Nano Dimensionality: A Way towards Better Li-Ion Storage

15
Send Orders for Reprints to [email protected] Nanoscience & Nanotechnology-Asia, 2013, 3, 21-35 21 Nano Dimensionality: A Way towards Better Li-Ion Storage Uttam Kumar Sen, Sudeep Sarkar, Pavan Srinivas Veluri, Shivani Singh and Sagar Mitra * Electrochemical Energy Laboratory, Department of Energy Science and Engineering, Indian Institute of Technology Bombay, Mumbai, 400076, India Abstract: Nanodimensional materials such as transition metal oxides, polyanionic based materials and metal fluorides can be used as cathode while metal nanoparticles, alloys and metal oxides are preferred as anode for next-generation lithium- ion batteries (LIBs) in order to obtain high reversible capacity, rate capability, safety, and longer cycle life. These nanomaterials can offer relatively short ionic and electronic pathways which leads to better transportation of both lithium ions and electrons to the particles core. This article emphasize on the effect of nanodimension on the electrochemical performance of cathode and anode materials. Their synthesis processes, electrochemical properties and electrode reaction mechanisms are briefly discussed and summarized. Furthermore, the article highlights recent past scientific works and new progresses in the field of LIBs. It also highlights the direction to overcome the existing issues of current lithium storage technology. In future, we may overcome all the existing issues of LIBs and can deliver excellent cathode and anode combinations to fulfill maximum practical efficiency with low cost and ultimate safety for high end applications. Keywords: Anode, cathode, lithium-ion battery, nanodimension, nanomaterials. 1. INTRODUCTION One of the major innovations of 21 st century is the fabrication of materials in nanodimensions, which led to the miniaturization in micro-electronics. The rapid development of consumer electronics needs high energy density and high power density materials for energy storage devices. Lithium- ion battery technology is an efficient power source owing to its high potential, energy density, and long cycle life. The concept of lithium rechargeable batteries was first demonstrated using layered TiS 2 as cathode and metallic lithium as anode. Layered structure of TiS 2 enables the insertion of Li-ions into the van der waals gap between sulphide layers [1]. Upon charge/discharge cycling process, the layered structure of TiS 2 is well maintained exhibiting good reversibility of the system [1]. Although extensive research had been done on several insertion compounds during 1970 and 1980s, they had not been commercialized until 1990s. High reactivity of lithium with the non-aqueous electrolytes facilitates passivation film on the surface of metallic lithium, which prevented the commercialization of LIBs [2, 3]. The surface passivation film prevents bulk material to further react with the electrolyte, leading to plating of lithium (dendrite formation) at low temperatures during charging. Dendrite formation, results in short circuiting of the cell, which imposes safety hazards due to local overheating. The problems associated with the metallic lithium anode lead to the quest for new anode materials. In 1983, R. Yazami et al. demonstrated the reversible electrochemical intercalation *Address correspondence to this author at the Electrochemical Energy Laboratory, Department of Energy Science and Engineering, Indian Institute of Technology Bombay, Mumbai, 400076, India; Tel: + 91-222576-7849; Fax: +91-22-2576-4890; E-mail: [email protected] of lithium in graphite [4]. The first commercial LIB was launched by Sony in 1991 in which graphite was used as anode, and LiCoO 2 as cathode [5]. Graphite came up as an alternative anode to metallic lithium, which can intercalate lithium reversibly at ~0.2 V vs. Li/Li + . Electrodes made of layered structure in which lithium ions act as a guest molecule during charge/discharge, offer significant advantages in terms of safety and cycle life (Schematic representation of LIB is shown in Fig. 1a). Electrochemical insertion/de-insertion in LIB follows three vital steps: (i) lithium-ion diffusion within electrode materials, (ii) charge transfer at electrode/electrolyte interface and (iii) lithium-ion percolation in electrolyte. The initial step is pivotal amongst all the other steps. Therefore, minimal diffusion length enhances the kinetics of lithium-ion within the material, which can be achieved by reducing dimension of the particles. Smaller dimensions can be achieved by modifying the morphology of the material such as (i) zero-dimensional spherical particles, (ii) one-dimensional rod/tube, and (iii) two-dimensional nanoplates or nanosheets [6]. Moreover, nanosized materials endowed with high surface area provide better electrode/electrolyte contact area, and also better strain accommodation, mechanical integrity during lithium insertion/deinsertion. It is a profound fact that by engineering nanostructuring of a material, reaction properties can significantly be improved. Though several reviews on different aspects of LIBs have appeared in the literature, however, present review is focused to demonstrate the effect of nanosized materials on the electrochemical performance of LIBs. 1.1. Advantages and Disadvantages of Nanodimensional Materials Particle size reduction and nanostructuring of materials have attracted great interest in recent years because of the 2210-6820/13 $58.00+.00 © 2013 Bentham Science Publishers

Transcript of Nano Dimensionality: A Way towards Better Li-Ion Storage

Send Orders for Reprints to [email protected]

Nanoscience & Nanotechnology-Asia, 2013, 3, 21-35 21

Nano Dimensionality: A Way towards Better Li-Ion Storage

Uttam Kumar Sen, Sudeep Sarkar, Pavan Srinivas Veluri, Shivani Singh and Sagar Mitra*

Electrochemical Energy Laboratory, Department of Energy Science and Engineering, Indian Institute of Technology

Bombay, Mumbai, 400076, India

Abstract: Nanodimensional materials such as transition metal oxides, polyanionic based materials and metal fluorides can

be used as cathode while metal nanoparticles, alloys and metal oxides are preferred as anode for next-generation lithium-

ion batteries (LIBs) in order to obtain high reversible capacity, rate capability, safety, and longer cycle life. These

nanomaterials can offer relatively short ionic and electronic pathways which leads to better transportation of both lithium

ions and electrons to the particles core. This article emphasize on the effect of nanodimension on the electrochemical

performance of cathode and anode materials. Their synthesis processes, electrochemical properties and electrode reaction

mechanisms are briefly discussed and summarized. Furthermore, the article highlights recent past scientific works and

new progresses in the field of LIBs. It also highlights the direction to overcome the existing issues of current lithium

storage technology. In future, we may overcome all the existing issues of LIBs and can deliver excellent cathode and

anode combinations to fulfill maximum practical efficiency with low cost and ultimate safety for high end applications.

Keywords: Anode, cathode, lithium-ion battery, nanodimension, nanomaterials.

1. INTRODUCTION

One of the major innovations of 21st century is the

fabrication of materials in nanodimensions, which led to the

miniaturization in micro-electronics. The rapid development

of consumer electronics needs high energy density and high

power density materials for energy storage devices. Lithium-

ion battery technology is an efficient power source owing to

its high potential, energy density, and long cycle life.

The concept of lithium rechargeable batteries was first

demonstrated using layered TiS2 as cathode and metallic

lithium as anode. Layered structure of TiS2 enables the

insertion of Li-ions into the van der waals gap between

sulphide layers [1]. Upon charge/discharge cycling process,

the layered structure of TiS2 is well maintained exhibiting

good reversibility of the system [1]. Although extensive

research had been done on several insertion compounds

during 1970 and 1980s, they had not been commercialized

until 1990s. High reactivity of lithium with the non-aqueous

electrolytes facilitates passivation film on the surface of

metallic lithium, which prevented the commercialization of

LIBs [2, 3]. The surface passivation film prevents bulk

material to further react with the electrolyte, leading to

plating of lithium (dendrite formation) at low temperatures

during charging. Dendrite formation, results in short

circuiting of the cell, which imposes safety hazards due to

local overheating.

The problems associated with the metallic lithium anode

lead to the quest for new anode materials. In 1983, R. Yazami

et al. demonstrated the reversible electrochemical intercalation

*Address correspondence to this author at the Electrochemical Energy

Laboratory, Department of Energy Science and Engineering, Indian Institute

of Technology Bombay, Mumbai, 400076, India; Tel: + 91-222576-7849;

Fax: +91-22-2576-4890; E-mail: [email protected]

of lithium in graphite [4]. The first commercial LIB was

launched by Sony in 1991 in which graphite was used as

anode, and LiCoO2 as cathode [5]. Graphite came up as an

alternative anode to metallic lithium, which can intercalate

lithium reversibly at ~0.2 V vs. Li/Li+. Electrodes made of

layered structure in which lithium ions act as a guest

molecule during charge/discharge, offer significant advantages

in terms of safety and cycle life (Schematic representation of

LIB is shown in Fig. 1a).

Electrochemical insertion/de-insertion in LIB follows

three vital steps: (i) lithium-ion diffusion within electrode

materials, (ii) charge transfer at electrode/electrolyte

interface and (iii) lithium-ion percolation in electrolyte. The

initial step is pivotal amongst all the other steps. Therefore,

minimal diffusion length enhances the kinetics of lithium-ion

within the material, which can be achieved by reducing

dimension of the particles. Smaller dimensions can be

achieved by modifying the morphology of the material such

as (i) zero-dimensional spherical particles, (ii) one-dimensional

rod/tube, and (iii) two-dimensional nanoplates or nanosheets

[6]. Moreover, nanosized materials endowed with high

surface area provide better electrode/electrolyte contact area,

and also better strain accommodation, mechanical integrity

during lithium insertion/deinsertion. It is a profound fact that

by engineering nanostructuring of a material, reaction

properties can significantly be improved.

Though several reviews on different aspects of LIBs have

appeared in the literature, however, present review is focused

to demonstrate the effect of nanosized materials on the

electrochemical performance of LIBs.

1.1. Advantages and Disadvantages of Nanodimensional Materials

Particle size reduction and nanostructuring of materials

have attracted great interest in recent years because of the

2210-6820/13 $58.00+.00 © 2013 Bentham Science Publishers

22 Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 Sen et al.

remarkable properties obtained from reducing dimensions of

such materials as predicted from size scaling laws [7]. Thus,

nanostructured materials have drawn a tremendous amount

of attention in electrochemical energy storage devices. The

intrinsic properties of bulk and nanomaterial are illustrated in

Fig. (1b).

The basic reasons for the improvement in the electro-

chemical behavior by nanostructured materials include:

(i) better kinetics owing to their large surface area for

Faradaic reaction leading to high rate performance, (ii) short

diffusion path length for lithium-ion transport and (iii) better

accommodation of strain due to lithium-ion insertion/

deinsertion.

Recently, M. Gaberscek et al. [8] showed an empirical

relation that the electrode resistance (Rm) as a function of

particle size (d).

Rm =A dn ,

where A and n are fitted parameter. M. Gaberscek et al. [8] showed that empirical formula follows the power law

with n = 2, which is generally valid for the low-conductivity

species. As, it is quite evident from the above relation that

the charge transfer resistance decreases with nanodimension

material leading to shorter diffusion path length for lithium-

ion transfer.

With introduction of nanodimensionality in electro-

chemical storage system, new challenges came into existence

such as: (i) undesirable electrode/electrolyte reactions due to

high surface area, (ii) low tap density which results in low

volumetric energy density and (iii) nanomaterials are

difficult to synthesize and handle. In addition, nanoparticles

tend to aggregate during prolong charge/discharge cycling,

ultimately reducing the calendar life of the electrode. To

avoid such disparity, many synthesis methods have been

reported to tailor the materials in reduced dimensions. In

addition, pseudo-capacitive behavior has been observed at

the reaction interface which has been reported for nanomaterials

having high surface area [9].

1.2. Synthesis Routes to Prepare Nanomaterials

As discussed above, there are several advantages of

nanomaterials over the bulk counter parts, but the major

difficulty lies in the synthesis of nanomaterials and their

control over the size and dimension. Due to innumerable

efforts by the researchers in the field of nanomaterials

synthesis, several techniques have evolved in the last two

decades. The available techniques for nanomaterials

synthesis are categorized into two classes (i) top down

approach and (ii) bottom up approach. High energy ball

milling is the most common and widely used top down

approach where bulk materials are broken down to nanosize.

Though ball milling is easy to use and can provide particle

size upto ~10 nm, there is no control over the size and

morphology. Moreover, a broader range of particle

distribution and incorporation of impurities discourage

researchers to adopt top down synthesis routes. Bottom up

synthesis process which is widely used for nanoparticle

synthesis, can be sub-categorized as gas phase synthesis,

liquid phase synthesis, solid phase synthesis, and biological

processes. In the gas phase synthesis methods, precursors are

in gas phase which interact with liquid and/or solid phase

materials. Different techniques have been developed namely;

gas state condensation, chemical vapour deposition, molecular

beam epitaxy, atomic layer deposition, combustion, pyrolysis,

metal oxide vapor epitaxy and ion implantation for

nanomaterials synthesis. In the liquid phase synthesis, the

precursors are in liquid media and the final product is

insoluble in that liquid. The commonly used liquid phase

synthesis methods are hydrothermal or solvothermal

synthesis, polyol synthesis, sol-gel synthesis, electro-

deposition and supercritical fluid expansion method. Liquid

phase synthesis methods were most common for metal

oxide and sulfides preparation. Among them, hydrothermal/

solvothermal synthesis is well established to prepare various

morphologies. Sometimes templates are also used to provide

pre-defined shape to the desired materials. The template

synthesis is classified into two categories depending on the

choice of the template i.e. hard and soft template method.

Fig. (1). Schematic representation of (a) Lithium-ion battery depicting CNT anode and layered metal oxide cathode, and (b) Intrinsic

properties of bulk vs. nanomaterials.

Nano Dimensionality: A Way towards Better Li-Ion Storage Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 23

The common hard templates are anodic aluminium oxide

(AAO), Silica and Polymer matrices (polystyrene spheres)

while organic surfactants, gas bubbles fall into the category

of soft templates [10-14]. Recently, ionic liquids have also

been used for the synthesis of specific morphologies of the

desired materials [15]. Newly introduced biological

processes are also useful to prepare nanoparticles using

microorganisms. Recently, metal oxides such as Co3O4,

FePO4 have also been prepared by virus assisted synthesis

and successfully used in lithium ion batteries [16, 17].

2. PROGRESS IN CATHODE MATERIALS

Since the introduction of LiCoO2 by Goodenough and co-

workers [5], this material remains as state-of-the-art material

for LIB for more than two decades. However, this material

leads to serious risk of explosion due to oxygen evolution

during charging of cathode (LixCoO2) at elevated

temperatures. Thus, present technology requires structural

and thermally stable materials with high gravimetric energy

density and safety.

Other promising classes of cathode materials are spinels

LiMn2O4 (space group: Fd-3m) [18], polyanionic based

phospo-olivine LiMPO4 [19], orthosilicates based Li2MSiO4

(where M= Fe, Co, Ni, Mn) [20], layered metal oxides such

as MoO3, V2O5, LiV3O8, and metal fluorides etc. These

materials are safer than commercialized LiCoO2 cathode.

2.1. Spinel LiMn2O4

Spinel LiMn2O4 structure is quite different from layered

LiMO2 structure (where M= Co, Mn, etc) [18]. It is a three

dimensional host in which Li and Mn ions occupy

tetrahedral (8a) and octahedral (16a) sites respectively. In the

framework of spinel LiMn2O4, MnO6 octahedra share edges

to build three dimensional rigid network with open channel

in <110> direction, wherein lithium-ions are mobile along

8a-16c-8a path [18]. Research in spinel LiMn2O4 is

stimulated owing to low cost and safety issues resulting from

electrochemical insertion/deinsertion processes. However,

capacity fading of spinel LiMn2O4 is prominent on lithium

insertion/deinsertion. Two possible causes of capacity fading

in spinel LiMn2O4 are (i) Jahn-Teller (J-T) distortion at 3V

region [21], (ii) dissolution of Mn-ions into the electrolyte

and decomposition of the electrolyte ~ 4 V region [22].

Extensive studies have been devoted towards various

strategies to improve the electrochemical activity of spinel

LiMn2O4, especially to stabilize the structural deformation

due to J-T effect. Few of the effective steps to improve the

activity of spinel LiMn2O4 are small amount of substitution

in Mn-site which is believed to occupy the 16d sites of the

Mn-ions and nanostructuring of LiMn2O4 which stabilizes

the crystal structure. Traditionally, spinel LiMn2O4 cathode

is synthesized via conventional solid state synthesis route at

elevated temperatures to get micron-sized particles which

show poor cyclic performance. For this reason, nanostructured

materials have been synthesized by various methods to

improve the performance of LiMn2O4. Nanoparticles of

LiMn2O4 prepared by hard template process with average

particle size of 15 nm have shown specific discharge capacity

of 113 mAh g-1

at 2C rate [23]. Similarly, mesoporous spinel

(Fig. 2a) with the composition Li1.12Mn1.88O4 has delivered

discharge capacity of 70 mAh g-1

at 30C (Fig. 2b) compared

to bulk material of same composition [24]. The improvement

of electrochemical performance in mesoporous LiMn2O4

is due to the 7 nm sized walls which accommodate the

strain occurring at 3 V due to cubic/tetragonal phase

transformation.

2.2. Phospho-Olivine LiMPO4

One of the most focused cathode materials in recent

literature is polyanionic based cathode (LiMPO4 where M =

Fe, Mn, Ni, Co) in particular LiFePO4 [19]. Since the

demonstration of reversible electrochemistry of LiFePO4,

lithium transition phospho-olivines LiMPO4 (where M = Fe,

Mn, Co, Ni) have attracted great attention among the

researchers. The profound advantages of Fe and Mn-based

olivine compounds are cost effectiveness, natural abundance

and low toxicity compared to Co and Ni based olivines [19].

Oxygen atoms in the polyoxyanionic are stabilized by strong

X-O (where X= P, Si, B) covalent bond [19]. Along with the

above advantages, polyanionic based cathode materials truly

suffer from serious issues like poor electronic conductivity

(LiFePO4:10-9

to 10-8

Scm-1

) [7], ionic conductivity

(LiFePO4:10-10

to 10-9

Scm-1

) [7, 25] and ionic diffusivity

(LiFePO4:10-17

to 10-12

cm2s

-1) [26]. M. M. Thackeray et al.

suggested that the poor electronic conductivity of pristine

LiFePO4 may be attributed to strongly distorted continuous

network of FeO6 octahedra [27]. Initially, it was thought that

lithium diffusion in LiFePO4 is two dimensional like in

LiMO2 layered oxides i.e. running parallel to the c-axis and

a-axis, respectively [19]. However, recent theoretical

calculation concluded that the lithium-ion diffusion in

LiFePO4 is one-dimensional i.e. parallel to b-axis, ac plane is

only active for lithium-ion insertion/deinsertion since other

migration pathways have much higher energy barriers [28,

29]. Better electrochemical performance can be achieved by

restricting the crystal growth along b-axis, thus reducing

diffusion path length of lithium-ions. Further strategies to

improve the electrochemical performance of LiFePO4 have

been demonstrated lately using nano LiFePO4 particles to

exhibit ultrafast lithium-ion storage [30]. Nanostructured

mesoporous LiFePO4 embedded in conductive interconnected

carbon networks resulted in high electronic conductivity to

facile mass transfer and charge transfer which delivers initial

discharge capacity of 115 mAh g-1

and retaining 91% of

initial capacity till 1000 cycles at 10C current rate [31]. In

another approach, hollow LiFePO4 spheres were synthesized

by CTAB (cetyltrimethylammonium bromide) assisted

precipitation reaction and exhibited superior electrochemical

performance of 133 mAh g-1

at 10 C and 100 mAh g-1

at

50 C current rates [32]. Beside LiFePO4 based cathode

materials, other olivine members like LiMnPO4 exhibit great

promise as future cathode materials. Even though its

prospect in electrochemical performance is high, the main

problem associated with LiMnPO4 is J-T active Mn3+

ion

[33]. Consequently, the lattice distortion caused by J-T

active Mn3+

ion is severe which leads to structural distortion.

However, its stability is improved by doping Fe ions in Mn-

site to form solid solution of LiMnyFe1-yPO4 with best

nominal composition at y=0.8, where it delivers 150 mAh g-1

capacity at 1C rate with particle size ranging from 25-60 nm

[33]. The substitution of Fe ions for some of the Mn ions of

24 Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 Sen et al.

LiMnPO4 leads to better electrochemical activity which is

attributed to defects in the mixed-transition-metal compound

[33].

2.3. Orthosilicates Based Li2MSiO4

Apart from phosphates, orthosilicate is another

alternative class of polyanionic cathode materials having

high capacity for application in next generation LIBs. The

silicates are quite rich in polymorphs, Li2MSiO4 has three

commonly reported polymorphs, two of the three known

polymorphs of Li2MSiO4 are orthorhombic (Pmnb and

Pmn21), whereas the third is monoclinic (P21/n) [34-36].

The respective cations can order within the tetrahedral sites

in different ways leading to the formation of various

structures which show complex polymorphism [37].

In Li2FeSiO4, lithium-ion can take two different paths for

migration between adjacent lithiums. Though, theoretical

studies suggest that lithium migration energy along c-axis is

less compared to b-axis. It confirms that lithium-ion

diffusion involves zig-zag paths through intervening vacant

octahedral sites [38]. Moreover, extraction of two lithium-

ions from host matrix delivers high theoretical capacity

(~333 mAh g-1

), which is possible if both redox couples i.e. M

2+/M

3+ and M

3+/M

4+ are electrochemically active [39].The

main disadvantage of silicate family is their low electronic

conductivity, which is about 3 orders of magnitude lower

than that of LiFePO4 [35]. However, improved performance

of polyanionic based cathode materials is reported in recent

literature by nanostructuring [7,30] and by the use of

conductive additives [40]. Murliganth et al. synthesized

nanostructured cathode materials using microwave-

solvothermal technique and have been able to produce

carbon coated Li2FeSiO4 nanoparticles of 20 nm (Fig. 3a)

resulting in decreased lithium-ion diffusion path. Such

improved morphology leads to better electrochemical

performance, which can be seen from the charge-discharge

profile of Li2FeSiO4/C and Li2MnSiO4/C [41]. Charge

profile of Li2FeSiO4/C depicts that accessibility of Fe3+

/Fe4+

redox couple is possible at elevated temperatures. The

Li2FeSiO4/C sample showed excellent rate performance and

cyclic stability with stable discharge capacities of 148 mAh

g-1

at 25 °C and 204 mAh g-1

at 55 °C. Li2MnSiO4/C showed

severe capacity fading especially at elevated temperatures.

Fig. (2). (a) TEM images of mesoporous Mn2O3, Mn3O4, and Li1.12Mn1.88O4 [24]. (b) Cycling performances of mesoporous Li1.12Mn1.88O4,

bulk Li1.05Mn1.95O4, and bulk Li1.12Mn1.88O4 between 3 and 4.3 V at a rate of 3000 mA g-1

(30C) [24], (c) Discharge profile for nano

LiFe0.9P0.95O4- at high rates [30], and (d) Micrographs of the hollow sphere of nanoparticles obtained from each step of the sequential

precipitation [32].

Nano Dimensionality: A Way towards Better Li-Ion Storage Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 25

This poor capacity retention of Li2MnSiO4 is due to the

presence of Mn3+

ions, which causes J-T distortion and

manganese dissolution as in case of spinel LiMn2O4.

Most recently, two dimensional nanosheets of Li2FeSiO4

were synthesized using supercritical fluid synthesis method

[42].The lateral dimensions of nanosheets are in the range of

100-300 nm with average thickness of about 3nm as shown

in Fig. (3b). The result demonstrates superior discharge

capacity of 340 mAh g-1

for Li2FeSiO4 and 350 mAh g-1

for Li2MnSiO4 at 45 ± 5 °C with 80 % capacity retention

after 20 cycles (Fig. 3c) [42]. It is reported that unique

mechanical, and electrical properties due to large surface

area and small atomic scale thickness leads to improved

electrochemical performance of the nanosheet structured

cathode materials [42].

2.4. Other Layered Metal Oxides

Recently, other layered metal oxides such as MoO3 have

also shown good lithium storage ability. The orthorhombic

phase of MoO3 ( -MoO3) exhibits 1.7 mole of Li+ storage

per mole of MoO3 in its first discharge curve but upon

cycling half of the stored Li was trapped inside the

unrecoverable sites which causes severe capacity fading.

Though this material can withstand high rate charge/discharge,

the main disadvantage of this material is the cyclic stability.

Recent study explains that using stainless steel current

collector and 1-dimensional MoO3, a stable capacity of 126

mAh g-1

after 100 cycle has been achieved against 86 mAh g-

1 for commercially available bulk material [43].

2.5. Vanadium Based Materials

Vanadium has rich chemistry/reactivity towards lithium

which helps to explore new active cathode materials.

Vanadium based materials can accommodate more than one

lithium-ion in host matrix delivering high specific capacity.

It is well established that by engineering nanostructure, the

insertion/deinsertion properties and rate performance of

V2O5 electrode will significantly increase in comparison to

bulk counterpart. Various nanostructures of orthorhombic

V2O5 have been studied in Detail for e.g.; V2O5 nanorods

[44], V2O5 nanotubes [45]. Beside V2O5 cathode material,

lithium-trivanadate (LiV3O8) [46, 47] is another member of

vanadate family which has recently gained interest as cathode

material for LIBs. Lithium-trivanadate can accommodate three

additional lithium ions delivering specific capacity of ~374

mAh g-1

[48]. However, Picciotto et al. suggested under

controlled reaction conditions, it is possible to lithiate LiV3O8

composition to a stoichiometric rock salt composition of

Li5V3O8 [49]. Pan et al. showed that LixV3O8 nanorods with

high first discharge capacity as high as 239 mAh g-1

at 1000

mA g-1

[50]. Q. Shi et al. prepared crystalline LiV3O8 with an

amorphous surrounding orientated along <100> direction

[51]. The LiV3O8 nanorod thin film was prepared by

adjusting the oxygen partial pressure while deposition

process using radio frequency (RF) magnetron sputtering.

Due to unique structure and amorphous surrounding in

LiV3O8 nanorods, the material delivered high discharge

capacity of 388 mAh g-1

at C/5 and 102 mA h g-1

at

40 C rate. The result revealed that the orientation of LiV3O8

Fig. (3). (a) FEG-SEM of Li2FeSiO4 nanosphere [41], (b) HR-TEM of Li2FeSiO4 nanosphere [41], and (c) Discharge capacity of Li2FeSiO4

nanosheet [42], and (d) HR-TEM of Li2FeSiO4nanosheet [42].

(c) (d)

(b)

26 Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 Sen et al.

nanorods along <100> direction provides shorter pathway

for ion transport and amorphous wrapping endures the

anisotropy of the surface during the electrochemical process

[51].

2.6. Metal Fluorides

Recently, high capacity storage has been demonstrated

with conversion based cathode materials. Conversion

materials like MFx(where M = Fe, Cu) [52], Li2O/CuO

[53] undergo reaction with lithium, which can accommodate

more than one lithium-ion in the host matrix leading to high

capacity. One of the most promising conversion based

cathodes is FeF3 (space group: R-3c) which reacts with three

lithium-ions to deliver a high capacity of 712 mAh g-1

at an

average potential of 2.7 V. However, due to insulating

nature, structural deformation, and micron size particles;

LeF3 upon cycling suffers from sluggish kinetics, which leads

to capacity fading [54]. Moreover, L. Li et al. demonstrated

superior electrochemical performance of FeF3 with one

dimensional nanowires compared to micron size particles.

FeF3 nanowires showed first discharge capacity of 543 mAh

g-1

and retained a capacity of 223 mAh g-1

after 50 cycles at

current rate of 50 mA g-1

[52].

3. PROGRESS IN ANODE MATERIALS

Unlike cathode materials where mainly insertion/

deinsertion dominates, anodes follow different lithiation/

delithiation mechanism such as intercalation/de-intercalation,

conversion reaction and alloying/dealloying mechanisms. All

the three reaction mechanisms are discussed here in the light

of nanodimensional materials.

3.1. Intercalation/Deintercalation Materials

3.1.1. Carbonaceous Materials

Since the commercialization of LIBs, intercalation

materials dominated the anode part. Graphite is the most

successful anode material since its commercialization by

Sony. Graphite acts as a host matrix where lithium-ion

intercalates/de-intercalates in the interlayer spacing of the

graphite layers. A variety of carbonaceous materials have

been demonstrated depending on their preparation,

processing, precursor, thermal and chemical treatments. In

fact, all carbonaceous materials are capable of storing

lithium, but the amount of lithium accommodation varies

from carbon to carbon depending on its structure,

crystallinity, surface area and surface modification in

addition to particle size and shape [55]. Some of the most

common carbonaceous materials used for the lithium

insertion are graphite [56, 57], coke [58, 59], mesophase

pitches [60, 61], carbon fibers, whiskers [62, 63], pyrolytic

carbon [64-66], carbon nanotubes (CNTs) [67, 68] and most

recently grapheme [69-72]. One lithium for every six carbon

atoms (LiC6) can be stored for perfectly layered graphite,

which corresponds to a specific discharge capacity of 372

mAh g-1

. Depending upon different factors like order of

graphitization, crystallinity and surface defects, practical

capacity of 360-380 mAh g-1

[55, 73] has been achieved with

micron sized graphite, whereas carbon fibers having

microporous structure exhibits 633 mAh g-1

capacity [74].

Even higher capacity was reported with ordered nanoporous

carbon (average pore size: 3.9 nm) which exhibits an initial

discharge capacity of 3100 mAh g-1

corresponding to Li8.4C6

with reversible capacities ranging from 850 to 1100 mAh g-1

[75]. The concept of nanostructuring in carbon materials

such as one, two and three-dimensional carbon architectures

is increasingly being explored for lithium storage. Due to

their unique structural features that support improved

kinetics, large surface area and high porosity which can

provide additional lithium storage sites, CNTs are being

projected as high performance anode materials. CNTs also

exhibit higher electrical conductivity (5 x 105 S/cm) [67, 76],

and thermal conductivity (2,000–4,000W mK-1

) [77] compared

to bulk carbon, which leads to improved safety of the device,

high strength resulting in better durable electrodes. Based on

the theoretical calculations it has been assumed that apart

from the normal mode of lithium storage as LixC6, an

additional curvature-induced lithium condensation is a

possibility inside the core of the nanotubes [78, 79]. Zhao et al. [80] and Meunier et al. [81] predicted using single walled

carbon nanotubes (SWCNTs) a lithium-rich composition of

Li2C6 or higher can be possible leading to capacity of 1,116

mAh g-1

. In practice, first discharge capacity of more than

1,000 mAh g-1

[67, 82] was reported with SWCNTs though

the reversible capacity was ranging between 400-500

mAh g-1

[67, 83-86]. On the other hand, Multiwalled carbon

nanotubes (MWCNTs) are attractive due to various lithium

storage sites such as spacing between the graphite layers,

local turbostratic disorders arising from their highly

defective structures and central core [87]. Recent results

shows that MWCNTs are capable of lithium intake more

than 1,400 mAh g-1

in the first discharge cycle with stable

reversible capacity ranging from 500-800 mAh g-1

[88-91].

CNTs allow to prepare flexible free-standing paper-like

electrodes without conductive additive, metal substrate or

binder. Free-standing electrodes prepared by SWCNTs

exhibited insertion capacities in the range of 400–460 mAh

g-1

[68, 83-85]. Most recently purified SWCNT electrodes

with titanium contacts have achieved a reversible capacity of

1050 mA h g1

[92]. Another carbon material having two

dimensional nanosheets known as graphene is also

considered as a potential anode material for LIBs owing

to its high electronic conductivity, mechanical stability,

chemically/thermally tolerant and non-toxic nature.

Graphene prepared by solution chemistry routes is easy to

functionalize, which permits its assembly into functional

nanostructures such as paper and thin films. Graphene

nanosheets having a large interlayer spacing (d002 4.0 Å)

can store more lithium than graphite and exhibits a practical

reversible capacity of 740–780 mAh g-1

[69]. Recently Pan

and co-workers obtained even higher reversible capacities of

794–1054 mAh g-1

with disordered Grapheme [93].

3.1.2. Layered Titanium Oxides

Titanium oxide is the other material which falls into the

elite class of intercalation compounds. Titanium oxide exists

in diverse crystal forms for example, rutile, anatase,

brookite, TiO2 (B), ramsdellite type Li2Ti3O7 and Li4Ti5O12

spinels. The main advantages of Titanium oxide are

negligible structural change that leads to superior cyclic life,

low cost and safety. The major obstacle of TiO2 for practical

usage is its poor electronic and lithium-ion conductivity and

low specific capacity. To overcome these issues various

Nano Dimensionality: A Way towards Better Li-Ion Storage Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 27

morphologies of TiO2 in nanodimensions have been

fabricated. Rutile TiO2 (P42/mnm) is the most common and

stable form of TiO2. Due to its unique structural

arrangement, Li+ diffusion in rutile is highly anisotropic

[94] Experimental as well as simulated results demonstrate

that the Li+ diffusion coefficient along the c-direction is

approximately 10-6

cm2 s

-1, while it is only 10

-15 cm

2 s

-1 in the

ab-plane, providing one dimensional lithium-ion diffusion in

rutile [95-97]. As a result, bulk rutile exhibits poor Li+

intercalation at room temperature due to its low electronic

conductivity along ab-plane. Enhancement in lithium-ion

storage capacity was found when bulk rutile was downsized

to nano dimensional rutile. For example first discharge

capacity increased from 110 mAh g-1

to 338 mAh g-1

when

particle size decreased from 300 nm to 30 nm. First

discharge capacity further increased to 378 mAh g-1

when

particle size reduced to 15 nm [98]. The achieved reversible

capacity was around 210 mAh g-1

, 130 mAh g-1

and 50 mAh

g-1

for 15 nm, 30 nm and 300 nm particle respectively.

Similarly, high initial discharge capacity of 234 mAh g-1

was

achieved for 6 nm anatase TiO2 compared to 210 mAh g-1

and 203 mAh g-1

for 15 nm and 30 nm particles respectively

[99]. The stable discharge capacity after 100 cycles was 125

mAh g-1

, 80 mAh g-1

and 71 mAh g-1

for 6 nm, 15 nm and 30

nm particles respectively. TiO2 (B) is most interesting

form of TiO2, unlike other crystal forms it exhibits

stable electrochemical activity in both bulk as well as

nanodimensional materials. It also exhibits maximum

reversible Li+ intake upto 0.85 Li/per mole of TiO2 compared

to rutile and anatase TiO2 as well as spinel Li4Ti5O12 [100].

Bulk TiO2 (B) can accommodate 0.7-0.8 moles of lithium

upon first discharge which is stabilized to 0.3-0.6 moles of

lithium per mole of TiO2 (B) upon cycling [101-104]. On the

other hand nanodimensional TiO2 (B) exhibits first discharge

capacity of 300-340 mAh g-1

(corresponding to 0.9-1.01 Li-

ion/ TiO2) and reversible second cycle capacity of 200 - 230

mAh g-1

[105-108]. A stable capacity of 100 mAh g-1

(after

100 cycle) was reported for bulk TiO2 (B), whereas nanowire

TiO2 (B) exhibits almost double the capacity of bulk

counterpart after 100 cycles at the same current density. A

stable capacity of 216 mAh g-1

for 200 cycles was reported

using 2-dimensional nanosheets of TiO2 at 10 C rate [105].

Mesoporous TiO2 (B) also exhibits a high stable capacity of

250 mAh g-1

with first cycle capacity of 310 mAh g-1

at 0.1

C rate [106].

3.2. Conversion Materials

The conversion reaction mechanism has brought great

interest since many important transitional metal oxides and

sulphides follow this mechanism during electrochemical

reaction with lithium. Recent trends show plenty of ongoing

work on the nanostructuring of existing conversion materials

as well as new conversion materials. The main advantages of

conversion electrode are moderate working potential (0.5-1.0

V), high capacity (800-1200 mAh g-1

) and high rate

capability. Conversion materials are the ideal candidate as

anode in terms of safety as well as capacity. Intercalation

materials such as TiO2 operates at high potential (1.5 V-1.7

V vs Li/Li+) that leads to low operating voltage (Ecell =

Ecathode- Eanode) and at the same time these materials also

exhibit low energy density. On the other hand carbon has the

safety issues due to its very low potential intercalation (0.1-

0.2 V vs. Li/Li+). In case of alloying reaction, though the

theoretical capacity is very high but the operating potential is

below 0.5 V vs. Li/Li+ which leads to safety issues (similar

to pure carbon). Only conversion electrode provides better

operating voltage (less than 1.0 V with fixed plateau) and

optimum specific capacity. Theoretical calculations based on

the equation (1) shows that for the existing cathode

materials, the optimum specific capacity of the anode part is

ranging between 1000-1200 mAh g-1

(Fig. 4). The capacity

contribution from the anode part is negligible after 1200

mAh g-1

.

(1)

Fig. (4). Variation of total cell capacity versus anode capacity for

existing cathode materials.

In 1981, Fe2O3 was reported to be used as reversible

electrode material through conversion reaction but only at

elevated temperatures [109]. The first use of conversion

reaction at room temperature was achieved by making nano-

dimensional transition metal oxides reported in the year 2000

[110]. After this discovery, plenty of materials mostly transition

metal oxides and sulphides were used as conversion anode

such as Fe2O3, Fe3O4, Co3O4, MoO3, MoO2, MoS2, WS2, CoS,

CuO. One dimensional nanorods, nanobelts, nanowires, two

dimensional nanosheets, nanoplates, three dimensional

hollow spheres and other nanostructures were successfully

reported using liquid phase and gas phase synthesis routes.

For example -Fe2O3 hollow spheres synthesized by quasi-

emulsion template method showed a reversible capacity of

900 mAh g-1

for 50 charge/discharge cycles [111]. Similarly

nanorods, nanotubes were prepared using template method

and exhibits even higher capacity ranging between 900 to

1000 mAh g-1

[112, 113]. Co3O4 also exhibits more than

900 mAh g-1

capacity by preparing nanotubes, nanocages,

nanoplates [114-116], but the real challenge with Co3O4 is its

stability issue. After first cycle, Co3O4 was converted to CoO

350

400

(375 mAh g-1) LiV3O8

200

250

300

(330 mAh g-1) Li2FeSiO4

ty (m

Ah

g-1)

LiFePO

100

150

200(170 mAh g-1)

(140 Ah -1)(148 mAh g-1)Tota

l Cap

acit LiFePO4

LiMn2O4 Li0.5CoO2

0 500 1000 1500 2000 2500 30000

50(140 mAh g )(148 mAh g )T

Anode Capacity (mAh g-1)

28 Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 Sen et al.

which causes structural changes that lead to severe capacity

fading upon cycling [116]. To overcome these issues several

methods like composite formation [117, 118], hollow

architecture formation [115, 119] as well as the direct use of

CoO have been tested. Nanodiscs of CoO reported with 900

mAh g-1

stable capacity at 200 mA g-1

current rate [120].

Despite having competitively low theoretical capacity (674

mAh g-1

), CuO has also been used as anode due to its

environment friendliness. High aspect ratio nanofibers of

CuO were prepared by electrospinning technique, which

exhibited very stable capacity of 452 mAh g1 after 100

cycles [121]. CuO particle threads with CNT delivered stable

capacity of 650 mAh g-1

at 0.1 C with 100 % capacity

retention [122]. At higher rates of 5 C a stable capacity of

580 mAh g-1

was achieved due to improved electronic

conductivity by composite formation with CNTs [122]. -

MoO3 is another interesting anode material as it exhibits

excellent cyclic stability and a relatively low conversion

potential (~ 0.5 V vs. Li/Li+), as a result -MoO3 is ideal

for high energy density battery applications. -MoO3 can

accommodate 6 Li+ through conversion reaction and exhibits

1117 mAh g-1

theoretical capacity. However, the major

drawback of this material is the first cycle irreversibility,

almost half of the Li-ion was found to be irreversible from

second cycle onwards. Ultra long nanobelts prepared by

hydrothermal method have been demonstrated as stable

capacity anode with 800 mAh g-1

capacity for more than 150

cycles at 0.2 C rate [123]. Most recently, 1000 mAh g-1

capacity has been achieved using porous MoO3 film

prepared by electrodeposition process [124]. Due to its

one dimensional characteristics - MoO3 nanobelts used to

prepare free standing electrodes which exhibited 1000 mAh

g-1

at a current density of 50 mA g-1

[125].

In last few years, metal sulphides such as MoS2, WS2,

CoS, Sb2S3 etc. are gaining more interest as anode material

due to its low polarisation loss (compared to oxides). MoS2

is the well studied system among sulphide anodes. MoS2 can

accommodate 4 Li+ during conversion reaction, but in reality

it takes around 6 lithium ions during 1st discharge cycle.

Extra two lithium ions are intercalated in the interlayer

spacing of MoS2. Bulk material exhibits constant capacity

reduction upon cycling (Fig. 5b). 2-dimensional nanobelts

prepared by two step hydrothermal method shows excellent

electrochemical stability and exhibit 880 mAh g-1

reversible

capacity at 200 mA g-1

(Fig. 5) [126]. In situ carbonization

and composite formation with graphene as well as CNTs

show even higher capacity of 1000 mAh g-1

or more with

good cyclic life [127, 128].

Though ample research has been done for conversion

electrodes leading to better cyclic stability, capacity retention,

and rate capability, but few basic issues are still unresolved

such as poor electronic conductivity, sluggish kinetics, 1st

cycle irreversible capacity loss and high polarisation loss. To

address these problems recent focus has been shifted

to better electrode making processes and efficient designs

Figs. (6,7). Preparing composite materials using conductive

additives such as graphene, CNTs and amorphous carbon has

resolved the problem of poor electronic conductivity. First

cycle irreversible loss has been reduced by redeveloping the

electrode fabrication process and the use of new and efficient

binders. The issues of poor kinetics have been reduced by

improving the electrode-current collector interface using

nano-architectured electrodes for example, Fe3O4 directly

deposited on Cu nano pillars on Cu substrate have been

demonstrated as stable and high rate capable conversion

anode materials [129]. Another route to prepare electrodes is

directly depositing active materials on conductive substrates

[130]. M. V. Reddy et al. deposited Fe2O3 nanofalkes on Cu

substrate, which has shown 600 mAh g-1

capacity for 100

cycles [131].

However, the issue of polarisation loss is still open as it

is the inherent property of the material. It was found that M-

X bond polarity is the main reason for the polarization loss

(Polarization loss was calculated as the difference between

the potential of the charge and discharge curve at the half

capacity). From experimental results it was found that

polarisation ( V) decreases form fluoride (~1.1 V) to oxide

(~0.9 V) to sulphide (~0.7 V) to nitride (~0.4 V) and

phosphieds (~0.4 V). Due to the low polarisation loss recent

Fig. (5). (a) FEG-SEM image of MoS2nano-walls and (b) Cyclic performance of MoS2nano-wall and bulk MoS2 [126].

1100M S N ll( )

700

800

900

1000 MoS2 Nano-wall

city

(mA

h g-1

)

Bulk MoS2

22 nm

(a) (b)

200

300

400

500

600

scha

rge

Cap

ac

0 10 20 30 40 500

100

200Di

Cycle Index

100 nm

Nano Dimensionality: A Way towards Better Li-Ion Storage Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 29

interest has shifted from oxide to phosphide and nitride

materials. Several phophide materials such as CoP3, NiP2,

FeP, VP2 and Zn3P2 have been envisioned as conversion

anodes. Full conversion of nickel phopshide with Li was

reported which shows 90 % reversible capacity. Monoclinic

NiP2 can reversibly react with 5 Li+ and delivers a capacity

of 1000 mAh g-1

[133]. By designing new electrodes in

which NiP2 is directly grown on Ni substrates, high

reversible capacities with increased rate capability can be

achieved as discussed earlier in case of Fe2O3. Hydrothermal

synthesis of hierarchical nanostructured Ni2P was reported

with reversible capacities at high current rates [134].

Graphene encapsulated Ni2P nanoparticles have been

reported with stable capacity of 500 mAh g-1

upto 50 cycles

Fig. (6). (a) Cu- nanostructured current collector before and (b) After Fe3O4deposites [129], (c) -Fe2O3nanoflakes [131], (d) -Fe2O3

nanorods at a low magnification and (e) End-view at a high magnification [113], (f) SEM and (g) TEM images of -Fe2O3 nanotubes [112],

(h) SEM and (i) STEM images of graphene encapsulated hollow Fe3O4 nanospheres [132].

Fig. (7). Cyclic performance of (a) Fe3O4 deposited on Cu nanorods [129], (b) -Fe2O3 nanoflakes [131], (c) -Fe2O3 nanorods [113], (d) -

Fe2O3 nanotubes [112] and (e) Graphene encapsulated Fe3O4 hollow nanoparticles [132].

y (m

Ahg‐

1 )

1Li+/4h 1Li+/2h

mAh

g‐1 )

(a) (b) (c)

scha

rge capa

city

Capa

city (m

Dis

Cycle number Cycle number

(d) (e)

30 Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 Sen et al.

[135]. Vanadium monophopshide (VP) has shown first

discharge capacity of 400 mAh g-1

and stable reversible

capacity of 250 mAh g-1

upto 250 cycles [136]. Copper

phosphide films deposited on Cu substrate have also been

demonstrated as conversion anode materials [137]. Metal

nitrides are also used as anode owing to low polarisation

loss. Nitrides have an additional advantage that the discharge

product i.e. Li3N is an excellent ionic conductor [138], which

provides additional conductivity during electrochemical

cycling. Transition metal nitrides having general formula of

Li3-xMxN (M= V, Cr, Fe, Co, Ni, Cu, Zn) [139-145] have

been visualized as conversion anode in the literature. CoN is

the most studied system in the nitride family. Porous

nanoparticles of CoN was prepared and tested against

lithium which show good stable performance in the range of

700-800 mAh g-1

[143]. Particularly Li2.6Co0.4N composition

shows a high and stable capacity of 760 mAh g-1

in the

voltage window of 0-1.4 V [146]. Recently Li3-xFexN

prepared in the form of a powder and exhibited a maximum

reversible capacity of 700 mAh g-1

for the composition

X=0.2 in the range of 0.05–1.3 V vs. Li/Li+

[147]. CrN and

VN have also drawn much attention due to its high

theoretical capacity. Though CrN has achieved 1000 mAh g-1

in its first discharge cycle but the stability is the main issue.

Stable reversible capacity of 500 mAh g-1

was achieved

using nanoparticles having average particle size of 20 nm

[142]. A thin film electrode prepared with VN was tested as

anode and exhibited superior stability of 800 mAh g-1

after

50 cycles [144]. Most recently metal hydrides are gaining

interest as polarization is suppressed in metal hydrides

compared to other conversion materials and establishes its

candidature as new competitor for anode materials.

Magnesium hydride (MgH2) has been reported as the first

candidate in the metal hydride category with 500 mAh g-1

reversible capacity [148].

3.3. Alloying Materials

Li can be electrochemically alloyed with a number of

metallic and intermetallic elements of groups IV and V, such

as Si, Sn, Ge, Pb, P, As, Sb, and Bi. Some of the other metals

such as Al, Au, In, Ga, Zn, Cd, Ag, and Mg, and metal

oxides such as SnO2 are also used for alloying reaction. The

main advantages of alloying reaction are high specific

capacity, flat plateau operating voltage and better safety

than carbon materials. But the main challenge for the

implementation of these anodes is the huge volume change

(200-400%) [149] during Li alloying/dealloying (discharge/

charge) processes, which causes severe cracking and

crumbling of the electrodes and loss of electrical contact

between individual particles resulting in large irreversible

capacity and severe capacity fading. Several strategies have

been used to overcome the problem of volume change and to

retain the structural integrity. Among all the strategies,

nanostructuring of the active material was largely used and

most effective in reducing the strain during lithiation/

delithiation processes Fig. (8). Recent studies illustrate that

by preparing nanostructures such as nanowires, nanorods and

hollow spheres the cycle stability can be improved

dramatically [150-155]. The structural beauty of these

nanostructures is that they can buffer against the volume

changes and reduces the strain on the material. It is

noteworthy to mention here that the nanocrystalline

materials could improve tensile and fatigue strength, as the

grain size decreases to nanometer scale leading to

dramatically increase in the yield and fracture strengths of

metals and alloys [156, 157]. Apart from nanodimensionality,

structural design and the formation of composites with soft

materials such as amorphous carbon and graphene also help

to reduce the effect of volume change as well as increase in

electronic conductivity. Si nanowires as alloying materials

have been first demonstrated by Cui et al. in 2008 in

which almost theoretical capacity was achieved in the first

cycle [150]. These nanostructures showed excellent

mechanical stability as well as electrochemical stability.

Above 3000 mAh g-1

reversible capacity was achieved after

10 cycles. Several other reports have used the concept of

nanoarchitectured electrode to make Si alloying anode.

Several solution based approaches have also been used to

coat Si nanowires with conductive carbon to increase the

cyclic stability. Carbon coated Si nanowires exhibited

reversible capacity above 1300 for more than 40 cycles

[158]. A unique core-shell strategy was introduced where

crystalline Si acts as support as well as conducting media in

the core while the amorphous Si at the shell is used for Li

storage. The core-shell strategy demonstrated excellent

stability for longer cyclic life, around 90 % capacity

retention for 100 cycles with reversible capacity around 1000

mAh g-1

[153]. As discussed above hollow structures have

their own advantages, Si hollow structures show even better

stable performance for longer cyclic life. Interconnected Si

hollow spheres having outer diameter of 200 nm using solid

Fig. (8). Schematic representation of alloying reaction during lithiation/delithiation.

Nano Dimensionality: A Way towards Better Li-Ion Storage Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 31

SiO2 templates, show first discharge capacity of 2725 mAh

g-1

while a reversible capacity of 1420 mAh g-1

was reported

after 700 cycles [159]. Similarly, Si nanotubes demonstrated

reversible discharge capacities over 3000 mAh g-1

with a

capacity retention of 89% for 200 cycles at a rate of 1 C

[155]. Even longer cycle life and better rate capability have

been achieved by surface modification of the nanotubes. The

surface modification or the clamping mechanism helps to

build a stable SEI layer on the outer surface of the wall [160-

162]. As a result, surface clamped Si nanotubes exhibit long

cyclic life as well as better rate capability. Recent literature

using this kind of approach demonstrated high specific

discharge capacity about 3000 mAh g-1

for 6000 cycles

with 88% capacity retention at C/5, and about 1000 mAh g-1

at 12 C [160].

In this review, we have documented several cathode and

anode materials which could be envisioned as future

electrodes of LIB technology. All discussed results are based

on half-cell configuration. For a full cell performance a good

combination of cathode and anode is needed for high energy

applications. Herein, Table 1 summarizes energy density for

possible combination of anode and cathode from the existing

materials.

4. CONCLUSIONS AND FUTURE PROSPECT

Advances in nanoscience and nanotechnology have given

us a great opportunity to design and fabricate new energy

storage materials for the high performance LIBs. Major

requirements such as high capacity along with longer cycle

life could be realized from the discussed advanced materials.

Nanostructured materials can provide high lithium-ion flux

across the interface, shorter diffusion length, number of

active sites for lithium storage and can accommodate high

volume change during charge/ discharge to enhance the

structural stability of the electrodes. The improved storage

and rate capability are closely related to their surface area,

crystallinity, morphology as well as the orientation of the

crystallites. The use of conversion based materials as anode

for LIBs is still far from reality because of their low

efficiency, poor safety, rate capability and polarization loss.

To realize wide-spread commercial applications, further

work is required to achieve controlled and large-scale

synthesis of nanostructures. It is also necessary to understand

the mechanisms of Li storage in nanomaterials and the

kinetic transport at the electrode/electrolyte interface since it

is quite different from the bulk counter parts.

Future direction in nanodimensional electrode materials

could be explored with new redox couples or different

lithium storage mechanisms and/or designing new electrode

architectures to fuel more electrons and ions to the active

materials. Better electrode-electrolyte interface could also

increase the efficiency of the electrode materials, since

most of the low dimensional materials react at the interface

with lithium it could be an interesting area of research. Few

issues like recycling of electrode materials, production of

nanomaterials using green chemistry should be considered.

Moreover, use of renewable resources is highly recommended

for electrode materials production.

CONFLICT OF INTEREST

The author(s) confirm that this article content has no

conflict of interest.

Table 1. Comparison of full cell energy density with some of the existing cathode and anode materials.

Cathode L iCoO2

[163]

3.9 V

140 mAh g-1

LiMn2O

4 [22]

4.1 V

120 mAh g-1

LiFePO4

[30]

3.44 V

160 mAh g-1

Li2FeSiO

4 [42]

3.1 V

340 mAh g-1

Anode Full Cell Combination Energy Density

Graphite [55]

0.1 V

372 mAh g-1

386 mWh g

-1

363 mWh g

-1

374 mWh g

-1

533 mWh g

-1

Graphene [69]

0.1 V

740 mAh g-1

447 mWh g

-1

413 mWh g

-1

439 mWh g

-1

699 mWh g

-1

TiO2

[106]

1.5 V

250 mAh g-1

215 mWh g

-1

211 mWh g

-1

189 mWh g

-1

230 mWh g

-1

Fe2O

3 [112]

0.8 V

1000 mAh g-1

380 mWh g

-1

353 mWh g

-1

364 mWh g

-1

583 mWh g

-1

MoS2

[126]

0.5 V

880 mAh g-1

411 mWh g

-1

380 mWh g

-1

398 mWh g

-1

638 mWh g

-1

Si [150] 0.4 V

3000 mAh g-1

468 m Wh g

-1

427 mWh g

-1

462 mWh g

-1

824 mWh g

-1

32 Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 Sen et al.

ACKNOWLEDGEMENTS

This work has been supported by National Centre for

Photovoltaic Research and Education (NCPRE) -Ministry of New and Renewable Energy, Govt. of India. Authors are

grateful for the support provided by Indian Institute of

Technology Bombay.

REFERENCES

[1] Whittingham, M. S. Electrical energy storage and intercalation

chemistry. Science, 1976, 192, 1126-1127.

[2] Aurbach, D.; Zinigrad, E.; Teller H.; Dan, P. Factors which limit

the cycle life of rechargeable lithium (metal) batteries. J.

Electrochem. Soc., 2000, 147, 1274-1279.

[3] Aurbach, D. Nonaqueous electrochemistry, Marcel Dekker, New

York, 1999.

[4] Yazami, R.; Touzain, Ph. A reversible graphite-lithium negative

electrode for electrochemical generators. J. Power Sources, 1983,

9, 365-371.

[5] Mizushima, K.; Jones, P. C.; Wiseman, P. J.; Goodenough, J. B.

LixCoO2 (0 < x < -1): A new cathode material for batteries of high

energy density. Mat. Res. Bull., 1980, 15, 783-789.

[6] Lee, K. T.; Cho, J. Roles of nanosize in lithium reactive

nanomaterials for lithium ion batteries. Nano today, 2011, 6, 28-41.

[7] Chiang, Y. M. Introduction and overview: physical properties of

nanostructured materials. J. Electroceram., 1997, 1, 205-209.

[8] Gaberscek, M., Dominko, R.; Jamnik, J. Is small particle size more

important than carbon coating? An example study on LiFePO4

cathodes. Electrochem. Commun., 2007, 9, 2778-2783.

[9] Maier, J. Nanoionics: ion transport and electrochemical storage in

confined systems, Nature Mater., 2005, 4, 805-815.

[10] Liu, D.; Nakashima, K. Synthesis of Hollow Metal Oxide

Nanospheres by Templating Polymeric Micelles with Core-Shell-

Corona Architecture. Inorganic Chemistry, 2009, 48, 3898-3900.

[11] Wang, Y.; Su, F.; Lee, J.Y.; Zhao, X.S. Crystalline carbon hollow

spheres, crystalline carbon-SnO2 hollow spheres, and crystalline

SnO2 hollow spheres: synthesis and performance in reversible Li-

ion storage. Chem. Mater., 2006, 18, 1347-1353.

[12] Zhong, Z.; Yin, Y.; Gates, B.; Xia, Y. Preparation of mesoscale

hollow spheres of TiO2 and SnO2 by templating against crystalline

arrays of polystyrene beads. Adv. Mater., 2000, 12, 206-209.

[13] Ye J.; Zhang H.; Yang R.; Li X.; Qi L. Morphology-controlled

synthesis of SnO2 nanotubes by using 1D silica mesostructures as

sacrificial templates and their applications in lithium-ion batteries.

Small, 2010, 6, 296-306.

[14] Ma Y.; Cheng F.; Chen J.; Zhao J.; Li C.; Tao Z.; Liang J. Nest-

like silicon nanospheres for high-capacity lithium storage, Adv.

Mater., 2007, 19, 4067-4070.

[15] Lian, J.; Duan, X.; Ma, J.; Peng, P.; Kim, T.; Zheng, W. Hematite

( -Fe2O3) with various morphologies: ionic liquid-assisted

synthesis, formation mechanism, and properties. ACS Nano, 2009,

3, 3749-3761.

[16] Nam, K.T.; Kim, D.W.; Yoo, P.J.; Chiang, C.-Y.; Meethong, N.;

Hammond, P.T.; Chiang, Y.M.; Belcher, A. M. Virus enabled

synthesis and assembly of nanowires for lithium ion battery

electrodes. Science, 2006, 312, 885-888.

[17] Lee, Y.J.; Yi, H.; Kim, W.-J.; Kang, K.; Yun, D.S.; Strano, M.S.;

Ceder, G.; Belcher, A.M. Fabricating genetically engineered high-

power lithium-ion batteries using multiple virus genes. Science,

2009, 324, 1051-1055.

[18] Thackeray, M. M.; David, W. I. F.; Bruce, P.G.; Goodenough, J. B.

Lithium insertion into manganese spinels. Mater. Res. Bull., 1983,

18, 461-472.

[19] Padhi, A. K.; Nanjundaswamy, K. S.; Goodenough, J. B.

Phospho olivines as positive electrode materials for rechargeable

lithium batteries. J. Electrochem. Soc.1997, 144, 1188-1194.

[20] Nytén, A.; Abouimrane, A.; Armand, M.; Gustafsson, T.; Thomas,

J.O. Electrochemical performance of Li2FeSiO4 as a new Li-battery

cathode material. Electrochem. Commun., 2005, 7, 156-160.

[21] Xia, Y.; Zhou, Y.; Yoshio, M. Capacity fading on cycling of 4 V

Li/LiMn2O4 cells. J. Electrochem. Soc., 1997, 144, 2593-2560.

[22] Jang, D. H.; Shin, Y. J.; Oh, S. M.; Dissolution of spinel oxides and

capacity losses in 4V Li/LixMn2O4cells, J. Electrochem. Soc., 1996,

143, 2204-2211.

[23] Cabana, J.; Valdés-Solís, T.; Palacín, M. R.; Oró-Solé, J.; Fuertes,

A.; Marbán, G.; Fuertes, A.B. Enhanced high rate performance of

LiMn2O4 spinel nanoparticles synthesized by a hard-template route.

J. Power Sources, 2007, 166, 492-498.

[24] Jiao, F.; Bao, J.; Hill, A. H.; Bruce, P. G. Synthesis of ordered

mesoporous Li-Mn-O spinel as a positive electrode for

rechargeable lithium batteries. Angew. Chem. Int. Ed., 2008, 47,

9711-9716.

[25] Wang, C. W.; Sastry, A. M.; Striebel, K. A.; Zaghib, K. J.

Extraction of layerwise conductivities in carbon-enhanced,

multilayered LiFePO4 cathodes. J. Electrochem. Soc., 2005, 152,

A1001-1010.

[26] Prosini, P. P.; Lisi, M., Zane, D.; Pasquali, M. Determination of the

chemical diffusion coefficient of lithium in LiFePO4. Solid State Ionics, 2002, 148, 45-51.

[27] Thackeray, M.; An unexpected conductor. Nature Mater., 2002, 1,

81-82.

[28] Morgan, D.; der Ven, A. V.; Ceder, G. Li conductivity in LixMPO4

(M=Mn, Fe, Co, Ni) olivine materials. Electrochem. Solid-State

Lett., 2004, 7, A30-A32.

[29] Islam, M.; Driscoll, D.; Fisher, C.; Slater, P. Atomic-scale

investigation of defects, dopants, and lithium transport in the

LiFePO4 olivine-type battery material. Chem. Mater., 2005, 17,

5085-5092.

[30] Kang, B.; Ceder, G. Battery materials for ultrafast charging and

discharging. Nature, 2009, 458, 190-193.

[31] Wang , G.; Liu, H.; Liu, J.; Qiao, S.; Lu, G.; M.; Munroe, P.; Ahn.

H. Mesoporous LiFePO4/C nanocomposite cathode materials for

high power lithium ion batteries with superior performance. Adv.

Mater., 2010, 22, 4944-4948.

[32] Lee, M.-H.; Kim, J.-Y., Song, H.-K. A hollow sphere secondary

structure of LiFePO4 nanoparticles. Chem. Commun., 2010, 46,

6795-6797.

[33] Martha, S. K.; Grinblat, J.; Haik, O.; Zinigrad, E.; Drezen, T.;

Miners, J. H.; Exnar, I.; Kay, A.; Markovsky, B.; Aurbach, D.

LiMn0.8Fe0.2PO4: An advanced cathode material for rechargeable

lithium batteries. Angew. Chem. Int. Ed., 2009, 48, 8559-8563.

[34] Nishimura, S.I.; Hayase, S.; Kanno, R.; Yashima, M.; Nakayama,

N.; Yamada, A. Structure of Li2FeSiO4. J. Am. Chem. Soc., 2008,

130, 13212-13213.

[35] Dominko, R.; Li2MSiO4 (M= Fe and/or Mn) cathode materials. J.

Power Sources, 2008, 184, 462–468.

[36] Sirisopanaporn, C.; Boulineau, A.; Hanzel, D.; Dominko, R.;

Budic, B.; Armstrong, A.R.; Bruce, P.G.; Masquelier, C. Crystal

structure of a New Polymorph. Inorg. Chem., 2010, 49, 7446–7451.

[37] Islam, M.S.; Dominko, R.; Masquelier, C.; Sirisopanaporn, C.;

Armstrong, A.R.; Bruce, P.G. Silicate cathodes for lithium

batteries: alternatives to phosphates. J. Mater. Chem., 2011, 21,

9811–9818.

[38] Armstrong, A.R.; Kuganathan, N.; Islam, M.S.; Bruce, P.G.

Structure and Lithium Transport Pathways in Li2FeSiO4 Cathodes

for Lithium Batteries. J. Am. Chem. Soc., 2011, 133, 13031–13035.

[39] Dompablo, M.E.A.; Armand, M.; Tarascon, J.M.; Amador, U. On-

demand design of polyoxianionic cathode materials based on

electronegativity correlations: An exploration of the Li2MSiO4

system (M = Fe, Mn, Co, Ni). Electrochem. Commun., 2006, 8,

1292–1298.

[40] Herle, P. S.; Ellis, B.; Coombs, N.; Nazar, L. F.; Nano-network

electronic conduction in iron and nickel olivine phosphates. Nature

Mater., 2004, 3, 147-152.

[41] Muraliganth, T.; Stroukoff, K. R.; Manthiram, A. Microwave-

Solvothermal Synthesis of Nanostructured Li2MSiO4/C (M = Mn

and Fe) Cathodes for Lithium-Ion Batteries. Chem. Mater., 2010,

22, 5754–5761.

[42] Rangappa, D.; Murukanahally, K.D.; Tomai, T.; Unemoto, A.;

Honma, I.; Ultrathin Nanosheets of Li2MSiO4 (M = Fe, Mn) as

High-Capacity Li-Ion Battery Electrode. Nano Lett., 2012, 12,

1146-1151.

[43] Sen, U. K.; Mitra, S. Electrochemical activity of -MoO3 nano-

belts as lithium-ion battery cathode. RSC Adv., 2012, 2, 11123-

11131.

[44] Takahshi, K.; Limmer, S. J.; Wang, Y.; Cao, G. Synthesis and

electrochemical properties of single-crystal V2O5 nanorod array by

template-based electrodeposition. J. Phys. Chem. B, 2005, 44, 662-

668.

Nano Dimensionality: A Way towards Better Li-Ion Storage Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 33

[45] Wang, Y.; Takahashi, K.; Shang, H.; Cao G.; Synthesis and

Electrochemical Properties of Vanadium Pentoxide Nanotube

Arrays. J. Phys. Chem. B, 2005, 109, 3085-3088.

[46] Wadsley, A. D. Crystal chemistry of non-stoichiometric

pentavalentvandadium oxides: crystal structure of Li1+xV3O8. Acta

Cryst., 1957, 10, 261-267.

[47] Sarkar, S.; Banda, H.; Mitra, S. High capacity lithium-ion battery

cathode using LiV3O8nanorods, Electrochemi. Acta, 2013, 99, 242-

252.

[48] West, K.; Christiansen, B. Z.; Østergard, M. J. L.; Jacobsen, T.

Vanadium oxides as electrode materials for rechargeable lithium

cells, J. Power Sources, 1987, 20, 165-172.

[49] Picciotto, L. A. de; Adendorff, K. T.; Liles, D. C.; Thackeray, M.

M. Structural characterization of Li1+xV3O8 insertion electrodes by

single-crystal X-ray diffraction, Solid State Ionics, 1993, 62, 297-

307.

[50] Pan, A.; Liu, J.; Zhang, J. -G.; Cao, G.; Xu, W.; Nie, Z.; Jie, X.;

Choi, D.; Arey, B. W.; Wang, C.; Liang, S. Template free synthesis

of LiV3O8 nanorods as a cathode material for high-rate

secondary lithium batteries, J. Mater. Chem., 2011, 21, 1153-1161.

[51] Shi, Q.; Liu, J.; Hu, R.; Zeng, M.; Dai, M.; Zhu, M. An amorphous

wrapped nanorod LiV3O8 electrode with enhanced performance for

lithium ion batteries, RSC Adv, 2012, 2, 7273-7278.

[52] Li, L.; Meng, F.; Jin, S. High-capacity lithium-ion battery

conversion cathodes based on iron fluoride nanowires and insights

into the conversion mechanism. Nano Lett., 2012, 12, 6030-6037.

[53] Li, T.; Ai, X. P.; Yang, H. X. Reversible Electrochemical

Conversion Reaction of Li2O/CuO Nanocomposites and Their

Application as High-Capacity Cathode Materials for Li-Ion

Batteries. J. Phys. Chem. C, 2011, 115, 6167-6174.

[54] Li, H.; Balaya, P.; Maier, J. Li-storage via heterogeneous reaction

in selected binary metal fluorides and oxides. J. Electrochem. Soc.,

2004, 151, A1878 A1885.

[55] Kumar, T. P; Kumari, T. S D.; Stephan A. M. Carbonaceous anode

materials for lithium-ion batteries-the road ahead. J. Ind. Inst. Sci.,

2009, 89, 393-424.

[56] Lee, Y. H.; Pan, K. C.; Lin, Y. Y.; Kumar, T. P.; Fey, G. T. K.

Lithium intercalation in graphites precipitated from pig iron melts.

Mater. Chem. Phys., 2003, 82 750-757.

[57] Zheng, T.; Liu, Y.; Fuller, E.W.; Tseng, S.; von Sacken, U.; Dahn,

J. R. Lithium insertion in high capacity carbonaceous materials. J.

Electrochem. Soc., 1995, 142 2581-2590.

[58] Shi, H. Coke vs. graphite as anodes for lithium-ion batteries. J

Power Sources, 1998, 75, 64-72.

[59] Qiu, W.; Zhou, R.; Yang, L.; Liu, Q. Lithium-ion rechargeable

battery with petroleum coke anode and polyanilinecatode. Solid State Ionics, 1996, 86-88, 903-906.

[60] Lin, J.-H.; Ko, T.-H.; Kuo, W.-S.; Wei, C.-H. Mesophase pitch

carbon coated with phenolic resin for the anode of lithium-ion

batteries. Energy Fuels, 2010, 24, 4090-4094.

[61] Takami, N.; Satoh, A.; Hara, M.; Ohsaki, T. Rechargeable lithium-

ion cells using graphitized mesophase pitch based carbon fiber

anodes. J. Electrochem. Soc., 1995, 142, 2564-2571.

[62] Endo, M.; Kim, Y. A.; Hayashi, T.; Nishimura, K.; Matusita, T.;

Miyashita, K.; Dresselhaus, M. S.; Vapor-grown carbon fibers

(VGCFs) basic properties and their battery applications, Carbon,

2001, 39, 1287-1297.

[63] Abe, H.; Zaghib, K.; Tatsumi, K.; Higuchi, S. Performance of

lithium-ion rechargeable batteries:

graphitewhisker/electrolyte/LiCoO2 rocking-chair system. J Power Sources, 1995, 54, 236-239.

[64] Gherghel, L.; Kubel, C.; Lieser, G.; Rader, H. J.; Mullen, K.

Pyrolysis in the mesophase: A chemist’s approach toward

preparing carbon nano and microparticles. J. Am. Chem. Soc., 2002, 124, 13130-13138.

[65] Stephan, A. M.; Kumar, T. P.; Ramesh, R.; Thomas, S.; Jeong, S.

K.; Nahm, K. S.; Pyrolitic carbon from biomass precursors as

anode materials for lithium batteries. Mater. Sci. Eng, 2006, A430,

132-137.

[66] Fey, G. T. K.; Lee, D. C.; Lin, Y. Y.; Kumar, T. P.; High-capacity

disordered carbons derived from peanut shells as lithium-

intercalating anode materials. Synth. Met., 2003, 139, 71-80.

[67] Landi, B. J.; Ganter, M. J.; Cress, C. D.; DiLeoand, R. A.;Raffaelle,

R. P. Carbon nanotubes for lithium ion batteries, Energy Environ. Sci., 2009, 2, 638-654.

[68] Ng, S. H.; Wang, J. Z.; Guo, Z. P.; Chen, J.; Wang, G. X.; Liu, H.

K. Single wall carbon nanotube paper as anode for lithium-ion

battery. Electrochim. Acta, 2005, 51, 23-28.

[69] Yoo, E.; Kim, J.; Hosono, E.; Zhou, H.; Kudo, T.; Honma, I. Large

reversible Li storage of grapheme nanosheet families for use in

rechargeable lithium ion batteries. Nano Lett., 2008, 8, 2277-2282.

[70] Wallace, G. G.; Wang, C. Y.; Li, D.; Too, C. O. Electrochemical

properties of graphene paper electrodes used in lithium batteries. Chem. Mater., 2009, 21, 2604-2606.

[71] Mukherjee, R.; Thomas, A. V.; Krishnamurthy, A.; Koratkar, N.

Photothermally reduced graphene as high-power anodes for

lithium-ion batteries. ACS Nano, 2012, 6, 7867-7878.

[72] Wu, Z.-S.; Ren, W.; Xu, L.; Li, F.; Cheng, H. -M. Doped graphene

sheets as anode materials with superhigh rate and large capacity for

lithium ion batteries. ACS Nano, 2011, 5, 5463-5471.

[73] Yishi, Y.; Nishida, T.; Suda, S.; Kobayashi, M. Hitachi chemical

company technical report. 2006, 47, 29.

[74] Sandi, G.; Carrado, K. A.; Winans, R. E.; Johnson, C. S.; Csencsits,

R. Carbons for lithium battery applications prepared using sepiolite

as inorganic template. J. Electrochem. Soc., 1999, 146, 3644-3648.

[75] Zhou, H.; Zhu, S.; Hibino, M.; Honma, I.; Ichihara, M. Lithium

storage in ordered mesoporous carbon (CMK-3) with high

reversible specific energy capacity and good cycling performance.

Adv. Mater., 2003, 15, 2107-2111.

[76] Zhang, Z.; Peng, J.; Zhang, H. Low-temperature resistance of

individual single-walled carbon nanotubes: a theoretical estimation.

Appl. Phys. Lett., 2001, 79, 3515-3517.

[77] Berber, S.; Kwon Y.-K.; Tománek, D. Unusually high thermal

conductivity of carbon nanotubes, Phys. Rev. Lett., 2000, 84, 4613-

4616.

[78] Zhao, M.; Xia, Y.; Liu, X.; Huang, B.; Li, F.; Ji, Y.; Song, C.

Curvature-induced condensation of lithium confined inside single-

walled carbon nanotubes: first-principles calculations. Phys. Lett.,

2005, A340, 434-439.

[79] Zhao, M.; Xia, Y.; Mei, L. Diffusion and condensation of lithium

atoms in single-walled carbon nanotubes. Phys. Rev. B, 2005, 71,

165413-165459.

[80] Zhao, J.; Buldum, A.; Han, J. First-principles study of Li

intercalated carbon nanotube ropes. Phys. Rev. Lett., 2000, 85,

1706-1709.

[81] Meunier, V.; Kephart, J.; Roland, C.; Bernholc, J. Ab initio

investigation of lithium diffusion in carbon nanotube systems.

Phys. Rev. Lett., 2002, 88, 075506-1-4.

[82] Gao, B.; Bower, C.; Lorentzen, J.D.; Fleming, L.; Kleinhammes,

A.; Tang, X.P.; McNeil, L.E.; Wu, Y.; Zhou, O. Enhanced

saturation lithium composition in ball-milled single-walled carbon

nanotubes. Chem. Phys. Lett., 2000, 327, 69-75.

[83] Claye, A. S.; Fischer, J. E.; Huffman, C. B.; Rinzler, A. G.;

Smalley, R. E. Solid-state electrochemistry of Li single wall carbon

nanotube system. J. Electrochem. Soc., 2000, 147, 2845-2852.

[84] Mukhopadhyay, I.; Kawasaki, S.; Okino, F.; Govindaraj, A.; Rao,

C. N. R.; Touhara, H. Electrochemical Li insertion into single-

walled carbon nanotubes prepared by graphite arcdischarge

method. Physica B, 2002, 323, 130-132.

[85] Frackowiak, E.; Beguin, F.; Electrochemical storage of energy in

carbon nanotubes and nanostructured carbons. Carbon, 2002, 40,

1775-1787.

[86] Landi, B. J.; Ganter, M.J.; Schauerman, C.M.; Cress, C.D.;

Raffaelle, R.P. Lithium Ion Capacity of Single Wall Carbon

Nanotube Paper Electrodes. J. Phys. Chem. C, 2008, 112, 7509-

7515.

[87] Wu, G. T.; Wang, C. S.; Zhang, X. B.; Yang, H. S.; Qi, Z. F.; He,

P. M.; Li, W. Z. Structure and lithium insertion properties of

carbon nanotubes. J. Electrochemical Society, 1999, 146, 1696-

1701.

[88] Welnaa, D.T.; Qu, L.; Taylor, B. E.; Dai, L.; Durstock M. F.

Vertically aligned carbon nanotube electrodes for lithium-ion

batteries. J. Power Sources, 2011, 196, 1455-1460.

[89] Lu, W.; Goering, A.; Qu, L.; Dia, L. Lithium-ion batteries based on

vertically-aligned carbon nanotube electrodes and ionic liquid

electrolytes. Phys. Chem. Chem. Phys., 2012, 14, 12099-12104.

[90] Oktaviano, H. S.; Yamada, K.; Waki, K. Nano-drilled multiwalled

carbon nanotubes: characterizations and application for LIB anode

materials. J. Mater. Chem., 2012, 22, 25167-25173.

34 Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 Sen et al.

[91] Wang, X. X.; Wang, J. N.; Zhang, Y. F. Preparation of Short

Carbon Nanotubes and Application as an Electrode Material in Li-

Ion Batteries. Adv. Funct. Mater., 2007, 17, 3613-3618.

[92] DiLeo, R. A.; Castiglia, A.; Ganter, M. J.; Rogers, R. E.; Cress, C.

D.; Raffaelle R. P.; Landi, B. J. Enhanced Capacity and Rate

Capability of Carbon Nanotube Based Anodes with Titanium

Contacts for Lithium Ion Batteries. ACS Nano, 2010, 4, 6121-6131.

[93] Pan, D.; Wang, S.; Zhao, B.; Wu, M. H.; Zhang, H. J.; Wang, Y.;

Jiao, Z. Lithium storage properties of disordered grapheme

nanosheets. Chem. Mater., 2009, 21, 3136-3142.

[94] Deng, D.; Kim, M. G.; Lee, J. Y.; Cho, J. Green energy storage

materials: nanostructured TiO2 and Sn-based anodes for lithium-ion

batteries, Energy Environ. Sci., 2009, 2, 818-837.

[95] Koudriachova, M. V.; Harrison N. M.; Leeuw, S. W. Diffusion of

Li-ions in rutile. An ab initio study, Solid State Ionics, 2003, 157,

35-38.

[96] Koudriachova, M. V.; Harrison N. M.; Leeuw, S. W. Density-

functional simulations of lithium intercalation in rutile, Phys. Rev.B, 2002, 65, 235423 1-12.

[97] Stashans, A.; Lunell, S.; Bergstroem, R. Theoretical study of

lithium intercalation in rutile and anatase, Phys. Rev. B, 1996, 53,

159-170.

[98] Jiang, C.; Honma, I.; Kudo, T.; Zhou, H.; Nanocrystalline rutile

TiO2 electrode for high-capacity and high-rate lithium

storage.Electrochem. Solid-State Lett., 2007, 10, A127-A129.

[99] Jiang, C.; Wei, M.; Qi, Z.; Kudo, T.; Honma, I.; Zhou, H. Particle

size dependence of the lithium storage capability and high rate

performance of nanocrystalline anatase TiO2 electrode. J. Power Sources, 2007, 166, 239-243.

[100] Armstrong, G.; Armstrong, A. R.; Canales, J.; Bruce, P. G.;

TiO2(B) nanotubes as negative electrodes for rechargeable lithium

batteries. Electrochem. Solid-State Lett., 2006, 9, A139-A143.

[101] Zachauchristiansen, B.; West, K.; Jacobsen, T.; Atlung, S. Lithium

insertion in different TiO2 modifications. Solid State Ionics, 1988,

28, 1176-1182.

[102] Zachauchristiansen, B.; West, K.; Jacobsen, T.; Skaarup, S.

Lithium insertion in isomorphousMO2(B) structures. Solid State

Ionics, 1992, 53, 364-369.

[103] Armstrong, A. R; Armstrong, G.; Canales, J.; Garcia, R.; Bruce, P.

G. Lithium-ion intercalation into TiO2-B nanowires. Adv. Mater., 2005, 17, 862-865.

[104] Inaba, M.; Oba, Y.; Niina, F.; Murota, Y.; Ogino, Y.; Tasaka, A.;

Hirota, K. TiO2(B) as a promising high potential negative electrode

for large-size lithium-ion batteries. J. Power Sources, 2009, 189,

580-584.

[105] Liu, S.; Jia, H.; Han, L.; Wang, J.; Gao, P.; Xu, D.; Yang, J.; Che,

S. Nanosheet-constructed porous TiO2-B for advanced lithium ion

batteries. Adv. Mater., 2012, 24, 3201-3204.

[106] Liu, H.; Bi, Z.; Sun, X. -G.; Unocic, R. R.; Paranthaman, M. P.;

Dai, S.; Brown, G. M. Mesoporous TiO2-B microspheres with

superior rate performance for lithium ion batteries. Adv. Mater.,

2011, 23, 3450-3454.

[107] Ren, Y.; Bruce, P. G. NanoparticulateTiO2 (B): an anode for

lithium-ion batteries. Angew. Chem. Int. Ed., 2012, 51, 2164-2167.

[108] Jang, H.; Suzuki, S.; Miyayama, M. Synthesis of open tunnel-

structured TiO2(B) by nanosheets processes and its electrode

properties for Li-ion secondary batteries. J. Power Sources, 2012,

203, 97-102.

[109] Thackeray, M.M.; Coetzer, J. A preliminary investigation of the

electrochemical performance of -Fe2O3 and Fe3O4 cathodes in

high-temperature cells. Mater. Res. Bull., 1981, 16, 591–597.

[110] Poizot, P.; Laruelle, S.; Grugeon, S.; Dupont, L.; Tarascon, J-M.

Nano-sized transition-metal oxides as negative-electrode materials

for lithium-ion batteries. Nature, 2000, 407, 496-499.

[111] Wang, B.; Chen, J.S.; Wu, H.B.; Wang, Z.; Lou, X.W.

Quasiemulsion-templated formation of -Fe2O3 hollow spheres

with enhanced lithium storage properties. J. Am. Chem. Soc., 2011,

133, 17146-17148.

[112] Wang, Z.; Luan, D.; Madhavi, S.; Li, C.M.; Lou, X.W. -Fe2O3

nanotubes with superior lithium storage capability. Chem. Commun., 2011, 47, 8061-8063.

[113] Lin, Y. M.; Abel, P. R.; Heller, A.; Mullins C. B. -Fe2O3 nanorods

as anode material for lithium ion batteries. J. Phys. Chem. Lett.,

2011, 2, 2885-2891.

[114] Wang, L.; Liu, B.; Ran, S.; Huang, H.; Wang, X.; Liang, B.;

Chenand, D.; Shen, G.;Nanorod-assembled Co3O4 hexapods with

enhanced electrochemical performance for lithium-ion batteries. J.

Mater. Chem., 2012, 22, 23541-23546.

[115] Li, Y.; Tan, B.; Wu, Y. Mesoporous Co3O4 nanowire array for

lithium ion batteries with high capacity and rate capability. Nano Lett., 2008, 8, 265-270.

[116] Yan, N.; Hu, L.; Li, Y.; Wang, Y.; Zhong, H.; Hu, X.; Kong, X.;

Chen, Q. Co3O4nanocages for high-performance anode material in

lithium-ion batteries. J. Phys. Chem. C, 2012, 116, 7227-7235.

[117] Wu, Z.-S.; Ren, W.; Wen, L.; Gao, L.; Zhao, J.; Chen, Z.; Zhou,

G.; Li, F.; Cheng, H.-M. Graphene anchored with Co3O4

nanoparticles as anode of lithium ion batteries with enhanced

reversible capacity and cyclic performance. ACS Nano, 2010, 4,

3187-3194.

[118] Li, B.; Cao, H.; Shao, J.; Li, G.; Qu, M.; Yin, G. Co3O4@graphene

composites as anode materials for high-performance lithium ion

batteries. Inorg. Chem., 2011, 50, 1628-1632.

[119] Tummala, R.; Guduru, R. K.; Mohanty, P. S. Binder free, porous

and nanostructured Co3O4 anode for Li-ion batteries from solution

precursor plasma deposition. J. Power Sources, 2012, 199, 270-

277.

[120] Sun, Y.; Hu, X.; Luo, W.; Huang, Y. Self-assembled mesoporous

CoO nanodisks as a long-life anode material for lithium-ion

batteries. J. Mater. Chem., 2012, 22, 13826-13831.

[121] Sahay, R.; Kumar, P.S.; Aravindan, V.; Sundaeamurthy, J.; Ling,

W.C.; Mhaisalkar, S.G.; Ramakrishna, S.; Madhavi, S. High aspect

ratio electrospun CuO nanofibers as anode material for lithium-ion

batteries with superior cycleability. J. Phys. Chem. C, 2012, 116,

18087-18092.

[122] Ko, S.; Lee, J.-I.; Yang, H. S.; Park, S.; Jeong, U. Mesoporous CuO

particles threaded with CNTs for high-performance lithium-ion

battery anodes. Adv. Mater., 2012, 24, 4451-4456.

[123] Wang, Z.; Madhavi, S.; Lou, X. W. Ultralong -MoO3 Nanobelts:

Synthesis and effect of binder choice on their lithium storage

properties, J. Phys. Chem. C, 2012, 116, 12508-12513.

[124] Zhao, G.; Zhang, N.; Sun, K. Electrochemical preparation of

porous MoO3 film with a high rate performance as anode for

lithium ion batteries. J. Mater. Chem . A, 2013, 1, 221- 224.

[125] Sun , Y.; Wang , J.; Zhao , B.; Cai, R. Ran, R.;Shao, Z. Binder-free

-MoO3 nanobelt electrode for lithium-ion batteries utilizing van

der Waals forces for film formation and connection with current

collector. J. Mater. Chem. A, 2013, 1, 4736-4746.

[126] Sen, U. K.; Mitra, S. High rate and high energy density lithium-ion

battery anode containing 2D MoS2 nanowall and cellulose binder.

ACS Appl. Mater. Interfaces, 2013, 5, 1240-1247.

[127] Zhou, X.; Wan L.-J.; Guo, Y.-G. Synthesis of MoS2 nanosheet-

graphenenanosheet hybrid materials for stable lithium storage.

Chem. Commun., 2013, 49, 1838-1840.

[128] Bindumadhavan, K.; Srivastava, S. K.; Mahanty, S. MoS2-

MWCNT hybrids as a superior anode in lithium-ion batteries.

Chem. Commun., 2013, 49, 1 823-1825.

[129] Taberna, P.L.; Mitra, S.; Poizot, P.; Simon, P.; Tarascon, J.-M.

High rate capabilities Fe3O4-based Cu nano-architectured

electrodes for lithium-ion battery applications. Nature Mater.,

2006, 5, 567-573.

[130] Mitra, S.; Poizot, P.; Finke, A.; Tarascon, J.-M. Growth and

electrochemical characterization vs Li of Cu-supported Fe3O4

electrodes made by electrodeposition. Adv. Funct. Materials, 2006,

16, 2281-2287.

[131] Reddy, M.V.; Yu, T.; Sow, C.-H.; Shen, Z.X.; Lim, C.T.; Rao,

G.V.S.; Chowdari, B.V.R. -Fe2O3 nanoflakes as an anode material

for Li-ion batteries. Adv. Funct. Mater., 2007, 17, 2792-2799.

[132] Chen, D.; Ji, G.; Ma, Y.; Lee, J. Y.; Lu, J. Graphene-encapsulated

hollow Fe3O4 nanoparticle aggregatesas a high performance anode

material for lithium ion batteries. ACS Appl. Mater. Interfaces,

2011, 3, 3078-3083.

[133] Gillot, F.; Boyanov, S.; Dupont, L.; Doublet, M.-L.; Morcrette, M.;

Monconduit, L.; Tarascon, J.-M. Electrochemical reactivity and

design of NiP2 negative electrodes for secondary Li-ion batteries.

Chem. Mater. 2005, 17, 6327-6337.

[134] Lu, Y.; Tu, J.P.; Xiang, J.Y.; Wang, X.L.; Zhang, J.; Mai, Y.J.;

Mao, S.X. Improved electrochemical performance of self-

assembled hierarchical nanostructured nickel phosphide as a

negative electrode for lithium ion batteries. J. Phys. Chem. C, 2011,

115, 23760-23767.

[135] Lu, Y.; Wang, X.; Mai, Y.; Xiang, J.; Zhang, H.; Li, L.; Gu, C.; Tu,

J.; Mao, S.X. Ni2P/Graphene Sheets as Anode Materials with

Nano Dimensionality: A Way towards Better Li-Ion Storage Nanoscience & Nanotechnology-Asia, 2013, Vol. 3, No. 1 35

Enhanced Electrochemical Properties versus Lithium. J. Phys.

Chem. C, 2012, 116, 22217-22225.

[136] Park, C.-M.; Kim, Y.U.; Sohn, H.-J. Topotactic Li insertion/

extraction in hexagonal vanadium monophosphide. Chem. Mater., 2009, 21, 5566-5568.

[137] Chandrasekar, M.S.; Mitra, S. Thin copper phosphide films as

conversion anode for lithium-ion battery applications. Electrochim.

Acta, 2013, 92, 47-54.

[138] Gregory, D. H.; O’Meara, P. M.; Gordon, A. G.; Hodges, J. P.;

Short, S.; Jorgensen, J. D. Structure of lithium nitride and transition

metal doped derivatives, Li3-x-yMxN (M = Ni, Cu): a powder

neutron diffraction study. Chem. Mater., 2002, 14, 2063-2070.

[139] Ducros, J. B.; Bach, S.; Ramos J. P. P.; Willmann, P. Comparison

of the electrochemical properties of metallic layered nitrides

containing cobalt, nickel and copper in the 1 V-0.02 V potential

range. Electrochem.Commun., 2007, 9, 2496-2500.

[140] Ducros, J. B.; Bach, S.; Ramos J. P. P.; and P. Willmann, Layered

lithium cobalt nitrides: a new class of lithium intercalation

compounds. J. Power Sources, 2008, 175, 517-525.

[141] Bach, S.; Ramos, J. P. P.; Ducros, J. B.; Willmann, P. Structural

and electrochemical properties of layered lithium nitridocuprates

Li3-xCuxN. Solid State Ionics, 2009, 180, 231-235.

[142] Das, B.; Reddy, M. V.; SubbaRao G. V.; Chowdari, B. V. R.

Synthesis and Li-storage behavior of CrN nanoparticles. RSC Adv., 2012, 2, 9022-9028.

[143] Das, B.; Reddy, M. V.; SubbaRao, G. V.; Chowdari, B. V. R.

Synthesis of porous-CoN nanoparticles and their application as a

high capacity anode for lithium-ion batteries. J. Mater. Chem., 2012, 22, 17505-17510.

[144] Sun, Q.; Fu, Z.-W.Vanadium nitride as a novel thin film anode

material for rechargeable lithium batteries. Electrochim. Acta.,

2008, 54, 403-409.

[145] Pereira, N.; Klein, L. C.; Amatucci, G. G. The electrochemistry of

Zn3 N 2 and LiZnN: a lithium reaction mechanism for metal nitride

electrodes. J. Electrochem. Soc., 2002, 149, A262-A271.

[146] Shodai, T.; Okada, S.; Tobishima, S.; Yamaki, J. Anode

performance of a new layered nitride Li3-xCoxN (x =0.2-0.6). J.

Power Sources, 1997, 68, 515-518.

[147] Yamada, A.; Matsumoto, S.; Nakamura, Y. Direct solid-state

synthesis and large-capacity anode operation of Li3-xFexN. J. Mater. Chem., 2011, 21, 10021-10025.

[148] Oumellal, Y.; Rougier, A.; Nazri, G.A.; Tarascon, J.-M.; Aymard,

L. Metal hydrides for lithium-ion batteries. Nature Mater., 2008, 7,

916-921.

[149] Wu, H.; Cui, Y. Designing nanostructured Si anodes for high

energy lithium ion batteries. Nanotoday, 2012, 7, 414-429.

[150] Chan, C. K.; Peng, H. L.; Liu, G.; McIlwrath, K.; Zhang, X. F.;

Huggins, R. A.; Cui, Y. High-performance lithium battery anodes

using silicon nanowires. Nat. Nanotechnol., 2008, 3, 31-35.

[151] Kim, H.; Cho, J. Superior lithium Electroactive mesoporous

Si@carbon core-shell nanowires for lithium battery anode material.

Nano Lett., 2008, 8, 3688-3691.

[152] Cui, L.-F.; Yang, Y.; Hsu, C.-M.; Cui, Y. Carbon silicon core shell

nanowires as high capacity electrode for lithium ion batteries. Nano Lett., 2009, 9, 3370-3374.

[153] Cui, L.-F.; Ruffo, R.; Chan, C. K.; Peng, H.; Cui, Y. Crystalline

amorphous core shell silicon nanowires for high capacity and high

current battery electrodes. Nano Lett., 2009, 9, 491-495.

[154] Peng, K.; Jie, J.; Zhang, W.; Lee, S.-T. Silicon nanowires for

rechargeable lithium-ion battery anodes. Appl. Phys. Lett. 2008, 93,

033105 1-3.

[155] Park, M.-H.; Kim, M. G.; Joo, J.; Kim, K.; Kim, J.; Ahn, J.; Cui,

Y.; Cho. J. Silicon nanotube battery anodes. Nano Lett., 2009, 9,

3844-3847.

[156] Kumar, K.S.; Swygenhoven, H. V.; Suresh, S. Mechanical behavior

of nanocrystalline metals and alloys. Acta Mater., 2003, 51, 5743-

5774.

[157] Meyers, M.A.; Mishra, A.; Benson. D. J. Mechanical properties of

nanocrystalline materials. Prog. Mater. Sci., 2006, 51, 427-556.

[158] Huang, R.; Fan, X.; Shen, W.C.; Zhu, J. Carbon-coated silicon

nanowire array films for high-performance lithium-ion battery

anodes. Appl. Phys. Lett., 2009, 95, 133119 1-3.

[159] Yao, Y.; McDowell, M.T.; Ryu, I.; Wu, H.; Liu, N.A.; Hu, L.B.;

Nix, W. D.; Cui, Y. Interconnected silicon hollow nanospheres for

lithium-ion battery anodes with long cycle life. Nano Lett., 2011,

11, 2949-2954.

[160] Wu, H.; Chan, G.; Choi, J.W.; Ryu, I.; Yao, Y.; McDowell,

M.T.; Lee, S.W.; Jackson, A.; Yang, Y.; Hu, L. B.; Cui, Y. Stable

cycling of double-walled silicon nanotube battery anodes through

solid-electrolyte interphase control. Nat. Nanotechnol., 2012, 7,

309-314.

[161] Hertzberg, B.; Alexeev, A.; Yushin, G. Deformations in Si-Li

anodes upon electrochemical alloying in nano-confined space. J.

Am. Chem. Soc., 2010, 132, 8548-8549.

[162] Wu, H.; Zheng, G.; Liu, N.; Carney, T. J.; Yang, Y.; Cui, Y.

Engineering empty space between Si nanoparticles for lithium-ion

battery anodes. Nano Lett., 2012, 12, 904-909.

[163] Li, G.; Yang, Z.; Yang, W.; Effect of FePO4 coating on

electrochemical and safety performance of LiCoO2 as cathode for

Li-ion batteries. J. Power Sources, 2008, 183, 741-748.

Received: March 20, 2013 Revised: May 16, 2013 Accepted: July 02, 2013