further synthesis. (iii) Flexible devices received much attention in recent years.

Converting these advanced electrospun nanofibers especially carbon nanofibers to yarn as

directly as electrode materials is an important issue which should be considered. Therefore,

resultant nanofibers must have desirable properties, such as mechanical behavior,

conductivity, and thermal properties, which needs further deeply investigated.

From the research work consolidated and presented in this review, it is clear that the

electrospinning technique demonstrates the potential to fabricate hollow nanofibers with

superior properties though many challenges still remain. However, the signs are very

promising, with the superlative performances of various hollow nanostructured materials in

advanced battery systems giving us confidence that such nanoengineered hollow

nanostructures are one of the few approaches to realizing the high-performance batteries of

the future. Further refinements to the synthesis protocols, and progress made in studying this

class of materials will likely drive this field forward with strong momentum toward

commercialization in the near term. It is our hope that this review has provided valuable

insight to researchers embarking on their own work to advance the state of the art in the

innovation and application of hollow nanofibrous materials.

Acknowledgement
We are thankful for the nancial support from the Singapore National Research Foundation
(NRF-CRP10-2012-06), the Fundamental Research Funds for the Central Universities
(NE2017004), and China Jiangsu Specially Appointed Professor.

76
References:

[1] B. Dunn, H. Kamath, J. M. Tarascon, Science 334 (2011) 928-935.


[2] M. Winter, R. J. Brodd, Chem. Rev. 104 (2004) 4245-4269.
[3] J. M. Tarascon, M. Armand, Nature 414 (2001) 359-367.
[4] B. Scrosati, J. Garche, J. Power Sources 195 (2010) 2419-2430.
[5] M. Yilmaz, P. T. Krein, Ieee Transactions on Power Electronics 28 (2013) 2151-2169.
[6] S. W. Kim, D. H. Seo, X. H. Ma, G. Ceder, K. Kang, Adv. Energy Mater. 2 (2012)
710-721.
[7] A. Manthiram, S. H. Chung, C. Zu, Adv. Mater. 27 (2015) 1980-2006.
[8] P. G. Bruce, S. A. Freunberger, L. J. Hardwick, J. M. Tarascon, Nat. Mater. 11 (2012)
19-29.
[9] F. Y. Cheng, J. Liang, Z. L. Tao, J. Chen, Adv. Mater. 23 (2011) 1695-1715.
[10] S. Peng, L. Li, H. B. Wu, S. Madhavi, X. W. Lou, Adv. Energy Mater. 5 (2015)
1401172.
[11] M. Xu, L. F. Fei, W. B. Zhang, T. Li, W. Lu, N. Zhang, Y. Q. Lai, Z. Zhang, J. Fang,
K. Zhang, J. Li, H. T. Huang, Nano Lett. 17 (2017) 1670-1677.
[12] M. Xu, L. Fei, W. Lu, Z. Chen, T. Li, Y. Liu, G. Gao, Y. Lai, Z. Zhang, P. Wang, H.
Huang, Nano Energy 35 (2017) 271-280.
[13] S. J. Peng, G. R. Jin, L. L. Li, K. Li, M. Srinivasan, S. Ramakrishna, J. Chen, Chem.
Soc. Rev. 45 (2016) 1225-1241.
[14] A. Greiner, J. H. Wendorff, Angew. Chem. Int. Ed. 46 (2007) 5670-5703.
[15] T. J. Sill, H. A. von Recum, Biomaterials 29 (2008) 1989-2006.
[16] M. M. Arafat, B. Dinan, S. A. Akbar, A. S. M. A. Haseeb, Sensors 12 (2012) 7207-
7258.
[17] R. Sridhar, R. Lakshminarayanan, K. Madhaiyan, V. A. Barathi, K. H. C. Limh, S.
Ramakrishna, Chem. Soc. Rev. 44 (2015) 790-814.
[18] V. Thavasi, G. Singh, S. Ramakrishna, Energy Environ. Sci. 1 (2008) 205-221.
[19] Q. Wei, F. Xiong, S. Tan, L. Huang, E. H. Lan, B. Dunn, L. Mai, Adv. Mater. 29
(2017) 1602300.
[20] L. Zhou, Z. Zhuang, H. Zhao, M. Lin, D. Zhao, L. Mai, Adv. Mater. 29 (2017)
1602914.
[21] L. Yu, H. B. Wu, X. W. D. Lou, Accounts of Chemical Research 50 (2017) 293-301.
[22] Y. Dai, W. Liu, E. Formo, Y. Sun, Y. Xia, Polymers for Advanced Technologies 22
(2011) 326-338.
[23] S. Ramakrishna, K. Fujihara, W. E. Teo, T. Yong, Z. Ma, R. Ramaseshan, Materials
Today 9 (2006) 40-50.
[24] Z. M. Huang, Y. Z. Zhang, M. Kotaki, S. Ramakrishna, Composites Science and
Technology 63 (2003) 2223-2253.
[25] S. Cavaliere, S. Subianto, I. Savych, D. J. Jones, J. Roziere, Energy Environ. Sci. 4
(2011) 4761-4785.
[26] L. T. H. Nguyen, S. Chen, N. K. Elumalai, M. P. Prabhakaran, Y. Zong, C. Vijila, S. I.
Allakhverdiev, S. Ramakrishna, Macromol. Mater. Eng. 298 (2013) 822-867.
[27] X. Lu, C. Wang, Y. Wei, Small 5 (2009) 2349-2370.
[28] S. Ramakrishna, R. Jose, P. S. Archana, A. S. Nair, R. Balamurugan, J. Venugopal, W.
E. Teo, Journal of Materials Science 45 (2010) 6283-6312.
[29] L. Li, S. Peng, Y. Cheah, Y. Ko, P. Teh, G. Wee, C. Wong, M. Srinivasan, Chemistry-
a European Journal 19 (2013) 14823-14830.
[30] T. Krishnamoorthy, V. Thavasi, G. M. Subodh, S. Ramakrishna, Energy Environ. Sci.
4 (2011) 2807-2812.
77
[31] J. Jiang, Y. Li, J. Liu, X. Huang, C. Yuan, X. W. Lou, Adv. Mater. 24 (2012) 5166-
5180.
[32] X. W. Lou, L. A. Archer, Z. Yang, Adv. Mater. 20 (2008) 3987-4019.
[33] V. Aravindan, J. Gnanaraj, Y. S. Lee, S. Madhavi, Chem. Rev. 114 (2014) 11619-
11635.
[34] W. Wei, Z. Wang, Z. Liu, Y. Liu, L. He, D. Chen, A. Umar, L. Guo, J. Li, J. Power
Sources 238 (2013) 376-387.
[35] Z. Dong, S. J. Kennedy, Y. Wu, J. Power Sources 196 (2011) 4886-4904.
[36] Z. Li, H. B. Wu, X. W. Lou, Energy Environ. Sci. 9 (2016) 3061-3070.
[37] A. Kraytsberg, Y. Ein Eli, J. Power Sources 196 (2011) 886-893.
[38] S. Peng, L. Li, J. K. Y. Lee, L. Tian, M. Srinivasan, S. Adams, S. Ramakrishna, Nano
Energy 22 (2016) 361-395.
[39] L. Li, S. Peng, H. B. Wu, L. Yu, S. Madhavi, X. W. Lou, Adv. Energy Mater. 5 (2015)
1500753.
[40] S. J. Peng, L. L. Li, Y. X. Hu, M. Srinivasan, F. Y. Cheng, J. Chen, S. Ramakrishna,
ACS Nano 9 (2015) 1945-1954.
[41] J. J. Xu, D. Xu, Z. L. Wang, H. G. Wang, L. L. Zhang, X. B. Zhang, Angew. Chem.
Int. Ed. 52 (2013) 3887-3890.
[42] K. R. Yoon, G. Y. Lee, J. W. Jung, N. H. Kim, S. O. Kim, I. D. Kim, Nano Lett. 16
(2016) 2076-2083.
[43] L. Persano, A. Camposeo, C. Tekmen, D. Pisignano, Macromol. Mater. Eng. 298
(2013) 504-520.
[44] V. Beachley, X. Wen, Materials Science and Engineering: C 29 (2009) 663-668.
[45] C. J. Thompson, G. G. Chase, A. L. Yarin, D. H. Reneker, Polymer 48 (2007) 6913-
6922.
[46] B. Dong, L. Gwee, D. Salas de la Cruz, K. I. Winey, Y. A. Elabd, Nano Lett. 10 (2010)
3785-3790.
[47] H. Yang, C. R. Lightner, L. Dong, ACS Nano 6 (2012) 622-628.
[48] R. Sahay, P. S. Kumar, R. Sridhar, J. Sundaramurthy, J. Venugopal, S. G. Mhaisalkar,
S. Ramakrishna, J. Mater. Chem. 22 (2012) 12953-12971.
[49] B. Zhang, F. Kang, J. M. Tarascon, J. K. Kim, Progress in Materials Science 76 (2016)
319-380.
[50] M. Inagaki, Y. Yang, F. Kang, Adv. Mater. 24 (2012) 2547-2566.
[51] C. Niu, J. Meng, X. Wang, C. Han, M. Yan, K. Zhao, X. Xu, W. Ren, Y. Zhao, L. Xu,
Q. Zhang, D. Zhao, L. Mai, Nature Communications 6 (2015) 7402.
[52] X. Zhang, V. Thavasi, S. G. Mhaisalkar, S. Ramakrishna, Nanoscale 4 (2012) 1707-
1716.
[53] G. Yang, X. Xu, W. Yan, H. Yang, S. Ding, Electrochim. Acta 137 (2014) 462-469.
[54] W. E. Teo, S. Ramakrishna, Nanotechnology 17 (2006) R89-R106.
[55] R. Ramaseshan, S. Sundarrajan, R. Jose, S. Ramakrishna, Journal of Applied Physics
102 (2007).
[56] E. N. Kumar, R. Jose, P. S. Archana, C. Vijila, M. M. Yusoff, S. Ramakrishna, Energy
Environ. Sci. 5 (2012) 5401-5407.
[57] S. H. Choi, G. Ankonina, D. Y. Youn, S. G. Oh, J. M. Hong, A. Rothschild, I. D. Kim,
ACS Nano 3 (2009) 2623-2631.
[58] J. Liu, K. Tang, K. Song, P. A. van Aken, Y. Yu, J. Maier, Nanoscale 6 (2014) 5081-
5086.
[59] J. W. Jung, C. L. Lee, S. Yu, I. D. Kim, J. Mater. Chem. A 4 (2016) 703-750.
[60] A. L. Yarin, E. Zussman, Polymer 45 (2004) 2977-2980.
[61] A. L. Yarin, Polymers for Advanced Technologies 22 (2011) 310-317.
78
[62] D. Li, Y. Xia, Nano Lett. 4 (2004) 933-938.
[63] J. T. McCann, M. Marquez, Y. Xia, Nano Lett. 6 (2006) 2868-2872.
[64] Y. W. Liu, Q. L. Ma, M. Yang, X. T. Dong, Y. Yang, J. X. Wang, W. S. Yu, G. X. Liu,
Chem. Eng. J. 284 (2016) 831-840.
[65] S. H. Zhan, D. R. Chen, X. L. Jiao, C. H. Tao, J. Phys. Chem. B 110 (2006) 11199-
11204.
[66] Y. Yu, L. Gu, C. L. Wang, A. Dhanabalan, P. A. van Aken, J. Maier, Angew. Chem.
Int. Ed. 48 (2009) 6485-6489.
[67] D. Li, J. T. McCann, Y. N. Xia, Small 1 (2005) 83-86.
[68] H. Chen, N. Wang, J. Di, Y. Zhao, Y. Song, L. Jiang, Langmuir 26 (2010) 11291-
11296.
[69] Y. Zhao, X. Y. Cao, L. Jiang, J. Am. Chem. Soc. 129 (2007) 764-765.
[70] A. V. Bazilevsky, A. L. Yarin, C. M. Megaridis, Langmuir 23 (2007) 2311-2314.
[71] Y. Yu, L. Gu, C. B. Zhu, P. A. van Aken, J. Maier, J. Am. Chem. Soc. 131 (2009)
15984-15985.
[72] Z. Li, J. T. Zhang, Y. M. Chen, J. Li, X. W. Lou, Nature Communications 6 (2015)
8850.
[73] K. Tang, Y. Yu, X. Mu, P. A. van Aken, J. Maier, Electrochem. Commun. 28 (2013)
54-57.
[74] X. Xia, X. J. Dong, Q. F. Wei, Y. B. Cai, K. Y. Lu, Express Polymer Letters 6 (2012)
169-176.
[75] M. Buyukyazi, S. Mathur, Nano Energy 13 (2015) 28-35.
[76] J. S. Cho, Y. J. Hong, Y. C. Kang, Chem. Eur. J. 21 (2015) 11082-11087.
[77] Y. J. Hong, J. W. Yoon, J. H. Lee, Y. C. Kang, Chemistry-a European Journal 21
(2015) 371-376.
[78] L. Lang, D. Wu, Z. Xu, Chemistry-a European Journal 18 (2012) 10661-10668.
[79] J. Zhao, Y. Cheng, X. Yan, D. Sun, F. Zhu, Q. Xue, Crystengcomm 14 (2012) 5879-
5885.
[80] Y. Cheng, B. Zou, J. Yang, C. Wang, Y. Liu, X. Fan, L. Zhu, Y. Wang, H. Ma, X. Cao,
Crystengcomm 13 (2011) 2268-2272.
[81] Y. B. Yin, J. J. Xu, Q. C. Liu, X. B. Zhang, Adv. Mater. 28 (2016) 7494-7500.
[82] Y. Sun, R. B. Sills, X. Hu, Z. W. Seh, X. Xiao, H. Xu, W. Luo, H. Jin, Y. Xin, T. Li, Z.
Zhang, J. Zhou, W. Cai, Y. Huang, Y. Cui, Nano Lett. 15 (2015) 3899-3906.
[83] K. Yoon, Y. Yang, P. Lu, D. H. Wan, H. C. Peng, K. S. Masias, P. T. Fanson, C. T.
Campbell, Y. N. Xia, Angew. Chem. Int. Ed. 51 (2012) 9543-9546.
[84] Y. Chen, X. Li, K. Park, L. Zhou, H. Huang, Y. W. Mai, J. B. Goodenough, Angew.
Chem. Int. Ed. 55 (2016) 1583115834.
[85] C. S. Yan, G. Chen, X. Zhou, J. X. Sun, C. D. Lv, Adv. Funct. Mater. 26 (2016) 1428-
1436.
[86] J. K. Yoo, J. Kim, Y. S. Jung, K. Kang, Adv. Mater. 24 (2012) 5452-5456.
[87] K. Won Sik, L. Byoung Sun, K. Dai Hong, K. Hong Chan, Y. Woong Ryeol, H.
Seong Hyeon, Nanotechnology 21 (2010) 245605.
[88] V. Aravindan, J. Sundaramurthy, P. Suresh Kumar, Y.-S. Lee, S. Ramakrishna, S.
Madhavi, Chem. Commun. 51 (2015) 2225-2234.
[89] P. S. Kumar, J. Sundaramurthy, S. Sundarrajan, V. J. Babu, G. Singh, S. I.
Allakhverdiev, S. Ramakrishna, Energy Environ. Sci. 7 (2014) 3192-3222.
[90] S. Jayaraman, V. Aravindan, P. Suresh Kumar, W. Chui Ling, S. Ramakrishna, S.
Madhavi, ACS Appl. Mater. Interfaces 6 (2014) 8660-8666.
[91] X. Miao, H. Ni, X. Sha, T. Wang, J. Fang, G. Yang, Journal of Applied Polymer
Science 133 (2016) 43022.
79
[92] D. Ma, Y. Li, P. Zhang, A. J. Cooper, A. M. Abdelkader, X. Ren, L. Deng, J. Power
Sources 311 (2016) 35-41.
[93] Y. Gu, F. Jian, J. Phys. Chem. C 112 (2008) 20176-20180.
[94] E. Hosono, T. Saito, J. Hoshino, Y. Mizuno, M. Okubo, D. Asakura, K. Kagesawa, D.
Nishio Hamane, T. Kudo, H. Zhou, CrystEngComm 15 (2013) 2592-2597.
[95] S. Jayaraman, V. Aravindan, P. Suresh Kumar, W. C. Ling, S. Ramakrishna, S.
Madhavi, Chem. Commun. 49 (2013) 6677-6679.
[96] N. Arun, V. Aravindan, S. Jayaraman, N. Shubha, W. C. Ling, S. Ramakrishna, S.
Madhavi, Nanoscale 6 (2014) 8926-8934.
[97] B. B. Wei, Y. B. Wu, F. Y. Yu, Y. N. Zhou, International Journal of Minerals,
Metallurgy, and Materials 23 (2016) 474-480.
[98] R. Hagen, H. Lorrmann, K. C. Mller, S. Mathur, Adv. Energy Mater. 2 (2012) 553-
559.
[99] D. Shao, J. Wang, X. Dong, W. Yu, G. Liu, F. Zhang, L. Wang, Journal of Materials
Science: Materials in Electronics 24 (2013) 4718-4724.
[100] H. G. Wang, D. L. Ma, Y. Huang, X. B. Zhang, Chem. Eur. J. 18 (2012) 8987-8993.
[101] G. H. An, D. Y. Lee, H. J. Ahn, ACS Appl. Mater. Interfaces 8 (2016) 19466-19474.
[102] S. Zhan, Y. Li, H. Yu, Journal of Dispersion Science and Technology 29 (2008) 702-
705.
[103] S. Zhan, Y. Li, H. Yu, Journal of Dispersion Science and Technology 29 (2008) 823-
826.
[104] L. L. Cui, X. W. Miao, Y. F. Song, W. Y. Fang, H. B. Zhao, J. H. Fang, Advances in
Manufacturing 4 (2016) 79-88.
[105] G. Yang, L. Wang, J. Wang, W. Yan, Mater. Lett. 177 (2016) 13-16.
[106] K. C. Soo, K. Cheong, P. T. Jun, S. J. Tae, Chemistry Letters 41 (2012) 1428-1429.
[107] C. Kim, B. R. Kim, J. T. Son, Journal of Electroceramics 33 (2014) 7-11.
[108] W. Zeng, W. Chen, Z. Li, H. Zhang, T. Li, Materials Research Bulletin 65 (2015) 157-
162.
[109] T. H. Hwang, Y. M. Lee, B.-S. Kong, J.-S. Seo, J. W. Choi, Nano Lett. 12 (2012) 802-
807.
[110] H. Wu, G. Zheng, N. Liu, T. J. Carney, Y. Yang, Y. Cui, Nano Lett. 12 (2012) 904-
909.
[111] X. Zhou, L. J. Wan, Y. G. Guo, Small 9 (2013) 2684-2688.
[112] C. Chen, K. Fu, Y. Lu, J. Zhu, L. Xue, Y. Hu, X. Zhang, RSC Advances 5 (2015)
30793-30800.
[113] H. Han, T. Song, J. Y. Bae, L. F. Nazar, H. Kim, U. Paik, Energy Environ. Sci. 4
(2011) 4532-4536.
[114] X. Sun, Y. Huang, M. Zong, H. Wu, X. Ding, Journal of Materials Science: Materials
in Electronics 27 (2016) 2682-2686.
[115] D. Pham Cong, J. S. Park, J. H. Kim, J. Kim, P. V. Braun, J. H. Choi, S. J. Kim, S. Y.
Jeong, C. R. Cho, Carbon 111 (2017) 28-37.
[116] J. S. Cho, Y. C. Kang, Small 11 (2015) 4673-4681.
[117] J. S. Cho, Y. J. Hong, Y. C. Kang, ACS Nano 9 (2015) 4026-4035.
[118] H. G. Wang, Y. Zhou, Y. Shen, Y. Li, Q. Zuo, Q. Duan, Electrochim. Acta 158 (2015)
105-112.
[119] H. Luo, K. Huang, B. Sun, J. Zhong, Electrochim. Acta 149 (2014) 11-17.
[120] M. E. Im, D. Pham-Cong, J. Y. Kim, H. S. Choi, J. H. Kim, J. P. Kim, J. Kim, S. Y.
Jeong, C. R. Cho, J. Power Sources 284 (2015) 392-399.
[121] Z. Cui, S. Wang, Y. Zhang, M. Cao, Electrochim. Acta 182 (2015) 507-515.
[122] Y. M. Chen, L. Yu, X. W. Lou, Angew. Chem. Int. Ed. 55 (2016) 5990-5993.
80
[123] C. Yan, G. Chen, J. Sun, X. Zhou, C. Lv, Phys. Chem. Chem. Phys. 18 (2016) 19531-
19535.
[124] G. H. An, J. I. Sohn, H. J. Ahn, J. Mater. Chem. A 4 (2016) 2049-2054.
[125] L. Li, S. Peng, J. Wang, Y. L. Cheah, P. Teh, Y. Ko, C. Wong, M. Srinivasan, ACS
Appl. Mater. Interfaces 4 (2012) 6005-6012.
[126] L. Li, S. Peng, Y. L. Cheah, J. Wang, P. Teh, Y. Ko, C. Wong, M. Srinivasan,
Nanoscale 5 (2013) 134-138.
[127] Y. J. Hong, J. S. Cho, Y. C. Kang, Chem. Eur. J. 21 (2015) 18202-18208.
[128] H. G. Wang, D. Liu, Y. Li, Q. Duan, Mater. Lett. 172 (2016) 64-67.
[129] S. Peng, L. Li, M. Srinivasan, Journal of Energy Chemistry 23 (2014) 301-307.
[130] W. Luo, X. Hu, Y. Sun, Y. Huang, J. Mater. Chem. 22 (2012) 8916-8921.
[131] J. Meng, C. Niu, X. Liu, Z. Liu, H. Chen, X. Wang, J. Li, W. Chen, X. Guo, L. Mai,
Nano Research 9 (2016) 2445-2457.
[132] S. M. Hwang, S. Y. Kim, J. G. Kim, K. J. Kim, J. W. Lee, M. S. Park, Y. J. Kim, M.
Shahabuddin, Y. Yamauchi, J. H. Kim, Nanoscale 7 (2015) 8351-8355.
[133] J. Zhu, Z. Xu, B. Lu, Nano Energy 7 (2014) 114-123.
[134] Y. M. Chen, X. Y. Li, K. Park, J. Song, J. H. Hong, L. M. Zhou, Y. W. Mai, H. T.
Huang, J. B. Goodenough, J. Am. Chem. Soc. 135 (2013) 16280-16283.
[135] Y. Chen, X. Li, X. Zhou, H. Yao, H. Huang, Y.-W. Mai, L. Zhou, Energy Environ. Sci.
7 (2014) 2689-2696.
[136] N. Kaerkitcha, S. Chuangchote, T. Sagawa, Nanoscale Research Letters 11 (2016) 186.
[137] J. Ning, M. Yang, H. Yang, Z. Xu, Materials & Design 109 (2016) 264-269.
[138] J. T. McCann, B. Lim, R. Ostermann, M. Rycenga, M. Marquez, Y. Xia, Nano Lett. 7
(2007) 2470-2474.
[139] J. T. McCann, M. Marquez, Y. Xia, J. Am. Chem. Soc. 128 (2006) 1436-1437.
[140] B. S. Lee, K. M. Park, W. R. Yu, J. H. Youk, Macromolecular Research 20 (2012)
605-613.
[141] B. S. Lee, S. B. Son, K. M. Park, W. R. Yu, K. H. Oh, S. H. Lee, J. Power Sources
199 (2012) 53-60.
[142] B. S. Lee, S. B. Son, K. M. Park, G. Lee, K. H. Oh, S. H. Lee, W. R. Yu, ACS Appl.
Mater. Interfaces 4 (2012) 6702-6710.
[143] B. S. Lee, H. S. Yang, H. Jung, S. K. Mah, S. Kwon, J. H. Park, K. H. Lee, W. R. Yu,
S. G. Doo, European Polymer Journal 70 (2015) 392-399.
[144] Y. Chen, Z. Lu, L. Zhou, Y. W. Mai, H. Huang, Nanoscale 4 (2012) 6800-6805.
[145] Y. Chen, Z. Lu, L. Zhou, Y. W. Mai, H. Huang, Energy Environ. Sci. 5 (2012) 7898-
7902.
[146] W. Li, Z. Yang, Y. Jiang, Z. Yu, L. Gu, Y. Yu, Carbon 78 (2014) 455-462.
[147] C. Zhan, Q. Xu, X. Yu, Q. Liang, Y. Bai, Z. H. Huang, F. Kang, RSC Advances 6
(2016) 41473-41476.
[148] F. Wu, L. Shi, D. Mu, H. Xu, B. Wu, Carbon 86 (2015) 146-155.
[149] D. Nan, Z. H. Huang, R. Lv, L. Yang, J. G. Wang, W. Shen, Y. Lin, X. Yu, L. Ye, H.
Sun, F. Kang, J. Mater. Chem. A 2 (2014) 19678-19684.
[150] J. Ryu, S. Choi, T. Bok, S. Park, Nanoscale 7 (2015) 6126-6135.
[151] 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.
[152] J. Wang, Y. Yu, L. Gu, C. Wang, K. Tang, J. Maier, Nanoscale 5 (2013) 2647-2650.
[153] J. Kong, W. A. Yee, Y. Wei, L. Yang, J. M. Ang, S. L. Phua, S. Y. Wong, R. Zhou, Y.
Dong, X. Li, X. Lu, Nanoscale 5 (2013) 2967-2973.
[154] D. Lee, B. Kim, J. Kim, S. Jeong, G. Cao, J. Moon, ACS Appl. Mater. Interfaces 7
(2015) 27234-27241.
81
[155] 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) 5989-5998.
[156] J. Wu, X. Qin, C. Miao, Y. B. He, G. Liang, D. Zhou, M. Liu, C. Han, B. Li, F. Kang,
Carbon 98 (2016) 582-591.
[157] Y. Xu, Y. Zhu, F. Han, C. Luo, C. Wang, Adv. Energy Mater. 5 (2015) 1400753.
[158] Z. Yang, Q. Meng, W. Yan, J. Lv, Z. Guo, X. Yu, Z. Chen, T. Guo, R. Zeng, Energy
82 (2015) 960-967.
[159] I. Meschini, F. Nobili, M. Mancini, R. Marassi, R. Tossici, A. Savoini, M. L. Focarete,
F. Croce, J. Power Sources 226 (2013) 241-248.
[160] G. Zhang, J. Zhu, W. Zeng, S. Hou, F. Gong, F. Li, C. C. Li, H. Duan, Nano Energy 9
(2014) 61-70.
[161] L. Xue, X. Xia, T. Tucker, K. Fu, S. Zhang, S. Li, X. Zhang, J. Mater. Chem. A 1
(2013) 13807-13813.
[162] Shilpa, A. Sharma, Electrochim. Acta 176 (2015) 1266-1271.
[163] P. S. Kumar, S. A. S. Nizar, J. Sundaramurthy, P. Ragupathy, V. Thavasi, S. G.
Mhaisalkar, S. Ramakrishna, J. Mater. Chem. 21 (2011) 9784-9790.
[164] P. Zhu, A. S. Nair, S. Peng, S. Yang, S. Ramakrishna, ACS Appl. Mater. Interfaces 4
(2012) 581-585.
[165] J. G. Kim, D. Shi, K. J. Kong, Y. U. Heo, J. H. Kim, M. R. Jo, Y. C. Lee, Y. M. Kang,
S. X. Dou, ACS Appl. Mater. Interfaces 5 (2013) 691-696.
[166] T. Yuan, B. Zhao, R. Cai, Y. Zhou, Z. Shao, J. Mater. Chem. 21 (2011) 15041-15048.
[167] L. Zuniga, V. Agubra, D. Flores, H. Campos, J. Villareal, M. Alcoutlabi, Journal of
Alloys and Compounds 686 (2016) 733-743.
[168] L. Li, X. Yin, S. Liu, Y. Wang, L. Chen, T. Wang, Electrochem. Commun. 12 (2010)
1383-1386.
[169] D. Zhou, W. L. Song, L. Z. Fan, ACS Appl. Mater. Interfaces 7 (2015) 21472-21478.
[170] D. Pham Cong, J. Y. Kim, J. S. Park, J. H. Kim, J. P. Kim, E. D. Jeong, J. Kim, S. Y.
Jeong, C. R. Cho, Electrochim. Acta 161 (2015) 1-9.
[171] H. Zhou, Z. Li, Y. Qiu, X. Xia, Journal of Alloys and Compounds 670 (2016) 35-40.
[172] F. Dong, L. Licheng, X. Weilin, L. Guangzhong, L. Zhiping, Z. Yingsan, X. Jie, X.
Chuanxi, Materials Research Express 1 (2014) 025012.
[173] L. Li, S. Peng, Y. Cheah, Y. Ko, P. Teh, G. Wee, C. Wong, M. Srinivasan, Chem. Eur.
J. 19 (2013) 14823-14830.
[174] L. Li, S. Peng, Y. Cheah, P. Teh, J. Wang, G. Wee, Y. Ko, C. Wong, M. Srinivasan,
Chem. Eur. J. 19 (2013) 5892-5898.
[175] S. Chaudhari, M. Srinivasan, J. Mater. Chem. 22 (2012) 23049-23056.
[176] T. Yuan, Y. Jiang, W. Sun, B. Xiang, Y. Li, M. Yan, B. Xu, S. Dou, Adv. Funct.
Mater. 26 (2016) 2198-2206.
[177] X. Yan, X. Tong, J. Wang, C. Gong, M. Zhang, L. Liang, Mater. Lett. 136 (2014) 74-
77.
[178] J. S. Cho, S. Y. Lee, H. S. Ju, Y. C. Kang, ACS Appl. Mater. Interfaces 7 (2015)
25641-25647.
[179] X. Liang, G. Jin, F. Liu, X. Zhang, S. An, J. Ma, G. Lu, Ceramics International 41
(2015) 13780-13787.
[180] Y. Jiang, M. Hu, D. Zhang, T. Yuan, W. Sun, B. Xu, M. Yan, Nano Energy 5 (2014)
60-66.
[181] Y. Wu, P. Zhu, X. Zhao, M. V. Reddy, S. Peng, B. V. R. Chowdari, S. Ramakrishna, J.
Mater. Chem. A 1 (2013) 852-859.
[182] C. Lv, J. Sun, G. Chen, C. Yan, D. Chen, Nano Energy 33 (2017) 138-145.

82
[183] W. Li, Y. H. Wu, J. W. Wang, D. Huang, L. Z. Chen, G. Yang, European Polymer
Journal 67 (2015) 365-372.
[184] R. H. Liu, L. N. Jin, D. Lu, G. Hu, L. He, J. Shan, Electrochim. Acta 114 (2013) 372-
378.
[185] P. Raghavan, X. H. Zhao, J. K. Kim, J. Manuel, G. S. Chauhan, J. H. Ahn, C. Nah,
Electrochim. Acta 54 (2008) 228-234.
[186] P. Raghavan, X. H. Zhao, J. Manuel, G. S. Chauhan, J. H. Ahn, H. S. Ryu, H. J. Ahn,
K. W. Kim, C. Nah, Electrochim. Acta 55 (2010) 1347-1354.
[187] N. Wu, Q. Cao, X. Y. Wang, S. Li, X. Y. Li, H. Y. Deng, J. Power Sources 196 (2011)
9751-9756.
[188] N. Wu, Q. Cao, X. Y. Wang, Q. Q. Chen, Solid State Ionics 203 (2011) 42-46.
[189] X. L. Tang, Q. Cao, X. Y. Wang, X. X. Peng, J. Zeng, Rsc Advances 5 (2015) 58655-
58662.
[190] X. E. Wang, H. J. Zhu, G. W. Greene, J. Y. Li, N. Iranipour, C. Garnier, J. Fang, M.
Armand, M. Forsyth, J. M. Pringle, P. C. Howlett, J. Mater. Chem. A 4 (2016) 9873-
9880.
[191] H. R. Jung, D. H. Ju, W. J. Lee, X. W. Zhang, R. Kotek, Electrochim. Acta 54 (2009)
3630-3637.
[192] P. Carol, P. Ramakrishnan, B. John, G. Cheruvally, J. Power Sources 196 (2011)
10156-10162.
[193] P. Raghavan, X. Zhao, C. Shin, D. H. Baek, J. W. Choi, J. Manuel, M. Y. Heo, J. H.
Ahn, C. Nah, J. Power Sources 195 (2010) 6088-6094.
[194] H. R. Jung, W. J. Lee, Electrochim. Acta 58 (2011) 674-680.
[195] Q. J. Wang, W. L. Song, L. N. Wang, Y. Song, Q. Shi, L. Z. Fan, Electrochim. Acta
132 (2014) 538-544.
[196] N. Yabuuchi, K. Kubota, M. Dahbi, S. Komaba, Chem. Rev. 114 (2014) 11636-11682.
[197] H. Pan, Y.-S. Hu, L. Chen, Energy Environ. Sci. 6 (2013) 2338-2360.
[198] J. S. Cho, J. K. Lee, Y. C. Kang, Scientific Reports 6 (2016) 23699.
[199] Y. Liu, N. Zhang, L. Jiao, J. Chen, Adv. Mater. 27 (2015) 6702-6707.
[200] C. Fang, Y. Huang, W. Zhang, J. Han, Z. Deng, Y. Cao, H. Yang, Adv. Energy Mater.
6 (2016) 1501727.
[201] L. Mu, S. Xu, Y. Li, Y. S. Hu, H. Li, L. Chen, X. Huang, Adv. Mater. 27 (2015) 6928-
6933.
[202] W. Li, L. Zeng, Z. Yang, L. Gu, J. Wang, X. Liu, J. Cheng, Y. Yu, Nanoscale 6 (2014)
693-698.
[203] H. Y. Kang, Y. C. Liu, K. Z. Cao, Y. Zhao, L. F. Jiao, Y. J. Wang, H. T. Yuan, J.
Mater. Chem. A 3 (2015) 17899-17913.
[204] Y. Zhou, W. Sun, X. Rui, Y. Zhou, W. J. Ng, Q. Yan, E. Fong, Nano Energy 21 (2016)
71-79.
[205] Y. Liu, N. Zhang, C. Yu, L. Jiao, J. Chen, Nano Lett. 16 (2016) 3321-3328.
[206] R. Alcntara, M. Jaraba, P. Lavela, J. L. Tirado, Chem. Mat. 14 (2002) 2847-2848.
[207] C. Zhu, X. Mu, P. A. van Aken, Y. Yu, J. Maier, Angew. Chem. Int. Ed. 53 (2014)
2152-2156.
[208] S. Peng, X. Han, L. Li, Z. Zhu, F. Cheng, M. U. Srinivansan, S. Adams, S.
Ramakrishna, Small 12 (2016) 1359-1368.
[209] Z. Hu, Z. Zhu, F. Cheng, K. Zhang, J. Wang, C. Chen, J. Chen, Energy Environ. Sci. 8
(2015) 1309-1316.
[210] Y. Xiao, S. H. Lee, Y. K. Sun, Adv. Energy Mater. (2016) 1601329.
[211] Y. J. Hong, J. H. Kim, Y. C. Kang, J. Mater. Chem. A 4 (2016) 15471-15477.

83
[212] M. Q. Zhao, X. F. Liu, Q. Zhang, G. L. Tian, J. Q. Huang, W. Zhu, F. Wei, ACS Nano
6 (2012) 10759-10769.
[213] A. Manthiram, Y. Fu, Y. S. Su, Accounts of Chemical Research 46 (2013) 1125-1134.
[214] Z. Li, J. Zhang, X. W. Lou, Angew. Chem. Int. Ed. 54 (2015) 12886-12890.
[215] Z. Yuan, H. J. Peng, J. Q. Huang, X. Y. Liu, D. W. Wang, X. B. Cheng, Q. Zhang,
Adv. Funct. Mater. 24 (2014) 6105-6112.
[216] A. Manthiram, Y. Fu, S. H. Chung, C. Zu, Y. S. Su, Chem. Rev. 114 (2014) 11751-
11787.
[217] L. Ji, M. Rao, S. Aloni, L. Wang, E. J. Cairns, Y. Zhang, Energy Environ. Sci. 4 (2011)
5053-5059.
[218] R. Cao, W. Xu, D. Lv, J. Xiao, J. G. Zhang, Adv. Energy Mater. 5 (2015) 1402273.
[219] J. Lu, L. Li, J. B. Park, Y. K. Sun, F. Wu, K. Amine, Chem. Rev. 114 (2014) 5611-
5640.
[220] F. Cheng, J. Chen, Chem. Soc. Rev. 41 (2012) 2172-2192.
[221] H. W. Park, D. U. Lee, P. Zamani, M. H. Seo, L. F. Nazar, Z. Chen, Nano Energy 10
(2014) 192-200.
[222] M. A. Rahman, X. Wang, C. Wen, J. Electrochem. Soc. 160 (2013) A1759-A1771.
[223] R. Padbury, X. Zhang, J. Power Sources 196 (2011) 4436-4444.
[224] Z. Peng, S. A. Freunberger, Y. Chen, P. G. Bruce, Science 337 (2012) 563-566.
[225] M. M. O. Thotiyl, S. A. Freunberger, Z. Peng, P. G. Bruce, J. Am. Chem. Soc. 135
(2013) 494-500.
[226] J. Huang, B. Zhang, Y. Y. Xie, W. W. K. Lye, Z. L. Xu, S. Abouali, M. Akbari
Garakani, J. Q. Huang, T. Y. Zhang, B. Huang, J. K. Kim, Carbon 100 (2016) 329-336.
[227] X. Han, Y. Hu, J. Yang, F. Cheng, J. Chen, Chem. Commun. 50 (2014) 1497-1499.
[228] G. Liu, H. Chen, L. Xia, S. Wang, L. X. Ding, D. Li, K. Xiao, S. Dai, H. Wang, ACS
Appl. Mater. Interfaces 7 (2015) 22478-22486.
[229] L. Li, L. Shen, P. Nie, G. Pang, J. Wang, H. Li, S. Dong, X. Zhang, J. Mater. Chem. A
3 (2015) 24309-24314.
[230] G. S. Park, J. S. Lee, S. T. Kim, S. Park, J. Cho, J. Power Sources 243 (2013) 267-273.
[231] W. H. Ryu, T. H. Yoon, S. H. Song, S. Jeon, Y. J. Park, I. D. Kim, Nano Lett. 13
(2013) 4190-4197.
[232] J. S. Lee, S. Tai Kim, R. Cao, N. S. Choi, M. Liu, K. T. Lee, J. Cho, Adv. Energy
Mater. 1 (2011) 34-50.

Dr. Linlin Li is currently an associate professor in Nanjing University of Aeronautics and


Astronautics (NUAA), China. She completed her Ph.D from Nanyang Technological
University of Singapore in 2015. She received her M.S. degree from Nankai University of
China in 2011. She is also a recipient of Chinese Government Award for Outstanding Self-
Financed Student Abroad (2014) and Ian Ferguson Postgraduate Fellowship (2014). Her

84
research interests focus on the synthesis of functional nanomaterials and their application in
energy storage devices such as rechargeable batteries, supercapacitors and advanced batteries.

Dr. Shengjie Peng received his PhD degree in Nankai University (P.R. China) in 2010. He is
now working as a senior research fellow in Prof. Seeram's group in National University of
Singapore. He has been working on functional nanomaterials in energy and catalysis for more
than ten years. He has coauthored over 75 peer-reviewed publications. His current research
interests include the design and development of nanomaterials and their applications in energy
and catalysis.

Mr. Jeremy Lee Kong Yoong graduated in 2014 from the National University of Singapore
with a B. Eng Degree in Materials Science and Engineering (First Class Honours). He now
works as a research engineer at the Lloyds Register Global Technology Centre and is a PhD
candidate under Professor Seeram as part of the NUS Centre for Nanofibers and
Nanotechnology. His current research interests lie mainly in the use of nanomaterials for
energy applications, as well as metal matrix nanocomposites.

85
Mr. Dongxiao Ji is a Ph.D candidate in the College of Textiles in Donghua University, China.
He is currently a visiting student in the Department of Mechanical Engineering at National
University of Singapore under the supervision of Prof. Seeram Ramakrishna. His research
mainly focuses on large-scale electrospinning technique and electrospun flexible materials for
self-standing electrocatalysts and energy storage electronic devices.

Prof. Dr. Madhavi Srinivasan is currently a Professor at the School of Materials Science and
Engineering, Nanyang Technological University (NTU), Singapore. She graduated from
Indian Institute of Technology (IIT), Chennai (India), and completed her Ph.D. from the
National University of Singapore. Her research interest is to enhance the performance of
energy storage devices such as lithium ion batteries, supercapacitors, and advanced batteries
with the help of multifunctional nanoscale materials to power printed electronics, to store
energy from renewable sources, and for powering electric vehicles. Her focus is on the
fabrication and investigation of nanoscale materials/architectures for electrochemical energy
storage devices.

Prof. Dr. Seeram Ramakrishna, FREng, is the Director of Center for Nanofibers &
Nanotechnology at the National University of Singapore. He received his PhD from the
University of Cambridge. He is among the most cited researchers working in the eld of
materials and electrospinning, and nanobers engineering. He is a Highly Cited Researcher in
Materials Science (www.highlycited.com). He authored 6 books and ~ 700 ISI listed journal
papers, which attracted ~ 48,000 citations and ~ 101 H-index. He received several awards and
recognitions worldwide. His research outcomes have been translated into products and
available in several countries.

86
Research Highlights:
The present advanced secondary batteries are reviewed.
The synthesis of electrospun nanofibers is emphasized.
The relationship between the structures and performance is discussed.
The remaining challenges are suggested as well.

87

Вам также может понравиться