You are on page 1of 204

University of Wollongong Theses Collection

University of Wollongong Theses Collection


University of Wollongong Year 

A study of electro materials for


lithium-ion batteries
Yueping Yao
University of Wollongong

Yao, Yueping, A study of electro materials for lithium-ion batteries, PhD thesis, In-
stitute for Superconducting and Electronic Materials, University of Wollongong, 2008.
http://ro.uow.edu.au/theses/88

This paper is posted at Research Online.


http://ro.uow.edu.au/theses/88
A STUDY OF ELECTRO MATERIALS

FOR LITHIUM-ION BATTERIES

A thesis submitted in fulfillment of


the requirements for the award of

Doctor of Philosophy

YUEPING (Jane) YAO, B.Sc. M. Eng

University of Wollongong

Institute for Superconducting & Electronic Materials


Faculty of Engineering

2008
CANDIDATE’S CERTIFICATE

This is to certify that the work presented in this thesis is original and was carried out by

the candidate in the laboratories of the Institute for Superconducting & Electronic

Materials (ISEM) and the Faculty of Engineering at the University of Wollongong, New

South Wales, Australia. It has not been submitted for a degree to any other university or

institution.

Yueping (Jane) Yao

Wollongong, Australia
2008

i
ACKNOWLEDGEMNENTS

I would firstly like to express my sincere gratitude and thanks to my research supervisors,

Professor H. K. Liu and Dr. K. Konstantinov, for their consistent academic supervision,

guidance and encouragement throughout may research. I also express my gratitude to the

Director of ISEM, Professor Shixue Dou, for support and financial assistance for my PhD

project. I am deeply indebted to the Australian Research Council (ARC) for the financial

support through the ARC Linkage project LP0453766.

I thank all the hard-working technical and administrative staff in ISEM and the Faculty of

Engineering, in particular Mr. Ron Kinnel, Mr. N. Mackie, and Mr. G. Tillman. They have

provided me with tremendous help in all my experiments. Thanks are also to Dr. T. Silver

for providing great help in proofreading this thesis.

Finally, I would like to express my gratitude to all other staff and to my fellow students at

the Institute for Superconducting and Electronic Materials for their constant

encouragement.

ii
Abstract
--------------------------------------------------------------------------------------------------------------------------

Abstract

Lithium-ion batteries (or rechargeable lithium batteries) are most advanced battery

technology for modern portable electronics such as mobile phones, notebook

computers and cameracorders. There are also big potentials for lithium-ion batteries

to be used for electric vehicles (EVs), hybrid electric vehicles (HEVs) and stationary

power storage. In particular, the later will bring a significant contribution to reduce

green-house gas emissions and address global warming and climate change. Materials

research plays a key role in the development of next generation of advanced

lithium-ion batteries with high energy density, high power density, and long cycle

life.

This PhD thesis describes my exploration on developing new anode materials and

cathode materials for lithium-ion batteries. I firstly investigated silicon based anode

materials since silicon has the highest theoretical lithium storage capacity of about

4200 mAh/g when forming Li21Si5 alloys. The reversible lithium storage mechanism

is totally different from that of graphite based anode. It relies on a process called

alloying and dealloying instead of intercalation and de-intercalation. However, the

formation of Li21Si5 alloys can induce more than 400% volume expansion. The

repeated expansion and shrinkage of the silicon electrode will cause cracking and

eventually failure of the battery system. A general strategy has been employed to

solve this problem. Firstly nanosize silicon powders were used to minimize the

volume expansion in local domains. Secondly, silicon particles were embedded in

carbon matrix to buffer the volume change during the reaction with lithium.

iii
Abstract
--------------------------------------------------------------------------------------------------------------------------

Si-mesocarbon microbeads (MCMB) composite anode materials were produced by

ball-milling. Si-MCMB composite electrodes demonstrated superior performance

(high capacity and satisfactory cyclability), compared to bare MCMB and bare

nano-Si electrodes. Silicon-amorphous carbon composite anode materials were also

prepared by carbon aerogel method, through which nanosize silicon particles are

homogeneous distributed in carbon matrix. A reversible capacity of 1450 mAh/g for

Si-C composite anodes was achieved. The good cyclability should be attributed to the

usage of nanosize Si powders and their homogeneous distribution in an amorphous

carbon matrix.

Carbon nanotubes have many unique and intriguing properties, including as anode

materials for lithium-ion batteries. Vertically aligned multiwalled carbon nanotubes

(VAMWCNTs) were prepared by chemical vapor deposition method. Nanosize

SnO2-MWCNTs composites were also synthesized. The VAMWCNTs have a typical

diameter of several tens of nanometers and consist of compartment structures. Cyclic

voltammetry measurements show that the carbon nanotubes are electrochemically

active to lithium insertion and extraction. A reversible lithium storage capacity of 950

mAh/g has been achieved for CNTs anodes. The solution-based chemical process

enables Sn2+ ions to penetrate into the inner cavity of the carbon nanotubes. The

SnO2-CNTs composite electrodes exhibited stable cyclability with a lithium storage

capacity of 410 mAh/g after fifty cycles.

Transition metal phosphides such as MnP4 and Zn3P2 were discovered to exhibit

interesting phenomena with respect to reversible lithium storage. They were

iv
Abstract
--------------------------------------------------------------------------------------------------------------------------

conventionally synthesized by solid state sintering at high temperature for long

periods. After sintering, the products were purified by acid etching, which is a very

tedious process. Crystalline iron phosphide (FeP4) powders were directly prepared by

a solvothermal synthesis technique. Cyclic voltammetry measurements demonstrated

the reversible reactivity of FeP4 anodes towards lithium insertion and extraction. The

FeP4 anode exhibited a stable lithium storage capacity about 700 mAh/g.

Lithium iron phosphate has been emerging as a new cathode material for lithium-ion

batteries with low cost. However, lithium iron phosphate has a very lost electronic

conductivity, inducing low rate capacity and preventing commercial application.

Various cation dopings have been studied with the goal to improve the overall

electrochemical performance of lithium iron phosphate. The synthesis, crystal

structure refinement, magnetic and electrochemical properties of a series of

LiMnxFe1-xPO4 cathode materials was investigated. A number of conductive

phosphides and manganese phosphates were found to be formed during the sintering

process with the effect of enhancing the electronic conductivity of the materials. The

effect of sintering temperature towards the crystal size of LiFePO4 and

Li0.95Mg0.05PO4 compounds were systematically investigated. LiFePO4 and

Li0.95Mg0.05PO4 samples exhibit a typical antiferromagnetic behaviour. This

antiferromagnetism could be induced by long range Fe-O-P-O-Fe triple exchange due

to the lack of direct Fe-O-Fe interactions. LiFePO4 and Li0.95Mg0.05PO4 electrodes

show specific capacity in the range 150 mAh/g – 160 mAh/g.

v
Contents
----------------------------------------------------------------------------------------------------------------------------

Contents

CANDIDATE’S CERTIFICATE i

ACKNOWLEDGEMENTS ii

ABSTRACT iii

CONTENTS vi

CHAPTER 1 INTRODUCTION 1

CHAPTER 2 LITERATURE REVIEW

2.1 Batteries 4

2.2 Rechargeable batteries 6

2.3 An introduction of lithium-ion batteries 7

2.4 Fundamentals of Lithium ion Batteries 10

2.5 Cathode Materials for Lithium-ion Batteries 14

2.5.1 LiCoO2 and doped LiMxCo1-xO2 cathode materials 15

2.5.2 LiNiO2 and doped LiMxNi1-xO2 cathode materials 19

2.5.3 Layered LiMnO2, orthorhombic LiMnO2 and their doped

compounds 23

2.5.4 Spinel LiMn2O4 and LiMxMn2-yO4 cathode materials 28

2.5.5 Lithium iron phosphate based cathode materials 36

2.5.5.1 Crystal structure of LiFePO4 38

vi
Contents
----------------------------------------------------------------------------------------------------------------------------

2.5.5.2 Enhancement of electronic conductivity and

improvement of electrochemical properties of

LiFePO4 cathode material 43

2.6 Anode Materials for Lithium-ion Batteries 46

2.6.1 Carbon based anode materials 47

2.6.2 Alloy based anode materials 52

2.6.3 Metal oxide anode materials 56

2.6.3.1 Tin oxide anode materials 56

2.6.3.2 Transition metal oxide anode materials 59

CHAPTER 3 EXPERIMENTAL

3.1 Materials and chemicals 62

3.2 Materials synthesis techniques 63

3.2.1 High energy ball milling 63

3.2.2 Carbon aerogel synthesis 63

3.2.3 Chemical vapor deposition (CVD) 64

3.2.4 Hydrothermal synthesis 64

3.2.5 Sol-gel synthesis 64

3.2.6 Solid-state reaction synthesis 65

vii
Contents
----------------------------------------------------------------------------------------------------------------------------

3.3 Materials characterization 65

3.3.1 Structure characterization 65

3.3.2 Material morphology observation 66

3.3.3 Magnetic property measurement 68

3.3.4 Electrochemical testing 68

3.3.4.1 Electrode fabrication 68

3.3.4.2 Cell assembly 69

3.3.4.3 Electrochemical testing 70

(i) Charge/discharge of cells 70

(ii) Cyclic voltammetry 70

CHAPTER 4 PREPARATION AND CHARACTERIZATION OF Si-

MCMB AND SI-CARBON COMPOSITE ANODE MATERIALS

4.1 Introduction 72

4.2 Experimental 73

4.2.1 Preparation of Si-MCMB composite 73

4.2.2 Preparation of Si-carbon composite 74

4.2.3 Electrochemical testing 75

4.3 Results and discussion 75

4.3.1 Si-MCMB composite anode materials 75

4.3.2 Si-carbon composite anode materials 87

4.4 Conclusions 94

viii
Contents
----------------------------------------------------------------------------------------------------------------------------

CHAPTER 5 MULTIWALLED CARBON NANOTUBES (CNTs) ANODE

MATERIALS

5.1 Introduction 96

5.2 Experimental 97

5.2.1 Synthesis of vertically aligned CNTs 97

5.2.2 Synthesis of SnO2-CNTs composites 99

5.2.3 Physical and structural characterization 99

5.2.4 Electrochemical measurement 100

5.3 Results and discussion 100

5.4 Conclusions 116

CHAPTER 6 SYNTHESIS AND ELECTROCHEMICAL

CHARACTERISATION OF IRON PHOSPHIDE AS ANODE MATERIAL

FOR LITHIUM-ION BATTERIES

6.1 Introduction 118

6.2 Experimental 119

6.3 Results and discussion 120

6.4 Conclusions 127

ix
Contents
----------------------------------------------------------------------------------------------------------------------------

CHAPTER 7 CHARACTERISATION OF OLIVINE-TYPE LiMnxFe1-xPO4

CATHODE MATERIALS

7.1 Introduction 128

7.2 Experimental 129

7.2.1 Materials synthesis 129

7.2.2 Magnetic measurement 130

7.2.3 X-ray diffraction and SEM analysis 130

7.2.4 Electrochemical testing 130

7.3 Results and discussion 131

7.4 Conclusions 138

CHAPTER 8 ELECTROCHEMICAL AND MAGNETIC

CHARACTERISTICS OF LiFePO4 AND Li0.95Mg0.05FePO4 CATHODE

MATERIALS

8.1 Introduction 141

8.2 Experimental 143

8.2.1 Materials synthesis 143

8.2.2 Magnetic measurement 144

8.2.3 Electrochemical testing 145

8.3 Results and discussion 145

8.3.1 XRD and SEM analysis 145

x
Contents
----------------------------------------------------------------------------------------------------------------------------

8.3.2 Magnetic properties of LiFePO4 and Li0.95Mg0.05PO4 150

8.3.3 Electrochemical performance of LiFePO4 and

Li0.95Mg0.05PO4 cathode materials 157

8.4 Conclusions 161

CHAPTER 9 GENERAL CONCLUSIONS AND SUGGESTIONS FOR

FUTURE STUDY

9.1 General conclusions

9.1.1 Silicon based composite anode materials 162

9.1.2 Carbon nanotubes and tin oxide – carbon nanotubes anode

materials 163

9.1.3 Novel iron phosphide anode materials 163

9.1.4 Lithium iron phosphate and manganese doped lithium iron

phosphate cathode materials 164

9.1.5 Electrochemical and magnetic properties of magnesium

doped lithium iron phosphate cathode materials 164

9.2 Recommendations for future studies 165

REFERENCES 167

FIGURE CAPTIONS AND TABLES 183

PUBLICATIONS 189

xi
Chapter 1 Introduction
--------------------------------------------------------------------------------------------------------------------------

Chapter 1 Introduction

Our modern societies currently are heavily reliant on fossil fuels to supply energy for

all our daily lives. Global warming and climate change are among the most serious

challenges facing us today. The main cause is increasing greenhouse gas

concentrations, primarily due to the burning of fossil fuels. To protect the health and

economic well-being of our global society, we must reduce greenhouse gas emissions

by using clean energy and renewable energy. The combustion of fossil fuels in

vehicles is responsible for about 22% of global CO2 emissions. The rechargeable

battery is an important system for electrochemical energy storage and conversion.

Lithium-ion batteries have the highest energy density among all the secondary battery

systems and represent state-of-the-art power sources for consumer electronics.

Although lithium-ion batteries possess a high energy storage capability, they still

cannot meet the requirements for large-scale energy storage and electric vehicles

applications. On the other hand, the global market for small rechargeable batteries is

on an impressive scale and growing, mainly due to the spectacular increase in the use

of mobile communications and portable electronics. All of these factors drive the

development of the next generation of rechargeable lithium batteries with excellent

performance characteristics, such as high energy density, long cycle life, and

absolutely safe operation. The electrochemical performance of lithium-ion batteries is

mainly determined by the physical, chemical, and structural properties of the cathode

materials and anode materials used in the battery system. This PhD thesis aims at

exploring and investigating a number of new and innovative anode materials and

cathode materials for lithium-ion batteries. A systematic study has been performed,

1
Chapter 1 Introduction
--------------------------------------------------------------------------------------------------------------------------

including materials preparation, structural characterisation, physical characterisation

and electrochemical testing.

Chapter 2 presents a literature review on lithium-ion batteries and electrode materials

for the lithium-ion battery system. The basic concepts of batteries, rechargeable

batteries and lithium-ion batteries are reviewed. Then, the properties and current

status of all cathode materials for lithium-ion batteries are summarised. These include

LiCoO2 and doped LiCoO2, LiNiO2 and doped LiNiO2, layered LiMnO2, spinel

LiMn2O4, LiFePO4, and their transition metal doped derivatives. Finally, a

comprehensive survey on anode materials for lithium-ion batteries is conducted. The

merits, disadvantages, and challenges of carbon based anode materials, alloy based

anode materials, and metal oxide materials are analysed.

Chapter 3 contains the experimental details of this study. The experimental

procedures are fully described, along with all materials, chemicals, and facilities used

in this work.

Chapter 4 presents the synthesis, structural characteristics, and electrochemical

performance of two types of anode materials: Si-mesocarbon microbead (MCMB)

composites and nanocrystalline Si-carbon composites. By combining the high

capacity element Si and carbon, the composite anode materials can deliver much

higher capacity than bare carbon and also maintain good cyclability.

Carbon nanotubes are a new emerging one dimensional materials with many unique

functions. In Chapter 5, the reversible lithium storage properties were investigated.

2
Chapter 1 Introduction
--------------------------------------------------------------------------------------------------------------------------

Vertically aligned carbon nanotubes were synthesised by a controlled chemical

vapour deposition (CVD) technique. Tin oxide – carbon nanotubes composites were

prepared to increase the specific capacity by utilizing the high capacity of SnO2.

Chapter 6 presents the preparation and electrochemical testing of iron phosphide

anode materials. Iron phosphide compounds were prepared by hydrothermal

synthesis. The as-prepared iron phosphide anode materials were electrochemically

active in lithium-ion cells and demonstrated a stable lithium storage capacity of about

700 mAh/g.

Chapter 7 describes the structural characteristics, magnetic properties and

electrochemical performance of a series of manganese doped lithium iron phosphate

cathode materials. The detailed lattice parameters of LiMnxFe1-xPO4 compounds

were obtained through Rietveld refinement on the X-ray diffraction patterns. The

undoped lithium iron phosphate cathode shows better electrochemical performance

than the doped cathode materials.

Chapter 8 presents the electrochemical and magnetic properties of lithium iron

phosphate and magnesium doped lithium iron phosphate cathode materials, in which

magnesium ions are doped into Li sites. It was found that the electrochemical

performance of lithium iron phosphate is closely related to the sintering temperature.

Finally, an overview and summary of this study are presented in Chapter 9, and

suggestions for future work are also proposed.

3
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Chapter 2 Literature Review

2.1 Batteries

Energy consumption/production relying on fossil fuels has a detrimental impact on

the world economy and environment. Electrochemical energy is regarded as an

alternative green energy/power source. There is an increasing need for efficient, high

energy power sources for a variety of applications. Batteries, in particular,

rechargeable battery systems, are important systems for electrochemical energy

storage and conversion. In his 2006 State of the Union Address, the U.S. president

George W. Bush claimed, “We must change how we power our automobiles. We will

increase our research on better batteries for hybrid and electric cars.” The discovery of

new electrode materials for batteries could dramatically increase energy density,

reducing the cost of electrochemical power systems, and so open up the capacity to

advance into an electric vehicle/hybrid electric vehicle (EV/HEV) age.

Batteries are self-contained units that store chemical energy and convert it directly

into electrical energy to power various devices. Batteries are generally categorized

into three classes: (i) Primary batteries. They are discharged once and discarded, (ii)

Rechargeable batteries. They can be repeatedly discharged and charged many times.

The batteries can be restored to their original condition by charging with a reverse

current flow, (iii) Specialty batteries. They are designed to meet a special purpose.

Specialty batteries are mainly used for military and medical applications. Batteries

provide power sources serving all aspects of both our daily lives and industry. The

-4-
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

estimated battery market is about US $50 billion worldwide. Basically, batteries

consist of a positive electrode (the cathode), an electrolyte, and a negative electrode

(the anode). The cathode is an electron acceptor, such as lead oxide, manganese

dioxide, or lithium cobalt oxide. The anode is an electron donor, such as lead, zinc, or

lithium. The electrolyte usually is a pure ionic conductor that physically separates the

anode from cathode, allowing only ionic transport but not electronic conduction. A

schematic of battery operation is illustrated in Fig. 2.1 [1]. Battery electrolytes are

usually used in the form of a porous electrically insulating material soaked with

electrolyte. Battery electrolytes can be divided into aqueous, nonaqueous, and

solid-state electrolytes. The aqueous electrolytes are normally salts of acids or bases,

and have high ionic conductivity, on the order of 1 S/cm. From the thermodynamic

and kinetic point of view, aqueous electrolytes have an electrochemical stability

limitation less than 2.0 V. The nonaqueous organic solvent-based electrolytes have

lower ionic conductivity, on the order of 10-2 – 10-3 S/cm. The operating voltage of an

organic electrolyte is limited to ~ 4.6 V. Solid-state electrolytes are referred to as

polymer-based electrolytes and usually have low ionic conductivity in the range of

10-4 – 10-6 S/cm.

Fig. 2.1 Schematic diagram of the battery and battery operation.

-5-
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

2.2 Rechargeable batteries

A rechargeable battery system is defined as a system that can be discharged and

charged again for many times over a long period. The chemical energy can be restored

by charging the battery using a reverse current. Rechargeable batteries can be

generally divided into lead acid batteries, lithium-ion batteries, and nickel cadmium

batteries, nickel metal hydride batteries. The energy densities of those recharge able

battery systems are compared in Fig. 2.2 [2]. Among all the rechargeable battery

systems, lithium-ion batteries have the highest energy density.

Fig. 2.2 Gravimetric and volumetric energy densities of different rechargeable battery

systems [2].

-6-
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

2.3 An introduction to lithium-ion batteries

In the past, the development of the lithium based battery system was motivated by the

fundamental properties of the element lithium. Lithium is the most electronegative

(-3.04 V) element versus the standard hydrogen electrode (SHE) and also the lightest

(MW = 6.94 g/mol, ρ = 0.53g/cm3). Therefore, a battery system with lithium as the

electrode could deliver high energy density. The earliest version of the lithium battery

was the Li/(CF)n battery with a cell potential of 2.8 – 3.0 V. It was proposed that

lithium initially intercalates into the carbon mono fluoride lattice and then forms

lithium fluoride as the final reaction product [3, 4].

Li + (CF)n → Lix(CF)n → C + LiF (2.1)

Later, in the early 1970s, various primary lithium batteries were developed and

demonstrated in the United States. The Li/SOCl2 cell was one of primary lithium

battery systems [5]. Sanyo also successfully developed and marketed lithium batteries

with the Li/MnO2 cell [6]. Most primary lithium batteries have been developed for

medical and military applications. For example, implantable cardiac defibrillators use

Li/Ag2V4O11 batteries [7, 8]. The development of rechargeable lithium batteries came

from the discovery of some inorganic compounds that can react with lithium ions in a

reversible way. This led to the development of battery systems that use lithium as

anode and layered transition metal dichalchogenides (TiS2), chalcogenides (NbSe3),

and molybdenum sulfide (MoS3) as cathode [9, 10]. The discovery of so-called

intercalation compounds was a crucial step in the development of rechargeable

-7-
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

lithium batteries. The intercalation materials host lithium ions inside their crystalline

structure. The lithium ions can be reversibly inserted into and extracted from the hosts.

This topotactical electrochemical reaction could occur reversibly without a major

phase change and with no damage to the material structure. From that time, a new

electrochemistry for energy storage started to spring up, involving more and more

solid state chemistry. However, many difficulties were encountered during the early

development of rechargeable lithium batteries. A stable electrolyte medium

containing lithium had to be identified. The electrolyte must consist of some lithium

salts dissolved in an organic solvent. A reversible positive electrode material needed

to be found.

Lithium metal was initially used as the negative electrode in the rechargeable lithium

battery system. However, lithium dendrites usually grow during repeated

charge/discharge cycling, resulting in short cycle life and cell shorting. A short circuit

could cause serious safety problems such as fire and explosion, and the eventually led

to the abandonment of lithium metal as the anode material in rechargeable lithium

batteries. Substituting Al for lithium metal solved the dendrite problem, but the

electrode could only be charged and discharged for a limited number of cycles due to

significant volume change [11]. Since then, many lithium alloys have been developed

as alternative anode materials. The energy densities of lithium alloys normally are

reduced by a factor of two or three, compared to pure lithium metal. Unfortunately,

the volume changes related to the insertion/extraction of Li into/from alloy matrices

are quite substantial (200 – 600 %) [12-16]. These cause a fast disintegration of the

alloy anodes by cracking and crumbling. As a result, short cycle life and low cycling

-8-
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

efficiency are induced. In the late 1980s, graphitic carbon was found to be

dimensionally stable for lithium insertion and extraction. Intercalation of lithium ions

corresponding to the composition LiC6 results in only about a 10% increase in the

layer distance [17, 18]. Therefore, graphite became the choice of anode materials for

the lithium-ion battery in place of lithium metal and lithium alloys.

The cathode material is also a critical electrode material for the development of

rechargeable lithium batteries. The search for cathode hosts has been continued for

more than two decades. Except for low-dimensional materials, the three dimensional

framework structure compound V6O13 was also found to function for lithium perfectly

intercalation and de-intercalation [19]. The family of lithium metal oxides LixMO2

(where M is Co, Ni, or Mn), was first reported as a source of cathode materials for

rechargeable lithium batteries by Mizushima and Goodenough et al. in 1980 [20].

This family of cathode compounds is the predominant choice in current commercial

lithium ion battery production.

The electrochemical performance of the lithium-ion battery is also strongly dependent

on the electrolyte. Many solutions were proposed, from well-known propylene

carbonate (PC) and ethylene carbonate (EC) to ether-based solutions, such as

1,3-dioxolane or 3-MeTHF [21]. In general, electrolytes for the lithium-ion battery

are based on organic systems, which consist of organic solvents and lithium salts. The

moisture is strictly controlled to avoid any side reaction. Polymer electrolytes have

also been developed to construct so-called all-solid-state lithium-ion batteries. Such a

-9-
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

system provides a fundamental safeguard for battery operation since no liquid

electrolyte is present inside the battery [22, 23].

The combination of an intercalation anode host and cathode host, as well as an

appropriate organic electrolyte, has been emerging in lithium-ion battery technology

since 1990. This battery technology is also called rocking-chair technology. Sony

Corporation successfully commercialized rechargeable lithium batteries with

C/LiCoO2 in 1991 [24]. This new class of rechargeable batteries had a potential

exceeding 3.6 V (almost three times that of alkaline systems) and energy densities of

120 – 150 Wh/kg (two to three times that of Ni-Cd and Ni-MH batteries). Today,

lithium-ion batteries are becoming the dominant power sources for portable electronic

devices, military applications, aerospace, and, especially, for wireless

telecommunication.

2.4 Fundamentals of Lithium ion Batteries

Lithium metal free lithium batteries were originally termed “rocking chair” batteries

(RCB) by Armand [25]. The basic concept of operation for the RCB system was

partially derived from concentration cells, which consist of essentially identical

electrodes in different reactant concentrations [26]. After the RCB concept was

revealed, this concept was demonstrated using transition metal compound anodes and

cathodes [27-28]. Later, new names for this technology appeared: lithium ion,

Shuttlecock and Swing Electrode System, etc., but the fundamental concept remained

the same.

- 10 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Fig. 2.3 The operating principle of the lithium-ion battery.

In a lithium-ion cell, the Li+ ions shuttle between the cathode and anode hosts during

the discharge and charge processes. The principles of lithium-ion battery operation

are shown in Fig. 2.3 [29]. Lithium ions are extracted from the cathode, go through

the electrolyte and separator and are inserted into the anode structure. The reverse

process happens during discharging. In order to achieve high cycling efficiency and

- 11 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

long cycle life, the movement of Li ions in anode and cathode hosts should not change

or damage the host crystal structure. The design of a lithium-ion battery system

requires careful selection of electrode pairs to obtain a high operating voltage (Vc).

The operation of lithium-ion batteries obeys the thermodynamic laws. Therefore, the

electrochemical reaction in lithium-ion batteries follows the Gibbs-Helmholtz

equation:

ΔG = ΔH - TΔS (2.2)

ΔG - free energy of the reaction

ΔH – reaction enthalpy

ΔS – reaction entropy

The free energy ΔG is also related to cell voltage for the electrochemical reactions

occurring at the two separate electrodes, based on Faraday’s law:

ΔG = -nFE (2.3)

Where n is the number of electrons transferred; F is the Faraday constant (96500); and

E is the cell voltage (the potential difference between positive electrode and negative

electrode).

A high Vc can be realized with an anode and cathode that have, respectively, smaller

and larger work functions φa and φc. The open-circuit voltage Voc of the cell can be

calculated from the following formula:

- 12 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Voc = (φc - φa) / e (2.4)

where e is the electronic charge. A schematic energy level diagram showing the work

functions is presented in Fig. 2.4.

E
Ec
ϕa
EF

φc Eg
Voc
EF
Ev

Cathode Electrolyte/separator Anode

Fig. 2.4 Schematic energy diagram of a cell at open circuit.

The Fermi energies (EF) of the anode and cathode must lie within the band gap (Eg) of

the electrolyte. Therefore, the anode and cathode materials are thermodynamically

stable in contact with the electrolyte, and there will be no side reduction or oxidation

of the electrolyte.

- 13 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

2.5 Cathode Materials for Lithium-ion Batteries

Cathode materials play an important role in the operation of lithium-ion batteries.

They provide lithium ion sources (Li+) for the Li-ion “shuttle” between the cathode

and the anode. Since carbon materials are widely used as anodes in lithium-ion

batteries, which have a potential close to that of Li/Li+ reference electrode, the voltage

of lithium-ion batteries is mainly determined by the potential of cathode materials.

Therefore, cathode materials must meet some strict requirements: (i) The cathode host

must be structurally stable for repeated lithium ion intercalation and de-intercalation.

The variation of the lattice structure should be as small as possible during the battery

operation process to maintain long cycle life. (ii) The cathode materials must have a

high potential relative to Li/Li+ reference electrode to enable high operating voltage

(high energy density). (iii) The cathode materials should contain as great an amount of

lithium as possible, and those lithium atoms must be electrochemically extractable

from the host crystal structure (high capacity). (iv) The cathode materials should have

high electronic conductivity and high lithium chemical diffusion coefficient to keep

high rate capacity (high power density). (v) The cathode materials should be low cost,

non-toxic, and easy to prepare. A large number of materials were investigated as

cathode materials for lithium-ion batteries. However, the number of different

compounds which are suitable as cathode hosts for lithium-ion batteries is quite

limited, due to the critical requirements such as high energy density, good cyclability,

and safety.

- 14 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Although LiCoO2 cathode materials are popularly used in commercial lithium-ion

battery production, currently, new layered and spinel structure cathode materials are

under extensive investigation worldwide. Their selection is mainly driven by reasons

such as reducing the cost, increasing environmental friendliness, and developing a

new generation of lithium-ion batteries for power storage and electric vehicle (EV)

applications. The layered and spinel structure cathode materials for lithium-ion

batteries mainly include layered structure LiCoO2, LiNiO2, LiMnO2, and spinel

LiMn2O4, as well as their doped derivatives. Recently, lithium iron phosphate

compounds have also emerged as new cathode materials for lithium-ion batteries.

2.5.1 LiCoO2 and doped LiMxCo1-xO2 cathode materials

In 1980, J.B. Goodenough et al. proposed and tested LiCoO2 as cathode for lithium

batteries. It was surprisingly found that LiCoO2 cathode had a high charge and

discharge voltage in the range of 3.6 – 3.8 V in the cell configuration of LixCoO2 / Li

[30]. After almost 10 years, SONY made the first batch of lithium-ion batteries with

the combination of a LiCoO2 cathode and a carbon anode [31]. Since then, lithium-ion

batteries have been under massive commercial production and dominated the

rechargeable battery market.

LiCoO2 compound has a layered α-NaFeO2 type structure (SP: R3m), with the

oxygen in a cubic close-packed array. This structure is closely related to the rock-salt

structure. The Li+ and Co3+ ions are located at octahedral 3(a) and 3(b) sites and

ordered on alternating (111) planes, inducing a hexagonal symmetry. The LiCoO2

- 15 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

crystal lattice has lattice parameters of a = 2.816 Å and c = 14.08 Å. Fig.2.5 shows a

schematic of the LiCoO2 crystal structure.

Fig. 2.5 The structure of layered LiCoO2.

T. Ohzuku et al [32] investigated the mechanism of the reaction at the LiCoO2

electrode in lithium cells using in situ X-ray diffraction (XRD) and concluded that the

reaction at the LiCoO2 electrode in lithium cells can be divided into three regions. In

the region 0 < x < 1/4 in Li1-xCoO2, the coexistence of two hexagonal phases was

observed. The reaction in terms of lithium insertion and extraction in this region is a

topotactic two-phase reaction. In the region 1/4 < x < 3/4 in Li1-xCoO2, the reaction

taken as a whole is a single phase reaction. In the region 3/4 < x < 1, the reaction is a

two-phase reaction [32]. Removal of lithium corresponding to 1 mole Li per formula

LiCoO2 delivers a theoretical capacity of about 274mAh/g. However, only part of the

- 16 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

lithium can be reversibly extracted and inserted due to structural restriction [33]. If the

charging voltage is limited to 4.2 - 4.25 V versus Li/Li+, LiCoO2 cathode can provide

a specific capacity of 150mAh/g over 1000 charge-discharge cycles without

significant specific capacity loss.

Electrochemically active LiCoO2 compounds can be prepared by many synthetic

techniques. Solid-state reaction is commonly used by sintering a mixture of cobalt

oxide (or cobalt carbonate) and lithium carbonate (or lithium hydroxide) at high

temperature. Some solution approaches such as sol-gel or organo-chemical synthesis

for the LiCoO2 compound have also been reported [5, 6]. In general, the different

synthesis methods using high temperature sintering have no significant impact on the

electrochemical performance of the LiCoO2 electrode. However, LiCoO2 material

synthesized at low temperature demonstrated a high specific capacity in the initial few

cycles, but with poor cycle life [36].

An interfacial layer, called the solid electrolyte interphase (SEI), usually is formed on

the surface of the LiCoO2 electrode during the charge/discharge process. The

formation of this interfacial film, caused by the reaction between the electrolyte and

the LiCoO2 compound, has an important influence on the electrochemical properties

of the LiCoO2 electrode. The characterization of this interfacial film has been

conducted by many researchers via a.c. impedance, electron microscopy and atomic

force microscopy(AFM) [37, 38].

- 17 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Since mineral resources of cobalt are very limited, it is better to partially replace Co

with other cheap transition metal elements in the LiCoO2 crystal structure to reduce

the cost of cathode materials. LiNiO2 can form a solid solution with LiCoO2 over the

entire composition range LiNixCo1-xO2 (0 ≤ x ≤ 1) [39]. R. Alcántara et al.

investigated both Co rich and Ni rich LiNixCo1-xO2 compounds. It was found that the

electrochemical behavior of LiNixCo1-xO2 solid solutions was significantly influenced

by the synthesis temperature. For a Ni rich composition (LiNi0.88Co0.12O2), the

electrochemical performance was poor, if prepared below 750 ºC. However, if the

sample was prepared at 750 ºC, it showed better reversibility, higher than that of a

Co-rich composition [40]. Tsutomu Ohzuku et al theoretically studied the solid- state

redox potentials of LiCoO2, LiCo1/2Ni1/2O2, and LiNiO2. They found that a one-to-one

mixture of Co and Ni in the layered structure induced a mixing effect on the redox

potentials. LiCo1/2Ni1/2O2 cathode has characteristic redox potential systems at 4.58 V,

4.05 V, and 3.58 V, respectively. The electron donating or accepting center in

LiCo1/2Ni1/2O2 is predominantly the Co at the potential of 4.58 V and Ni at the

potential of 3.58 V [41]. Al, which is not a transition metal, can also substitute for Co

in the LiCoO2 structure, because the ionic radii of Al3+ (0.54 Å) and Co3+ (0.55 Å) are

very close, and α-LiAlO2 is isostructural with LiCoO2. LiAlxCo1-xO2 compounds (0.1

≤ x ≤ 0.3) have been successfully synthesized by solid-state reaction. The redox

potential increases, due to the substitution of Al in LiCoO2. A high initial capacity of

160mAh/g was observed for LiAl0.15Co0.75O2 cathode with stable cyclability [42].

The capacity and cyclability of LiCoO2 cathode can be further improved by coating a

metal oxide or phosphate on the surface of the cathode material particles. The coated

- 18 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

LiCoO2 cathode demonstrated an increased capacity of 170 mAh/g when cycled

between 2.75 V and 4.4 V. The mechanism could be attributed to minimizing the

reactivity of Co4+ with the acidic HF in the electrolyte in the fully charged state [43,

44]. It has also been identified that a coating of high-fracture-toughness metal oxide

such as ZrO2, Al2O3, TiO2 or B2O3 on LiCoO2 particles can effectively suppress the

change in the lattice constants during the lithium intercalation and de-intercalation

processes, and therefore suppresses the phase transition. The capacity retention of

LiCoO2 cathode can be dramatically enhanced through such a coating [45, 46].

The use of LiCoO2 as the cathode material in large scale lithium prototype batteries

has been demonstrated by several institutions. Cylindrical 1.5 - 2 Ah and 8 Ah cells,

having a flat plate construction inside, were assembled and tested [47]. Cylindrical

“C” size cells showed high capacity and a cycle life of more than 100 cycles.

Although LiCoO2 compound has excellent electrochemical properties as a cathode

material in lithium ion batteries, some drawbacks have to be addressed and cannot be

ignored. The major drawback is the high cost for LiCoO2 due to the high price and

limited natural abundance of Co. Another problem is the toxicity of Co, which causes

a serious environmental impact when batteries are disposed of. These two factors

forbid the application of LiCoO2 as the cathode material in any large batteries for

electric vehicle and power storage applications.

- 19 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

2.5.2 LiNiO2 and doped LiMxNi1-xO2 cathode materials

LiNiO2 is an attractive candidate for lithium ion batteries because the natural

resources of nickel are relatively abundant. The theoretical capacity of LiNiO2

cathode is close to that of LiCoO2, i.e. 275 mAh/g. However, it can deliver a practical

specific capacity in the range of 185 – 210 mAh/g by charging up to 4.1 – 4.2 V vs.

LiNiO2 cathode suffers from a high capacity fading rate and poor thermal stability in

the highly oxidized state (Ni4+). LiNiO2 is isostructural with LiCoO2. The lattice

dimensions of the LiNiO2 unit cells (ah = 2.9 Å, ch = 14.2 Å, ch/ah = 4.9 Å in the

hexagonal setting) are very close to the parameters of a cubic unit cell (ac = ch / 2 3

= 4.1 Å), implying that nickel and lithium ions can easily displace each other in the

crystal structure. Once nickel ions occupy the Li ion position (3a) in the LiNiO2

structure, Ni3+/Ni4+ ions will block Li+ diffusion, resulting in a low lithium diffusion

coefficient and low capacity of the electrode [48].

Poor cycle ability of LiNiO2 electrode has been reported, but the cycle life of the

LiNiO2 electrode is greatly influenced by the way the compound is prepared. If

properly synthesized, the cycle life of LiNiO2 can be improved to an acceptable level.

No evolution of the structure of the LiNiO2 compound was observed even after 1200

cycles. Solid-state reactions are generally used to synthesize the LiNiO2 compound

under appropriate temperature and strongly oxidizing conditions. Unlike to LiCoO2,

which can be prepared by sintering a mixture of almost any Li, Co and O sources, the

precursors for the preparation of LiNiO2 have to be carefully chosen. The typical

approach is to heat a mixture of hydroxides (e.g. nickel hydroxide and lithium

- 20 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

hydroxide), nickel hydroxide and lithium nitride, nickel oxide and lithium hydroxide

or by the use of Na2O2 and Ni sources, followed by ion exchange with LiNO3 at

elevated temperature. With careful control of all of the above factors, LiNiO2

compound with good electrochemical performance as cathode materials for lithium

batteries could be obtained. Li/LiNiO2 and C/LiNiO2 cells have been assembled using

liquid electrolyte. Results using personal area network (PAN) based hybrid polymer

electrolyte were also obtained [49]. Long cycle life, excellent high temperature

performance and low self-discharge rates have all been demonstrated using

LiNiO2/coke cells [50].

Layered LiNiO2 compounds have been synthesized using a variety of conditions

including varied nominal Li: Ni ratios, different sintering temperatures and different

atmospheres [51]. When sintered at low temperatures such as 650 ºC or 700 ºC, the

impurity phases Li2O and NiO were present in the final product. It was also revealed

that too much excess Li (Li:Ni = 1.14:1) also induces the presence of Li2O impurity.

An appropriate amount of excess lithium in the precursors can compensate for the loss

of the volatile Li at high temperature, and therefore is necessary. The decomposition

product Li1-xNi1+xO2 has a disordered cation distribution, with nickel ions partially

occupying the lithium ion sites in the layered structure, resulting in very poor

electrochemical performance. So, the synthesis temperature for LiNiO2 should be kept

below 860 ºC to avoid any decomposition and oxygen evolution. The properly

prepared LiNiO2 cathode showed a discharge capacity of 180 mAh/g with good

reversibility. A.c. impedance analysis was performed on LiNiO2 cathode tostudy the

kinetic processes of lithium insertion and extraction. LiNiO2 compound synthesized

- 21 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

in air has a high charge-transfer resistance (RCT) and low exchange current density.

Oxygen atmosphere is favorable for preparing ordered LiNiO2 compound.

Many efforts have been made to improve the electrochemical performance of LiNiO2

cathode by partial substitution of Ni by other metal elements such as Mg, Al, Mn, Cr,

Fe and Ti respectively [52 - 54]. LiCoxNi1-xO2 solid solutions ( x = 0, 0.25, 0.5, 0.75

and 1) were prepared by sintering the precursor reagents LiOH.H2O, NiO, and CoO at

750 ºC for 24 h under oxygen flow [55]. Electrochemical testing showed that LiNiO2

cathode had a higher initial capacity, but less stability on cycling comparing to

LiCoO2 cathode. The cyclability of LiNi0.5Co0.5O2 and LiNi0.25Co0.75O2 cathodes are

satisfactorily good, with a capacity fading rate of 0.3 and 0.18mAh/g per cycle

respectively. The replacement of Ni by Co in LiCoxNi1-xO2 cathode can improve the

cyclability, but with a sacrifice of the initial capacity.

The structural integrity of the LiNiO2 cathode could be preserved via inert Al3+ cation

doping to prevent overcharge of the electrode, which is beneficial for extending the

cycle life of the cell. LiAlxNi1-xO2 ( x = 0, 0.1, 0.2 and 0.25) solid solutions were

synthesized by sintering LiOH.H2O, NiO, and Al powders at 750 ºC under an oxygen

stream for 24 – 40 h [56]. Single phase LiAlxNi1-xO2 solid solutions were obtained.

The Al doped LiAlxNi1-xO2 cathodes demonstrated almost the same charge and

discharge behavior as the LiNiO2 cathode. Both the initial charge and discharge

capacities of LiAlxNi1-xO2 cathodes decreased with increasing Al3+ of dopant cations.

However, the cyclability of the cathode was improved for doped LiAlxNi1-xO2

cathodes in the voltage window of 3.0 – 4.3 V. This could be because the effects of

- 22 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Al3+ dopant prevent Li+ extraction in the voltage range of 4.1 – 4.3 V, thus avoiding

the two-phase region for 0.75 < x < 1 in LiAlxNi1-xO2.

LiAlxNi1-xO2 ↔ □1-xLixAlxNi1-xO2 + (1-x) Li+ + (1-x) e- (2.5)

The fully charged state of LiAlxNi1-xO2 is □1-xLixAlxNi1-xO2, in which the valence

electrons of Al3+ and Ni4+ are 3S0, so that this material is an electronic insulator. Li+

ions cannot be further electrochemically extracted from the structure because no

electrons can be removed from either Al3+ or Ni4+. Residual Li+ and Al3+ are useful to

suppress the dimensional change of the interlayer distance due to the extraction of Li+

ions. This effect contributes to maintaining the structural integrity of the cathode and

enhancing the cyclability.

2.5.3 Layered LiMnO2, orthorhombic LiMnO2 and their doped compounds

The LiMnO2 system contains some of the most intensively investigated cathodic

materials for rechargeable lithium batteries, due to its low-cost, non-toxicity, and high

energy density. Theoretically, if all the lithium can be reversibly extracted from and

inserted into the LiMnO2 structure, it can deliver a specific capacity of about 270

mAh/g. LiMnO2 compound has two types of crystal structure e.g. a layered structure

(monoclinic C2/m, or m-LiMnO2) and an orthorhombic structure (Pmnm or

o-LiMnO2). These two structures have to be synthesized under different synthetic

conditions. The layered LiMnO2 (m-LiMnO2) is not thermodynamically stable at

elevated temperature. Therefore, it can not be prepared by the normal high

- 23 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

temperature sintering process. Layered LiMnO2 was successfully prepared in 1996 by

an ion exchange technique from NaMnO2 [57, 58]. Firstly, layered NaMnO2 was

synthesized by solid-state reaction between stoichiometric Na2CO3 and Mn2O3 at 700

– 730 ºC under flowing argon. Then, LiMnO2 was obtained by refluxing NaMnO2

with excess LiCl in n-hexanol at 145 – 150 ºC, during which Li+ ions replace Na+ ions.

Neutron diffraction confirmed the monoclinic layered structure with unit cell

dimensions: a = 5.4387(7) Å, b = 2.80857(4) Å, c = 5.3878(6) Å, and β = 116.006(3)

º. The orthorhombic LiMnO2 has an ordered rock-salt structure in which LiO6 and

MnO6 octahedra are arranged in corrugated layers. It can be easily prepared via

solid-state reaction at high temperature with lattice parameters: a = 2.805 Å, b = 5.746

Å, and c = 4.574 Å [59].

Both m-LiMnO2 and o-LiMnO2 are electrochemically active as cathodes in

lithium-ion cells. It was found that up to 0.95 lithium per formula of LiMnO2 can be

extracted from m-LiMnO2 cathode on the initial charging [31]. When lithium ions are

removed from LiMnO2, the Jahn-Teller Mn3+ ions are oxidized to Mn4+, inducing the

loss of the monoclinic distortion. X-ray diffraction has identified that the layered

LiMnO2 is converted to a spinel structure during the charge and discharge cycling

[60]. The mechanism for this structural conversion on cycling could be related to the

fact that the random distribution of lithium and manganese in a cubic-close-packed

array of oxygen ions is much closer to the spinel structure, and also, the spinel

structure represents a uniform, isotropic distribution of manganese ions in a

close-packed oxygen lattice and thus is thermodynamically stable. On the other hand,

o-LiMnO2 has much better electrochemical performance than that of layered

- 24 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

m-LiMnO2 cathode. The low-temperature (~100 ºC) synthesized o-LiMnO2 cathode

showed an initial reversible capacity of 170 mAh/g and exhibited little capacity loss

with cycle number when cycled either the 3 V or the 4 V plateau, but rapid capacity

loss when cycled over both voltage plateaus [61]. It was also found that the o-LiMnO2

was gradually converted to the single phase spinel structure on cycling [62]. The

layered LiMnO2 structure can be efficiently stabilized via replacement of Mn by other

metal elements such as Al, Co, Ni, Cr, etc.. The doped m-LiMxMn1-xO2 compounds

can be prepared via solid-state reaction by sintering at high temperature, which is an

important advantage over the ion-exchange approach. Ab initio calculations showed

that LiAlO2 has a theoretical intercalation voltage plateau of 5.4 V vs. Li/Li+ reference

electrode, therefore Al doping can potentially increase the intercalation voltage

plateau and cathode energy density. Y. Jang et al. first reported the stabilization and

electrochemical performance of Al doped LiMnO2 cathode material [63]. Compounds

with two compositions, LiAl0.05Mn0.95O2 and LiAl0.25Mn0.75O2, were synthesized. It

was found that m-LiMnO2 becomes the major phase when oxygen partial pressures

are below 10-5atm. Energy dispersive X-ray mapping using scanning and transmission

electron microscopy (SEM and TEM) confirmed the homogeneous distribution of Al

and Mn through the LiAlxMn1-xO2 particles. The LiAl0.25Mn0.75O2 cathode exhibited

148mAh/g reversible capacity at C/5 rate and 182 mAh/g at the C/15 rate with

excellent cyclability when cycled in the voltage range of 2.0 V – 4.5 V.

The stability of Co doped LiMnO2 was predicted by ab initio calculations using the

scalar-relativistic, spin-polarized Green’s function technique based on the full charge

exact muffin-tin orbital (EMTO) method [64]. The calculations were performed for

- 25 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

three (pmmn, C2/m, and R3m) crystallographic allotropes of LiCoxMn1-xO2 in

ferromagnetic (FM), antiferromagnetic (AF), and disordered local moment (DLM)

states. The substitution of 10 at. % of Mn by Co makes the AF monoclinic structure

less stable than the orthorhombic structure in the DLM state. When there is 30 at. %

Co in LiCoxMn1-xO2, the FM orthorhombic structure has the lowest energy. Finally,

with ~ 60 at.% manganese replaced by Co, the desired layered R3m LiMnO2 structure

is stabilized. It is a thermodynamically stable equilibrium structure that can be easily

prepared by solid-state reaction. However, such a prediction does not represent the

successful preparation of Co doped LiMnO2 with the layered structure via soft

chemical approaches. P.G. Bruce et al. [65, 66] have successfully synthesized

LiCoxMn1-xO2 (0 ≤ x ≤ 0.5) using a solution based route coupled with ion exchange.

Five compositions of LiCoxMn1-xO2 ( x = 0.1, 0.2, 0.3, 0.4, and 0.5) were examined by

powder neutron diffraction. They were all identified as layered structure with the

space group R3m. The lattice parameter a decreases with increasing of Co content,

while c increases with increasing Co content. The Co doped LiMnO2 cathodes could

be charged and discharged over a wide voltage range of 2.6 – 4.8 V. The

LiCo0.1Mn0.9O2 cathode yielded an initial discharge capacity of 210mAh/g. As a

general trend, with increasing Co dopant level, LiCoxMn1-xO2 cathodes exhibit lower

initial capacity, but better cyclability. Through analyzing the incremental capacity

plots, they found that the transformation to spinel is relatively facile for LiCoxMn1-xO2

with low Co content (< 30 at. %), but much slower above this composition. This could

be responsible for the increased stability as the Co doping content increases. They

also concluded that Co doping can significantly modify the structure chemistry and

electrochemistry of layered LiMnO2, that 10 at. % Co doping can efficiently prevent

- 26 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

the capacity loss on the first cycle, and that higher Co level doping can suppress the

transition to a spinel-like structure, resulting in long cycle life.

LiNixMn1-xO2 cathode materials have also been extensively investigated. The

synthesis and electrochemical performance of LiNixMn1-xO2 solid solutions were

reported by Dahn et al. in 1992 [67]. The electrochemical properties deteriorated with

increasing Mn content. The optimum electrochemical behavior was found for the

composition of LiNi0.5Mn0.5O2 by different research groups [68 – 70]. X-ray

photoelectron spectroscopy (XPS) and magnetic property measurements show the

existence of Ni2+ and Mn4+ rather than Ni3+ and Mn3+. Noguchi et al. [71]

systematically studied the phase diagram and electrochemical properties of

Li-Ni-Mn-O compounds. The LiNixMn1-xO2 solid solutions between LiNiO2 and

Li2MnO3 can be prepared by the usual methods. They are rhombohedral (R3m) when

x ≤ 0.3, and have monoclinic symmetry (C2/m) when x ≥ 0.5. Ex situ XRD and

chronopotentiograms showed that complex phase transitions can be completely

suppressed with x ≥ 0.5. The electrochemical properties of the two compositions,

Li1.04Mn0.1Ni0.86O2 and Li1.13Mn0.26Ni0.61O2, were tested at a current density of 0.55

mA/cm2 and exhibited average discharge capacities of 165 mAh/g and 150 mAh/g

respectively.

The layered LiMnO2 structure can also be effectively stabilized at room temperature

by Cr doping. The groups of both Davidson and Ammundsen have described the

successful synthesis and excellent electrochemical properties of Cr doped LiMnO2

cathode materials [72 - 76]. The Li1.2Cr0.4Mn0.4O2 cathode compound was prepared

- 27 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

by reacting stoichiometric amounts of co-precipitated mixed hydroxides of chromium

and manganese with lithium carbonate at 800 ºC in argon atmosphere. X-ray

diffraction confirmed the layered hexagonal structure in the R3m space group, with

unit cell parameters of a = 2.886 Å and c = 14.372 Å. k edge X-ray absorption near

edge specter (XANES) showed that Cr is trivalent and in octahedral sites, and Mn is

tetravalent and in regular octahedral coordination. In Li1.2Cr0.4Mn0.4O2, the k edge

extended X-ray absorption fine structure (EXAFS) spectra supporteded a structural

model for Li1.2Cr0.4Mn0.4O2, in which the transition metal layers consist of clusters of

[Li1/3Mn2/3] and clusters of transition metal, predominantly Cr. The Li1.2Cr0.4Mn0.4O2

cathodes exhibited very stable cyclability, both at room temperature (20ºC) and at an

elevated temperature of 55ºC, with average capacities of 135 mAh/g and 170 mAh/g,

respectively. The LiCrxMn1-xO2 (0 ≤ x ≤ 0.5) compands and their delithiated /

relithiated derivatives were investigated via X-ray absorption spectroscopy (XAS)

and micro-Raman spectroscopy analysis [77]. It was identified that the trivalent Cr

ions are fixed to the octahedral site in the (Mn, Cr) O2 layer before and after

electrochemical cycling. The Cr doping can reduce the local structural variation of

manganese in LiCrxMn1-xO2 upon lithium extraction/insertion.

2.5.4 Spinel LiMn2O4 and LiMxMn2-yO4 cathode materials

The spinel-type oxide LiMn2O4 and its derivatives have been studied for more than a

decade as cathode materials for lithium ion batteries. The lithium insertion and

extraction in LiMn2O4 spinel provides an average voltage of 4 V versus Li/Li+, which

is similar to LiCoO2 and LiNiO2. Significant efforts have been made to extend the

- 28 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

capacity of the 4 V plateau to the 3 V plateau without compromising the specific

capacity. Spinel LiMn2O4 has the advantages over the Co- and Ni- containing

systems of being inexpensive, environmentally benign, and having a larger thermal

stability domain especially when overcharged. However, the theoretical capacity of

LiMn2O4 is only 148 mAh/g. The high rate of capacity fading on cycling is another

serious drawback for LiMn2O4 spinel electrodes. Fortunately, the problems related to

the capacity fading can be circumvented by doping Mn with other transition metals M

to suppress the Jahn-Teller (Mn3+) effect. Therefore, LiMn2O4 spinel is still an

attractive cathode candidate for lithium ion batteries. Some doped LiMxMn2-xO4

spinels have already been used in commercial production.

Spinel LiMn2O4 belongs to the cubic system with a space group of Fd3m, in which Li+,

Mn3+/4+ and O2- ions occupy tetragonal 8a and octahedral 16d and 32e position in the

cubic-closed-packed (ccp) oxygen array. This structure provides a three dimensional

network of face-sharing tetrahedra and octahedra for lithium ion diffusion. In the [M2]

O4 spinel framework, 75% of the metal cations occupy alternate layers between the

ccp oxygen planes; the remaining 25% of the metal cations are located in the adjacent

layers. Therefore, sufficient metal M cations exist in every layer to provide a

sufficiently high binding energy to maintain an ideal ccp oxygen array when lithium

ions are extracted from the structure. The unit cell of the spinel structure expands and

contracts isotropically [78]. The electrodes with such a structure are likely to be more

stable under electrochemical cycling than those in which there is a large anisotropic

expansion or contraction of the lattice parameters. For example, the spinels

Li[M1.67Li0.33]O4 (M = Mn, Ti) are extremely tolerant to cycling, because the volume

- 29 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

variation of the cubic unit cell is less than 1% within controlled compositional limits

[79, 80].

The LiMn2O4 spinel electrode offers a 4 V voltage versus lithium in the compositional

range 0 ≤ x ≤1 in LixMn2O4. Lithium ions can further insert into LiMn2O4 to form

Li1+xMn2O4 at 3 V against lithium. For Li1+xMn2O4, the electrode consists of two

phases: cubic phase LiMn2O4, and tetragonal phase Li2Mn2O4. The tetragonal phase

grows at the expense of the cubic phase. In the LiMn2O4 spinel structure, the strong

Mn-O bond maintains the structural integrity of the electrode during lithium ion

insertion and extraction. However, the Jahn-Teller distortion effect starts at Mnn+ =

3.5 (c/a = 1) and reaches a maximum at the valence of Mnn+ = 3 (c/a = 1.16), which

reduces the symmetry of the cubic spinel phase to tetragonal phase Li2Mn2O4 and

increases c/a to 1.16 concomitantly. Such distortion accompanying the volume

change of the unit cell would cause the crystal structure of the electrode to be

degraded and the electrochemical properties to deteriorate.

In practice, when the discharge voltage is close to 3.0 V, it is plausible that a few

spinel particles may result in an overall composition, Li1+xMn2O4, in which the

average Mn oxidation state falls below 3.5. A Jahn-Teller distortion might occur at the

particle surface, inducing the growth of tetragonal phase. The incompatibility

between the oxygen arrays of the cubic and tetragonal phases could then cause a

highly crystalline electrode particle to be degraded at the crystal surface, which would

damage the structural integrity and particle-to-particle contacts. Therefore, the

electronic conductivity and lithium diffusivity will deteriorate. Fortunately, this

- 30 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

damaging effect can be reduced by suppressing the onset of the Jahn-Teller distortion

by modifying the composition of the spinel electrode to keep the average Mn

oxidation state slightly above 3.5 at the end of the 4.0 V discharge plateau.

To combat the Jahn-Teller distortion effect, doping the LiMn2O4 spinel structure with

low valence ions is an effective approach. Since the average valence of Mn in

LiMn2O4 is + 3.5, when partially replacing Mn with other low valence ions, the

average valence of Mn will be increased above + 3.5. Correspondingly, at the end of

the 4 V discharge (lithium insertion into the spinel structure), the average valence of

Mn will be kept slightly above +3.5. Because the Jahn-Teller distortion starts at Mnn+

= 3.5, low valence ion doping can suppress the Jahn-Teller effect in LiMn2O4 spinel.

However, the theoretical capacity of LiMn2O4 spinel decreases with increasing

amounts of dopant ions. It is believed that the redox reaction accompanying lithium

insertion and extraction relies on the valence change of Mn in LiMn2O4 spinel [80].

Doping Mn with other ions inevitably reduces the amount of lithium which can be

reversibly extracted from the spinel structure. So, a trade-off has to be made to achieve

the optimum balance between suppressing the Jahn-Teller effect and keeping an

appropriate capacity.

A series of alien cation doped spinel compounds with the general formula of

LiMδMn2-2δO4 (M = Co, Zn and Mg; δ = 0.02, 0.03, and 0.05) were synthesized [79].

In this type of doped LiMδMn2-2δO4 spinel, one fraction of M ion substitutes for 2

fractions of Mn ions. Therefore, the average valence of Mn can be maximally

increased to suppress the Jahn-Teller distortion effect. The resulting compound

- 31 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

LiMδMn2-2δO4 is cation deficient. The triple or double valence of the M cations

partially substitute for Mnn+ ions at 16d sites. Thus, there are unoccupied 16d sites in

LiMδMn2-2δO4 spinels. It is expected that more Li ions can be accommodated by such

partial substitution. Table 1 summarizes the structural and electrochemical parameters

of LiMδMn2-2δO4 (M = Co, Zn, and Mg) cathode materials.

Table 2.1 The structural and electrochemical parameters of LiMδMn2-2δO4 cathodes

[79].

Compounds Lattice Theoretic Fully Oxidised Theoretic Initial The av. Rate
Constants Mn oxidation Composition al discharge of capacity
(A) state in ( n+Mn=4) Capacity capacity fade
LiMδMn2-2δO4 (mAh/g) (mAh/g) (mAh/g per
cycle)

LiMn2O4 8.2495 3.5 Mn2O4 154 125 0.5

LiCo0.02Mn1.96O4 8.2458 3.54 Li0.1Co0.02Mn1.96O4 139 120 0.15

LiCo0.03Mn1.94O4 8.2422 3.56 Li0.15Co0.03Mn1.94O4 131 116 0.06

LiCo0.05Mn1.9O4 8.2418 3.61 Li0.25Co0.05Mn1.9O4 116 109 0.08

LiZn0.02Mn1.96O4 8.2460 3.55 Li0.12Zn0.02Mn1.96O4 134 118 0.32

LiZn0.03Mn1.94O4 8.2432 3.58 Li0.18Zn0.03Mn1.94O4 126 111 0.24

LiZn0.05Mn1.9O4 8.2426 3.63 Li0.3Zn0.05Mn1.9O4 108 101 0.24

LiMg0.02Mn1.96O4 8.2410 3.55 Li0.12Mg0.02Mn1.96O4 136 114 0.28

LiMg0.03Mn1.94O4 8.2400 3.58 Li0.18Mg0.03Mn1.94O4 127 108 0.25

LiMg0.05Mn1.9O4 8.2362 3.63 Li0.3Mg0.05Mn1.9O4 109 98 0.22

As is shown, not all of the lithium can be electrochemically extracted from the doped

LiMxMn2-2xO4 spinel structure, because complete removal of lithium would result in an

- 32 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

average oxidation state of Mnn+> 4, which is chemically impossible. Residual lithium

could also help to suppress the chemical decomposition of the electrode at a high level

of the charge. The reasons for the degradation of spinel electrodes on cycling have

been identified as: (i) the structural damage to the electrode due to the volume change

in the unit cell caused by Jahn-Teller distortion in the 3V discharge regime; (ii) the

disproportionation reaction of Mn3+ from the cathode with the electrolyte: Mn3+ →

Mn2+ + Mn4+; and (iii) instability of the organic electrolyte at high voltage. The dopant

effect can combat (i) to suppress Jahn-Teller distortion and preserve the structural

integrity of the spinel electrode during cycling. Since we use LiPF6 electrolyte in

EC+DMC, which can resist oxidation even at 5 V, cause (iii) could be excluded. Co

doped spinel electrodes have demonstrated excellent cyclability. It is possible that the

Co dopant effect prevents the disproportionation reaction of Mn3+ with the electrolyte

[81]. Wakihara et al. proposed that the improvement in cycling behavior of the Co

doped LiMn2O4 compound could be attributed to enhancement in the stability of the

spinel skeleton structure. This is because the binding energy of CoO is larger than that

of MnO [82].

The Co and Cr doped LiMn2O4 spinels synthesized by solid-state reaction have shown

the greatest improvement in cycle life without dramatically sacrificing the initial

capacity [83, 84]. However, the solid-state reaction produces heterogeneous spinels

with uneven distribution of the dopant cations. LiCrxMn2-xO4 spinels (x = 0, 0.02,

0.04, 0.06 and 0.1) have been synthesised by the Pechini method, in which the cations

can be mixed on an atomic scale [85 – 87]. With more homogeneous distribution of

the dopant Cr ions in the spinel structure, the electrochemical properties of LiMn2O4

- 33 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

spinel possibly could be optimised. Neutron diffraction was performed on

LiCrxMn2-xO4 spinels to determine their structural characteristics. Using XRD

analysis alone, it is not possible to provide sufficient evidence to judge definitely

whether there is a mixture of Li+ and Mn3+/4+ at the 8a and 16d sites and whether

dopant Cr ions are present instead of Mn ions at 16d sites. Consequently, Lu et al,

performed neutron diffraction on three samples: pure LiMn2O4, LiCr0.02Mn1.98O4, and

LiCr0.04Mn1.98O4. Because of the distinctive properties of thermal neutrons, light

elements such as Li can be easily located [88].

Fig. 2.6 shows the neutron diffraction patterns obtained for pure LiMn2O4 and

LiCr0.04Mn1.96O4 spinels. The Rietveld method and LHPM program were employed to

refine the neutron diffraction patterns [89]. The determination of the crystal structure

of a compound depends very much on calculating the diffraction pattern, assuming a

model of the crystal structure, and comparing the calculated diffraction pattern with

the observed one [90]. For undoped LiMn2O4, there is a substantial difference

between the experimental data and the calculated pattern. When we assumed that

there was some mixing of Li+ and Mn3+/4+ at 8a and 16d sites, when it was assumed

that all of the Li ions were in 8a sites and the Mn ions were in 16d sites, the best fit

was obtained. During the refinement of the LiCrxMn2-xO4 structure (x = 0.02 and

0.04), when we assumed that there were some Cr ions in 8a Li sites, the differences

between the experimental data and the calculated pattern were obvious. Only when it

was assumed that all Cr ions were in 16d Mn sites, did the calculated intensities best

fit the observed ones, and the differences were almost negligible. It was clearly

- 34 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

elucidated that there was no mixing of Li and Mn ions in the LiMn2O4 spinel structure

and that the dopant Cr ions occupy 16d Mn sites in doped LiCrxMn2-xO4 spinels.

Cr doped LiCrxMn2-xO4 spinels demonstrated substantially improved capacity

retention on cycling compared to the undoped LiMn2O4 spinel. The Cr doping effect

could also modify the surface states of the spinel electrode in the highly charged state,

reducing the self-discharge rate.

1.0 (a) LiMn2O4 Observed data


Calculated pattern
0.8
Intensity (Normalised)

0.6

0.4

0.2

0.0

-0.2
20 30 40 50 60 70 80 90 100 110 120 130
2 Theta (degree)

1.0 (b ) LiC r 0.04 M n 1.96 O 4 O bserved data


C alculated pattern
0.8
Intersity (Normalised)

0.6

0.4

0.2

0.0

-0.2
20 30 40 50 60 70 80 90 100 110 120 130

2 T heta (degree)

Fig. 2.6 Neutron diffraction patterns of (a) LiMn2O4 (b) LiCr0.04Mn1.96O4.

- 35 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

2.5.5 Lithium iron phosphate based cathode materials

Iron based oxides have long been pursued as cathode materials for lithium-ion

batteries, because iron (Fe) is abundant, inexpensive, and less toxic than Co, Ni, or

Mn. LiFePO2 was initially prepared by ion-exchange from α-NaFePO2, but showed

little capability for reversible lithium intercalation and de-intercalation [91]. Later,

attention has been shifted to compounds with a NASICON (sodium super ionic

conductor) framework, allowing high mobility for lithium ion diffusion. This class of

compounds are built on corner-sharing MO6 octahedra (where M is Fe, Ti, V, or Nb)

and XO4n- tetrahedral anions (where X is S, P, As, Mo or W). The choice of the

transition metal M has a significant effect on the cell voltage, as it is primarily

dependent on the redox couple of the metal atom present in the structure. The

transition-metal redox couples are actually stabilized, and the potentials are made

more positive by the presence of the XO4 tetrahedra in the structure, due to strong

polarization of the oxygen atoms towards the X-cation and lowering of the covalent

component in the M-O bond by the inductive effect.

In 1997, J.B. Goodenough et al. discovered the electrochemically active lithium iron

phosphate olivine compounds [92 -93]. This is a new cathode material with

potentially low cost natural abundance of raw materials and environmentally

friendliness. It is expected that lithium iron phosphate will have a significant impact

in electrochemical energy storage. It has the following characteristics:

• A high theoretical capacity (170 mAh/g);

- 36 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

• A high and relatively stable open-circuit voltage of 3.4V relative to lithium

metal

• Easy to synthesize and inexpensive to fabricate;

• Little hygroscopicity, which makes it easy to handle;

• Excellent stability during cycling when used with common organic electrolyte

systems (Since LiFePO4 does not react with the electrolyte) and long cycle life

(up to 2000, cycles compared to 400 cycles with LiCoO2);

• Good overcharge tolerance compared to other cathode materials, so that on

overcharge, there is less heat released and thus less risk of fire or explosion;

• Good capacity retention during cycling;

• Non-toxic and safe in operation;

• Environmentally benign;

• Low cost.

However, lithium iron phosphate cathode material also has several disadvantages,

which impede its industrial applications. These include: (i) very low electronic

conductivity in its pure form, on the order of 10-9 S cm-1; (ii) slow lithium diffusion in

the solid phase; and (iii) Low rate capacity. Those drawbacks can be overcome by

various strategies, including carbon coating, conductive additives, alien cation doping,

and synthesis of nanosize crystallites.

- 37 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

2.5.5.1 Crystal structure of LiFePO4

LiFePO4 has an ordered orthorhombic olivine crystal structure. Olivines belong to a

general class of ‘polyanion’ compounds having compact tetrahedral anion structural

units (XO4) n- (X = P, S, As, Mo or W) with strong covalent bonding, combined with

MO6 octahedra (M = Mn, Fe, Co, Ni, etc.) to produce a 3D network [94, 95]. Amongst

the various polyanion compounds that are known, this olivine structure has the most

extensive interconnection of octahedra. For a general M2XO4 olivine, the M atoms

occupy half of the octahedral sites and X atoms one-eighth of the tetrahedral sites of a

hexagonal close-packed (hcp) oxygen array [96]. A diagram of an olivine structure is

given in Fig. 2.7 [2].

According to Fig. 2.7, the two octahedral sites (M1 and M2) are crystallographically

distinct and have different sizes, which favors ordering in MM’XO4 olivines

containing M and M’ ions of different sizes and charges. In the actual structure

(Figure 1b), the M2 octahedron has mirror symmetry with M1, the M-O bond

distances are greater than in M1, and it is slightly distorted compared to the ideal

model. The M1 sites form linear chains of edge-sharing octahedra running parallel to

the c-axis in alternate a-c planes, and the M2 sites form zigzag planes of corner-shared

octahedra running parallel to the c-axis in the other a-c planes [16]. Each M1 site

shares its edges with two M2 sites and two X sites, while there is one edge shared by a

M2 site with an X site. The distortion of the array is related to the repulsion of positive

units across the shared edges.

- 38 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Fig. 2. 7 The crystal structure of olivine LiFePO4 along [001]. Left: Expanded view of

the framework built on FeO6 octahedra and PO4 tetradedra. Right: Restricted view of

Li, Fe, and P distribution between two distorted h.c.p. (hexagonal close packed)

oxygen-dense layers (PTd[LiFe]octO4). LiO6 octahedra share edges and Li ions may

diffuse along [010] and [001].

On the removal of lithium, FePO4 phase is formed. LiFePO4 and FePO4 have

essentially the same crystal structure. FePO4 is isostructural with heterosite.

Therefore, it is a two-phase system with LiFePO4 in equilibrium with FePO4. The

conversion of the crystal structure between LiFePO4 and FePO4 is illustrated in Fig.

2.8 [97].

- 39 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Fig. 2.8 Crystal structures of orthorhombic LiFePO4 (left) and trigonal quartz-like

FePO4 (right).

The striking similarity in crystal structure similarity between LiFePO4 and FePO4 is

responsible for the superior reversibility of the cells on continuous cycling. The

connections among the different types of atoms are also shown. According to Fig. 2.8,

both phases have the same space group. On extraction of lithium from LiFePO4 [in

terms of lithium extraction/reinsertion, the material should be written as Li1-xFePO4 (0

< x < 1), where x is the number of lithium ions that are extracted from a unit cell], the

host matrix structure changes slightly, with a contraction in the a and b lengths but a

small increase in the c length. The volume is reduced by 6.81% and the density by

2.59%. Even though the overall changes to the FePO4 framework are only slightly

displacive, there is still a limit to the reversible first-order transition reaction between

the two phases [98]. This is because the transition would require a cooperative elastic

deformation of the framework. In other words, the principal change in the framework

- 40 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

on delithiation is an adjustment due to the coulombic repulsion between the O2- ion

sheets occupying the interface between delithiated planes.

Fig. 2.9 displays a 3D structural model of LiFePO4. The unit cell contains a total of 28

atoms [99, 100]. In the structure, the cations that occupy the M2 sites (Fe sites in

LiFePO4) form a corner-sharing network of octahedra in the (010) plane, and the

cations on M1 (Li) sites form octahedra edge-sharing chains in the [100] direction.

For LiFePO4, there are three different O sites, and the crystal can be viewed as

interconnected but distorted LiO6 and FeO6 octahedra, sharing oxygen vertices with

PO4 tetrahedra [3]. In particular, the O-O distance in the PO4 tetrahedron is small and

three of its edges are shared by the octahedra. Figure 4d shows that the FeO6

octahedra are linked through common corners in the bc plane and the LiO6 octahedra

form edge-sharing chains in the b direction. One FeO6 octahedron has common edges

with two LiO6 octahedra, while the PO4 groups share one edge with an FeO6

octahedron and two edges with LiO6 octahedra.

The separation between the octahedra and the tetrahedra hinders the ionic

conductivity, because the PO4 tetrahedra will bridge the two adjacent FeO6 octahedra

planes, which constrains the free volume in which the Li+ ions can move. Thus the

electrode supports only small current densities at room temperature compared to other

electrode materials. If the current density is increased, it will not reduce the open

circuit voltage Voc; rather it decreases, reversibly, the cell capacity (i.e. reducing the

current restores the capacity). Also, the low value of the electronic conductivity is

believed to be due to the large electron effective mass in the structure. Nevertheless,

- 41 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

because of having lithium in a continuous chain of edge-shared octahedra and

interstitial voids, reversible extraction/insertion of lithium is still possible. In fact,

ideally up to 0.9 Li+ ions/formula unit can be reversibly intercalated when electronic

conductors (i.e. carbon and dopant atoms) are introduced. The ions will diffuse

through the two-dimensional pathway provided by the interconnected octahedra and

tetrahedra [101-102].

- 42 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Fig. 2.9 Model displaying the connection among FeO6, LiO6, and PO4 structural

units (a); Structure viewed along the c-axis (b).

2.5.5.2 Enhancement of electronic conductivity and improvement of

electrochemical properties of LiFePO4 cathode material

Good electronic conductivity of LiFePO4 cathode materials can be effectively

achieved by carbon coating. Carbon can be added into the bulk of the LiFePO4

material via either in situ (formed during the preparation process of the active

material) or ex situ (mixed with carbon powders) methods. Graphite was found to be

the best carbon source, and it offers the highest electronic conductivity for LiFePO4.

When uniformly dispersed throughout the LiFePO4 material, it optimizes its

- 43 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

morphology and enhances the kinetics of the electrochemical process by improving

electron transfer.

Carbon is often deposited on the surface of each LiFePO4 particle and forms a thin

layer, and so it is commonly referred to as carbon coating. This coating is porous, so

that the lithium ions from the electrolyte solution can penetrate it to reach the surface

of the LiFePO4 particle with little resistance. Without carbon, the utilization of the

pure material, even at relatively low discharge rates, is poor, due to the voltage

reaching the cutoff potential much before any transport limiting factors such as

particle size become important; thus the electrode becomes ohmically limited. It has

been estimated that for solid-state synthesis techniques, approximately 20 wt% of the

cathode needs to be carbon in order for LiFePO4 to reach its theoretical capacity [12].

However, if using advanced methods to produce LiFePO4 such as chemical sol-gel

synthesis and hydrothermal synthesis, only 5 wt % carbon is sufficient to improve the

performance of the electrode [103-105].

Transition metal cation doping is another strategy to increase the electronic

conductivity of LiFePO4 cathode material. In comparison with LiFePO4, the doped

LiFePO4 materials were reported to reach conductivity values of > 10-2 S cm-1 at

room temperature, show near-theoretical energy density at low charge/discharge

rates, and retain significant capacity with little polarization at rates as high as 6000

mA g-1 (40C rate). Experiments have been carried out to study LiFePO4 samples

doped with various cations, including Mg2+, Al3+, Ti4+, and Zr4+.

- 44 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Electrical conductivity measurements have been performed on all the different

samples sintered at various temperatures and sectioned from pellets. They clearly

show that all doped compositions Li1-xMxFePO4 (M = Mg, Al, Ti, Nb or W) have

room temperature conductivities in excess of 10-3 S cm-1 [106]. Another experiment

found that a pressed Li0.99Zr0.01FePO4 pellet sintered at high temperature had an

electronic conductivity of 2 x 10-2 Scm-1 [106, 107].

The study also examined the relationship between the conductivity and the dopant

concentration. For example, in the Nb sample, it was observed that doping levels of

0.1, 0.5 and 1.0 atom% Nb resulted in room temperature conductivities of 1.1 x 10-3,

4.1 x 10-2, and 2.2 x 10-2 S cm-1 respectively, for samples fired at 700°C. At higher

doping levels of 2 and 4 atom%, Nb-enriched impurity phases appeared, and the

conductivity was reduced to 2.8 x 10-3 and ~10-6 S cm-1 respectively. Thus, it seemed

that there was a lack of proportionality between the dopant level and the conductivity

within the solid solution [108].

The electrochemical performance of LiFePO4 can also be improved by reduced

crystal size, such as in the nanometer domain. Smaller crystals mean a higher surface

area for lithium insertion/extraction reactions and higher specific capacity due to

more utilization of the active material. Nanosize LiFePO4 powders have been

synthesized by heating amorphous LiFePO4. The amorphous precursor was obtained

by lithiation of FePO4 synthesized by spontaneous precipitation from equimolar

aqueous solutions of Fe (NH4)2(SO4)2.6H2O and NH4H2PO4, using hydrogen peroxide

as the oxidizing agent. The as-prepared LiFePO4 powders have a relatively high BET

- 45 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

specific surface area of 16.5m2g-1. Nanocrystalline LiFePO4 showed very good

electrochemical performance, with a capacity near the full theoretical capacity (170

mAh/g) when cycled at the C/10 rate [109]. Nanocomposites of LiFePO4 and

conductive carbon were prepared, leading to enhanced electrochemical accessibility

of the Fe redox centers in this insulating material. The resultant LiFePO4/C composite

achieves 90% theoretical capacity at C/2 rate, with very good rate capability and

excellent stability [110]. It was also found that the nanophase phosphide network that

is created in situ within the grain boundaries of the insulating LiMPO4

nanocrystallites forms an efficient electrical conduit. Metal phosphide formation can

significantly enhance the electronic conductivity of LiFePO4 cathode materials [111].

2.6 Anode Materials for Lithium-ion Batteries

Lithium metal (lithium foil) was traditionally used for anodes in lithium primary

batteries (non-rechargeable). When lithium is used as anode in rechargeable lithium

batteries, lithium is often deposited in a dendrite form during repeated

charge/discharge cycles. These lithium dendrites are porous, have high surface area,

and are very reactive in organic electrolytes. Once lithium dendrites grow and

penetrate the separator after a certain number of cycles, it will cause short circuiting

and lead to fire or explosion. Therefore, this problem has driven research to find

alternative anode materials for rechargeable lithium batteries.

The basic requirements for anode materials used in rechargeable lithium batteries

include:

- 46 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

(i) The potential of lithium insertion and extraction in the anode versus

lithium the need to keep as low as possible.

(ii) The amount of lithium which can be accommodated by the anode material

should be as high as possible to achieve a high specific capacity.

(iii) The anode hosts should endure repeated lithium insertion and extraction

without any structural damage to obtain long cycle life.

2.6.1 Carbon based anode materials

Carbon materials are currently used in commercial lithium battery production. The

most widely used carbon anode materials are various types of graphitized carbon.

Graphite exists in various crystalline and amorphous forms. Carbon anode materials

can be produced from a variety of sources, such as natural graphite, oil, pitch, coal tar,

benzene, polymer resins, hydrocarbon gas, etc. In general, carbon anodes can be

categorized into:

(1) Graphite.

(2) Hard carbon - non-graphitized glass-like carbon which does not graphitized

even when heat treated at high temperature.

(3) Soft carbon – changing with heat treatment.

Schematics of the structures of these three types of carbons are shown in Fig. 2.10.

- 47 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Please see print copy for Figure 2.10

Fig. 2.10 Schematics of the structural of three types carbons [29].

Graphite consists of ABAB layers (graphene layers), which are held together by van

der Waals forces. Each graphene layer contains sp2 bond. Lithium ions can intercalate

into the interlayer between graphene layers to form LiC6 intercalation compounds.

The theoretical capacity of graphite is 372 mAh/g (850 mAh.cm-3), based on the

reaction to form the LiC6 compound.

Li+ + 6C + e- LiC6 (2.6)

The insertion compound LiC6 has a golden color and belongs to the graphite

intercalation compounds (GICs) of stage1, where the stage number corresponds to the

number of graphite layers which separate two successive intercalated planes.

The intercalation of lithium into graphite and the related electrochemical properties

have been well investigated [112-113]. Graphite synthesized at 3000 °C has a

capacity of 400 mAh/g in charging and delivers a capacity of 290 mAh/g on

- 48 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

discharging. This is because a thin lithium ion conductive film (solid electrolyte

interface, SEI) is formed on the surface of the graphite anode in the initial

charge/discharge cycle. Dahn et al. described three classes of lithium-ion battery

relevant carbons [114-115]. One is graphitic carbons, which are normally prepared

by heating so-called soft carbon precursors to temperatures above ~ 2400°C, where

well graphitized materials result. The second one is hydrogen containing carbons,

which are prepared from pyrolyzed organic precursors at temperatures near 700°C.

The third class of carbons are hard carbons or non-graphitized carbons. The structure

and chemistry of carbons strongly depends on the procedures of heating the organic

precursors. During the early stages of pyrolysis in inert gas (below 600 °C), organic

compounds decompose and emit gases that contain carbon, such as CO and CH4. The

remaining C atoms condense into planar aromatic structures (graphene sheets). If the

decomposing precursor forms a semifluid state, then these planar sheets can align in a

more parallel fashion that leads to easy graphitization upon heating to very high

temperatures. Such precursors yield “soft” or graphitizable carbons. However, if the

organic precursor is sufficiently cross-linked, then a fluid state is not realized during

decomposition and the planar aromatic structure cannot align. These carbon materials

are difficult to graphitize at high temperature and thus are called “hard” or

nongraphitizable carbons. All carbon materials formed by heating up to about 1300 –

1500 °C (so called soft carbon) have been shown to consist of small domains of

graphite-like hexagonal layers of carbon atoms with a diameter of 1-1.5 nm and

parallel stacking of two or three layers. These have been called basic structure units

(BSUs) [116,117].

- 49 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Highly orientated pyrolytic graphite (HOPG) is a typical example of synthetic

graphite having a high degree of plane orientation. The second classification scheme

relates to the axial orientation of layers, which consists of coaxial and radial cases. All

of the fibrous carbon materials have this scheme of orientation. The third scheme is

the orientation of the BSUs around a reference point (point orientation), in which two

cases have to be differentiated, radial and concentric. Most of the spherical carbon

particles have this texture.

The spherical graphite particles are obtained from pitches, so-called mesophase

spheres (or mesocarbon microbeads, MCMB), and also by pressure carbonization of a

mixture of polyethylene with a small amount of polyvinyl chloride, called carbon

spherules. These particles were found to have a radial arrangement of BSUs.

The structure of carbon materials strongly depends on the precursors and the heat

treatment conditions (temperature and pressure). The average interlayer spacing

between neighbouring hexagonal carbon layers, d002, the crystallite sizes along the c

and a axes, Lc and La, and the probability of graphitic AB stacking between adjacent

layers, P1, are commonly used as fundamental structural parameters. In general, the

degree of graphitization, P1, and the crystallite thickness, Lc(002), increase with the

heat treatment temperature.

The lithium insertion behavior in hard carbon electrodes such as coke is somewhat

different from that observed in graphite electrodes. In contrast with graphite, the

cycling profile of hard carbon electrode shows no evidence of staging plateaus but

- 50 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

rather continuous charge/discharge curves associated with lithium intercalation that

slope between 1.2 V and 0.2 V [118]. It is believed that the insertion of lithium in

hard carbon electrodes corresponds to the filling of micro-pores in the carbon by

clusters of lithium. This mechanism would be expected to produce weakly bound

lithium (relative to that in Li metal) and to show a very low plateau, consistent with

experimental results. Hard carbon materials consisted primarily of small single layers

of carbon arranged more or less like a house of cards. Lithium could be absorbed on

both surfaces of these single sheets, leading to more lithium per carbon than in

intercalated graphite, which would have the theoretical maximum capacity of Li2C6 or

740 mAh/g. In the structurally disordered coke electrodes, the intercalation process

does not promote formation of the staging phases, and due to its lack of crystallinity,

these electrodes are not as sensitive to the nature of the electrolyte as is the case for the

graphite electrodes. In fact, coke electrodes are successfully used in commercial

lithium-ion batteries, such as the Sony Camcorder battery.

Hydrogen-containing carbons can deliver high lithium intercalation capacities up to

800 – 900 mAh/g. The hydrogen could be playing a crucial role in the mechanism of

lithium insertion. The reversible capacity increases with an increasing H/C atomic

ratio in this material. It has been suggested that lithium atoms can bind in the vicinity

of H atoms in the hydrogen-containing carbons. The inserted lithium could transfer

part of its 2s electrons to a nearby hydrogen, resulting in a corresponding change to

the H-C bond, which would cause changes in the relative atomic positions of the C

and H atoms. Hydrogen-carbon electrodes do not maintain their large capacities over

long term cycling [119, 120].

- 51 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Carbon anode materials are still under intensive investigation to improve their

specific capacity and cycle life. Some super carbons prepared by pyrolsyzing special

precursors have demonstrated a capacity of more than 1000 mAhg-1 [121], but with

short cycle life. Nanocrystalline carbons prepared via high energy ballmilling had a

lithium-insertion capacity of up to 2500 mAh/g in the first charge. Unfortunately,

only half of this capacity was reversible during the first discharge. Such large

irreversible capacity in the first cycle is unacceptable for any practical application

[122, 123]. Nevertheless, it is expected that carbon materials will reach a capacity of

600 – 700 mAh/g with satisfactory cycle life and reversible capacity.

2.6.2 Alloy based anode materials

Many metallic elements (M) can react with lithium to form LixMy alloys. The

gravimetric capacities of some lithium-alloy forming metal elements are presented in

Fig. 2.11. Pure lithium has the largest specific capacity of all the metallic elements.

Silicon also has very high specific capacity when forming Li4.2Si alloys. However,

these metal anodes experienced a dramatic volume increase upon reacting with

lithium, causing cracking which eventually leads to the failure of the electrode.

In order to circumvent the problem of volume change upon lithium insertion, alloy

powders with small particle size have been synthesized by a chemical reaction

method. Powders of Sn, SnSb, and SnAg (typical grain size: 200-400 nm) were

prepared by J.Besenhard et al.,

- 52 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

5000

4500

4000

Specific capacity (mAh/g)


3500

3000

2500

2000

1500

1000

500

0
Li C(LiC6) Si As Al Sn Sb Pb
(Li21Si5) (Li As)
3
(LiAl) (Li21Sn5) (Li3Sb) (Li22Pb5)

Fig. 2.11 Gravimetric capacity of lithium-alloy forming elements

These fine-grained alloy elements have demonstrated excellent cycling performance

[124]. A model for lithium insertion in alloys with small particle size has been

proposed. Many pores and cavities exist in a loosely packed small particle size

metallic matrix. During lithium insertion, even 100% volume expansion of individual

particles will not crack the electrode as the absolute changes in dimensions are still

small. In other words, the pores and cavities have the effect of buffering the expansion

of individual particles. During the extraction of lithium from LixM particles, the

expanded particles do not contract very much, thus the overall dimensions of the

electrode remain almost constant in the subsequent cycles. Wolfenstine [125] made a

prediction of the critical grain size below which microcracking does not occur, based

on volume change as a result of Li insertion into brittle Li-alloys. The predicted

- 53 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

critical grain size was less than the unit cell size for single-phase materials, which

suggests that decreasing the particle/grain size cannot fundamentally solve the

mechanical instability problem associated with Li-alloys. Some potential solutions to

solve the mechanical instability problems were proposed: (i) incorporating within the

Li-alloys a ductile Li-ion conducting metal or polymer matrix, and (ii) embedding the

alloys within a matrix which places them under compressive stress [126]. The use of

intermetallic alloys is one choice to provide a ductile matrix for lithium alloying.

The intermetallic alloys MM’ have emerged as a new class of anode materials for

lithium-ion batteries. The operating voltage of these alloy anode materials is a few

hundred millivolts above lithiated graphite, which is generally considered to be safer

than carbon anodes. In the intermetallic compound MM’, M is an “inactive” element,

and M’ is an “active” element which can react with lithium to form LixM’ alloy. The

reaction is supposed to proceed as follow:

xLi+ + MM’ + e- LixM’ + M (2.7)

The inactive M matrix is generated simultaneously. So, the reactions between lithium

and MM’ intermetallic alloys are generally described as displacement reactions, in

which domains of lithium alloy LixM’ are created within the inactive M’ matrix.

It is well known that LixM’ alloy systems undergo several phase changes when the

active element M’ alloys with lithium to form a series of LixM’ alloys. For example,

Sn can alloy with lithium to form various LixSn alloys: Li5Sn2, Li7Sn2, Li13Sn5, and

eventually, if heavily lithiated, Li22Sn5. Severe volume expansion and contraction are

- 54 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

expected during charge/discharge cycling. These dramatic volume changes cause

cracking of alloy particles, damage the integrity of the electrode, and limit the cycle

life of the alloy electrode. This problem can be significantly improved by finely

dispersing the active lithium-alloying element in an inactive composite matrix. The

inactive matrix must be electrochemically conducting and mechanically ductile. It is

believed that the inactive matrix provides structural stability to the composite

electrode and combats the volume change of the active element. Various intermetallic

MM’ alloys show improved cycling performance compared to the pure

lithium-alloying metal. For example, some intermetallic alloys such as Cu6Sn5 [127],

FeSn2, and FeSn [128] have demonstrated much improved cycling behavior

compared to pure Sn. These alloys were prepared by either powder sintering or

ball-milling. When lithium ions insert themselves into these anode materials, a

displacement reaction occurs in which the intermetallic structure is broken down to

form a series of LixSn alloys within the inert Cu or Fe matrix. In situ x-ray diffraction

studies combined with Mossbauer spectroscopy have confirmed the formation of

metallic Cu and Fe in Cu6Sn5 and FeSn2 electrodes after cycling. This displacement

reaction is partially reversible. The capacity fade on cycling has been attributed to the

agglomeration of the inactive phase into larger grains.

Nanocrystalline alloys provide an approach to further improve the electrochemical

performance of intermetallic MM’ alloy anode materials. Since there exist countless

pores and cavities within nanocrystalline materials, these cavities provide some free

room for the volume expansion of the active element when reacting with lithium to

form LixM’ alloys. The generated inactive matrix is also nanocrystalline in nature,

- 55 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

which will be compatible with the original MM’ alloys if the starting materials are

nanocrystalline.

Theoretically, the intermetallic MM’ alloys can be any combination of active

element, M’ and inactive element M, depending on their ability to react with lithium

to form lithium-alloy. Therefore, there will be many intermetallic alloys which have

the potential to be used as anode materials for lithium-ion batteries. Furthermore, the

intermetallic alloys can be ternary, or multi-element. This means that there can be

two, three, or more different active elements and vice versa for inactive elements.

Thus the possibilities for intermetallic alloys are many. Intensive research work is

required to explore and classify appropriate intermetallic alloys for lithium-ion

batteries.

2.6.3 Metal oxide anode materials

2.6.3.1 Tin oxide anode materials

The tin-based amorphous composite oxide (TCO) anode materials were first

developed by Fujifilm Celltec [129, 130]. The TCO anode yields a specific capacity

for reversible lithium adsorption more than 50 percent higher than those of the carbon

families. It provides a gravimetric capacity of > 600 mAh/g for reversible Li

adsorption and release, which corresponds in terms of reversible capacity per unit

volume to more than 2200 mAh cm-3. The tin-based composite oxide active material

has a basic formula represented by SnMxOy, where M is a group of glass-forming

- 56 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

metallic elements whose total stoichiometric number is equal to or greater than that of

tin (x ≥ 1), and is typically comprised of a mixture of B (III), P (V), and Al (III). In the

oxide structure, Sn (III) forms the electrochemically active center for Li insertion and

potential development, and the other metal group provides an electrochemically

inactive network of – (M-O)- bonding that delocalizes the Sn (II) active center. To

confer high reversibility in Li storage and release, the Sn-O framework is thus

anisotropically expanded by incorporating glass-forming network elements – B, P,

and Al, in view of the enhancement of Li-ion mobility in the anisotropic glass

structure, which is favorable for ionic diffusion and release. The TCO

SnB0.5P0.4Al0.42O3.6 was prepared by mixing SnO, B2O3, Sn2P2O7, and Al2O3, sintering

at the high temperature of 1100 °C for more than 10 hours, and then quenching to

room temperature [130]. TCO anode can be coupled successfully with several

cathode materials, such as LiCoO2, LiNiO2, LiMn2O4, and LiMnO2. Such lithium-ion

batteries have been tested to have an energy density of more than 420 wh/l and good

cyclability. The major drawback for the TCO anode is the large irreversible capacity

in the first cycle (~ 40 percent of the first charge), which requires excess cathode

material to compensate.

Various tin oxides have been investigated as anode materials for lithium-ion batteries,

including SnO, SnO2, Li2SnO3, and SnSiO3 glass. All of these oxides can react

reversibly with lithium with a capacity from 1200 mAh/g to 1500 mAh/g in the first

charge, but with large irreversible capacity. In situ x-ray diffraction measurement

revealed that these tin oxides and tin oxide composites follow a similar mechanism for

lithium insertion. When lithium ions are inserted into these oxides, Li2O and Sn are

- 57 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

generated simultaneously. Then the lithium alloys with Sn to form a series of LixSn

alloys. The subsequent cycle is a process of alloying/de-alloying of lithium with tin.

This process is shown as follows.

6.4Li + SnO Li2O + Li4.4Sn Li2O + 4.4Li +Sn (2.8)

8.4Li + SnO2 2Li2O + Li4.4Sn 2Li2O + 4.4Li + Sn (2.9)

6.4Li + SiSnO3 Li2O + SiO2 + Li4.4Sn

Li2O + SnO2 + 4.4Li +Sn (2.10)

8.4Li + Li2SnO3 3Li2O + 4.4Li + Sn (2.11)

The formation of a Li2O and SiO2 matrix may act to retard the aggregation of tin

atoms into large coherent regions. When large tin regions form, the large volume

differences between coexisting bulk Li-Sn phases may introduce cracking and

crumbling of the electrode structure. Therefore, capacity will be lost.

The electrochemical performance of the SnO2 electrode can be improved by Mo

doping. It was found that the Mo doping influences the habit of growth of the SnO2

crystallites and facilitates growth along the [hk0] direction. The presence of Mo in the

Li-Sn alloy as it is formed may favor Sn atom dispersion, which contributes to the

improvement of the reversibility of the Li reaction [131,132].

- 58 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

Much work has been done on TCO and tin oxides. Although TCO glass and some

nanocrystalline tin oxides have demonstrated excellent rechargeability, the problem

of large irreversible capacity related to oxygen in the first cycle is still not solved,

which prevents their commercial application.

2.6.3.2 Transition metal oxide anode materials

It was a surprise finding that MO-type compounds (where M is Co, Ni, Fe, Cu or Mn)

with a rock-salt structure and containing metal elements (M) that do not alloy with Li,

exhibited capacities two to three times those of carbon with 100% capacity retention

for up to 100 cycles. Anodes made of nanoparticles of transition-metal oxides (MO,

where M is Co, Ni, Cu, or Fe) have demonstrated electrochemical capacities of 700

mAh/g, with 100% capacity retention for up to 100 cycles and high recharging rates.

Fig. 2.12 shows the charge/discharge profiles and cyclability of MO anodes in

lithium-ion cells.

(a)

- 59 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

(b)

Fig. 2.12 Electrochemical properties of MO anodes in lithium-ion cells. (a) The

voltage – composition profiles for MO/Li cells cycled between 0.01 V and 3 V of

CoO,NiO and FeO anodes at the C/5 rate. (b) Cyclability of Co3O4, CoO, FeO and

NiO anodes, with the data, also included demonstrating that the behaviour is not

specific to divalent oxides. The inset shows the rate capacity of a CoO electrode.

The reaction mechanism between Li and MO oxides differs from the classical Li

intercalation/de-intercalation or Li-alloying/de-alloying process. The whole process

involves the formation and decomposition of Li2O accompanying the reduction and

oxidation of metal particles (in the range of 1 – 5 nanometers). For example, the

reaction of CoO anode in a lithium-ion cell can be described as:

CoO + 2Li+ + 2e- ↔ Li2O + Co (2.12)

- 60 -
Chapter 2 Literature Review
-------------------------------------------------------------------------------------------------------------------------

2Li ↔ 2Li+ + 2e- (2.13)

CoO + 2Li ↔ Li2O + Co (2.14)

Bright-field imaging and selected area electron diffraction (SAED) analysis

confirmed the above reversible process. Furthermore, magnetic measurements and in

situ extended X-ray absorption fine structure (EXAFS) measurements also revealed

the reversible nature of the process between CoO and Li ions. Since then, many metal

oxides have been investigated as anode materials for lithium-ion batteries [134-135].

It is expected that the use of transition-metal nanoparticles to enhance surface

electrochemical reactivity will lead to further improvements in the electrochemical

properties of lithium-ion batteries.

- 61 -
Chapter 3 Experimental
--------------------------------------------------------------------------------------------------------------------------

Chapter 3 Experimental
3.1 Materials and chemicals

The materials, chemicals and reagents used in this study are listed in table 3-1.

Table 3.1 Materials and chemicals

Raw materials / Formula Purity Supplier


Chemicals

Lithium foil Li 99.999% Hosen Corporation,


Japan
LP30 electrolyte EC(Ethylene Merck
carbonate):DMC
(dimethyl carbonate)
=1:1 w/w
LP31 electrolyte EC:DMC=2:1 w/w Merck
CR2032 coin-cell Stainless steel Hosen Corporation,
hardwares Japan
Polyvinylidene fluoride (-CH2CF2-)n Aldrich
(PVDF)
Dimethyl phthalate C6H4-1,2-(CO2CH3)2 Aldrich
(DMP)
N-methyl-2-pyrrolidinone Aldrich
(NMP)
Carbon black powders Aldrich
Nanocrystalline Silicon Si (99.999%) Nanostructured &
Amorphous Materials
Inc., USA
Mesocarbon microbeads C 100% Hosen Corporation,
(MCMB-10 graphitizing) Japan
Resorcinol 98% Aldrich

Formaldehyde 37.6% in methanol Aldrich


Argon gas Ar 99.99% BoC Gas , Australia
Camphor Aldrich
Iron carbonyl Aldrich
Red phosphorus P 99% Aldrich
Iron chloride FeCl2.4H2O 99% Aldrich
Lithium hydroxide LiOH.H2O 99.9% Aldrich
Iron oxylate FeC2O4.2H2O 99 Aldrich
Manganese nitrate Mn(NO3)2.6H2O 99.9% Aldrich
Polyacrylic acid Aldrich
Citric acid Aldrich
Mixture gas (10% 10% H2 in Ar BOC Gas Limited,
hydrogen in argon) Australia
Ammonia phosphate NH4.H2PO4 97% Aldrich

- 62 -
Chapter 3 Experimental
--------------------------------------------------------------------------------------------------------------------------

Magnesium acetate Mg(C2H3O2)2 99% Aldrich

Lithium carbonate Li2CO3 99.99% Aldrich


PTFE Membrane Hosen Corporation,
(separator) Japan
Silicon powder (325 Si 99%
mesh)
Acetone Aldrich
Ethanol (absolute) 100% Aldrich
Multiwalled carbon C Carbon
nanotubes (MWCNTs) Nanotechnology Inc.,
USA

Tin chloride SnCl2 99.9% Aldrich

3.2 Materials synthesis techniques

3.2.1 High energy ball milling

Si-MCMB composites were prepared by high energy ball milling. The ball-milling

facility used was a planetary ball-milling machine (Labtechnics, Australia). Stainless

steel balls with a diameter of 10 mm were used. The weight ratio between the material

powders and the balls was 1: 8-10. Home-made stainless steel milling jars were

employed, equipped with ventilation valve to allow evacuation of the milling-jar and

purging with argon gas. During the ball-milling process, a speed of 120 - 140 rpm was

maintained.

3.2.2 Carbon aerogel synthesis

Carbon aerogel was synthesized to prepare Si-Carbon composites. Carbon gel was

formed by reacting resorcinol and formaldehyde in a pH range of 6.5 – 7.4, in which

the pH value was adjusted with an ammonia solution. Simultaneously, nanosize

silicon powders were added and dispersed by ultrasonic probe and magnetic stirring.

- 63 -
Chapter 3 Experimental
--------------------------------------------------------------------------------------------------------------------------

The gel formation was accomplished by conducting the reaction at 85 °C. Finally,

carbon gel was sintered at high temperature (650 °C) to be converted into amorphous

carbon.

3.2.3 Chemical vapor deposition (CVD)

Multiwalled carbon nanotubes were synthesized by the catalytic chemical vapor

deposition (CVD) approach. The CVD reaction chamber consists of three different

temperature zones controlled by 3 independent programmable temperature

controllers. Camphor was evaporated and carried by the flowing argon gas, and then

deposited on a silicon substrate downstream. The surface of the silicon wafer was

pre-deposited with iron carbonyl as catalyst. After deposition, carbon nanotubes were

harvested from the silicon substrates.

3.2.4 Hydrothermal synthesis

Nanocrystalline iron phosphide (FeP4) was synthesized by a hydrothermal process. A

Teflon lined autoclave (Parr Instrument Company, USA) was used with a maximum

temperature limitation of 300 °C. The ingredient reagents phosphorus, iron chloride

and de-ionized water were loaded in to an Autoclave. The autoclave was heated at 180

°C for 24 hours. After cooling down to room temperature, the reaction product was

taken out, washed thoroughly with distilled water, and dried in a vacuum oven.

- 64 -
Chapter 3 Experimental
--------------------------------------------------------------------------------------------------------------------------

3.2.5 Sol-gel synthesis

Lithium iron phosphate and Mn and Mg doped lithium iron phosphate cathode

materials were prepared by the sol-gel synthesis method. Polyacrylic acid and citric

acid were used as complexing agents for the formation of a complex gel. The sol

solution was heated at 85 °C under vigorous stirring until forming a viscous gel was

formed. The gel was sintered at 500 °C to decompose the organic reagents into

carbon. The final products were obtained by sintering the intermediate product at 750

°C under argon atmosphere.

3.2.6 Solid-state reaction synthesis

Lithium iron phosphate and alien cation doped lithium iron phosphate compounds

were also synthesized via solid-state reaction. Stoichiometric quantities of reaction

reagents were hand-mixed with ethanol using an agate mortar and pestle. Then the

mixture was dried in a vacuum oven. The dried mixtures were pre-sintered in alumina

boats at 500 °C. Then, the pre-sintered mixtures were mixed again via mortar and

pestle. The re-mixed reagents were sintered again at 750 °C for 12 hours. The final

products were ground and then used for characterization and electrochemical testing.

3.3 Materials characterization

3.3.1 Structure characterization

The phases and crystal structure of all material samples were identified and

characterized by X-ray diffraction using a Philips PW-1730 diffractometer (40 kV,

25mA) with a monochromatized Cu Kα radiation (λ = 1.5418 Å) at a scan rate of

- 65 -
Chapter 3 Experimental
--------------------------------------------------------------------------------------------------------------------------

1°/min and step size of 0.02 °. The crystallite sizes were calculated using the Scherrer

equation:

D = 0.9λ / β cosθ

Where λ is the x-ray wavelength, β is the observed full width at half maximum

(FWHM), and θ is the Bragg angle.

All X-ray diffraction patterns were analyzed using the Difftech Traces V6 software

program and checked with the International Centre for Diffraction Data (ICDD)

database. X-ray diffraction patterns were also refined by the Rietveld method using

the Rietical software program.

3.3.2 Material morphology observation

The morphologies of all the as-prepared powders were observed and analyzed using a

scanning electron microscope (SEM, JEOL JSM-6460A with a LAB-6 filament).

Surface element analysis and element mapping were performed using Energy

Dispersive X-ray Spectroscopy (EDXS) apparatus attached to the SEM. The SEM

facility is shown in Fig. 3.1. The morphologies and nanosize particles were analyzed

using transmission electron microscopy (TEM, JEOL 2011 TEM) and high resolution

TEM (HRTEM). The TEM apparatus is shown in Fig. 3.2. TEM samples were

prepared by dispersing powders in ethanol via ultrasonic vibration to form a

homogeneous suspension. Then, droplets of the suspension were dropped onto a

copper mesh coated with holey amorphous carbon films.

- 66 -
Chapter 3 Experimental
--------------------------------------------------------------------------------------------------------------------------

Fig. 3.1 JEOL JSM-6460A SEM facility.

Fig. 3.2 JEOL 2011 TEM facility.

- 67 -
Chapter 3 Experimental
--------------------------------------------------------------------------------------------------------------------------

3.3.3 Magnetic properties measurement

The magnetic properties of lithium iron phosphate and transition metal doped lithium

iron phosphate compounds were measured using a Quantum XL Magnetic Properties

Measurement System (MPMS). The measurements of the temperature dependence of

the magnetization of the samples under zero magnetic field cooling (ZFC) and applied

magnetic field cooling (FC) were carried out in a magnetic field of 5000 Oe and over

a temperature range of 5 K – 300 K. The Curie and Néel temperatures, as well as the

spin state of the magnetic cations, were calculated from the inverse magnetic

susceptibility (1/χ), based on the equation:

χ = C/ (T-Өp) + χo (3.1)

Where χo is the temperature independent susceptibility and Өp is the Pauli-Weiss

temperature.

3.3.4 Electrochemical testing

3.3.4.1 Electrode fabrication

The electrode materials were hand mixed with 10 wt% carbon black and 5-10 wt%

PVDF binder in NMP solvent by mortar and pestle to form viscous slurry. Electrodes

were prepared by coating slurries either of copper foil substrate (anode) or on

aluminum foil. After coating, the electrodes were dried in a vacuum oven for 12 h at

120 o C and pressed between two steel plates at 3.0 × 106 Pa/cm2. The area of each

electrode was 0.785 cm2.

- 68 -
Chapter 3 Experimental
--------------------------------------------------------------------------------------------------------------------------

3.3.4.2 Cell assembly

CR2032 coin cells were assembled for all electrochemical testing. A microporous

plastic film (Cellgard 2500, Cellgard Co., USA) was used as separator. The

electrolyte was 1M LiPF6 dissolved in a 50:50 volume percent (v/o) mixture of

ethylene carbonate (EC) and diethyl carbonate (DEC), and a 300 μm thick, 0.78 cm2

lithium foil for the negative electrode was also used. Coin cells were assembled in an

argon-filled glove box (Unilab, Mbraun, Germany) with both water and oxygen

concentrations less than 0.1 ppm, as shown in Fig. 3.3.

Fig. 3.3 Atmosphere controllable glove box for assembling CR2032 coin cells.

- 69 -
Chapter 3 Experimental
--------------------------------------------------------------------------------------------------------------------------

3.3.4.3 Electrochemical testing

(i) Charge/discharge of cells

Charge/discharge tests were carried out by using a battery test device

(charge/discharge unit, BT-5 Cell-Tester, Neware Technology Inc., China) with a

computer (PC) and software. Cell-Testers operate in the voltage range of -5 V - +5 V,

with a maximum current of 1 mA and accuracy of 1/10000. The system is capable of

switching between charge and discharge automatically, according to the cut off

potentials set.

(ii) Cyclic voltammetry

Fig. 3.4 CHI 660 B and 660C Electrochemical Work Stations.

- 70 -
Chapter 3 Experimental
--------------------------------------------------------------------------------------------------------------------------

Cyclic voltammetry (CV) is one of the most reliable electrochemical approaches to

elucidate the nature of the electrochemical process, and to provide insights into the

nature of processes beyond the electron-transfer reaction. It was used in the present

work to study the electrochemical properties of electrode materials. CV is performed

by scanning the voltage between two chosen cut-off voltages at a given sweep rate,

while simultaneously, the response current arising from the electron transference is

measured and recorded. In a cyclic voltammogram, the positive current represents the

lithium de-intercalation process, and the negative current the lithium intercalation

process. Cyclic voltammograms were obtained by measuring the I-V response at a

scan rate of 0.01-0.1 mV/s and a cut-off voltage of 0.0 V – 2.0 V for anode materials

and 2.5 V – 4.4 V for cathode materials. CHI 660B and CHI660C Electrochemical

Work Stations (CHI Instruments, USA) were employed for all CV measurement

(shown in Fig. 3.4).

- 71 -
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

Chapter 4

Preparation and characterization of

Si-MCMB and Si-carbon composite anode

materials

4.1 Introduction

Silicon has a high theoretical lithium storage capacity of 4000mAh/g when forming

Li21Si5 alloys. Dahn et al. [136] synthesized nanodispersed silicon in carbon using

chemical vapour deposition (CVD). Although the Si-C anodes demonstrated a

reversible capacity of 500mAh/g, it is difficult to control the morphology of Si and C

using the CVD approach. Lee et al. [137] dispersed crystalline silicon (325 mesh) in a

sol-gel graphite matrix by ball-milling and achieved a reversible capacity of 832.2

mAh/g in the first cycle. Nano-Si-carbon composites which demonstrated a high

reversible capacity of 1700 mAh/g [138] have been prepared by hand mixing nano-Si

and carbon black. The lithium storage properties of nanostructured Si and Si film have

also been investigated [139, 140]. Since mesocarbon microbeads (MCMB) anode

materials have the best cyclability among all the various types of carbon anode

materials, the combination of MCMB and Si could result in Si-MCMB composite

anode materials with high capacity and satisfactory rechargeability. Based on this

hypothesis, we prepared nanocrystalline Si-MCMB composite materials by

72
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

high-energy ballmilling. Their electrochemical properties as anodes in lithium-ion

cells were systematically evaluated.

On the other hand, the performance of Si anodes can be significantly improved by

preparing nanostructured electrode. This is because the mechanisms of fracture will

be changed when the crystals in the material have sizes on the tens of nanometers

scale. We synthesized nanostructured Si-C composites by dispersing nanocrystalline

Si in carbon aerogel. The electrochemical properties of nanostructured Si-C were

systematically evaluated. The improved electrochemical performance of Si-C could

originate from the unique mechanisms behind the lithiation and de-lithiation

processes in nanostructured Si-C composites.

4.2 Experimental

4.2.1 Preparation of Si-MCMB composite.

Nanocrystalline Si with an average particle size of 80 nm, was obtained from

Nanostructured & Amorphous Materials Inc.,USA (which was prepared by laser

driven silane gas reaction). MCMB (MCMB-10 graphitizing) was supplied by Osaka

Gas Co. (Japan). MCMB powders have an average particle size of 10 μm.

Nanocrystalline Si-MCMB composites were prepared by high-energy ball milling

using a planetary ball-milling machine (Labtechnics, Australia). The nanocrystalline

Si content in the composite is 20 % by weight. We chose this percentage based on the

results of a previous investigation on ball-milled Sn-graphite [141] and ball-milled

73
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

Si-graphite [142]. The mixtures were ball-milled for 5, 10 and 20 hours respectively

to obtain 3 batches of Si-MCMB composites. The as-prepared Si-MCMB composites

were characterised by SEM (Leica / Cambridge Stereoscan 440 scanning electron

microscope) and x-ray diffraction with Cu Kα radiation ( MO3xHF22, MacScience,

Japan).

4.2.2 Preparation of Si-carbon composite

Nanocrystalline Si with an average particle size of 80 nm (which were prepared by

laser driven silane gas reaction). was purchased from Nanostructured & Amorphous

Materials Inc., USA. Resorcinol (98%, Sigma-Aldrich) and formaldehyde (37.6% in

methanol, Sigma-Aldrich) were used for preparing carbon aerogel. A typical carbon

gel was formed by mixing 0.29 M resorcinol and 0.57 M formaldehyde. The pH value

was adjusted to be in the range of 6.5 – 7.4 by adding NH3.H2O solution. The mixture

solution was put in an ampoule, sealed, and heated on a hot plate. The temperature

was maintained at 85 °C. The solution changed progressively from clear to milk white

to yellow to orange as the reaction progressed. When the solution became viscous,

nanocrystalline Si powders were added, and dispersed through magnetic stirring. The

ampoule was kept at 85 °C for 10 hours, and then carbon aerogel was formed, with Si

dispersed inside the gel. The obtained gel was then sintered at 650 °C under flowing

argon to yield an Si-C composite, containing 40% carbon by weight. The carbon

content was measured by thermogravimetric analysis (TG).

74
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

4.2.3 Electrochemical testing.

The electrochemical properties of Si-MCMB composites and Si-carbon composites

were measured via coin cell testing. The Si-MCMB electrodes were made by

dispersing 84 wt % active materials, 8 wt % carbon black and 8 wt % polyvinylidene

fluoride (PVDF) binder in dimethyl phthalate solvent to form homogeneous slurry.

The slurry was spread onto a copper foil. The coated electrodes were dried in a

vacuum oven at 120 °C for 12 hours and then pressed to enhance the contact between

the active materials and the conductive carbons. The CR2032 coin cells were

assembled in an argon filled glove-box (Mbraun, Unilab, Germany) using lithium

metal foil as the counter electrode. The electrolyte was 1 M LiPF6 in a mixture of

ethylene carbonate (EC) and dimethyl carbonate (DMC) (1:1 by volume, provided by

MERCK KgaA, Germany). The cells were galvanostatically discharged and charged

in the voltage range of 0.01 – 3 V at a current density of 0.05 mA/cm2. Cyclic

voltammetry (CV) was performed on a potentiostat (Model M362, EG & G Princeton

Applied Research, USA), at a scanning rate of 0.1 mV/s.

4.3 Results and discussion

4.3.1 Si-MCMB composite anode materials

MCMB particles have a spherical shape. Fig. 4.1 (a) shows an SEM image of MCMB

particles, which each consist of many small graphite particles. The images of the

Si-MCMB composites ball milled for different times are shown in Fig. 4.1 (b), (c) and

75
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

(d). After ball milling for 5 and 10 hours, the spherical agglomerates were partially

broken. But most of the particles still retained a spherical shape. However, after 20

hours of extensive ball milling, most of the spherical MCMB agglomerates were

destroyed and formed finely ground graphite debris. Energy dispersive spectroscopy

(EDS) was performed on the bulk of the Si-MCMB composites and confirmed the

presence of Si. Fig. 4.2 shows the X-ray diffraction (XRD) pattern of Si-MCMB. As

a comparison, the XRD patterns of bare MCMB and nanocrystalline Si are also

presented in the insets. The intensities of the diffraction lines of nanocrystalline Si are

much weaker than for MCMB graphite, indicating the nanocrystalline nature of the Si.

The ball-milling process caused slight broadening of the (002) peak of MCMB

graphite. The full width at half maximum (FWHM) of the (002) diffraction peak of

the MCMB were calculated to be 0.248 °, 0.250 °, 0.278 °, and 0.296 ° for pristine

MCMB, 5 h ball-milled Si-MCMB, 10 h ball-milled Si-MCMB, and 20 h ball-milled

Si-MCMB, respectively. No SiC phase was detected by x-ray diffraction. Therefore,

the mechanically added nanocrystalline Si formed Si-MCMB composites.

76
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

(a)

(b)

77
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

(c)

(d)

Fig. 4.1 SEM images of (a) bare MCMB graphite; (b) 20 wt% nano Si-MCMB ball
milled for 5 hours; (c) 20 wt% nano Si-MCMB ball-milled for 10 hours; and (d) 20
wt% nano-Si-MCMB ball-milled for 20 hours.

78
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

Fig. 4.2 X-ray diffraction pattern of 20 wt% nano-Si-MCMB ball milled for 10 hours.
The insets show the XRD patterns of pristine MCMB and nanocrystalline Si for
comparison.

Cyclic voltammogramms of MCMB and Si-MCMB electrodes in lithium-ion cells, in

which a lithium foil was used as the counter electrode, and reference electrodes are

shown in Fig. 4.3 (a) and (b) respectively. In Fig. 4.3(a), the potential of the lithium

ion insertion peak in the cyclic voltammetric curve of the MCMB electrode is very

close to 0.0 V versus the Li/Li+ reference electrode, whereas the potential of lithium

extraction is in the range of 0.3 – 0.4 V versus the Li/Li+ reference electrode. In Fig.

4.3 (b), one pair of additional reduction and oxidation peaks appeared in the CV curve

of the Si-MCMB electrode, which are located at around 0.2 V and 0.51 V,

respectively versus the Li/Li+ reference electrode. This pair of redox peaks should

correspond to the lithiation and de-lithiation of Si in the MCMB matrix. The results of

79
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

CV measurements clearly demonstrated that Si participated in the reaction with

lithium ions.

400
1st scanning cycle
300 25th scanning cycle
200
Current i (μA)

100

-100

-200

-300

-400

-500

0.0 0.5 1.0 1.5 2.0 2.5


+
Potential (V vs. Li /Li)

(a)
2000

1500

1000 1st scanning cycle


15th scanning cycle
500
Current i / μA

-500

-1000

-1500

-2000

0.0 0.5 1.0 1.5 2.0 2.5


+
Potential ( V vs. Li /Li )

(b)
Fig. 4.3 Cyclic voltammograms of (a) MCMB electrode,
(b) nano-Si-MCMB electrode (sample E).

80
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

The specific capacity and cyclability of nanocrystalline Si-MCMB composite

electrodes were measured by constant current charge/discharge testing. Table I shows

the discharge and charge capacities in the first cycle for different sample electrodes.

The bare MCMB electrode (sample A) delivered a reversible capacity of 325 mAh/g

in the first cycle with high efficiency. The bare nanocrystalline Si electrode (sample

B) delivered a very high lithium storage capacity of 3752 mAh/g, which is very close

to its theoretical capacity of 4000 mAh/g. However, this nanocrystalline Si anode had

a very high irreversible capacity of about 1821 mAh/g in the first cycle. The formation

of Li4.2Si alloy induces a 323% volume increase [143]. Such a high volume increase

could create microcracks and therefore destroy the integrity of the electrode, causing

high irreversible capacity and poor cyclability. Sample C is crystalline Si (2

μm)-MCMB composite prepared by ball milling for 10 hours. The sample C anode

shows a lower specific capacity and poorer rechargeability than for nanocrystalline Si

– MCMB electrodes (samples D, E and F). The chemical reactivity of coarse

crystalline Si is much lower than that of nanocrystalline Si. Correspondingly, the

contribution to lithium storage capacity from Si would be smaller. Because of the

effect of the coarse particle size, the volume change in local domains in the sample C

anode could be much larger than for nanocrystalline Si when embedded in MCMB.

Therefore, it is not surprising that sample C demonstrated lower specific capacity and

higher irreversible capacity on cycling.

81
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

Table 4.1 Specific capacity in the first cycle for bare MCMB, bare nanocrystalline Si,

and Si-MCMB electrodes

Samples Qdischarge Qcharge Qirreversible Efficiency


(mAh/g) (mAh/g) (mAh/g) (%)

Sample A (bare MCMB) 355 325 25 91.5

Sample B (bare 3752 1931 1821 51.5


nanocrystalline Si)
Sample C (20 wt % Si (2 687 382 305 55.6
μm) +MCMB, ballmilled
for 10 h)

Sample D (20 wt % Si 1105 936 239 84.7


(nanocrystalline) +
MCMB, ballmilled for 5 h)

Sample E (20 wt % Si 1175 1066 115 90.7


(nanocrystalline) +
MCMB, ballmilled for 10
h)

Sample F (20 wt % Si 1273 1012 261 79.5


(nanocrystalline) +
MCMB, ballmilled for 20
h)

It was found that the ball milling time has a significant impact on the specific capacity

and rechargeability of Si-MCMB composite anodes. Sample E (ball milled for 10

hours) shows an optimal lithium storage capacity and capacity retention on cycling.

We intended to use the ball-milling process to disperse nanocrystalline Si in the

MCMB matrix. The longer the ball-milling time, the better the dispersing effect that

will be obtained. However, ball-milling inevitably breaks up the MCMB spherical

particles and destroys the graphite structure, therefore degrading the electrochemical

82
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

properties of MCMB electrodes. Long-time ball-milling induces a high surface area in

the MCMB powders. The irreversible capacity in the first cycle is considered to come

from the formation of a passivating film, which consumes lithium at the electrode

material/electrolyte interface [144-146]. From SEM observations, we found that the

spherical MCMB particles were severely damaged after 20 hours of ball milling.

Therefore, the sample F (ball milling for 20 hours) anode exhibits a very high

irreversible capacity. To achieve better electrochemical performance for

nanocrystalline Si-MCMB anodes, Si has to be homogeneously dispersed in the

MCMB matrix without damaging the spherical MCMB structure. Sample E seems to

meet this requirement.

In order to combat the problems caused by volume increase in the alloying process,

reducing the particle size of Si is an efficient way to minimize the volume change in

local domains due to the formation of LixSi alloys. Another approach to overcome this

problem is to embed the active element in an inactive matrix, to let the inactive matrix

absorb or buffer the volume expansion. The intermetallic MM’ alloys and tin

composite oxides (TCO) are exactly designed to utilise this principle. We embedded

nanocrystalline Si particles in a MCMB matrix through moderate ball milling. When

lithium ions are inserted into Si-MCMB composites, nano-Si reacts with Li to form

LixSi alloys, and MCMB graphite reacts with Li to form LixC6. The intercalation of Li

in MCMB graphite causes only minor changes in the interlayer spacing and stacking

order. MCMB graphite is almost dimensionally invariable during lithium insertion

and extraction. Because Si particles are nanosize in nature, the volume increase in the

local environment is small, and can be easily absorbed by the ductile MCMB graphite

83
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

matrix surrounding the nano-Si clusters. As a result, the volume change in the

macro-domain is small and negligible for Si-MCMB electrodes. During the lithium

extraction process, a similar phenomenon occurs. This allows the integrity of the

electrode to be preserved for repeated lithium insertion and extraction. A schematic

model of the lithiation and de-lithiation processes in Si-MCMB composite has been

proposed and is presented in Fig. 4.4.

Fig. 4.4 Schematic model of the lithiation and de-lithiation process as in Si-MCMB

composites. The volume increases due to the formation of LixSi alloys are small in

local domain and effectively buffered by the MCMB matrix.

84
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

Fig. 4.5 shows the discharge/charge profiles in the first cycle for bare nanocrystalline

Si (sample B) and nano-Si-MCMB (sample E) anodes. The nanocrystalline Si

electrode shows a flat discharge plateau in the 0.2 – 0.05 V range, corresponding to

the reduction peak in the CV curve (Fig.4.3 (b)), indicating the highly reactive nature

of nanocrystalline Si powders. In contrast, the Si-MCMB (sample E) anode shows a

parabolic curve between 0.5 V and 0.2 V, and a flat plateau between 0.1 and 0.0 V,

which is typical for composite electrode materials. The cyclabilities of all the sample

anodes are shown in Fig. 4.6 The bare MCMB anode is very stable on cycling, but

with limited capacity. Although the bare nanocrystalline Si anode (sample B) had a

high initial capacity, its capacity decreased very quickly on cycling. Therefore, it is

not suitable to use nanocrystalline Si alone as an electrode in Li-ion cells. In general,

nanocrystalline Si-MCMB composite electrodes demonstrated superior performance

(high capacity and satisfactory cyclability), compared to bare MCMB and bare

nano-Si electrodes. In particular, sample E shows stable cyclability, similar to the bare

MCMB electrode, but with much higher capacity. We believe that the electrochemical

performance of nano-Si-MCMB composites can be further optimised through tuning

the material processing.

85
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

2.0

1.5

Voltage (V)
1.0

0.5

0.0

0 500 1000 1500 2000 2500 3000 3500 4000


Capacity (mAh/g)

(a)

2.0

1.5
Voltage (V)

1.0

0.5

0.0

0 200 400 600 800 1000 1200 1400


Capacity (mAh/g)

(b)

Fig. 4.5 The discharge/charge profiles of (a) bare nano-Si electrode

(b) nano-Si-MCMB electrode (sample E).

86
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

4000

3500 Sample A (MCMB)


Sample B (Nano-Si)

Specific capacity (mAh/g)


3000
Sample C ( Si (2μm)-MCMB)
Sample D (Nano-Si-MCMB, ball milled for 5 h)
2500
Sample E (Nano-Si-MCMB, ball milled for 10h)
2000 Sample F (Nano-Si-MCMB, ball milled for 20 h)

1500

1000

500

-500
0 5 10 15 20 25
Cycle number

Fig. 4.6 The discharge capacity of the samples versus cycle number

(Current density: 0.05 mAcm-2).

4.3.2 Si-carbon composite anode materials

Resorcinol reacts with formaldehyde to initially form chains of hydroxymethyl

derivatives of resorcinol, and these then further react with each other to form a

cross-linked 3-dimensional network (called RF gel). Fig. 4.7 shows a schematic

diagram of the reaction of resorcinol with formaldehyde. In the resorcinol structure,

an enhanced electron density exists in the 2, 4, and 6 ring positions. However, the

number 2 position is sterically hindered by the adjacent hydroxyl groups. So, the

reaction occurs primarily in the 4 and 6 positions [147,148]. Through cross-linking,

RF gel forms an interconnected bead structure (3-dimensional network). The

nanocrystalline Si was homogeneously embedded in this RF gel network. After heat

87
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

treatment at 650 °C, the gel was carbonized to become amorphous carbon. The final

products were Si-C composites.

Fig. 4.7 A schematic diagram of the R-F gel reaction.

Fig. 4.8 shows an X-ray diffraction pattern of typical Si-C composites. All diffraction

lines are indexed to Si. No diffraction lines from crystalline carbon (graphite) were

observed, indicating the amorphous nature of carbon in the composites. A TEM image

of nanocrystalline Si is shown in Fig. 4.9 (a). The pristine Si powders contain

nanosize Si particles with a spherical shape. The size of individual particles ranges

from 20 nm to 80 nm. Some amorphous Si powders are also present. Fig. 4.9 (b)

shows a TEM image of nano Si-C composite powders. It clearly demonstrates that Si

88
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

powders are surrounded by amorphous carbon. Spot EDX (energy dispersive x-ray)

analysis confirmed that the spherical black crystals in Fig. 4.9 (b) are Si. Therefore,

nanocrystalline Si particles are uniformly embedded in the amorphous carbon matrix

through the carbon aerogel synthesis process.

(111)
Intensity (a.u.)

(220)

(311)

(400)
20 30 40 50 60 70
2θ (degree)

Fig. 4.8 X-ray diffraction pattern of Si-C composites.

The electrochemical performance of the nanostructured Si-C composites was

systematically measured. Cyclic voltammograms of Si-C electrodes in a lithium-ion

cell, in which lithium foil was used as the counter electrode and reference electrode,

are shown in Fig. 4.10. The initial potential of the Si-C electrode is about 3.0 V in the

open-circuit state. In the first scanning cycle, there are two broad cathodic peaks

located at 0.60 V and 0.38 V, respectively, which disappear from the second cycle.

These reduction peaks could be attributed to the formation of a solid-electrolyte

89
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

interphase (SEI) passivation layer on the surface of the composite electrode, due to the

reaction of lithium with electrolyte. Once the SEI layer is formed, it will be stable

under subsequent lithium insertion and extraction, which is evidenced by the

disappearance of these two reduction peaks from the second cycle. Two anodic peaks,

located at 0.38 V and 0.50 V, gradually evolve from the first scanning cycle and

become more distinct after the fourth cycle, which corresponds to the extraction of

lithium ions from the Si-C electrode. One additional cathodic peak at 0.20 V appears

from the fourth cycle. The two pairs of redox peaks remain stable from the fourth

scanning cycle until the twentieth cycle. We suggest that several cycles are required to

activate the Si-C composite electrode under a scanning rate of 0.1 mV/S. The

observed two redox peaks should be attributed to the reaction of lithium with Si to

form LixSi alloys as indicated in equation (4.1).

xLi + Si ↔ LixSi (4.1)

This observation in cyclic voltammetry is similar to that for thin-film Si electrodes

[139].

90
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

(a)

(b)

Fig. 4.9 (a) TEM image of nanocrystalline Si powders,

(b) TEM image of nanostructured Si-C composites.

91
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

600
1st scanning cycle
400 2nd scanning cycle
4th scanning cycle
200 20th scanning cycle
Current (μA) 0

-200

-400

-600

-800

-1000

-1200
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
+
Potential vs. Li/Li (V)

Fig. 4.10 Cyclic voltammograms of Si-C composite electrode.

The specific capacity and cyclability of nanostructured Si-C composite electrodes

were measured by constant discharge/charge testing. Fig. 4.11 shows the

discharge/charge profiles of a Si-C electrode at the rate C/10. The Si-C electrode

delivered a discharge capacity of about 2000 mAh/g in the first cycle, but with a

irreversible capacity of about 550 mAh/g. Part of this irreversible capacity was used to

form a SEI passivation layer on the surface of the electrode. The reversible capacity

was improved from the second cycle. In the first cycle, there is a discharge plateau in

the voltage range of 0.8 – 0.5 V, which corresponds to the formation of the SEI layer.

However, this discharge plateau disappears from the second cycle. This observation is

in good agreement with CV measurement. A stable discharge plateau between 0.30 V

and 0.02 V is evolved from the second cycle and remains stable.

92
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

3.0
1st cycle
2.5 2nd cycle

Voltage (V) 2.0

1.5

1.0

0.5

0.0

0 250 500 750 1000 1250 1500 1750 2000


Capacity (mAh/g)

Fig. 4.11 The discharge/charge profiles of Si-C composite electrode.

2000
Discharge capacity (mAh/g)

1500

1000

500

0
0 5 10 15 20 25 30 35 40 45 50
Cycle number

Fig. 4.12 The discharge capacity versus cycle number for Si-C electrode.

93
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

The results of the cycling tests are shown in Fig. 4.12, with indicates good cyclability

of the nanostructured S-C composite electrodes. It is well known that a 300% volume

increase accompanies to the lithium alloying process with Si to form Li4.4Si. This

dramatic volume variation could pulverize the electrode, inducing poor cyclability.

We resolve this problem by utilizing nanosize silicon powders and embedding Si in an

amorphous carbon matrix. Through the synthesis of carbon aerogel, Si powders are

homogeneously embedded in a three-dimensional carbon network matrix. Since Si

particles are nanosize in nature, the volume increase in the local sites is small and can

be easily cushioned by the carbon matrix surrounding the Si nanoclusters. Therefore,

the volume change in the macro-domain will be negligibly small for the Si-C

composite electrode, which allows the integrity of the electrode to be preserved for

repeated lithiation and de-lithiation processes. As a result, the Si-C composite

electrode demonstrates a high lithium storage capacity and good cycle life.

4.4 Conclusions

Nanocrystalline Si-MCMB composite anode materials were prepared by ball milling.

They are intended to utilise the high capacity of Si and the stable cyclability of

MCMB. The cyclic voltammetry measurements showed that Si participates in the

reaction with lithium in Li-ion cells. In this investigation, we achieved a reversible

capacity of 1066 mAh/g for nanocrystalline Si-MCMB composites and fairly good

cyclability. A reaction model was proposed for the lithiation and de-lithiation

processes in the Si-MCMB electrodes in lithium-ion cells.

94
Chapter 4 Preparation and characterization Si-MCMB and Si-carbon composite anode materials
--------------------------------------------------------------------------------------------------------------------------

We also prepared nanostructured Si-C composite anode materials by dispersing

nanocrystalline Si powders in carbon aerogel, followed by heat treatment for

carbonization. Cyclic voltammetry measurements showed two pairs of redox peaks

associated with the lithiation and de-lithiation processes of LixSi alloys. A reversible

capacity of 1450 mAh/g for Si-C composite electrodes was achieved. The good

cyclability should be attributed to the usage of nanosize Si powders and their

homogeneous distribution in an amorphous carbon matrix.

95
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

Chapter 5

Multiwalled carbon nanotubes (CNTs)

anode materials

5.1 Introduction

Since the discovery of carbon nanotubes (CNTs) by Iijima [149], tremendous effort

has been devoted to the preparation, characterisation and applications of carbon

nanotubes, for both multiwalled CNTs (MWCNTs) and single walled CNTs

(SWCNTs) [150, 151]. Carbon nanotubes can be produced using various techniques

such as electric arc discharge [152], laser vaporization [153], and chemical vapour

deposition [154]. Carbon nanotubes have unique structural, mechanical, electronic,

and electrical properties. They are an attractive system with potential uses in a wide

range of applications, such as nanoscale electronics [155], chemical filters, energy

storage, and electron field emission technology [156, 157].

Carbon nanotubes have been proposed for application as lithium storage materials in

lithium-ion batteries. Carbon nanotubes have been reported with a reversible lithium

storage capacity in the range of 80 – 600 mAh/g [158 – 160]. The electrochemical

performance of carbon nanotubes strongly depends on their structure, morphology,

and disorder, and on the quality of the carbon nanotube bundles. Intermetallic alloy

(MM’) anodes rely on the reaction between lithium ions and the active element M to

96
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

form LixM alloys. Transition metal oxides have been reported to exhibit high Li-ion

storage capacity in lithium-ion cells, which is based on the reversible reaction for the

formation of Li2O and metal oxide. Tin oxide (SnO2) has been extensively

investigated as an alternative anode materials for Li-ion batteries. The

electrochemical reaction of SnO2 anode in a Li-ion cell involves two steps: (i) the

formation of lithium oxide, and (ii) lithium alloying with Sn. The second step is an

alloying/de-alloying process, giving a maximum lithium storage capacity of 781

mAh/g. However, the significant volume variation during lithiation induces poor

cyclability [161, 162].

Vertically aligned carbon nanotubes were grown in this work on a planar quartz

substrate by a thermal CVD approach using camphor as the carbon source. The

electrochemical properties of these carbon nanotubes for lithium storage in

lithium-ion cells were systematically evaluated. The preparation, characterization,

and electrochemical performance of composite anode materials formed from CNTs

and SnO2 in lithium-ion cells are described in this chapter.

5.2 Experimental

5.2.1 Synthesis of vertically aligned CNTs

The vertically aligned carbon nanotubes were grown on planar quartz substrates with

a size of 20 × 20 mm by catalytic chemical vapour deposition (CVD), using

nanocrystalline iron as the catalyst. Fig. 5.1 shows a schematic diagram of the CVD

97
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

quartz reactor, which consists of three different temperature zones heated by resistive

heating coils. The quartz substrate was placed in the hot zone and heated to 900 °C.

The low temperature zone was maintained at 350 °C. A quartz boat containing iron

carbonyl and another quartz boat containing camphor were also prepared. A small

piece of nickel wire was attached to the quartz boats for magnetic attraction. When the

temperatures in the heating zones reached the designated temperatures, the quartz

boat containing iron carbonyl was first moved into the low temperature zone with a

permanent magnet. The iron carbonyl was then evaporated and deposited on the

quartz substrate in the hot zone, to serve as a catalyst. The second quartz boat

containing camphor was then brought into the low temperature zone. The camphor

was then evaporated and decomposed in the high temperature zone, inducing carbon

nanotubes to grow on the quartz substrate. An argon gas flow was applied in the

opposite direction to the reaction gas at a speed of 20 ml/min.

Fig. 5.1 A schematic diagram of the CVD quartz reactor.

98
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

5.2.2 Synthesis of SnO2-CNTs composites

Multiwalled carbon nanotubes (MWNTs, Carbon Nanotechnology Inc., USA), SnCl2

(99.9%, Sigma-Aldrich), and ethanol were mixed homogeneously with a high speed

mixer. The molar ratio of Sn:C was 0.1:1 in the mixture, with a nominal weight

percentage of 55.6 wt% SnO2 in the final product. The mixed suspension was put into

a vacuum oven heated to 90 ºC to evaporate the ethanol. The dried mixture was then

heated to 600 ºC to form SnO2. The calcined product was washed with de-ionized

water to remove Cl- by filtering, and then dried at 100 ºC under vacuum to obtain a

SnO2-CNT composite.

5.2.3 Physical and structural characterization

X-ray diffraction (XRD) measurements were carried out on a Philips PW 1730

diffractometer with Cu Kα radiation. The overall configuration and morphology of

the CNTs were observed by SEM (Leica/Cambridge Stereoscan 440 scanning

electron microscope). The detailed structure of the CNTs was examined by HRTEM

using a 300 KV JEOL JEM-3000F transmission electron microscope with field

emission. Thermogravimetric analysis (TGA) was employed to determine the

percentage of SnO2 in the composites.

99
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

5.2.4 Electrochemical measurements

The electrochemical performance of the vertically aligned carbon nanotubes and

SnO2-CNT composites was tested by fabricating CR2032 coin cells. The working

electrode was prepared by mixing the active materials with 10 wt% carbon black and

10 wt% polyvinylidene fluoride (PVDF) binder in N-methyl-2-pyrrolidinone (NMP)

to form a slurry with appropriate viscosity. The slurry was then spread onto copper

foil, which was then cut out into φ12 mm electrodes. The thickness of each electrode

was approximately 60 – 80 μm, with a loading of 1.5 mg. The coin cells were

assembled in an argon filled glove box (Mbraun, Unilab, Germany), using a

hand-operated closing tool (Hosen Co., Japan). The electrolyte was 1 M LiPF6 in a

mixture of ethylene carbonate (EC) and dimethyl carbonate (DMC) (1:1 by volume,

provided by MERCK KGaA, Germany). The cells were galvanostatically charged

and discharged in the voltage range of 0 – 2.0 V versus Li/Li+. Cyclic voltammetry

(CV) measurements were carried out using a CHI660B Electrochemical Work Station

at a scanning rate of 0.1 mVs-1.

5.3 Results and discussion

The configuration of the CNTs is shown in Fig. 5.2 (a) (b) (c) and (d). The as-prepared

CNTs are well-aligned over the quartz substrate. The CNTs have grown in a vertical

direction to the substrate and have a uniform distribution. The length of the CNTs is

about 70 - 80 μm. Fig. 5.2 (b) shows a vivid view of the vertically aligned CNTs

obtained by cutting the CNT layer with a razor. CNTs were then peeled off from the

100
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

substrate for further SEM study. Fig. 5.2 (c) shows the CNT bundles, which

demonstrate the feature of very long CNTs entangled with each other. Only very tiny

quantities of carbon clusters are present in Fig. 5.2 (c), indicating the high yield rate of

CNTs. Fig. 5.2 (d) shows a magnified view of the dispersed CNTs, which further

demonstrates the high quality of the CNTs obtained.

TEM and HRTEM analysis was performed on the CNTs grown on the quartz

substrate. The CNTs were peeled off and dispersed on a carbon coated microgrid. Fig.

5.3 (a) shows a TEM image of the CNTs. Except for some stand-alone multiwalled

carbon nanotubes, the CNTs are entangled together, forming bundles. Some catalyst

particles are encapsulated inside the CNTs. Spot EDS analysis confirmed that the

black dots are Fe catalyst. The outer diameter of the CNTs is about 40 – 60 nm. Fig.

5.3 (b) shows the HRTEM image of a single multiwalled carbon nanotube.

Compartment structures were observed in this carbon nanotube, which consists of a

few curved chambers. The walls of the CNT consist of approximately 10 graphite

sheets with clean fringes.

101
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

(a)

(b)

102
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

(c)

(d)

Fig. 5.2 SEM images of vertically aligned carbon nanotubes.

103
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

(a)

(b)

Fig. 5.3 (a) TEM image of carbon nanotubes,

(b) HRTEM image of a carbon nanotube.

104
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

Cyclic voltammetry was conducted on the carbon nanotube electrodes to determine

the electrochemical reactivity of the CNTs in lithium-ion cells, in which lithium metal

was used as both the working electrode and the reference electrode. Fig.5.4 shows

cyclic voltammograms of the carbon nanotube electrodes. The characteristics of the

CV curves demonstrated that lithium ions can reversibly intercalate and de-intercalate

in carbon nanotubes. The lithium insertion potential is very low, close to 0.0 V versus

the Li+/Li reference electrode. However, the potential for lithium extraction is in the

range of 0.15 - 0.30 V. A slight voltage hysteresis between lithium insertion and

extraction in the CNT electrodes is similar to the situation with other carbonaceous

materials [164]. The intensity of the redox peaks remained stable for many scanning

cycles, indicating good reversibility of the carbon nanotubes for lithium insertion and

extraction.

60
2nd scanning cycle
40
20th scanning cycle

20
Current i (μA)

-20

-40

-60

-80

-100
-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
+
Potential (V) versus Li/Li

Fig.5.4 Cyclic voltammograms of the carbon nanotube electrode,

Scanning rate: 0.1mV/s.

105
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

Fig. 5.5 shows the first discharge and subsequent discharge curves of a carbon

nanotube electrode. The first discharge curve shows a plateau at about 0.8 V, which is

associated with the formation of a SEI layer on the surface of the carbon nanotube

electrode. Carbon nanotubes usually have a large specific surface area. The

decomposition of the electrolyte and the formation of lithium organic compounds

occur on the surface of carbon nanotubes during the first discharge process. The

capacity used for SEI formation is not reversible [165]. In the subsequent cycles, the

discharge plateau at 0.8 V disappeared. A reversible lithium storage capacity of 950

mAh/g was obtained from the second cycle and remained stable, indicating good

cyclability. Multiwalled carbon nanotubes consist of graphite sheets rolled into closed

concentric cylinders. The concentric tubes are separated by van der Waals gaps of ~

3.4 Å, a typical interlayer spacing in turbostratically disordered graphite. Assuming

that lithium ions can only combine with every second hexagon on the external surface

of a rolled graphene sheet, the limiting stoichiometry would be LiC6, with a

theoretical capacity of 372 mAh/g. However, we obtained a reversible capacity of

950 mAh/g. We suggest that there are four possibilities for lithium intercalation in the

carbon nanotube electrode: (i) lithium ions intercalate into intratubal sites (the hollow

core); (ii) lithium ions intercalate into intertubal sites (the sites between individual

nanotubes ) through diffusion in the bundles; (iii) lithium ions intercalate into carbon

nanoclusters (since small quantities of carbon clusters have been identified in carbon

nanotubes); (iv) lithium ions intercalate into layered graphite structures, including

both nanotube structures and nanocluster structures. All these four possibilities could

contribute to the high lithium insertion capacity of carbon nanotube electrodes [166].

The lithium storage capacity versus cycle number is shown in the inset in Fig. 5.5.

106
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

After the first cycle, the carbon nanotube electrode exhibited stable cyclability for

repeated lithium insertion and extraction.

2.0
1500

1000

1.5
500
Voltage (V)

0
5 10 15 20 25 30
1.0

0.5

3D
0.0 2D 1D

0 200 400 600 800 1000 1200 1400 1600


Discharge capacity (mAh/g)

Fig. 5.5 Discharge curves of the carbon nanotube electrode. The inset shows the

discharge capacity versus cycle number for the carbon nanotube electrode.

The amount of SnO2 in the composite was determined by TGA analysis using air as

the carrying gas. The composite contains 55 wt % SnO2, which is consistent with the

nominal percentage. Fig. 5.6 shows the X-ray diffraction pattern of SnO2-CNT

composite. The SnO2 diffraction lines are clearly indexed, with broad diffraction

peaks indicating a nanocrystalline nature. The diffraction lines of CNTs are weak and

not distinguishable. The particle size of SnO2 was calculated to be in the range of 10 –

20 nm using the Scherrer equation. The distribution and morphology of the SnO2

particles were analyzed from TEM and HRTEM observations.

107
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

(110)
Intensity (a.u.)

(101)

(211)
20 30 40 50 60 70
Angles (2θ)

Fig. 5.6 X-ray diffraction pattern of SnO2-CNT composites.

Fig. 5.7 (a) presents a TEM image showing a general view of the SnO2-CNT

composite, which clearly demonstrates that nanosize SnO2 particles are distributed

homogeneously on the carbon nanotube matrix. Fig. 5.7 (b) further shows a high

magnification TEM image of the composite, from which it was found that some SnO2

particles formed, inside carbon nanotubes. Therefore, we can conclude that SnO2

particles were formed not only on the outside walls of CNTs (inter-), but also inside

CNTs (intra-), when synthesized by the chemical treatment process.

108
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

(a)

(b)

Fig. 5.7 TEM images of SnO2-CNT composites:

(a) a general view, (b) HRTEM image.

109
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

(a)

(b)

110
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

(c)

(d)

Fig. 5.8 HRTEM images of SnO2-CNT composites.

111
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

Fig. 5.8 shows high resolution TEM images of SnO2-CNT composites. As shown in

Fig. 5.8 (a), a large number of SnO2 nanocrystals were formed, and they became

attached to the walls of the carbon nanotubes. Fig. 5.8 (b) shows the distribution of

SnO2 nanoparticles on a single carbon nanotube. The HRTEM images are shown in

Fig. 5.8 (c) and (d). We can see that the carbon nanotubes are multiwalled, with an

inner diameter of about 10 - 20 nm. The nature of the SnO2 nanocrystals is

demonstrated in further detail in Fig. 5.8 (d), from which the sizes of the SnO2

nanocrystals were measured to be in the range of 2 – 5 nm. Lattice images of SnO2

nanocrystals are clearly visible, with a 3.4 Å interplanar distance. Although those

SnO2 nanocrystals are of very small size, of the same order as quantum dots, they

have attained a well ordered crystalline state.

The electrochemical reactivity of SnO2-CNT composite electrode was examined by

cyclic voltammetry (CV). Fig. 5.9 shows the CV curves of SnO2-CNT composite

electrodes. In the first cycle, there is a strong reduction peak at 0.85 V, which is

associated with the reaction between Li+ and SnO2:

4Li+ + SnO2 → Sn + 2Li2O (5.1)

The strength of this reduction peak decreased substantially in the second scanning

cycle and gradually disappeared in the subsequent scanning cycles. Following the

formation of Sn and Li2O, there is a reduction peak at 0.15 V, which is associated with

lithium alloying with Sn:

112
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

xLi+ + Sn + xe- → LixSn 0 ≤ x ≤ 4.4 (5.2)

The final reduction peak at 0.0 V is related to lithium insertion in carbon nanotubes:

xLi+ + C + xe- → LixC (5.3)

During the charging process, three oxidation peaks appeared on the CV curves. The

first oxidation peak (a small hump), which is related to the extraction of Li from

carbon nanotubes, is located at 0.1 V, indicating that a limited amount of Li was

inserted and extracted from the carbon nanotubes. The strongest oxidation peak is at

0.55 V. This is related to the de-alloying of lithium from LixSn alloy:

LixSn → xLi+ + Sn + xe- (5.4)

The last oxidation peak at 1.25 V is weak and decreases gradually. We suggest that

this oxidation line originates from the extraction of lithium from Li2O, which is

similar to the mechanism for transition metal oxides.

The specific capacity and cyclability of SnO2-CNT composite electrodes were tested

via galvanostatic charge and discharge cycling at the C/2 current rate using CR2032

coin cells. The voltage window was set between 0.0 V and 2.0 V vs. the Li/Li+ counter

electrode. Fig. 10 shows the charge and discharge curves of SnO2-CNT electrode in

the first and second cycles. In the first cycle, there is an irreversible capacity of about

390 mAh/g, which was caused by the formation of both Li2O and an SEI film on the

113
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

surface of the electrode. However, the irreversible capacity decreased dramatically in

the second cycle. The cyclability of SnO2-CNTs is shown in Fig. 5.11. In the first ten

cycles, the lithium storage capacity decreased continuously, but the SnO2-CNT

electrode demonstrated stable cyclability thereafter. The electrode retained a lithium

storage capacity of 410 mAh/g after fifty cycles. The charging efficiency (the ratio of

charge capacity to discharge capacity) is shown in the inset in Fig. 5.11. The

SnO2-CNTs electrode demonstrated a very high charging efficiency (> 98%) after ten

cycles.

4.0x10-4
E

2.0x10-4
F
D
Current / A/cm2

0.0

-2.0x10-4
20th
2nd

-4.0x10-4
1st
-6.0x10-4 A
C
B
-8.0x10-4
0.0 0.5 1.0 1.5 2.0
Potential vs. Li/Li+ / V

Fig. 5.9 Cyclic voltammograms of SnO2-CNT composite electrode.

Scanning rate: 0.1 mV/s.

114
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

2.5

2.0
Voltage (V)
1.5

1.0

0.5

1st
0.0 2nd

0 200 400 600 800 1000


Discharge capacity (mAhg -1 )

Fig. 5.10 The first and second discharge and charge curves of
SnO2-CNT composite electrode.

1100

1000 100
Discharge capacity / mAhg-1

90
80
Efficiency / ( /100)

900 70
60

800 50
40
30

700 20
10

600 0 10 20 30
Cycle number
40 50

500

400

300

200

100

0
0 10 20 30 40 50
Cycle number

Fig. 5.11 Discharge capacity vs. cycle number for SnO2-CNT composite electrode,

the inset presents the charging efficiency vs. cycle number.

115
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

From TEM observation, we know that SnO2 crystals are very tiny (only a few

nanometers). They are well distributed and stick on both the inner and the outer

carbon nanotube walls. Due to the one dimensional characteristics of carbon

nanotubes, a network structures is formed when pasting SnO2-CNTs into a thick-film

electrode. Therefore, excellent electronic conductivity of the electrode is expected.

On the other hand, the volume expansion during the lithium reaction with SnO2 is

relatively small and limited due to the size of the SnO2 nanocrystals. Such a small

volume change can be effectively cushioned by the carbon nanotube matrix. We

believe that these two factors contributed to the good cyclability of the SnO2-CNT

electrodes.

5.4 Conclusion

Vertically aligned multiwalled carbon nanotubes were successfully prepared by

chemical vapour deposition. The carbon nanotubes have a typical diameter of several

tens of nanometers and consist of compartment structures. Cyclic voltammetry

measurements showed that the carbon nanotubes are electrochemically active to

lithium insertion and extraction. A reversible lithium storage capacity of 950 mAh/g

has been achieved for the as-prepared carbon nanotubes.

SnO2-CNT composite anode materials have been synthesized via chemical treatment

followed by heating at high temperature. The solution-based chemical process

enables Sn2+ ions to penetrate into the inner cavity of the carbon nanotubes. SnO2

nanocrystals were observed to be homogeneously distributed on the carbon nanotube

116
Chapter 5 Multiwalled carbon nanotubes (CNTs) anode materials
--------------------------------------------------------------------------------------------------------------------------

matrix by TEM and HRTEM analysis. The SnO2-CNT composite electrodes

exhibited stable cyclability, with a lithium storage capacity of 410 mAh/g after fifty

cycles.

117
Chapter 6 Synthesis and electrochemical characterisation of iron phosphide as anode material for
lithium-ion batteries
--------------------------------------------------------------------------------------------------------------------------

Chapter 6

Synthesis and electrochemical

characterisation of iron phosphide as anode

material for lithium-ion batteries

6.1 Introduction

Lithium-ion batteries using a carbon anode and a lithium cobalt oxide cathode are the

state-of-the-art power sources for modern electronic devices [167–170]. Exploration

of new anode materials with high energy density is critical for the development of the

next generation of lithium-ion batteries. Nanosize electrode materials are expected to

result in significant advances in the search for high capacity anode materials. Systems

recently developed as non-carbonaceous anode materials include tin-based oxides,

intermetallic alloys, transition metal oxides, etc.. These anode materials have

demonstrated higher capacities than carbon anodes, but suffer from large irreversible

capacity and poor cycle life. For example, there have been intensive investigations on

tin, tin based alloys, and tin oxide composites [171,141,172]. The formation of Li4.4Sn

results in a theoretical capacity of 991mAh/g for elemental Sn. However, this alloying

process is accompanied by a 259% volume increase, causing the disintegration of the

electrode. This phenomenon can be partially alleviated by embedding Sn in an

inactive matrix to create a buffering effect. Silicon has the highest known theoretical

capacity of 4000 mAh/g when forming Li4.2Si alloys, far greater than that of carbon

118
Chapter 6 Synthesis and electrochemical characterisation of iron phosphide as anode material for
lithium-ion batteries
--------------------------------------------------------------------------------------------------------------------------

and Sn. However, this alloying process is also associated with a 300% volume

dilation, pulverizing the brittle electrode and inducing poor cyclability [173 - 175].

Transition metal phosphides such as MnP4 and Zn3P2 were discovered to exhibit

interesting phenomena with respect to reversible lithium storage [176, 177].

Transition metal phosphides have conventionally been synthesized by solid state

sintering at high temperature over long periods. After sintering, the products were

purified by acid etching [178-180]. Here we report the direct synthesis of transition

metal phosphides using a hydrothermal method. Crystalline iron phosphide (FeP4)

powders were prepared by a solvothermal synthesis technique.

6.2 Experimental

Red phosphorus powders (99%, Aldrich) and FeCl2.4H2O (99%, Aldrich) were

weighed out in appropriate amounts according to the stoichiometry to synthesize FeP4

compounds and loaded into a Teflon lined autoclave. De-ionised water was added as

solvent. The autoclave was inserted into an oven and kept at the reaction temperature

of 180 ˚C for 24 hours. The resulting product was filtered by vacuum suction and

washed with water, then dried at 110 ˚C. X-ray diffraction was performed on the

prepared FeP4 powders using Cu-Kα radiation (Philips PW1730 X-ray

diffractometer). The micro-morphology of the FeP4 powders was studied using a

scanning electron microscope (JEOL JSM-6460A).

119
Chapter 6 Synthesis and electrochemical characterisation of iron phosphide as anode material for
lithium-ion batteries
--------------------------------------------------------------------------------------------------------------------------

The electrochemical evaluation of the FeP4 anode materials was accomplished by

assembling CR2032 coin cells. The electrodes were made by dispersing 92wt %

active materials and 8wt % polyvinylidene fluoride (PVDF) binder in n-methyl

pyrolidone (NMP) solvent to form homogeneous slurry. The slurry was then spread

onto an Al foil to form the electrodes. The coated electrodes were dried in a vacuum

oven at 120 °C for 12 hours and then pressed to enhance the contact between the

active materials and the conductive carbon. The cells were assembled in an argon

filled glove-box (Mbraun, Unilab, Germany) using lithium metal foil as the counter

electrode. The electrolyte was 1 M LiPF6 in a mixture of ethylene carbonate (EC) and

dimethyl carbonate (DMC) (1:1 by volume, provided by MERCK KgaA, Germany).

The cells were galvanostatically charged and discharged over a voltage range of 0.02

– 1.5 V at a C/5 rate. Cyclic voltammetry (CV) measurements were performed using a

CH Instruments electrochemical workstation (CHI 660A) at a scanning rate of 0.1

mV/s.

6.3 Results and discussion

We intended to synthesize FeP4 compounds by using a hydrothermal synthesis

method. Crystalline FeP4 has a monoclinic structure (space group: p21/c). Fig. 6.1

shows the X-ray diffraction pattern of the as-prepared iron phosphide compound.

Most of the diffraction lines could be indexed as the monoclinic phase, but some

small amounts of unknown impurities were also present. In general, the diffraction

peaks are weak and broad, indicating the nanosize nature of the synthesized FeP4

powders.

120
Chapter 6 Synthesis and electrochemical characterisation of iron phosphide as anode material for
lithium-ion batteries
--------------------------------------------------------------------------------------------------------------------------

350

(-112)
(012)
300

(121)
Intensity (a.u.)

(-121)
(100)

(031)
250

200

(140)

(042)
150

100

50
20 25 30 35 40 45 50 55 60 65 70 75
Angles 2θ (degree)

Fig. 6.1 X-ray diffraction pattern of the as-prepared FeP4 powders

SEM image of the FeP4 powders is shown in Fig. 6.2. The FeP4 powders

demonstrated the characteristics of a mixture of small crystalline particles and

amorphous cotton-like phases. The sizes of the individual crystals were less than 100

nm. Those individual crystals were agglomerated together to form large composite

particles. The observed amorphous phases could be impurity phases. SEM

observations revealed that the as-prepared FeP4 compounds produced via

hydrothermal reaction are nanosize. We performed energy dispersive spectroscopy

(EDS) analysis on the powder sample, which was spread onto a carbon tape without

any coating. The atomic ratio of Fe: P is 1:3.98, which confirmed the correct

stoichiometry of the prepared sample. Therefore, the as-prepared powders are

121
Chapter 6 Synthesis and electrochemical characterisation of iron phosphide as anode material for
lithium-ion batteries
--------------------------------------------------------------------------------------------------------------------------

primarily FeP4 compound with small amount of impurities. Conventionally, transition

metal phosphides have been synthesized via a tin-flux method at high temperature.

Under solvothermal conditions, red phosphorous becomes active, acting both as a

reducing agent and a reactant. The possible reaction is proposed as equation (6.1).

FeCl2 + P + H2O → FeP + HCl + O2 (6.1)

The reaction product FeP4 is a precipitate, which can be easily separated via filtration.

Fig. 6.2 SEM image of the FeP4 powders.

Cyclic voltammetry measurements (CV) were performed on the FeP4 electrode using

lithium metal foil as both a pseudo-reference and a counter electrode in a CR2032

122
Chapter 6 Synthesis and electrochemical characterisation of iron phosphide as anode material for
lithium-ion batteries
--------------------------------------------------------------------------------------------------------------------------

coin cell at a scanning rate of 0.1 mV/s in the voltage range of 0.0 V – 3.0 V. Fig. 6.3

shows the CV curves of FeP4. During the cathodic scanning process, there are three

well-distinguished reduction peaks at 1.50 V, 0.55 V, and 0.05 V respectively, while

in the anodic scanning process, there are two well-defined anodic peaks centred at 1.1

V and 1.3 V. Previously, it was proposed that the lithium insertion and extraction in

MnP4 proceed as MnP4 ↔ Li7MnP4, accompanied by a topotactic first–order

transition between different related crystal structures [176]. Our observed reaction

phenomena could be much more complicated.

0.30

0.25

0.20 1st scanning cycle


0.15
20th scanning cycle

0.10
Current (mA)

0.05

0.00

-0.05

-0.10

-0.15

-0.20

-0.25

-0.30
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
+
Potential vs. Li/Li

Fig. 6.3 Cyclic voltammograms of FeP4 anode in lithium-ion cell.

123
Chapter 6 Synthesis and electrochemical characterisation of iron phosphide as anode material for
lithium-ion batteries
--------------------------------------------------------------------------------------------------------------------------

Based on the CV curves, we propose the following reaction mechanisms.

Lithium insertion process:

Li + FeP4 → LixFeP4 (6.2)

Li + LixFeP4 → LiFeP4 + 3P (6.3)

Li + P → Li3P (6.4)

Lithium extraction process:

Li3P → Li+ + P (6.5)

LiFeP4 + P → FeP4 + Li+ (6.6)

However, the confirmation and further analysis of the reactions between lithium ions

and iron phosphides require in situ X-ray diffraction measurements.

124
Chapter 6 Synthesis and electrochemical characterisation of iron phosphide as anode material for
lithium-ion batteries
--------------------------------------------------------------------------------------------------------------------------

2.5

2.0

1.5
Voltage (V)

1.0

0.5

0.0 2nd 1st


20th

0 200 400 600 800 1000 1200


-1
Capacity (mAhg )

Fig. 6.4 Charge/discharge profiles of FeP4 anode.

The lithium storage capacity and reversibility of the FeP4 anodes were tested via

galvanostatic charge and discharge testing Fig. 6.4. The first discharge capacity of

FeP4 electrode reaches as high as 1130 mAh/g, which is equivalent to the intake of

4.33 mole lithium per formula FeP4. However, about 410 mAh/g of the capacity is

irreversible in the first charging process. After the first cycle, FeP4 electrode is fairly

stable for repeated charge and discharge, with a capacity of about 700 mAh/g. As

shown in Fig. 6.4, the FeP4 electrode showed the same charging and discharging

profile after the first cycle. During the charging process (lithium extraction), there are

two voltage plateaus, centred at 1.1 V and 1.3 V, respectively, which match the

observed oxidation peaks on the CV curves Fig. 6.3. During the discharging process

125
Chapter 6 Synthesis and electrochemical characterisation of iron phosphide as anode material for
lithium-ion batteries
--------------------------------------------------------------------------------------------------------------------------

(lithium insertion), the voltage of the electrode drops sharply from 2.0 V to 0.8 V, and

then gradually descends to a plateau centered at 0.5 – 0.6 V, followed by another steep

descent to 0.0 V. The 0.5 V plateau matches the reduction peak at 0.5 V that was

observed in CV measurements. There is no voltage plateau in the discharge profile

matching the reduction peaks at 1.5 V and 0.02 V. The cyclability of FeP4 with respect

to lithium insertion and extraction is shown in Fig. 6.5. In general, FeP4 electrode

exhibited stable cyclability after the first cycle, with a capacity of about 700 mAhg-1.

1200

1000
Discharge capacity (mAhg )
-1

800

600

400

200

0
0 10 20 30 40 50
Cycle number

Fig. 6.5 The discharge capacity of FeP4 anode vs. cycle number.

126
Chapter 6 Synthesis and electrochemical characterisation of iron phosphide as anode material for
lithium-ion batteries
--------------------------------------------------------------------------------------------------------------------------

6.4 Conclusion

We prepared nanosize FeP4 compound via the hydrothermal method. The as-prepared

products contain a majority of monoclinic FeP4 phase with some small amount of

impurities. SEM observation shows nanosize particles of FeP4 powders. Cyclic

voltammetry measurements demonstrated the reversible reactivity of FeP4 anodes

towards lithium insertion and extraction. The FeP4 anode exhibited a stable lithium

storage capacity of about 700 mAhg-1.

127
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

Chapter 7

Characterisation of olivine-type

LiMnxFe1-xPO4 cathode materials

7.1 Introduction

Lithium polyanion compounds with the general formula LixMy(XO4)z (M: transition

metal; X: P, S, As, Mo,or W) have attracted worldwide interest as new cathode

materials for lithium-ion batteries [92,93,181 - 183]. Among them, LiFePO4 and

LiMnPO4 compounds are particularly interesting. Both LiFePO4 and LiMnPO4 have

an orthorhombic olivine structure (space group: Pnma), in which Li, Fe (Mn) and P

atoms occupy octahedral 4a, octahedral 4c,and tetrahedral 4c sites. The strong P-O

covalency stabilizes the anti-bonding Fe3+/Fe2+ (Mn3+/Mn2+) state through the Fe-O-P

(or Mn-O-P) inductive effects to generate a high operating potential. In the olivine

crystal structure, the oxygen atoms are in a slightly distorted, hexagonal close-packed

arrangement. The FeO6 (MnO6) octahedra share corners in the bc plane, and LiO6

octahedra form an edge sharing chain in the b-direction. The polyanion with strong

P-O covalency provides a stable crystal structure for lithium insertion and extraction,

inducing good cycle life and excellent safety when the compound is used as electrode

in lithium-ion cells. However, the separation of the FeO6 octahedra by PO4 polyanions

dramatically reduces the electronic conductivity of the material, causing poor rate

capacity of the electrode. The electronic conductivity of LiFe(Mn)PO4 olivines can be

128
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

effectively improved by various techniques, such as carbon coating, supervalence

cation doping and synthesis of nanocrystalline grains [184 - 189].

Paddi et al. [1] have reported on the electrochemical characteristics of LiMnxFe1-xPO4

(x = 0.25, 0.50, 0.75, and 1.0), and found that the capacity rapidly decreases at x >

0.75. Yamada et al. [190–192] also performed intensive research on LiMnxFe1-xPO4

systems. They discovered that Mn-rich LiMnxFe1-xPO4 (x > 0.8) is not suitable as a

cathode material, due to the large anisotropic distortion of Mn3+. We report here the

synthesis, crystal structure refinement, and magnetic and electrochemical properties

of a series of LiMnxFe1-xPO4 cathode materials. We have identified a number of

conductive phosphides and manganese phosphates that were formed during the

sintering process, which could enhance the electronic conductivity of the materials.

7.2 Experimental

7.2.1 Materials synthesis

LiMnxFe1-xPO4 (x = 0, 0.1, 0.3, 0.5, 0.7, 0.9 and 1) compounds were synthesised by a

sol-gel method. LiOH.H2O (99.9%, Aldrich), FeC2O4.2H2O (99%, Aldrich), and

Mn(NO3)2.6H2O (99.9%, Aldrich) were used as reactants. The stoichiometric

amounts of reactants were dissolved in de-ionised water, in which polyacrylic acid

and citric acid were added as complexing agents for the formation of the gel. The

solution was heated at 85 °C under vigorous stirring until a viscous gel was formed.

The as-formed gel was calcined at 500 °C for 4 hours to decompose the organic

129
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

components under flowing argon gas. The decomposed precursors were then sintered

at 750 °C for 15 hours under a flowing gas mixture (10% H2 in Ar). The reducing

atmosphere was employed during the sintering process in order to prevent the

oxidation of Fe2+ and Mn2+ cations.

7.2.2 Magnetic measurement

The temperature-dependent dc magnetic susceptibility was measured using a

Quantum Design MPMS XL SQUID magnetometer in a magnetic field of 2000 Oe

from 5 K to 300 K with a 2 K step under field cooling (FC) conditions.

7.2.3 X-ray diffraction and SEM analysis

Powder X-ray diffraction patterns were obtained on a Sintag (X1) diffractometer

with Cu Kα radiation. The diffraction intensity was measured from 20 ° to 120 ° using

a step size of 0.02 ° with a counting time of 10 s per step. Rietveld refinement was

performed using the GSAS (General Structure Analysis System) program to obtain

the crystal structure parameters. Scanning electron micrographs were obtained with a

JEOL JSM-6460A scanning electron microscope (SEM).

7.2.4 Electrochemical testing

Electrochemical characterisation was performed by assembling CR2032 coin cells for

130
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

cyclic voltammetry and galvanostatic charge/discharge measurements. The electrodes

were made by dispersing 80 wt% active material, 10 wt% carbon black and 10 wt%

polyvinylidene fluoride (PVDF) in n-methyl pyrrolidone (NMP) to form a slurry. The

slurry was then coated onto an Al foil. The coated electrodes were dried in a vacuum

oven and then pressed at 1200 kg/cm2. The thicknesses of the electrodes were about

100 µm, and the mass of the active materials was about 2 mg. The coin cells were

assembled in an argon filled glove box (Mbraun, Unilab, Germany) with lithium foil

as the counter electrode. The electrolyte was 1 M LiPF6 in a 1: 1 mixture of ethylene

carbonate (EC) and dimethyl carbonate (DMC).

7.3 Results and discussion

X-ray diffraction patterns of the LiMnxFe1-xPO4 (x = 0, 0.1, 0.3, 0.5, 0.7, 0.9 and 1.0)

samples are shown in Fig. 7.1 (a). The main diffraction lines are indexed as an

orthorhombic olivine crystal structure (triphylite, space group: Pnma). We performed

Rietveld refinement on the X-ray diffraction patterns using the GSAS program [193].

Fig. 7.1 (b) shows a typical Rietveld refinement on the XRD pattern of the

LiMn0.5Fe0.5PO4 sample. Several small fractions of impurities have been identified

from the Rietveld refinement, which are listed in Table 7.1.

It is worth noting that the impurity phases in LiFePO4 are Fe2P, FeP, and Li4P2O7.

With increasing Mn content, the impurity phases changed to Mn2P2O7. Iron

phosphides are usually formed during the sintering process under a reducing

131
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

atmosphere [194]. The selected bond distances and angles for all the LiMnxFe1-xPO4

samples are described in Table 7.2.

(a) LiFePO4
(b) LiMn0.1Fe0.9PO4
(c) LiMn0.3Fe0.7PO4
(d) LiMn0.5Fe0.5PO4
(e) LiMn0.7Fe0.3PO4
(f) LiMn0.9Fe0.1PO4
(g) LiMnPO4
Intensity (a.u.)

(g)
(f)
(e)
(d)
(c)
(b)
(a)

20 25 30 35 40 45 50 55 60 65 70 75
2θ / degree

(a)
Intensity (a.u.)

LiFePO4
Mn2P2O7
Fe2P
FeP

10 20 30 40 50 60 70 80 90 100 110 120


2Θ / degree

(b)

Fig. 7.1 (a) X-ray diffraction patterns of LiMnxFe1-xPO4, (b) Rietveld refinement

on X-ray diffraction pattern of LiMn0.5Fe0.5PO4 sample.

132
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

Table 7.1 Phase composition and lattice parameters of olivine phase for

LiMnxFe1-xPO4 compounds

Sample Phase composition Lattice parameters of olivine phase

Fraction
Phase * a, Å b, Å c, Å V, Å3

Olivine 7.3366
Li3PO4 0.34292
LiFePO4 10.32559(7) 6.00463(4) 4.68908(3) 290.729(3)
Fe2P 0.70963
FeP 0.079816

Olivine 11.714
LiFe0.9
Fe2P 0.092737 10.3366(2) 6.0130(1) 4.69465(8) 291.791(9)
Mn0.1PO4
FeP 0.23027

Olivine 46.819
LiFe0.7
Fe2P 0.44832 10.3548(3) 6.0279(2) 4.7031(1) 293.56(1)
Mn0.3PO4
FeP 0.63682

Olivine 78.076
LiFe0.5 Fe2P 0.17978
10.3873(1) 6.05132(7) 4.71797(6) 296.557(6)
Mn0.5PO4 FeP 3.0864
Mn2P2O7 1.9555

Olivine 20.368
LiFe0.3
FeP 0.064183 10.4160(2) 6.07332(9) 4.73074(8) 299.262(8)
Mn0.7PO4
Mn2P2O7 1.1980

LiFe0.1 Olivine 17.375


10.4323(3) 6.0913(2) 4.7375(1) 301.05(2)
Mn0.9PO4 Mn2P2O7 0.30770

Olivine 6.9780
LiMnPO4 10.44723(9) 6.10279(5) 4.74778(5) 302.705(5)
Mn2P2O7 1.1349

*Phase fraction in the sample is the phase fraction coefficient in the GSAS program.

133
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

Table 7.2 Selected interatomic distances (Å) and angles (deg) for LiMnxFe1-xPO4

samples

LiMn0.1 LiMn0.3 LiMn0.5 LiMn0.7 LiMn0.9


LiFePO4 LiMnPO4
Fe0.8PO4 Fe0.7PO4 Fe0.5PO4 Fe0.3PO4 Fe0.1PO4

Fe(Mn)-O(1) 2.198(4) 2.202(3) 2.221(4) 2.234(4) 2.252(4) 2.262(4) 2.239(2)

Fe(Mn)-O(2) 2.097(4) 2.113(3) 2.114(4) 2.117(4) 2.122(4) 2.159(4) 2.11(1)

Fe(Mn)-O(3)×2 2.237(2) 2.252(2) 2.261(2) 2.272(2) 2.275(2) 2.271(2) 2.28(1)

Fe(Mn)-O(3) ×2 2.051(3) 2.075(2) 2.080(2) 2.099(2) 2.118(2) 2.116(2) 2.129(1)

Fe(Mn)-P 2.838(2) 2.846(1) 2.846(2) 2.853(2) 2.859(2) 2.853(2) 2.865(2)

P-O(1) 1.534(3) 1.530(3) 1.533(3) 1.543(3) 1.556(3) 1.534(3) 1.553(2)

P-O(2) 1.545(4) 1.551(3) 1.558(4) 1.561(4) 1.583(4) 1.556(4) 1.570(1)

P-O(3) ×2 1.563(3) 1.550(2) 1.551(2) 1.541(2) 1.539(3) 1.553(3) 1.548(1)

P-Li×2 2.658(1) 2.658(1) 2.658(1) 2.662(1) 2.664(1) 2.655(1) 2.656(2)

Li-O(1)×2 2.163(2) 2.170(2) 2.171(2) 2.178(3) 2.174(3) 2.206(3) 2.213(1)

Li-O(2)×2 2.093(2) 2.078(2) 2.080(2) 2.090(2) 2.088(2) 2.078(2) 2.115(1)

Li-O(3)×2 2.221(3) 2.178(2) 2.174(2) 2.170(3) 2.163(3) 2.174(3) 2.160(1)

O(2)-Fe(Mn)-O(3) 90.58(8) 90.03 89.95(7) 89.71(8) 89.89(8) 90.09(8) 89.57(0)

O(3)-Fe(Mn)-O(3) 120.8(1) 119.1(1) 118.71(13) 118.7(1) 118.4(2) 118.42(8) 118.00(1)

O(1)-P-O(2) 112.2(2) 113.6(2) 113.0(2) 112.5(2) 112.1(2) 112.9(2) 112.34(0)

O(2)-P-O(3) 107.8(1) 105.8(1) 105.7(1) 105.6(1) 105.3(1) 105.9(2) 106.4(0)

O(3)-P-O(3) 102.5(2) 103.5(2) 104.3(2) 104.6(2) 104.6(2) 104.5(2) 104.8(1)

134
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

In general, Fe (Mn) – O bond distances increase with increasing Mn content.

However, the O-Fe-O and O-P-O bond angles vary with changing Mn content, but

without any specific trend. Refinement of atomic position occupancies gives a

relatively large value for the Li in LiMnPO4 compared to the stoichiometric

compound. Therefore, taking into account the phase composition of the LiMnPO4

sample, we suggest the following decomposition reaction of LiMnPO4:

(1+x) LiMnPO4 → x/2 Mn2P2O7 + Li1+xMnPO4+x/2 (7.1)

The variations in the lattice parameters of the olivine phases are shown in Table 7.1.

As the Mn content increases, the lattice parameters of the unit cell expand along all

the axes; a, b, and c, inducing an increase in the volume of the unit cell. This is in good

agreement with the fact that the Mn2+ ion (r = 0.84 Å) has a larger radius than that of

the Fe2+ ion (r = 0.74 Å) in octahedral coordination.

Fig. 7.2 shows SEM images of the LiFePO4, LiMn0.5Fe0.5PO4, and LiMnPO4 samples.

In general, all three samples have similar micromorphology and similar grain sizes,

regardless of the composition of the compounds. The average grain size is in the range

of 2 – 5 μm. As the Mn content increases, the micromorphologies of the materials

become cotton wool-like. For the LiMnPO4 sample, this characteristic becomes

especially salient. Thermogravimetric analysis (TGA) was performed on all the

samples. Our as-prepared samples contain about 2 % carbon, which was induced by

the composition of the organic precursors.

135
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

(a)

(b)

136
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

(c)

Fig. 7.2 SEM images of LiMnxFe1-xPO4 sample powders:

A: LiFePO4 sample; B: LiMn0.5Fe0.5PO4 sample; C: LiMnPO4 sample.

The temperature dependencies of the molar magnetic susceptibilities of the LiFePO4

and LiMn0.5Fe0.5PO4 samples are shown in Fig. 7.3. The magnetic susceptibilities of

these two samples show a maximum at 50 ± 2 °K, demonstrating a paramagnetic –

antiferromagnetic transition with Curie-Weiss behaviour above the Néel temperature.

These two samples show similar dependencies of the magnetic susceptibilities on

temperature. A ferromagnetic transition exists at 225 °K for Mn doped samples,

which is caused by the ferromagnetic impurities Fe2P and FeP. In the triphylite

structure of LiFe(Mn)PO4 olivines, Fe2+ (Mn2+) ions occupy octahedral 4c sites.

Following the super and super-super exchange rule, the interactions between Fe(Mn)

137
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

– O – Fe(Mn) induce antiferromagnetism, as shown by the magnetic measurements.

However, there is no direct Fe(Mn) – O – Fe(Mn) bonding, only Fe(Mn) – O – P – O

– Fe(Mn) bonding. Therefore, magnetic interactions in LiFe(Mn)PO4 could be long

range through Fe(Mn) – O – P – O – Fe(Mn) triple exchange [195-196].

0.026

0.024

0.022
χ emu/mol LiFePO4

0.020

0.018

0.016

0.014

0.012

0.010

0.008
0 50 100 150 200 250 300
Temperature (K)

(a)

0.040

0.035

0.030
χ emu/mol LiFe0.5Mn0.5PO4

0.025

0.020

0.015

0.010

0.005
0 50 100 150 200 250 300 350
Tem perature (K)

(b)

Fig. 7.3 Temperature dependencies of the molar magnetic susceptibilities of

(a) LiFePO4 sample, (b) LiMn0.5Fe0.5PO4 sample.

138
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

200

180

160
Discharge capacity (mAh/g) 140

120 LiFePO4
LiMn0.1Fe0.9PO4
LiMn0.3Fe0.7PO4
100
LiMn0.5Fe0.5PO4
LiMn0.7Fe0.3PO4
80 LiMn0.9Fe0.1PO4

60

40

20

0
0 10 20 30 40 50
Cycle number

Fig. 7.4 Discharge capacity vs. cycle number for LiMnxFe1-xPO4 sample electrodes.

Current rate: C/8; Voltage range: 2.75 V – 4.30 V.

The electrochemical performance of all LiMnxFe1-xPO4 samples was tested via

constant current charge/discharge testing. Fig. 7.4 shows the specific capacities and

cyclabilities of all samples except for LiMnPO4 electrode. LiMnPO4 was found to

almost completely lack electrochemical activity. The cyclability measurements

exhibit a general trend that the specific capacity and cycle life decrease with

increasing Mn content. For the pure LiMnPO4 sample, the electrochemical reactivity

is very limited. Therefore, we concluded that the substitution of Mn for Fe in

LiMnxFe1-xPO4 cathode materials degrades the electrochemical performance.

139
Chapter 7 Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials
--------------------------------------------------------------------------------------------------------------------------

7.4 Conclusions

LiMnxFe1-xPO4 (x = 0. 0.1, 0.3, 0.5, 0.7, 0.9 and1.0) compounds were prepared via a

sol-gel synthetic method. Rietveld refinement on X-ray diffraction patterns has

identified a variety of impurities. LiMnxFe1-xPO4 materials show antiferromagnetic

characteristics. The electrochemical reactivity and cycle life degrade with increasing

Mn content for LiMnxFe1-xPO4 samples. Therefore, undoped LiFePO4 is the best

choice as a cathode material for Li-ion batteries.

140
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

Chapter 8

Electrochemical and magnetic

characteristics of LiFePO4 and

Li0.95Mg0.05FePO4 cathode materials

8.1 Introduction

Because of their high energy density lithium-ion batteries have become the dominant

power sources for portable electronic devices [2,197-199]. The current commercial

lithium-ion batteries use LiCoO2 as the cathode material, which limits their

applications to small batteries due to the high cost and toxicity of Co. In the past

decade, a number of new cathode materials, such as layer structured LiMnO2 [200,

201], LiMn2O4 spinels [202, 203], and LiMxNi1-xO2 compounds [204, 205], have been

developed as alternative cathode materials for lithium-ion batteries. However, none of

them are in popular industrial scale applications so far.

A group of lithium transition metal “polyanion” compounds, incorporating

polyanions XO43- (X = S, P, As and Mo) have been extensively investigated as new

cathode materials for lithium-ion batteries [206–210]. Among them, LiFePO4 olivines

have attracted particular interest because of the low cost and environmental

141
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

friendliness of iron. LiFePO4 has a lithium intercalation/de-intercalation potential of

3.4 V – 3.5 V vs. Li/Li+ due to the tuning effect of the large polyanion [PO4]3+.

LiFePO4 is an ideal cathode materials for large-scale lithium-ion batteries for electric

vehicles (EVs) and stationary power storage. LiFePO4 has an ordered olivine

structure (space group: Pnma), in which Li, Fe, and P atoms occupy octahedral 4a,

octahedral 4c, and tetrahedral 4c sites. The oxygen atoms are arranged in a slightly

distorted, hexagonal close-packed arrangement. The FeO6 octahedra share common

corners in the bc plane, and the LiO6 octahedra form an edge-shares chain in the

b-direction. The separation of the FeO6 octahedra by PO4 polyanions significantly

reduces the electrical conductivity of the material. This causes poor rate capacity and

low utilisation of Li in the LiFePO4 host structure. Extensive investigations have been

made to improve the performance of LiFePO4, including carbon coating, addition of

conductive copper/silver powders, dispersion of high surface area carbon black,

supervalence cation doping, and synthesis of nanosize grains [211-218].

The formation of a nano-network for electronic conduction in LiFePO4 and LiNiPO4

olivines by creating conductive iron phosphides has been reported as one efficient

approach to improve the electronic conductivity [219]. However, the electrochemical

performance of these olivines has not been described. Here we show the

electrochemical and magnetic properties of LiFePO4 and doped Li0.95Mg0.05FePO4

compounds.

142
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

8.2 Experimental

8.2.1 Materials synthesis

LiFePO4 and Li0.95Mg0.05FePO4 were prepared by a sol-gel preparation route.

Li(OH).H2O (99.9%, Aldrich), FeC2O4.2H2O (99%, Aldrich), NH4.H2PO4 (97%,

Aldrich), and Mg(C2H3O2)2 (99%, Aldrich) were used as reactants. The

stoichiometric reactants were dissolved in de-ionised water, in which polyacrylic acid

and active acid were added as complexing agents for the formation of the gel. The

solutions were heated and maintained at 85 °C under vigorous stirring until a viscous

gel was formed. The as-formed gel was heated to 500 °C to decompose the organics

under flowing argon gas. The decomposed precursors were further sintered at 700 °C,

800 °C, and 850 °C, respectively, under a flowing gas mixture (10% H2 in Ar). A

slightly reducing atmosphere was employed during the sintering process in order to

prevent the oxidation of Fe2+ cations. Six samples were prepared, which are shown in

Table 8.1.

X-ray analysis and SEM observation. X-ray diffraction was performed on the

prepared lithium iron phosphates to determine the phase purity using Cu Kα radiation

(MO3xHF22, MacScience, Japan). The morphology of lithium iron phosphate

powders was studied using a scanning electron microscope (JEOL JEM-3000).

143
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

Table 8.1 Sintering temperature, label, and average crystal size of lithium iron
phosphates

Sintering Average crystal


Sample Label
Temperature( °C) Size (μm)

LiFePO4 700 °C LFP-700 0.6

LiFePO4 800 °C LFP-800 4.0

LiFePO4 850 °C LFP-850 8.0

Li0.95Mg0.05FePO4 700 °C LFPMg-700 0.8

Li0.95Mg0.05FePO4 800 °C LFPMg-800 4.5

Li0.95Mg0.05FePO4 850 °C LFPMg-850 10

8.2.2 Magnetic measurements

The magnetic properties of LiFePO4 and Li0.95Mg0.05PO4 samples sintered at 700 °C

and 850 °C were studied using a Quantum Design MPMS XL SQUID magnetometer.

The dc susceptibility was measured in a magnetic field of 5000 Oe from 5 K to 300 K.

144
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

8.2.3 Electrochemical testing.

Electrochemical characterisation was performed by assembling CR2032 coin cells for

galvanostatic charge/discharge. The electrodes were made by dispersing 80 wt%

active material, 10 wt% carbon black, and 10 wt% polyvinylidene fluoride (PVDF) in

n-methyl pyrrolidone (NMP) to form a slurry. The slurry was then coated onto Al

foilto form the electrodes. The coated electrodes were dried in a vacuum oven and

then pressed at 1200 Kg/cm2. The coin cells were assembled in an argon filled glove

box (Mbraun, Unilab, Germany) with lithium foil as the counter electrode. The

electrolyte was 1 M LiPF6 in a 1: 1 mixture of ethylene carbonate (EC) and dimethyl

carbonate (DMC).

8.3 Results and discussion

8.3.1 XRD and SEM analysis

X-ray diffraction was performed on the LiFePO4 and Li0.95Mg0.05PO4 powders. Fig.

8.1 shows the XRD patterns of LiFePO4 and Li0.95Mg0.05PO4 samples sintered at

different temperatures. Phase pure LiFePO4 and Li0.95Mg0.05PO4 were obtained after

sintering at both 700 °C, and 800 °C.

145
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

o
(a) Sintered at 700 C
o
(b) Sintered at 800 C
o
(c) Sintered at 850 C

Intensity / a.u. Fe2P


FeP4
(c)

(b)

(a)

20 25 30 35 40 45
2θ / degree

(a)

o
(a) Sintered at 700 C
o
(b) Sintered at 800 C
o
(c) Sintered at 850 C

Fe2P
Intensity (a.u.)

(c)

(b)

(a)

20 25 30 35 40 45
2θ / degree

(b)

Fig. 8.1 X-ray diffraction patterns of (a) LiFePO4 and (b) Li0.95Mg0.05FePO4.

146
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

All diffraction lines were indexed to an orthorhombic crystal structure (triphylite,

space group: Pnma). For LiFePO4 sintered at 850 °C, the main phase is still triphylite

LiFePO4, but there are substantial impurity phases such as FeP4 and Fe2P, as indicated

in Fig 1(a). However, for Li0.95Mg0.05PO4 sintered at 850 °C, the main phase is not

triphylite. There are numerous impurity phases that cannot be totally identified. The

strongest diffraction peak was identified as Fe2P. The x-ray diffraction analysis

confirmed a previous report [219] that iron phospides are formed in LiFePO4 and

Li0.95Mg0.05PO4 compounds when sintered at high temperature. The formation of iron

phosphides occurs more in Li0.95Mg0.05PO4 than in LiFePO4. Thermogravimetry

analysis was performed on the LiFePO4 and Li0.95Mg0.05PO4 samples. All the samples

contained 2 – 3 wt% carbon, which was induced by the sol-gel synthesis process.

Fig. 8.2 (a)–(c) shows SEM images of Li0.95Mg0.05PO4 powders sintered at 700 °C,

800 °C, and 850 °C, respectively. The sample sintered at 700 °C has a small grain size

of about 0.5 – 1 μm. When sintered at 800 °C, the grain size increased to 3 – 5 μm,

while the 850 °C sintered sample exhibits the biggest grain size of 5 – 10 μm. The

crystal size of the lithium iron phosphates was significantly influenced by the

sintering temperature. Fig. 8.3 shows the EDS elemental mapping of Li0.95Mg0.05PO4

powders. Fig. 8.3 A is the secondary electron image. A very uniform element

distribution was observed for the main elements Fe, P, and O. For reasons of the

distributions of the trace dopant Mg and carbon are not shown. Trace dopant Mg2+

ions are also observed to be homogeneously distributed in the crystal structures. This

should be attributed to the chemical synthesis process. The ingredient Li+, Fe2+, Mg2+,

147
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

and PO43- ions were dissolved in water and mixed on the atomic scale in solution.

Following the subsequent gel formation and sintering, a uniform element distribution

in the as-prepared materials is expected.

(a)

(b)

148
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

(c)

Fig. 8.2 SEM images of Li0.95Mg0.05FePO4 samples sintered at

(A) 700 °C (B) 800 °C and (C) 850 °C.

149
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

Fig. 8.3 EDS elemental mapping by SEM. (A) Secondary electron image,

(B) Element Fe mapping, (C) Element P mapping, (D) Element O mapping.

8.3.2 Magnetic properties of LiFePO4 and Li0.95Mg0.05PO4

The temperature dependences of the molar magnetic susceptibilities of LiFePO4

samples sintered at 700 °C and 850 °C are shown in Fig. 8.4 (a) and (b) respectively.

The reciprocal susceptibilities are shown in the inset. The magnetic susceptibility of

LiFePO4 shows a maximum at 50 ± 2 °K, demonstrating a

paramagnetic-antiferromagnetic phase transition. This critical temperature is also

150
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

called the Néel temperature (θN). A typical Curie-Weiss law behaviour was followed

above the Néel temperature. The two samples show similar dependences of the

magnetic susceptibilities on temperature, but the sample sintered at 850 °C shows

lower molar susceptibility. The magnetic moments in LiFePO4 are attributed to Fe2+

in the LiFePO4 crystal structure. As detected by x-ray diffraction analysis, a small

amount of iron phosphides were formed in LiFePO4 when sintering was at 850 °C, but

these have no contribution to the antiferromagnetic behaviour. Fig. 8.5 (a) and (b)

shows the molar magnetic susceptibilities of Li0.95Mg0.05FePO4 powders sintered at

700 °C and 850 °C, respectively. The Li0.95Mg0.05FePO4 sample sintered at 700 °C

shows the same magnetic behaviour as LiFePO4, with a typical anti-ferromagnetie

behavior below the Néel temperature. However, the Li0.95Mg0.05FePO4 sample

sintered at 850 °C shows ferromagnetic behaviour, which is totally different from its

counterpart sintered at 700 °C. The main phases for Li0.95Mg0.05FePO4 sintered at 850

°C are iron phosphides. Most of these phosphides (Fe2P, Fe3P, FeP4, and Fe75P15C4)

are ferromagnetic [220]. Therefore, the measured ferromagnetic behaviour also

confirmed the formation of substantial iron phosphides when sintering at high

temperature (850 °C).

151
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

0.026
110
0.024 100

1/χ emu mol Fe


90
80
0.022

-1
70
60
50
0.020 40
χ emu/mol Fe

30
20
0.018 10
0
0 50 100 150 200 250 300
Temperature (K)
0.016

0.014

0.012

0.010

0.008
0 50 100 150 200 250 300
Temperature (K)

(a)

0.026

0.024 120
1/χ emu mol Fe

100
0.022
-1

80
χ emu/ mol Fe

0.020 60

0.018 40

0 50 100 150 200 250 300

Temperature (K)
0.016

0.014

0.012

0.010

0.008
0 50 100 150 200 250 300
Temperature (K)

(b)

Fig. 8.4 Temperature dependencies of molar magnetic susceptibilities of LiFePO4

samples (a) LFP-700 and (b) LFP-850. The insets are the reciprocal magnetic

susceptibilities.

152
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

0.026
120

0.024

1/χ emu mol Fe


100

0.022

-1
80
χ emu/mol Fe

0.020 60

0.018 40

0 50 100 150 200 250 300

0.016 Temperature (K)

0.014

0.012

0.010

0.008
0 50 100 150 200 250 300
Temperature (K)

(a)

0.16

0.14

0.12
χ emu/mol Fe

16
0.10
1/χ emu mol Fe

14

12
0.08
-1

10

0.06 8

6
0 50 100 150 200 250 300

0.04 Temperature (K)

0.02
0 50 100 150 200 250 300
Temperature (K)

(b)

Fig. 8.5 Temperature dependencies of molar magnetic susceptibilities

of Li0.95Mg0.05FePO4samples: (a) LFPMg-700 and (b) LFPMg-850.

The insets are the reciprocal magnetic susceptibilities.

153
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

The effective magnetic moment was calculated using the paramagnetic formula μ =

(8χT)-1/2 [221], which is shown in Fig. 8.6. The measured μ values for the three

anti-ferromagnetic samples are in the range of 4.6 – 4.9 μB at room temperature. The

calculated μ for the ferromagnetic Mg doped Li0.95Mg0.05FePO4 sample is 12.61 μB,

which is much higher than that of the other three samples. The Curie temperature θC

was obtained by linear fitting of the reciprocal magnetic susceptibility above the Néel

temperature. Table 8.2 shows the magnetic parameters of the four samples. All three

anti-ferromagnetic samples have the same Néel temperature. The Mg doped sample

sintered at 700 °C shows a different Curie temperature of θC = - 76 K. The

ferromagnetic sample has a positive θC value of 110.5 K.

In the triphylite structure of LiFePO4, Fe2+ ions occupy octahedral 4c sites with

coordinates x = 0.28, y = 0.25, and z = 0.96. Due to the influence of the octahedral

crystal field, the five 3d orbitals of the Fe2+ ion split into three t2g and two eg

configurations. There are six electrons in 3d orbitals for the Fe2+ ion. Among them,

four electrons are on the three t2g orbitals (dxy, dxz, and dyz), in which three electrons

are spin up and the fourth electron spin down. The magnetic moments for the paired

two electrons in one of the t2g orbitals cancel each other. Therefore, there are a net two

spin-up electrons in t2g orbitals. Two electrons occupy two eg orbitals (dx2-y2 and dz2),

in which one electron is in each orbital with spin-up configuration. The Fe2+ ion

presents net four spin-up electrons (high spin), with an expected magnetic moment of

4μB on the basis of the spin-only value. However, the measured magnetic moment is

in the range of 4.7 – 4.9 μB, which is higher than the theoretical value. Since the

154
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

angular momentum for the Fe2+ ion is not equal to 0, the Fe2+ ion could have a strong

spin-orbit coupling, inducing anisotropy and a non-quenched angular momentum.

The orbital momentum could contribute to the measured magnetic moment. The

magnetic structure of LiFePO4 has been identified to be along the b-axis,

corresponding to a spin sequence (+--+) with moments along [010]. Following the

super and super-super exchange rules, the interactions between Fe-O-Fe induce

antiferromagnetism, as exhibited by the magnetic measurements. However, in the

orthorhombic phase, there is no Fe-O-Fe bonding, only Fe-O-P-O-Fe bonding. So, the

magnetic interaction in LiFePO4 could be long-range through Fe-O-P-O-Fe triple

exchange [222–225].

4
o
LiFePO4 700 C sintered
μ / μB

o
LiFePO4 850 C sintered
3 o
Li0.95Mg0.05FePO4 700 C sintered

0
0 50 100 150 200 250 300
Temperature (K)

Fig. 8.6 The effective magnetic moment per mole of iron of LiFePO4 and

Li0.95Mg0.05FePO4 samples.

155
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

Table 8.2 Magnetic parameters of LiFePO4 and doped Li0.95Mg0.05FePO4 samples.

Effective
Magnetic Néel Curie
Properties Magnetic
Temperature Temperature
Moment μeff (μB)
TN (K) θC (K)
at 25°C
Sample

LiFePO4
50 ± 2 -96.2 4.77
700 °C sintered

LiFePO4
50 ± 2 -96.2 4.55
850 °C sintered

Li0.95Mg0.05FePO4
50 ± 2 -76.0 4.57
700 °C sintered

Li0.95Mg0.05FePO4
~ +110.5 12.61
850 °C sintered

156
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

8.3.3 Electrochemical performance of LiFePO4 and Li0.95Mg0.05PO4 cathode

materials

Electrochemical properties of LiFePO4 and Li0.95Mg0.05PO4 cathodes were

characterised via cyclic voltammetry (CV) and constant current charge/discharge

testing. Fig. 8.7 shows typical voltammograms of LiFePO4 and Li0.95Mg0.05PO4

electrodes. In Fig. 8.7(a), the reduction and oxidation peak positions for the LFP-700

and LFP-800 samples are the same, at 3.30 V and 3.51 V vs Li/Li+ respectively. There

is a shift of the redox potentials for the LFP-850 sample to 3.12 V and 3.60 V vs

Li/Li+. The shift of the redox potential could be related to the formation of iron

phosphite impurities in the LFP-850 sample. Fig. 8.7 (b) shows the CV curves of

Li0.95Mg0.05PO4 electrodes. The redox potentials for doped samples are different from

those of the undoped one, indicating that the Mg doping has an influence on the redox

potential of LiFePO4 cathode materials. Cyclic voltammetry measurements on the

LFPMg-850 sample failed, indicating that the sample electrode has no

electrochemical reactivity due to the change of the main phase into iron phosphides as

a result of sintering at high temperature.

157
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

300

250

200

150

100
Current i (μA)
50

-50 LFP-700 sample


LFP-800 sample
-100 LFP-850 sample
-150

-200

-250

-300
2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8
Potential E vs. Li/Li+ (V)

(a)

300

250

200

150

100
Current i (μA)

50

-50 LFPMg-700 sample


LFPMg-800 sample
-100

-150

-200

-250

-300
2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8
+
Potential E vs. Li/Li (V)

(b)

Fig. 8.7 shows typical voltammograms of LiFePO4 and Li0.95Mg0.05PO4 electrodes.

158
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

The capacities and cyclabilities of the LiFePO4 and Li0.95Mg0.05FePO4 electrodes

were tested by galvanostatic charge/discharge testing at a C/8 rate. Fig. 8.8 (a) and (b)

shows the voltage profiles in the first cycle for the LFP-700 and LFPMg-700 sample

electrodes, respectively, the cells were charged and discharged in the voltage range of

2.75 V – 4.2 V vs. Li/Li+. The profiles exhibit a flat charge and discharge plateau at

3.45 – 3.55 V, which matches the oxidation and reduction peaks in the CV curves.

The LiFePO4 electrode delivered a specific discharge capacity of 160 mAh/g in the

first cycle, whereas the Li0.95Mg 0.05FePO4 cathode shows a lower discharge capacity

of 150 mAh/g in the first cycle, which is due to the partial substitution of Li+ by

non-active Mg2+. The cyclabilities of all the sample electrodes are shown in Fig. 8.9.

5.0

4.5

4.0
Voltage (V)

3.5

3.0

2.5

2.0
0 50 100 150 200
Capacity (mAh/g)

(a)

159
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

5.0

4.5

4.0
Voltage (V)

3.5

3.0

2.5

2.0
0 20 40 60 80 100 120 140 160 180
Capacity (mAh/g)

(b)

Fig. 8.8 Charge/discharge curves in the first cycle for (a) LiFePO4 and (b)

Li0.95Mg0.05FePO4 sample electrodes. Current rate: C/8.

200
Discharge capacity (mAh/g)

150

100 LFP-700
LFP-800
LFP-850
LFPMg-700
LFPMg-800
50

0
0 10 20 30 40 50
Cycle number

Fig. 8.9 Discharge capacity vs. cycle number for LiFePO4 and

Li0.95Mg0.05 FePO4 sample electrodes. Current rate: C/8.

160
Chapter 8 Electrochemical and magnetic characteristics of LiFePO4 and Li0.95Mg0.05FePO4 cathode

materials

-----------------------------------------------------------------------------------------------------------

In general, the as-prepared LiFePO4 and Li0.95Mg0.05FePO4 cathode materials sintered

at different temperatures demonstrated a stable capacity in the range of 150

mAh/g–160 mAh/g at the C/8 rate. The rate of degradation with cycling for the

LiFePO4 and Li0.95Mg0.05FePO4 sample electrodes increases with increasing sintering

temperature, which is due to the influence of the increased crystal size. The previously

proposed “Radial Model” and “Mosaic Model” for LiFePO4 can exactly explain the

observed phenomena [226, 227]. The process of lithium insertion and extraction in

LiFePO4 is a two-phase co-existing process. Therefore, the crystal size has a critical

influence on the transport of lithium ions and electrons into and out of the individual

crystals. The larger the crystal size, the greater the inefficiency of the full conversion

of LiFePO4 to FePO4 and back again, inducing a loss of capacity with cycling.

8.4 Conclusions

The sintering temperature has a significant influence on the crystal size of LiFePO4

and Li0.95Mg0.05PO4 compounds. Typical antiferromagnetic behaviour was

demonstrated for LiFePO4 and Li0.95Mg0.05PO4 samples with a Néel temperature TN =

50 ± 2 K. This antiferromagnetism could be induced by long range Fe-O-P-O-Fe

triple exchange, due to the lack of direct Fe-O-Fe interactions. LiFePO4 and

Li0.95Mg0.05PO4 compounds are electrochemically active with a specific capacity

between 150 mAh/g and 160 mAh/g.

161
Chapter 9 Conclusions and suggestions for future study
--------------------------------------------------------------------------------------------------------------------------

Chapter 9

Conclusions and suggestions

for future study

9.1 General conclusions

Studies on the performance of various anode materials and cathode materials have

been presented in the previous chapters. The outcomes are summarized as follows.

9.1.1 Silicon based composite anode materials

This part of the work was intended to utilise the high capacity of Si and the stable

cyclability of mesocarbon microbeads (MCMB). Si-MCMB composites were

prepared via high energy ball-milling. A reversible capacity of 1066 mAh/g and fairly

good cyclability was been achieved for nanocrystalline Si-MCMB composite anode

materials. A reaction model was proposed for the lithiation and de-lithiation process

in the Si-MCMB electrodes in lithium-ion cells. Nanostructured Si-C composite

anode materials were produced by dispersing nanocrystalline Si powders in carbon

aerogel, followed by heat treatment to induce carbonization in argon atmosphere.

Cyclic voltammetry measurements showed two pairs of redox peaks associated with

the lithiation and de-lithiation processes of LixSi alloys. A reversible capacity of 1450

mAh/g for Si-C composite electrodes was achieved. The good cyclability should be

162
Chapter 9 Conclusions and suggestions for future study
--------------------------------------------------------------------------------------------------------------------------

attributed to the usage of nanosize Si powders and the buffering effect provided by the

amorphous carbon matrix.

9.1.2 Carbon nanotubes and tin oxide – carbon nanotubes anode materials

Vertically aligned multiwalled carbon nanotubes were synthesised by chemical

vapour deposition using iron carbonyl as the catalyst. A reversible lithium storage

capacity of 950 mAh/g has been achieved for the as-prepared vertically aligned

carbon nanotubes. SnO2-CNT composite anode materials were prepared via chemical

treatment followed by sintering at high temperature. The solution-based chemical

process enabled Sn2+ ions to penetrate into the inner cavity of the carbon nanotubes.

SnO2 nanocrystals were observed to be homogeneously distributed on the carbon

nanotube matrix by TEM analysis. The SnO2-CNT composite electrodes exhibited

stable cyclability, with a lithium storage capacity of 410 mAh/g after fifty cycles.

9.1.3 Novel iron phosphide anode materials

Nanosize FeP4 compound was synthesised via the hydrothermal method. The

as-prepared products contain a majority of monoclinic FeP4 phase with some small

amount of impurities. Cyclic voltammetry measurements demonstrated the reversible

reactivity of FeP4 anodes towards lithium insertion and extraction. The FeP4 anode

exhibited a stable lithium storage capacity of about 700 mAh/g.

163
Chapter 9 Conclusions and suggestions for future study
--------------------------------------------------------------------------------------------------------------------------

9.1.4 Lithium iron phosphate and manganese doped lithium iron phosphate

cathode materials

A series of LiMnxFe1-xPO4 (x = 0, 0.1, 0.3, 0.5, 0.7, 0.9 and 1) compounds were

synthesised by a sol-gel preparation route. X-ray diffraction and Rietveld structure

refinement revealed that the lattice parameters (a, b, and c) of orthorhombic

LiMnxFe1-xPO4 increase with increasing Mn content. Magnetic susceptibility

measurements showed antiferromagnetic behaviour for the LiMnxFe1-xPO4 samples.

The electrochemical properties of LiMnxFe1-xPO4 as cathodes in lithium-ion cells

were tested via cyclic voltammetry and galvanostatic charge/discharge

measurements. The electrochemical performance of LiMnxFe1-xPO4 degrades with

increasing Mn content.

9.1.5 Electrochemical and magnetic properties of magnesium doped lithium iron

phosphate cathode materials

LiFePO4 and Li0.95Mg0.05PO4 compounds were synthesised via the sol-gel route

followed by sintering at high temperature. Iron phosphides, which are electronic

conductors, were formed when sintering took place at 850 °C. Magnetic susceptibility

measurements on the powder samples show antiferromagnetic behaviour with TN =

50 ± 2 K for LiFePO4 and Li0.95Mg0.05PO4 sintered at temperatures below 850 °C. The

measured effective magnetic moments are in the range of 4.6 – 4.9 μB at room

temperature. The LiFePO4 and Li0.95Mg0.05PO4 cathodes show a stable

electrochemical capacity in the range of 150 – 160 mAh/g on cycling. The cyclability

164
Chapter 9 Conclusions and suggestions for future study
--------------------------------------------------------------------------------------------------------------------------

deteriorates with increasing sample sintering temperature due to the increased crystal

size.

9.2. Recommendations for future studies

As an important energy storage system, lithium ion batteries will continue to attract

worldwide investigation to improve their performance. Anode materials and cathode

materials play a critical role in determining the overall properties of lithium-ion

batteries. Therefore, exploration of new anode materials and cathode materials will

certainly provide synergy for the development of new generation lithium-ion batteries

to meet the challenging requirements of large-scale energy storage, electric vehicles,

and portable electronic devices.

Alloy based anode materials are a new direction with the potential to double or triple

the current carbon based anode materials used in lithium-ion batteries. In particularly,

tin and silicon or their alloy have high theoretical lithium storage capacities.

However, the problems related to volume expansion and contraction during the

lithiation and de-lithiation processes need to be overcome before any practical usage

is possible. Nanosize metal oxides are also candidate anode materials which can

provide high capacity.

Without a doubt, lithium iron phosphate is a very promising cathode material for

lithium ion batteries, particularly for large-scale lithium ion batteries, which are safe

in operation and cheap in terms of cost. However, low electronic conductivity induces

165
Chapter 9 Conclusions and suggestions for future study
--------------------------------------------------------------------------------------------------------------------------

low rate capacity and short cycle life. Carbon coating, metal ion doping, and the

synthesis of nanocrystalline materials could effectively combat this problem.

166
References
--------------------------------------------------------------------------------------------------------------------------

References
[1] M.Winter and R.J. Brodd, Chem. Rev. 104 (2004) 4245.

[2] J.-M. Tarascon and M.Armand, Nature 414 (2001) 359.

[1] M.S. Whittingham, J. Electrochem. Soc. 122 (1975) 526.

[2] N. Watanabe, M. Fukuba, U.S. Patent 3,536,532, 1970.

[3] A. Robinson, Science, 184 (1974) 554.

[4] Sanyo, Lithium Battery Calculator, Model CS-8176L.

[5] E.S. Takeuchi, W.C. Thiebolt, J. Electrochem. Soc. 135 (1988) 2691.

[6] A. Crespi, C. Schmidt, J. Norton, K. Chen, P.J. Skarstad, J. Electrochem. Soc. 148

1 (2001) A 30.

[7] K.M. Abraham, D.M. Pasquariello, and F.J. Martin, J. Electrochem. Soc. 133

(1986) 661.

[8] J.J. Auborn, Y.L. Barberio, K.J. Hanson, D.M. Schleich, and M.J. Martin, J.

Electrochem. Soc.134 (1987) 580.

[9] B.M.L. Rao, R.W. Francis, H.A. Christopher, J. Electrochem. Soc. 124 (1977)

1490 – 1492.

[10] P. Sanchez, C. Belin, C. Crepy and A. De Guibert, J. Appl. Electrochem. 19

(1989) 421.

[11] A.S. Baranski, W.R. Fawset, J. Electrochem. Soc. 129 (1982) 901.

[12] J.O. Besenhard, P. Komenda, A. Paxines, E. Wudy and M. Josowices, Solid State

Ionics 18&19 (1986) 823.

[13] A. Anani, S.C. Baker, R.A. Huggins, J. Electrochem. Soc. 134 (1987) 3098.

[14] A. Anani, S.C. Baker, R.A. Higgins, J. Electrochem. Soc.135 (1988) 2103.

167
References
--------------------------------------------------------------------------------------------------------------------------

[15] M. Mohri, N. Yanagisawa, Y. Tajima, H. Tanaka, T. Mitate, S. Nakajima, M.

Yoshida, Y. Yoshimoto, T. Suzuki, H. Wada, J. Power Sources 26 (1989) 545 –

551.

[16] R.Yazami, P. Touzain, J. Power Sources 9 (1983) 365.

[17] D.W. Murphy, P.A. Christian, Science 205 (1979) 651 – 656.

[18] K. Mizushima, P.C. Jones, P.J. Wiseman and J.B. Goodenough, Mater. Res. Bull.

17 (1980) 783.

[19] J.L. Godman, R.M. Mank, J.H. Young, V.R. Koch, J. Electrochem. Soc., 127

(1980) 1461.

[20] G. Feuillade, P. Perche, J. Appl. Electrochem. 5 (1975) 63-69.

[21] M. Armand, Solid State Ionics 69 (1994) 309.

[22] T. Nagaura, K. Tozawa, Prog. Batteries Solar Cells 9 (1990) 209.

[23] D.E.Fenton, J.M.Parker, P.V.Wright, Polymer 14 (1973) 589.

[24]G.B.Apetecchi, F.Croce, B.Scrosatti, J. Power Sources 66 (1997) 77.

[25] M. Armand, in: Materials for Advanced Batteries, eds. D. W. Murphy, J.

Broadhead and B.C.H. Steele (Plenum Press, New York, 1980) p. 145.

[26] K.O. Hever, J. Electrochem. Soc. 115 (1968) 826.

[27] M. Lazzari and B. Scrosati, J. Electrochem. Soc.127 (1980) 773.

[28] S. Morzilli, B. Scrosati and F. Sgarlata, Electrochim. Acta 30 (1985) 1271.

[29] M.Wakihara, Mater. Sci. Eng. 33 (2001) 109.

[30] K. Mizushima, P.C. Jones, P.J. Wiseman and J.B. Goodenough, Mater. Res. Bull.

15 (1980) 783.

[31] T. Nagaura, K. Tozawa, Prog. Batteries Solar Cells 9 (1990) 209.

[32] T. Ohzuku and A. Ueda, J. Electrochem. Soc.141 (1994) 2972.

168
References
--------------------------------------------------------------------------------------------------------------------------

[33] C. Delmas, Mater. Sci. Eng. B3 (1989) 97.

[34] B. Garcia, J. Farcy, J.P. Pereira-Ramos, J. Perichon and N. Baffier, J. Power

Sources 54 (1995) 373.

[35] R.Yazami, N. Lebrun, M. Bonneau and M. Molteni, J. Power Sources 54 (1995)

389.

[36] J. N. Reimers, E. Rossen, C.D. Jones and J.R. Dahn, Solid State Ionics 61 (1993)

335.

[37] M.G.S.R. Thomas, P.G. Bruce and J.B. Goodenough, Solid State Ionics 17

(1985) 13.

[38] Y. Ma, M.M. Doeff, S.J. Visco and L.C. De Jonghe, J. Electrochem. Soc. 140

(1993) 2726.

[39] C. Saadoune, I. Delmas, Solid State Ionics 53 – 56 (1992) 370.

[40] R. Alcántara, P. Lavela, J.L. Tirado, R. Stoyanova and E. Zhecheva, J.

Electrochem. Soc. 145 (1998) 730.

[41] T. Ohzuku and A. Ueda, J. Electrochem. Soc. 144 (1997) 2780.

[42] H.T. Huang, G.V. Subba Rao, B.V.R. Chowdari, J. Power Sources, 81-82 (1999)

690.

[43] J. Cho, G. Kim, Electrochem. Solid State Lett. 2 (1999) 253.

[44] J. Cho, J.G. Lee, B. Kim and B. Park, Chem. Mater. 15 (2003) 3190.

[45] L. Liu, Z. Wang, H. Li, L. Chen, and X. Huang, Solid State Ionics 152 (2002)

341.

[46] J. Cho, Y.J. Kim, T.J. Kim and B. Park, Angew. Chem. Int. Ed. 40 (2001) 3367.

[47] K.W. Beard, W.A. Depalma and J.P. Buckley, Proc. 1992 IEEE 35th Intern.

Power Sources Symp., IEEE, NJ. USA, 1992, P. 201.

169
References
--------------------------------------------------------------------------------------------------------------------------

[48] T. Ohzuku, A. Ueda and M. Nagayama, J. Electrochem. Soc. 140 (1993) 1862.

[49] R. Koksbang, I.I. Olsen and D. Shackle, Solid State Ionics 69 (1994) 320.

[50] M. Broussely, F. Perton, P. Biensan, J.M. Bodet, J. Labat, A. Lecerf, C. Delmas,

A. Rougier, J.P. Peres, J. Power Sources 54 (1995) 109.

[51] G. X. Wang, S. Zhong, D.H. Bradhurst, S.X. Dou and H.K. Liu, J. Power Sources

76 (1998) 141.

[52] B. Garcia, J. Farcy, J.P. Pereira-Ramos, J. Perichon and N. Baffier, J. Power

Sources 54 (1995) 373.

[53] W. Li, J. N. Reimers, J.R. Dahn, Solid State Ionics 67 (1993) 123.

[54] R.V. Moshtev, P. Zlatilova, V. Manev and A. Sato, J. Power Sources 54 (1995)

329.

[55] G.X.Wang, J. Horvat, D.H. Bradhurst, H.K. Liu and S.X. Dou, J. Power Sources

85 (2000) 279.

[56] G.X.Wang, S. Zhong, D.H. Bradhurst, S.X. Dou and H.K. Liu, Solid State Ionics

116 (1999) 271.

[57] A.R Armstrong and P.G. Bruce, Nature 381 (1996) 499.

[58] F. Capitaine, P. Gravereau, C. Delmas, Solid State Ionics 89 (1996) 197.

[59] L. Croguennec, P. Deniard and R. Brec, J. Electrochem. Soc. 144 (1997) 3323.

[60] G. Vitins and K. West, J. Electrochem. Soc. 144 (1997) 2587.

[61] I. Koetschon, M.N. Richard, J.R. Dahn, J.B. Sonpart and J.C. Rousche, J.

Electrochem. Soc. 9 (1995) 2906.

[62] R. J. Gummow and M.M. Thackeray, J. Electrochem. Soc. 141 (1994) 1178.

[63] Y-II Jang, B.Y. Huang, Y.-M. Chiang, and D.R. Sadoway, Electrochem.

Solid-state Lett.1 (1998) 13.

170
References
--------------------------------------------------------------------------------------------------------------------------

[64] A.I.Landa, C.-C. Chang, P.N. Kumta, L. Vitos, and I.A. Abrikosov, Solid State

Ionics 149 (2002) 209.

[65] A. R. Armstrong, R. Gitzendanner, A.D. Robertson, P.G. Bruce, Chem.

Commun. (1998) 1833.

[66] A.R. Armstrong, A.D. Robertson, R. Gitzendanner, and P.G. Bruce, J. Solid

State Chemistry 145 (1999) 549.

[67] E. Rossen, C.D.W.Jones, and J.R. Dahn, Solid State Ionics 57 (1992) 311.

[68] M.E. Spahr, P. Novák, B. Schnyder, O. Haas, R.J. Nesper, J. Electrochem. Soc.

145 (1998) 1113.

[69] Z. Lu, D.D. MacNeil, J.R. Dahn, Electrochem. Solid State Lett. 4 (2001) A191.

[70] T. Ohzuku, Y. Makimura, Chem. Lett. (2001) 744.

[71] L.Q. Zhang, H. Noguchi, M. Yoshio, J. Power Sources 110 (2002) 57.

[72] C. Storey, I. Kargina, Y. Grincourt, I.J. Davidson, Y.C. Yoo, D.Y. Seung, J.

Power Sources, 97 -98 (2001) 711.

[73] M. Balasubramanian, J. McBreen, I.J. Davidson, P.S. Whitfield, I. Kargina, J.

Electrochem. Soc. 149(2002) A176.

[74] Y. Grincourt, C. Storey, I.J. Davidson, J. Power Sources 97-98 (2001) 711.

[75] B. Ammundsen, J. Paulsen, I. Davidson, R.-S. Liu, C.-H. Shen, J.-M. Chen,

L.-Y. Lee, J.-F. Lee, J. Electrochem. Soc. 149 (2002) A431.

[76] S.J. Hwang, H.S. Park, J.H. Choy, Solid State Ionics 151 (2002) 275.

[77] G.X. Wang, Z.P. Guo, X.Q. Yang, J. McBreen, H.K. Liu and S.X. Dou, Solid

State Ionics 167 (2004) 183.

[78] P. Arora, B.N. Popov and R.E. White, J. Electrochem. Soc.145 (1998) 807.

171
References
--------------------------------------------------------------------------------------------------------------------------

[79] G.X.Wang, D.H. Bradhurst, H.K. Liu, S.X. Dou, Solid State Ionics 120 (1999)

95.

[80] M.M. Thackery, J. Electrochem. Soc. 142 (1995) 2558.

[81] R. Bittihn, R. Herr and D. Hoge, J. Power Sources 43-44 (1993) 223.

[82] M.Wakihara, Li Guoha, H. Ikuta, T. Uchida, Solid State Ionics 86-88 (1996) 907.

[83] A.D. Robertson, S.H. Lu, W.F. Averill and W.F. Howard, Jr., J. Electrochem.

Soc.144 (1997) 3500.

[84] Li Guohua, H. Ikuta, T. Uchida and M. Wakihara, J. Electrochem. Soc.143

(1996)178.

[85] M. Pechini, U.S. Pat. 330697, 1967.

[86] W. Liu, G.C. Farrington, J. Electrochem. Soc. 143 (1996) 897.

[87] K. Nomura, S. Tanase, Solid State Ionics 98 (1997) 229.

[88] C. J. Howard and S. J. Kennedy, Materials Forum 18 (1994) 155.

[89] R.J. Hill and C.J. Howard, Australian Atomic Energy Commission Report,

AAEC/M112.

[90] E.H. Kisi, Materials Forum 18 (1994) 135.

[91] Y. Takeda, K. Nakahara, M. Nishijima, N. Imanashi, O. Yamamoto, M. Takano,

and R. Kanno, Mater. Res. Bull. 29 (1994) 659.

[92] A.K. Padhi, K.S. Nanjundaswamy, J.B. Goodenough, J. Electrochem. Soc.144

(1997) 1188.

[93] A.K. Padhi, K.S. Nanjundaswamy, C. Masquelier, S. Okada, J.B. Goodenough,

J. Electrochem. Soc.144 (1997) 1609.

[94] Y.N. Song, P.Y. Zavalij, M. Suzuki, M.S. Whittingham, Inorg. Chem. 41 (2002)

5778 – 5786.

172
References
--------------------------------------------------------------------------------------------------------------------------

[95] Y.N. Xu, S.Y. Chung, J.T. Bloking, Y.M. Chiang, W.Y. Ching, Electrochemical

and Solid-state Letters 7 (2004) A131.

[96] M.S. Islam, D.J. Driscoll, C.A.J. Fisher, P.R. Slater, Chem. Mater. 17 (2005)

5085.

[97] M.S.Whittingham, Chem. Rev. 104 (2004) 4271.

[98] S.Y. Chung, J.T. Bloking, Y.M. Chiang, Nat. Mater. 1 (2002) 123.

[99] J. Molenda, A.Stoklosa, T. Bak, Solid State Ionics 36 (1989) 53.

[100] G.Rousse, J. Rodrigues-Carvajal, S. Patoux, C. Masquelier, Chem. Mater. 15

(2003) 4082.

[101] A. Deb, U. Bergmann, E.J. Cairns, S.P. Cramer, J. Synchrotron Rad. 11 (2004)

497.

[102] A.Yamada, H. Koizumi, N. Sonoyama, R. Kanno, Electrochem. Solid-State

Lett. 8 (2005) A409.

[103] H. Huang, S.C. Yin, L. F. Nazar, Electrochem. Solid State Lett. 4 (2001) A170.

[104] K. Zaghib, J. Shim, A. Guerfi,; P. Charest, K.A. Striebel, Electrochem. Solid

State Lett. 8 (2005) A207.

[105] A.D. Spong, G. Vitins, J. R. Owen, J. Electrochem. Soc. 152 (2005) A2376.

[106] D.X. Gouveia, V. Lemos, J. A.C. de Paiva, A.G. Souza Filho, J. Mendes Filho,

S. M. Lala, L. A. Montoro, J. M. Rosolen, Phys. Rev. B, 72 (2005) 24105.

[107] G.X. Wang, S. Bewlay, S. A. Needham, H.K. Liu, R.S. Liu, V. A. Drozd, J. F.

Lee, J. M. Chen, J. Electrochem. Soc. 153 (2006) A25.

[108] H. Liu, Q. Cao, L.J. Fu, C. Li, Y.P. Wu, H.Q. Wu, Electrochem. Commun. 8

(2006) 1553.

[109] P.P. Prosini, M. Carewska, S. Scaccia, P. Wisniewski, S. Passerini and M.

173
References
--------------------------------------------------------------------------------------------------------------------------

Pasquali, J. Electrochem. Soc. 149 (2002) A886.

[110] H. Huang, S.-C. Yin, and L.F. Nazar, Electrochem. Solid-state Lett. 4 (2001)

A170.

[111] P.S. Herle, B. Ellis, N. Coombs and L.F. Nazar, Nat. Mater. 3 (2004) 147.

[112] Z.X.Shu, R.S. McMillan, J.J. Murry, J. Electrochem. Soc.140 (1993) 922.

[113] J.M.Tarascon, D. Guyomard, Electrochim. Acta 38 (1993) 1221.

[114] T. Ohzuku, Y. Iwakoshi, K. Swai, J. Electrochem. Soc. 140 (1993) 2490.

[115] J.R. Dahn, A.K. Sleigh, H. Shi, J.N. Reimers, Q. Zhong, B.M. Way,

Electrochim. Acta 38 (1993) 1179.

[116] K. Sato, M. Noguchi, A. Demachi, N. Oki, M. Endo, Science 264 (1994) 556.

[117] J.R. Dahn, T. Zheng, Y.H. Liu, J.S. Xue, Science 270 (1995) 590.

[118] J.R. Dahn, A.K. Sleigh, H. Shi, B.M. Way, W.J. Wegdanz, J.N. Reimers, Q.

Gong and U. von Sacken, in G. Pistoia (ed.), Lithium Batteries, Elsevier

Science, Amsterdam, 1994.

[119] T. Zheng, Y. Liu, E.W. Fuller, S. Tseng, U.V. Sacken and J.R. Dahn, J.

Electrochem. Soc. 142 (1995) 2581.

[120] T. Enoki, S. Miyajima, M. Sano, H. Inokuchi, J. Mater. Res. 5 (1990) 435.

[121] G.T.-K. Fey and C.L. Chen, 10th International Meeting on Lithium Batteries,

Abstract No. 17, Como, Italy, May 28 – Jun 2, 2000.

[122] C.S. Wang, G.T. Wu, X.B. Zhang, Z.F. Qi and W.Z. Li, J. Electrochem.

Soc.145 (1998) 2751.

[123] F. Salrer-Disma, C. Lenair, B. Beaudoin, L. Aymard, J-M. Tarascon, Solid

State Ionics 98 (1997) 145-158.

[124] J.O. Besenhard, J. Yang and M. Winter, J. Power Sources 68 (1997) 87-90.

174
References
--------------------------------------------------------------------------------------------------------------------------

[125] J. Wolfenstine, J. Power Sources 79 (1999) 111-113.

[126] K.D. Kepler, John T. Vaughey, and M.M. Thackeray, Electrochemical and

Solid-State Letters 2 (1999) 3070309.

[127] O. Mao, R.A. Dunlap, and J.R. Dahn, J. Electrochem. Soc. 146 (1999) 405-423.

[128] B. Dipietro, M. Patriarca, B. Scrosati, J. Power Sources 8 (1982) 289.

[129] US patent 5, 478, 671(1995).

[130] Y. Idota, T. Kubota, A. Matsufuji, Y. Maekawa, T. Miyasaka, Science 276

(1997) 1395.

[131] Ian A. Courtney and J.R. Dahn, J. Electrochem. Soc. 144 (1997) 2045.

[132] J. Morales and L.Sanchez, J. Electrochem. Soc. 146 (1999) 1640-1642.

[133] P. Poizot, S. Laruelle, S. Grugeon, L. Dupont and J-M. Tarascon, Nature 407

(2000) 496.

[134] P. Balaya, H. Li, L. Kienle, J. Maier, Adv. Funct. Mater. 13 (2003) 621.

[135] G. Binotto, D. Larcher, A.S. Prakash, R. H. Urbina, M.S. Hegde, J.-M.

Tarascon, Chem. Mater. 19 (2007) 3032.

[136] A. M. Wilson and J.R. Dahn, J. Electrochem. Soc. 142 (1995) 326.

[137] J. Niu and J. Y. Lee, Electrochem. Solid-State Lett. 5 (2002) A107.

[138] H. Li, X.J. Huang, L.Q. Chen, Z. G. Wu and Y. Liang, Electrochem. Solid-State

Lett. 2 (1999) 547.

[139] J. Gratz, C.C. Ahn, R. Yazami, and B. Fultz, Electrochem. Solid-State Lett. 6

(1999) A194.

[140] S. Ohara, J. Suzuki, K. Sekine, T. Takamura, J. Power Sources, 119-121 (2003)

591.

175
References
--------------------------------------------------------------------------------------------------------------------------

[141] G.X. Wang, M.J. Lindsay, L. Sun, D.H. Bradhurst, S.X. Dou and H.K. Liu, J.

Power Sources, 97 – 98 (2001) 211.

[142] C.S. Wang, G.T. Wu, X.B. Zhang, Z.F. Qi and W.Z. Li, J. Electrochem. Soc.

145 (1998) 2751.

[143] M. Winter, J. O. Besenhard, Electrochim. Acta, 45 (1999) 31.

[144] J.O. Besenhard, J. Yang and M. Winter, J. Power Sources, 68 (1997) 87.

[145] F. Disma, L. Aymard, L. Dupont and J.-M. Tarascon, J. Electrochem. Soc.143

(1996) 3959.

[146] F.Disma, C. Lenain, B. Beaudoin, L. Aymard and J.-M. Tarascon, Solid State

Ionics, 98, 145 (1997).

[147] R.W. Pekala, Journal of Materials Science 24 (1989) 3221.

[148] D.D. Werstler, Polymer 27 (1986) 757.

[149] S. Iijima Nature 354 56 (1991) 8.

[150] R.Saito, G.Dresselhaus, MS.Dresselhaus Physical Properties of Carbon

Nanotubes. London: Imperial College Press. 1998.

[151] AS.Claye, JE.Fischer, CB.Huffman, AG.Rinzler, RE.Smalley. J Electrochem

Soc 147 (2000) 2845-2852.

[152] TW.Ebbesen, PM.Ajayan, Nature 358 220 (1992) 2.

[153] M.Yudasaka, T.Komatsu, T.Ichihashi, S.Iijima. Chem. Phys. Lett 278 102

(1997) 8.

[154] J.Kong, H.Soh, AM.Cassel, CF.Quate, H.Dai. Nature 395 (1998) 878-881.

[155] PL.Mceuen Nature 393 15 (1998) 7.

[156] GL.Che, BB.Lakshmi, ER.Fisher, CR.Martin. Nature 393 346 (1998) 8.

176
References
--------------------------------------------------------------------------------------------------------------------------

[157] CJ.Lee, JH.Park, SY.Kang, JH.Lee. Chemical Physics Letters 326 (2000) 175 -

180.

[158] E.Frackowiak, S.Gautier, H.Gaucher, S.Bonnamy, F.Beguin. Carbon 37:61

(1999) 9.

[159] GT.Wu, CS.Wang, XB.Zhang , HS.Yang, ZF.Qi, PM.He, WZ.Li. J

Electrochem Soc 146 (1999) 1696-1701.

[160] A.Funabiki, M.Inaba, Z.Ogumi. J Power Sources 68 (1997) 227 - 231.

[161] I.A.Courtney, J.R.Dahn, J. Electrochem. Soc.144 (1997) 2045

[162] I.A.Courtney, J.R.Dahn, J. Electrochem. Soc.144 (1997) 2943.

[163] S.A. Needham, G.X. Wang, K. Konstantinov, Y. Tournayre, Z. Lao, H.K. Liu,

Electrochemical and Solid-State Letters 9, (2006) A 315.

[164] M.Jean, C.Desnoyer, A.Tranchant, R.Messina. J. Electrochem. Soc 142 (1995)

2122.

[165] Y.Matsumura, S.Wang, J.Mondori. J.Electrochem.Soc 142 (1995) 2914 - 8.

[166] R.Dominko, D.Arcon, A.Mrzel, A.Zorko, P.Cevc, P.Venturini, et al. Adv Mater

14 (2002) 1531 - 4.

[167] J.B. Goodenough, A. Manthiram, and B. Wnetrzewski, J. Power Sources 43/44

(1993) 269.

[168] J-M. Tarascon and M. Armand, Nature 414 (2001) 359.

[169] J.L. Tirado, Materials Science and Engineering R 40 (2003) 103.

[170] M. Stanley Whittingham, Chem. Rev. 104 (2004) 4271.

[171] K. Sato, M. Noguchi, A. Demachi, N. Oki, and M. Endo, Science 264 (1994)

556.

177
References
--------------------------------------------------------------------------------------------------------------------------

[172] O. Mao, R.L. Turner, I.A. Countney, B.D. Fredericksen, M.I. Buckett, L.J.

Krause, J.R. Dahn, Electrochem. Solid-state Lett. 2 (1999) 3 – 5.

[173] J.O. Besenhard, M. Hess and P. Komenda, Solid State Ionics 40/41 (1990) 525.

[174] G.X. Wang, J.H. Ahn, Jane Yao, Steve Bewlay and H.K. Liu, Electrochem.

Commun. 6 (2004) 689.

[175] G.X. Wang, Jane Yao and H.K. Liu, Electrochem. Solid-state Lett. 7 (2004)

A250.

[176] D.C.S. Souza, V. Pralong, A.J. Jacobson and L.F. Nazar, Science 296 (2002)

2012.

[177] M.P. Bichat, J.L. Pascal, F. Gillot, F. Favier, Chem. Mater. 17 (2005) 6761.

[178] W. Jeitschko and D.J. Braun, Acta Cryst. B 34 (1978) 3196 – 3201.

[179] C. Perrier, H. Vincent, P. Chaudouet, B. Chenevier and R. Madar, Materials

Research Bulletin 30 (1995) 357.

[180] A. Dommam, R.E. Marsh and F. Hulliger, Journal of the Less-common Metals

152 (1989) 1.

[181] O. Garcia-Moreno, M. Alvarez-Vega, F. Garcia-Alvarado, J. Garcia-Jaca, J.M.

Gallardo-Amores, M.L. Sanjuan and U. Amador, Chem. Mater. 13 (2001) 1570

- 1576.

[182] C. Delacourt, P. Poizot, M. Morcrette, J.-M. Tarascon, and C. Masqulier, Chem.

Mater. 16 (2004) 93 – 99.

[183] C. Delacourt, P. Poizot, J.-M. Tarascon, and C. Masquelier, Nature Materials 4

(2005) 254 – 259.

178
References
--------------------------------------------------------------------------------------------------------------------------

[184] E.M. Bauer, C. Bellitto, M. Pasquali, P.P. Prosini, and G. Righini,

Electrochemical and Solid-State Letters, 7, A85 (2004).

[185] G.X. Wang, Steve Bewlay, Jane Yao, J.H. Ahn, S.X. Dou, and H.K. Liu,

Electrochemical and Solid-State Letters, 7, (2004) A503.

[186] F. Croce, A.D. Epifanio, J. Hassoun, A. Deptula, J. Olczac, and B. Scrosati,

Electrochemical and Solid-State Letters, 5, (2002) A47.

[187] P.P. Prosini, D. Zane, and M. Pasquali, Electrochimica Acta, 46, (2001) 3517.

[188] S.Y. Chung, J.T. Bloking, and Y.M. Chiang, Nature Materials, 1 (2002) 123.

[189] P.S. Herle, B. Ellis, N. Coombs, and L.F. Nazar, Nature Materials, 3 (2004)

147.

[190] A. Yamada, Y. Kudo, and K.Y. Liu, J. Electrochem. Soc. 148 (2001) A747 –

A754.

[191] A. Yamada and S.C. Chung, J. Electrochem. Soc. 148 (2001) A960 – A967.

[192] A. Yamada, Y. Kudo, and K.Y. Liu, J. Electrochem. Soc. 148 (2001) A1153 –

A1158.

[193] A.C. Larson and R.B. Von Dreele, General Structure Analysis System, Los

Alamos National Laboratory, Los Alamos, NM, 1994.

[194] P.S. Herle, B. Ellis, N. Coombs, and L.F. Nazar, Nat. Mater. 3 (2004) 147.

[195] R.P. Santoro and R.E. Newnham, Acta Cryst. 22 (1967) 344.

[196] G. Rousse, J.R. Carvajal, S. Patoux, and C. Masquelier, Chem. Mater. 15 (2003)

4082.

179
References
--------------------------------------------------------------------------------------------------------------------------

[197] M. Walkihara, Materials Science and Engineering R, 33 (2001) 109 – 134.

[198] J.L. Tirado, Materials Sceince and Engineering R, 40 (2003) 103 – 136.

[199] M.S. Whittingham, Chem. Rev., 104 (2004) 4271 – 4301.

[200] A.R. Armstrong and P.G. Bruce, Nature, (1996) 499 – 500.

[201] B. Ammundsen, J. Desilvestro, T. Groutso, D. Hassell, J.B. Metson, E. Regan,

R. Steiner and P.J. Pickering, J. Electrochem. Soc. 147 (2000) 4078.

[202] M.M. Thackeray, J. Electrochem. Soc. 142 (1995) 2558.

[203] G.H. Li, H. Ikuta, T. Uchida, M. Wakihara, J. Electrochem. Soc. 143 (1996)

178.

[204] I. Saadoune, C. Delmas, J. Mater. Chem., 6 (1996) 193.

[205] A.R. Armstrong, A.J. Paterson, A.D. Robertson, P.G. Bruce, Chem. Mater., 14

(2002) 710.

[206] A.K. Padhi, K.S. Nanjundaswamy, and J.B. Goodenough, J. Electrochem. Soc.,

144 (1997)1188 – 1194.

[207] A.K. Padhi, K.S. Nanjundaswamy, C. Masquelier, and J.B. Goodenough, J.

Electrochem. Soc.144 (1997) 1609 – 1613.

[208] K. Amine, H. Yasuda, and M. Yamachi, Electrochem. Solid State Lett. 3 (2000)

178 – 179.

[209] H. Huang, S.C. Yin, T. Kerr, N. Taylor, and L.F. Nazar, Adv. Mater., 14 (2002)

1525 – 1528.

[210] C. Delacourt, P. Poizot, M. Morcrette, J.-M. Tarascon, and C. Masquelier,

Chem. Mater., 16 (2004) 93 – 99.

180
References
--------------------------------------------------------------------------------------------------------------------------

[211] H. Huang, S.-C. Yin, and L.F. Nazar, Electrochem. Solid State Lett. 4 (2001)

A170 – A172.

[212] A.Yamada, S.C. Chung, and K. Hinokuma, J. Electrochem. Soc. 148 (2001)

A224 – A229.

[213] F. Croce, A. D’Epifanio, J. Hassoun, A. Deptula, T. Olczac, and B. Scrosati,

Electrochem. Solid State Lett. 5 (2002) A47 – A50.

[214] S. Franger, F. Le Cras, C. Bourbon, and H. Rouault, Electrochem. Solid State

Lett. 5 (2002) A231 – A233.

[215] J. Barker, M.Y. Saidi, and J.L. Swoyer, Electrochem. Solid State Lett. 6 (2003)

A53 – A55.

[216] G.X. Wang, S. Bewlay, J. Yao, J.H. Ahn, S.X. Dou, and H.K. Liu, Electrochem.

Solid State Lett. 7 (2004) A503 – A506.

[217] S. Franger, C. Bourbon, and F. Le Cras, J. Electrochem. Soc. 151 (2004) A1024

– A1027.

[218] C.H. Mi, X.B. Zhao, G.S. Cao, and J.P. Tu, J. Electrochem. Soc.152 (2005)

A483 – A487.

[219] P.S. Herle, B. Ellis, N. Coombs, and L.F. Nazar, Nature Materials, 3, (2004)

147 – 151.

[220] B.S. Berry and W.C. Pritchet, Solid State Commun. 26 (1978) 827 – 829.

[221] Y.N. Song, P.Y. Zavalij, N.A. Chernova, and M.S. Whittingham, Chem. Mater.

17 (2005) 1139 – 1147.

181
References
--------------------------------------------------------------------------------------------------------------------------

[222] R.P. Santoro and R.E. Newnham, Acta Cryst. 22 (1967) 344 – 347.

[223] Y.N. Song, P.Y. Zavalij, M. Suzuki, and M.S. Whittingham, Inorg. Chem. 41

(2002) 5778 – 5786.

[224] G. Rousse, J. Rodriguez-Carvajal, S. Patoux, and C. Masquelier, Chem. Mater.,

15 (2003) 4082 – 4090.

[225] S.Y. Chung, J.T. Blocking, and Y.M. Chiang, Nature Materials, 1 (2002) 123 -

128.

[226] A.S. Andersson, B. Kalska, L. Häggström, J.O. Thomas, Solid State Ionics, 130

(2000) 41 – 52.

[227] A.S. Andersson, J.O. Thomas, J. Power Sources, 97 – 98 (2001) 498 – 502

182
Figure and table Captions
--------------------------------------------------------------------------------------------------------------------------

Figure and table Captions

Figure 2.1 Schematic diagram of the battery and battery operation.

Figure 2.2 Gravimetric and volumetric energy densities of different

rechargeable battery systems.

Figure 2.3 The operating principle of the lithium-ion battery.

Figure 2.4 Schematic energy diagram of a cell at open circuit

Figure 2.5 The structure of layered LiCoO2.

Figure 2.6 Neutron diffraction patterns of (a) LiMn2O4 (b) LiCr0.04Mn1.96O4

Figure 2.7 The crystal structure of olivine LiFePO4 along [001]. Left:

Expanded view of the framework built on FeO6 octahedra and PO4

Tetradedra. Right: Restricted view of Li, Fe and P distribution

between two distorted h.c.p. (hexagonal close packed)

oxygen-dense layers (PTd [LiFe] octO4). LiO6 octahedra share edges

and Li ions may diffuse along [010] and [001].

Figure 2.8 Crystal structures of left: orthorhombic LiFePO4 and right: trigonal

quartz-like FePO4

Figure 2.9 a. Model displaying the connection among FeO6, LiO6 and PO4

structural units; b Structure viewed along the c-axis.

Figure 2.10 Structural schematics of three types of carbons.

Figure 2.11 Gravimetric capacity of lithium-alloy forming elements.

Figure 2.12 Electrochemical properties of MO anodes in lithium-ion cells. (a)

The voltage – composition profiles for MO/Li cells cycled between

183
Figure and table Captions
--------------------------------------------------------------------------------------------------------------------------

0.01 V and 3 V at a rate of C/5. (b) Cyclability of MO anodes, the

data also shows for a Co3O4/Li cell, demonstrating the behaviour is

not specific to divalent oxides. The inset is the rate capacity of a

CoO electrode.

Table 2.1 The structural and electrochemical parameters of LiMδMn2-2δO4

cathodes

Figure 3.1 JEOL JSM-6460A SEM facility

Figure 3.2 JEOL 2011 TEM facility

Figure 3.3 Atmosphere controllable Glove-box for assembling CR2032 coin

cells

Figure 3.4 CHI 660 B and 660C Electrochemical Work Stations

Table 3.1 Materials and chemicals

Figure 4.1 SEM image of (a) bare MCMB graphite (b) 20 wt% nano

Si-MCMB ball milled for 5 hours. (c) 20 wt% nano Si-MCMB ball

milled for 10 hours (d) 20 wt% nano-Si-MCMB ball milled for 20

hours

Figure 4.2 X-ray diffraction pattern of 20 wt% nano-Si-MCMB ball milled for

10 hours.

Figure 4.3 Cyclic voltammograms (a) MCMB electrode (b) nano-Si-MCMB

electrode (sample E)

Figure 4.4 Schematic model of the lithiation and de-lithiation process in

Si-MCMB composites. The volume increases due to the formation

184
Figure and table Captions
--------------------------------------------------------------------------------------------------------------------------

of LixSi alloys are small in local domain and effectively buffered by

MCMB matrix.

Figure 4.5 The discharge/charge profiles of (a) bare nano-Si electrode (b)

nano-Si-MCMB electrode (sample E).

Figure 4.6 The discharge capacity versus cycle number (Current density: 0.05

mAcm-2).

Figure 4.7 A schematic diagram of the R-F gel reaction.

Figure 4.8 X-ray diffraction pattern of Si-C composites.

Figure 4.9 (a) TEM image of nanocrystalline Si powders. (b) TEM image of

nanostructured Si-C composites.

Figure 4.10 Cyclic voltammetry of Si-C composite electrode.

Figure 4.11 The discharge/charge profiles of Si-C composite electrode.

Figure 4.12 The discharge capacity versus cycle number for Si-C electrode.

Table 4.1 Specific capacity in the first cycle for bare MCMB, bare

nanocrystalline Si, and Si-MCMB electrodes

Figure 5.1 A schematic diagram of the CVD quartz reactor.

Figure 5.2 (a) (b) (c) and (d) SEM images of vertically aligned carbon

nanotubes.

Figure 5.3 (a) TEM image of carbon nanotubes. (b) HRTEM image of a

carbon nanotube.

Figure 5.4 Cyclic voltammograms of the carbon nanotube electrode. Scanning

rate: 0.1 mV/s.

185
Figure and table Captions
--------------------------------------------------------------------------------------------------------------------------

Figure 5.5 Discharge curves of the carbon nanotube electrode (The inset is the

discharge capacity versus cycle number for the carbon nanotube

electrode).

Figure 5.6 X-ray diffraction pattern of SnO2-CNT composites

Figure 5.7 TEM images of SnO2-CNT composites.

Figure 5.8 HRTEM images of SnO2-CNT composites.

Figure 5.9 Cyclic voltammograms of SnO2-CNT composite electrode.

Scanning rate: 0.1mV/s.

Figure 5.10 The first and second discharge and charge curves of SnO2-CNT

composite electrode.

Figure 5.11 Discharge capacity vs. cycle number for SnO2-CNT composite

electrode. The inset presents the charging efficiency vs. cycle

number.

Figure 6.1 X-ray diffraction pattern of the as-prepared FeP4 powders

Figure 6.2 SEM image of FeP4 powders.

Figure 6.3 Cyclic voltammograms of FeP4 anode in lithium-ion cell.

Figure 6.4 Charge/discharge profiles of FeP4 anode.

Figure 6.5 The discharge capacity of FeP4 anode vs. cycle number.

Figure 7.1 (a) X-ray diffraction patterns of LiMnxFe1-xPO4 (b) Rietveld

refinement on X-ray diffraction pattern of LiMn0.5Fe0.5PO4 sample.

Figure 7.2 SEM images of LiMnxFe1-xPO4 sample powders, A: LiFePO4

sample; B: LiMn0.5Fe0.5PO4 sample; C: LiMnPO4 sample.

Figure 7.3 Temperature dependencies of molar magnetic susceptibilities of

186
Figure and table Captions
--------------------------------------------------------------------------------------------------------------------------

(a) LiFePO4 sample (b) LiMn0.5Fe0.5PO4 sample

Figure 7.4 Discharge capacity vs cycle number for LiMnxFe1-xPO4 sample

electrodes. Current rate: C/8; Voltage range: 2.75 V – 4.30 V.

Table 7.1 Phase composition and lattice parameters of olivine phase for

LiMnxFe1-xPO4 compounds.

Table 7.2 Selected interatomic distances (Å) and angles (deg) for

LiMnxFe1-xPO4 samples.

Figure 8.1 X-ray diffraction patterns of (a) LiFePO4 and (b) Li0.95Mg0.05FePO4.

Figure 8.2 SEM images of Li0.95Mg0.05FePO4 samples sintered at (A) 700 °C,

(B) 800 °C, and (C) 850 °C.

Figure 8.3 EDS elemental mapping by SEM.

Figure 8.4 Temperature dependencies of molar magnetic susceptibilities of

LiFePO4, samples (a) LFP-700 and (b) LFP-850. The insets are

reciprocal magnetic susceptibilities.

Figure 8.5 Temperature dependencies of molar magnetic susceptibilities of

Li0.95Mg0.05FePO4 samples (a) LFPMg-700 and (b) LFPMg-850.

The insets are reciprocal magnetic susceptibilities.

Figure 8.6 The effective magnetic moment per mole of iron of LiFePO4 and

Li0.95 Mg0.05 FePO4 samples.

Figure 8.7 shows typical voltammograms of LiFePO4 and Li0.95Mg0.05PO4

electrodes.

Figure 8.8 Charge/discharge curves in the first cycle for (a) LiFePO4 and (b)

Li0.95 Mg0.05 FePO4 sample electrodes. Current rate: C/8.

187
Figure and table Captions
--------------------------------------------------------------------------------------------------------------------------

Figure 8.9 Discharge capacity vs. cycle number for LiFePO4 and

Li0.95Mg0.05FePO4 sample electrodes. Current rate: C/8.

Table 8.1 Sintering temperature, label and average crystal size of lithium iron

phosphates.

Table 8.2 Magnetic parameters of LiFePO4 and doped Li0.95Mg0.05FePO4

samples.

188
Publications
------------------------------------------------------------------------------------------------------------------------

Publications

1. J. Yao, J.S Park, K. Konstantinov, G.X.Wang, J.H. Ahn, J. Wang and H.K. Liu
“Electrochemical performance of nanocrystalline SnO2-carbon nanotube
composites as anode in lithium-ion cells” Journal of Nanoscience and
Nanotechnology (in press).

2. J. Yao, K. Konstantinov, G.X. Wang, and H.K. Liu, “Electrochemical and magnetic
characterization of LiFePO4 and Li0.95 Mg0.05 FePO5 cathode materials” Journal of
Solid State Electrochemistry 11(5) (2007) 177 -185.

3. J. Yao, S. Bewlay, V.A. Drozd, R.S. Liu, X.L. Wang, H.K. Liu and G.X. Wang,
“Characterisation of olivine-type LiMnxFe1-xPO4 cathode materials” Journal of
Alloys and compounds 425 (2006) 362 – 366.

4. G.X. Wang, J. Yao, H.K. Liu, S.X. Dou and J.H. Ahn, “Growth and lithium storage
properties of vertically aligned carbon nanotubes” Metals and Materials
International 12(5) (2006) 413 – 416.

5. G.X. Wang, S. Needham, J. Yao, J.Z. Wang, R.S. Liu and H.K. Liu, “A study on
LiFePO4 and its doped derivatives as cathode materials for lithium-ion batteries”
Journal of Power Sources 159 (2006) 282 - 286.

6. G.X. Wang, Y. Chen, L. Yang, J. Yao, S. Needham, H.K. Liu and J.H. Ahn,
“Synthesis of nanocrystalline transition metal and oxides for lithium storage”
Journal of Power Sources 146 (2005) 487 - 491.

189
Publications
------------------------------------------------------------------------------------------------------------------------

7. G.X. Wang, S. Bewlay, J. Yao, J.H. Ahn, H.K. Liu and S.X. Dou,
“Characterisation of LiMxFe1-xPO4 (M=Mg, Zr, Ti) cathode materials prepared by
the sol-gel method” Electrochemical and Solid-state Letters 7(12) (2004) A503 –
A506.

8. G.X. Wang, J.H. Ahn, J. Yao, S. Bewlay and H.K. Liu, “Nanostructured Si-C
composite anodes for lithium-ion batteries” Electrochemistry Communication 6(7)
(2004) 689 – 692.

9. G.X. Wang, J. Yao and H.K. Liu “Characterization of Nanocrystalline Si-MCMB


Composite Anode Materials” Electrochemical and Solid-State Letters 7(8) (2004)
A250 –A253.

10. G.X. Wang, J. Yao, J.H. Ahn, H.K. Liu and S.X. Dou, “Electrochemical
properties of nanosize Sn-coated graphite anodes in lithium-ion cells” Journal of
Applied Electrochemistry 34 (2004) 187 – 190.

11. G.X. Wang, S. Bewlay, J. Yao, H. K. Liu, and S. X. Dou, “Tungsten Disulfide
Nanotubes for Lithium Storage”Electrochemical and Solid-State Letters 7 10
(2004) A321-A323.

190

You might also like