synthesis and characterization of new positive electrode

174
HAL Id: tel-03080208 https://tel.archives-ouvertes.fr/tel-03080208 Submitted on 17 Dec 2020 HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers. L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés. Synthesis and characterization of new positive electrode materials for Li-ion high energy applications Yagmur Celasun To cite this version: Yagmur Celasun. Synthesis and characterization of new positive electrode materials for Li-ion high energy applications. Chemical and Process Engineering. Université Grenoble Alpes [2020-..], 2020. English. NNT : 2020GRALI047. tel-03080208

Upload: others

Post on 24-Oct-2021

3 views

Category:

Documents


0 download

TRANSCRIPT

HAL Id: tel-03080208https://tel.archives-ouvertes.fr/tel-03080208

Submitted on 17 Dec 2020

HAL is a multi-disciplinary open accessarchive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come fromteaching and research institutions in France orabroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, estdestinée au dépôt et à la diffusion de documentsscientifiques de niveau recherche, publiés ou non,émanant des établissements d’enseignement et derecherche français ou étrangers, des laboratoirespublics ou privés.

Synthesis and characterization of new positive electrodematerials for Li-ion high energy applications

Yagmur Celasun

To cite this version:Yagmur Celasun. Synthesis and characterization of new positive electrode materials for Li-ion highenergy applications. Chemical and Process Engineering. Université Grenoble Alpes [2020-..], 2020.English. �NNT : 2020GRALI047�. �tel-03080208�

THÈSE

Pour obtenir le grade de

DOCTEUR DE L’UNIVERSITE GRENOBLE ALPES

Spécialité : 2MGE : Matériaux, Mécanique, Génie Civil, Electrochimie

Arrêté ministériel : 25 mai 2016

Présentée par

Yagmur CELASUN

Thèse dirigée par Sébastien MARTINET, Ingénieur de Recherche

Co-encadrée par Jean-François COLIN, Ingénieur de Recherche

et David PERALTA, Ingénieur de Recherche

préparée au sein du Laboratoire LM (CEA/LITEN) Grenoble dans l'École Doctorale I-MEP2 - Ingénierie Matériaux, Mécanique, Environnement, Energétique, Procédés, Production

Synthèse et caractérisation de nouveaux matériaux d'électrode positive pour des applications Li-ion à haute énergie Thèse soutenue publiquement le « 30 Septembre 2020 » devant le jury composé de :

Madame Fannie, ALLOIN Directeur de Recherche, LEPMI, Université Grenoble Alpes, Présidente de jury

Monsieur Sylvain FRANGER Professeur des Universités, ICMMO, Université Paris-Sud, Rapporteur

Monsieur Philippe MOREAU Professeur des Universités, IMN, Université de Nantes, Rapporteur

Monsieur Sébastien MARTINET Ingénieur de recherche, CEA Grenoble LITEN, Directeur de thèse Et des membres invités : Monsieur Jean-François COLIN Ingénieur de recherche, CEA Grenoble LITEN, Invité

Monsieur David PERALTA Ingénieur de recherche, CEA Grenoble LITEN, Invité

Acknowledgments

This thesis study was often exciting, sometimes difficult and rarely stressful for me. I have

always felt that I have been so lucky because I have had a great thesis director and talented

supervisors. I would like to thank my thesis director, Sebastien Martinet, for his valuable

insights and fruitful discussions during my project meetings. I always found him as a very

professional and gentleman person. His vision and attitude also guided me how to become a

charismatic leader. I would also like to thank to my project supervisors, Jean François Colin

and David Peralta for their valuable supports during my thesis. Jean François’s deep

knowledge in crystalline materials and David’s practical synthesis ideas shaped this project as

it is today and produced valuable outcomes. With their supervision, I learned how to continue,

where to stop and when I should push forward. I always admired their respect to individuals,

thinking style and way of criticism. Though I have not seen them before my interview, the

harmony between us was always great. I thank again for their leadership in my project and I

will never forget the time I spent with them.

As like other PhD theses, we found interesting materials, but then struggled in the sea of

unknowns. At that time, Prof. Claude Delmas, who is a precious visitor of CEA LITEN, shared

his rich knowledge with us and showed different aspects of the problems. I would like to thank

him for our fruitful discussions during our team meetings.

When I needed a practical information or opinion about any method or knowledge, I have got

some hints from my lovely office mate, Carole Bourbon. I would like to thank her for her joyful

attitude and the great time that we spent together. I wish to her happy retirement in 2020!

I would also like to thank Anass Benayad, for his significant help for XPS measurements and

discussions on the results, Laurent Gonon for his help for Raman Spectroscopy, Celine

Barchasz and Alice Robba for the discussions about lithium sulfur system. I would also like to

extend my thanks to all my colleagues in LITEN/STB, Sebastien Launois, Vasily Tarnopolskiy,

Ognjen Hajndl, Camille Guerin, Laurent Bernard, Guillaume Lefevre and Guillaume Tonin,

Gaelle Besnard, Damien Guegan, Benjamin Mercier and Oumaima Chaouachi for their

friendship and personal support during my thesis.

2

I would like to thank to Parviz Hajıyev for his valuable support in my personal and

professional life. I never forget our discussions during lunches and coffee breaks in CEA. We

always found crazy ideas to create new materials and also shared the problems that we faced

during the thesis.

I also would like to thank to my friends, Mauro, Alessandro, Taki, Krishna, Nono, Youcef,

Abdel, Tom, Shu Chen and Nora from “Lets Speak French” activity. I could improve my

French speaking skills during our weekly meetings, and I spent wonderful time when we were

having a barbecue, playing football in the grass, skiing Rhone Alpes, watching French movies

and sightseeing small French cities.

At stage of the submitting the thesis manuscript, the stress level increased. Sarp Mamıkoglu

helped me to correct the language of the manuscript at final stage and encouraged me to get

rid of the stress. I would like to thank to him for his support.

I would also like to thank CEA LITEN, especially to our department manager, Sebastien

Patoux, and lab manager, Didier Bloch, for their financial support for this thesis study. With

their support, I could have participated to various conferences, presented my thesis work,

obtained necessary scientific formations for my project and improved my French speaking

skills.

My special thanks go to my family for their encouragements, advices and prays during my

thesis study. With their precious support, I have always believed in myself to achieve the

things that I started, and I could have strived against negativity, depression and

procrastination.

3

Table of Contents

General Introduction .......................................................................................................................... 6

Chapter I. State of the art ................................................................................................................ 10

I.1. Rechargeable Lithium-Ion Batteries ........................................................................................ 10

I.1.a. Li-ion Battery Working Principle and Characteristics ..................................................... 10

I.1.b. Positive Materials for Li-ion Batteries ................................................................................ 13

I.1.c. High Energy Positive Materials ........................................................................................... 14

I.2. Cation Disordered Rocksalts as Positive Materials .............................................................. 16

I.2.a. The first cation disordered rocksalts: stoichiometric compositions ............................... 16

I.2.b. Lithium diffusion mechanism through cation disordered rocksalts ............................. 18

I.3. Li-rich Cation Disordered Rocksalt Oxides ........................................................................... 21

I.3.a. Li3NbO4 based cation disordered rocksalt oxides............................................................. 21

I.3.b. Li2TiO3 based cation disordered rocksalt oxides .............................................................. 24

I.3.c. Li4MoO5 based cation disordered rocksalt oxides ............................................................ 25

I.4. Li-rich Cation Disordered Rocksalt Sulfides ......................................................................... 27

I.4.a. Li2TiS3 based cation disordered rocksalt sulfides ............................................................. 27

I.4.b. Li3NbS4 based cation disordered rock salt sulfides .......................................................... 29

I.5. Redox Activity in Disordered Rocksalts ................................................................................. 30

I.6. Conclusion .................................................................................................................................... 35

Chapter II: Experimental Procedures ............................................................................................. 36

II.1. Synthesis Methods .................................................................................................................. 36

II.2. Structural Characterizations .................................................................................................. 39

II.3. Electrochemical Characterizations ....................................................................................... 51

Chapter III: Development of New Li2.2NiTi0.2Nb0.6O4 Positive Material in Disordered

Rocksalt Structure ............................................................................................................................. 57

III.1. Synthesis Improvement .......................................................................................................... 61

III.1.a. Simple tempering (long firing) and quenching strategies ............................................ 63

III.1.b. Effect of washing (type of solvent) .................................................................................. 66

III.1.c. Effect of molten salt to precursor ratio ............................................................................ 68

III.1.d. Molten salt synthesis with a different eutectic salt (LiCl-KCl) .................................... 70

III.1.e. Effect of calcination temperature and duration ............................................................. 73

4

III.2. Understanding Studies ........................................................................................................... 76

III.2.a. Ex Situ X-ray Diffraction Analysis ................................................................................... 78

III.2.b. In Situ X-ray Diffraction Analysis .................................................................................... 79

III.3. Conclusion................................................................................................................................. 84

Chapter IV: Developing Li-rich Rocksalt Metal Sulfides-Selenides: Li2TiSexS3-x ................ 86

IV.1. Disordered Li2TiS3 as a Positive Electrode .......................................................................... 87

IV.1.a. Mechanochemical Synthesis of Li2TiS3 Powders ............................................................ 87

IV.1.b. Electrochemical Performance of Li2TiS3 .......................................................................... 92

IV.1.c. Air Sensitive Property of Li2TiS3 ....................................................................................... 95

IV.2. Selenium Substituted Li2TiS3 .............................................................................................. 100

IV.2.a. Synthesis Improvements ................................................................................................. 101

IV.2.b. Electrochemical performances of Li2TiSexS3-x ............................................................... 109

IV.3. Conclusion............................................................................................................................... 120

Chapter V: Characterization of Li2TiSexS3-x Compositions ...................................................... 122

V.1. Operando Neutron Diffraction of Li2TiS3 ........................................................................... 122

V.2. Understanding Degradation: Study of the Structural Property by ex situ XRD ......... 126

V.2.a. Ex situ XRD and SEM Analyses of Li2TiS3 electrodes ................................................... 126

V.2.b. Ex situ XRD and SEM Analyses of Li2TiSexS3-x electrodes ............................................ 128

V.3. XPS Analysis of Li2TiSexS3-x Compositions ........................................................................ 132

V.3.a. Methodology: Analysis of XPS Spectra .......................................................................... 132

V.3.b. Survey analysis of Li2TiSexS3-x samples ........................................................................... 134

V.3.c. High Resolution XPS Analysis of Li2TiSexS3-x Powders ................................................ 137

V.3.d. XPS Analysis of Li2TiSexS3-x Pristine Electrodes ............................................................ 141

V.3.e. XPS Analysis on Li2TiSexS3-x Electrodes at Different States of Charge ....................... 144

V.4. Discussion on the XPS Results ............................................................................................. 148

V.5. Conclusion ................................................................................................................................ 150

General Conclusion and Perspectives ......................................................................................... 151

Bibliography ..................................................................................................................................... 156

Appendix ........................................................................................................................................... 170

5

General Introduction

6

General Introduction

In recent years, energy requirements have increased in-line with population growth in

the world. To preserve our resources and to create a sustainable world that is less

dependent on fossil fuels, application of renewable energy technologies has been

expanded. However, sustainability and efficiency of these technologies highly depend

on the success of energy storage systems (1,2).

Moreover, we are now faced with global warming and its result, climate change.

Changing our lifestyle, such as building smart homes and clean transportation are seen

as possible solutions to reduce our carbon footprint. Many automakers have already

started new initiatives, some of them even became actual battery producers (3). Such

attempts are linked to vast investments into battery research and manufacturing,

therefore batteries are becoming a hot topic again, after their first implementation in

cell phones in 1990s.

Figure 0 - 1: Cost breakdown of Li-ion cell (4)

However, electric vehicles are still expensive, since the battery pack is the single major

cost component in an electric vehicle. To reduce electric vehicle costs, US Department

of Energy has set a mission for EV producers that the cost of battery packs should be

reduced to 125 $.kWh-1 until 2022 (3). In order to realize this mission, new battery

chemistries should be discovered, volume of production should be augmented, cell

7

design should be optimized and manufacturing costs should be decreased. In fact, the

most significant contributor to battery costs are raw material costs (50-70%). Among

them, positive electrode is reported to have the highest percentage of costs (27%) (4)

(Figure 0 - 1). NMC [Lithium Nickel Manganese Cobalt Oxide (LiNixMnyCozO2)], NCA

[Lithium Nickel Cobalt Aluminium Oxide (LiNixCoyAlzO2)] are common positive

materials for electrical vehicles, while LCO [Lithium Cobalt Oxide (LiCoO2)] is

common positive electrode for portable applications. Such materials contain

considerable amount of cobalt and the price of cobalt is now an important concern

because of its scarcity, cost and mining conditions in Democratic Republic of Congo.

Thus, research on new cathode material focus on Co-poor or even Co-less materials.

However, it is not so easy since Co is usually needed to stabilize the layered structure,

therefore new chemistries should be discovered.

Figure 0 - 2: Specific energy of materials from material to pack level (5)

As electro-mobility applications have volume, weight and size constraints, a high

energy density battery cell is mandatory (Figure 0 - 2). The positive material (cathode)

is the limiting factor of the battery, so its specific energy has to be increased (5).

Currently there are only a few materials that allow battery systems reaching high

specific energies (250 Wh.kg-1). Among them, we can cite Li-rich NMC, Ni-rich NMC

and high-voltage spinel materials. Despite their high capacity, low cycling stability and

8

safety problems associated to structural deformations are also reported for these

materials. This highlights the importance of discovering new high energy materials.

Recently, disordered rocksalt type positive electrodes appear to be promising because

disordered rocksalt oxides offer capacities exceeding 250 mAh.g-1 at high average

potential (3.6 V vs Li+/Li), while disordered rocksalt sulfides offer excellent capacities

(>400 mAh.g-1) at relatively low average potential (2.3 V vs Li+/Li). Consequently, their

energy density is much higher than today’s commercial materials. This attracts

research interest; thus, many groups are now working on this subject to bring new

materials into the market.

In this work, we study two different disordered rocksalt families: disordered rocksalt

oxides and sulfides materials. We apply different methods to synthesize each family

and optimize our processes to have better electrochemical performances. Then, we

perform various characterizations to understand their electrochemical and structural

properties.

This thesis work is divided in five chapters:

The first chapter introduces Li-ion technology and current high-energy positive

materials briefly. Thereafter, it describes disordered rocksalt positive materials and

their properties in detail. It also explains why stoichiometric compositions failed before

and what Li-rich compositions offer today. Finally, it shows promising electrochemical

properties of Li-rich disordered rocksalts.

The second chapter contains the detail of the synthesis processes, sample preparations

and procedure of electrochemical, morphological and structural characterizations.

The third chapter contains synthesis optimizations to improve electrochemical

performance of disordered rocksalt Li2.2NiTi0.2Nb0.6O4. As the model material is

prepared, ex situ and in situ X-ray diffraction (XRD) studies are performed to

understand the structural evolution upon cycling. Finally, this chapter explains why

electrochemical performances of such materials are limited at the end of the first cycle.

9

The fourth chapter describes preparation of disordered rocksalt Li2TiS3 powders with

mechanochemical synthesis. To solve cycling stability problem of Li2TiS3, selenium

substituted materials are prepared with new synthesis process and their

electrochemical performances are compared with bare Li2TiS3.

The fifth chapter contains structural characterizations, such as operando neutron

diffraction and ex situ XRD to investigate lithium insertion/extraction mechanism in

Li2TiS3 during cycling. To elucidate redox processes in both Li2TiS3 and selenium

substituted materials, powders and electrodes are analyzed with ex situ X-ray

photoelectron spectroscopy (XPS).

In the last part of the thesis, we summarize our findings on disordered rocksalt oxide

and sulfide materials. By referring our experience in those materials, we present a

perspective and share our suggestions for future studies.

10

Chapter I. State of the art

I.1. Rechargeable Lithium-Ion Batteries

Rechargeable lithium-ion batteries (LIBs) have become widely used power sources not

only for portable devices but also for electrical vehicles. Differently from primary

batteries, the electrochemical reaction is reversible, therefore the cells can be recharged

several times and the delivered energy can be stored (6).

Traditional LIBs contain four main components: positive electrode, negative electrode,

non-aqueous electrolyte and separator. Transition metal oxides or phosphate based

materials are generally used as positive electrodes (LiCoO2, LiNi1/3Co1/3Mn1/3O2,

LiMn2O4 and LiFePO4), graphite is commonly used as negative electrode, non-aqueous

electrolytes contain lithium salts (LiPF6, LiClO4, LiTFSI, etc.) and carbonate based

solvents (ethylene carbonate, propylene carbonate, etc.) and separators are made of

polymers (polyethylene, polypropylene). The electrolyte is a medium that can only

allow ion transport, whereas it is an insulator for electron transport (6).

I.1.a. Li-ion Battery Working Principle and Characteristics

A typical LIB works following the principle that upon charging of the battery, Li ions

are extracted from the positive electrode, move inside the electrolyte and insert into

the negative electrode (Figure I - 1). During lithium extraction, oxidation processes take

place at the positive electrode and electrons flow through the external circuit from

positive to negative electrode, where reduction processes happen. Thus, electricity is

produced by the electrochemical reaction inside the cell. The inverse process happens

during the discharge of the battery.

11

Figure I - 1: Rechargeable Li ion battery: transfer of lithium ions from positive

electrode (LixMO2) to negative electrode (LixC6) during charge process

When the electrochemical cell is charging, oxidation process at positive electrode (i.e.,

LiMO2) can be represented by

LiMO2 → Li1-xMO2 + xLi+ + xe- eq. I - 1

While, reduction process at negative electrode (i.e., graphite) can be represented by

C6 + xLi+ + xe- → LixC6 eq. I - 2

The overall reaction is the sum of two half-cell reactions,

LiMO2 + C6 ⇄ Li1-xMO2 + LixC6 eq. I - 3

The free energy of the system, which is the driving force of the redox processes, can be

expressed based on the assumption that cell is at equilibrium,

∆𝐺 = 𝑛. 𝐹. 𝑈𝑡ℎ𝑒𝑜𝑟𝑒𝑡𝑖𝑐𝑎𝑙 eq. I - 4

where n is number of electrons exchanged, F is Faraday constant (96485 C.mol-1) and

𝑈𝑡ℎ𝑒𝑜𝑟𝑒𝑡𝑖𝑐𝑎𝑙 is the theoretical voltage between electrodes. At equilibrium state, there is

no current flow, therefore open circuit voltage (𝑈𝑜𝑐𝑣) is approximately equal to

12

theoretical voltage. During charge and discharge, current evolves; therefore, working

voltage, which is the actual operating voltage of the battery, differs from the open

circuit voltage.

Among rechargeable storage systems, Li-ion battery has the highest specific energy,

which is the product of the capacity and average discharge voltage. To maximize the

specific energy of the cell, active materials should provide large discharge capacity

and/or can be cycled at high discharge potentials. The theoretical capacity can be

defined as the charge (or lithium ions) storage ability of active materials, which is

calculated by

𝑄𝑡ℎ =

𝑛.𝐹

3600.𝑀𝑤

eq. I - 5

where Qth is theoretical capacity (mAh.g-1), n is the number of electrons (or lithium

ions) exchanged during charge-discharge, F is Faraday constant (96485 C.mol-1) and

Mw is the molecular mass of active material (g.mol-1). The capacity of the active material

is strongly affected by the C-rate (1/h), which can be defined as a measure of the

current rate at which a battery is discharging and can be represented by

𝐶 = 𝑐ℎ𝑎𝑟𝑔𝑒 (𝑜𝑟 𝑑𝑖𝑠𝑐ℎ𝑎𝑟𝑔𝑒) 𝑐𝑢𝑟𝑟𝑒𝑛𝑡

𝑡ℎ𝑒𝑜𝑟𝑒𝑡𝑖𝑐𝑎𝑙 𝑐𝑎𝑝𝑎𝑐𝑖𝑡𝑦 =

𝑖

𝑄 eq. I - 6

where i is current in A, theoretical capacity is expressed in Ah. For example, a cell

charging at a C/10 rate needs 10h to reach its full capacity.

For industrial applications, Li-ion cells should have long cycle life, which is usually

defined by the number of charge-discharge cycles that the cell can offer before its

capacity falls below 80% of its nominal capacity.

13

To develop new high performance and affordable Li-ion batteries, some important

parameters must be considered, such as

1. Selection of widely available and low-cost precursors for the active materials

2. Development of new high-energy materials with cost effective and scalable

production methods

3. Improvement on battery cycle (shelf) life

4. Understanding of battery safety and failure mechanisms

5. Discovery of easy recycling methods

I.1.b. Positive Materials for Li-ion Batteries

In 1980s, TiS2 was reported as the first intercalating compound by Whittingham (7) et

al. after it demonstrated the ability to intercalate and de-intercalate lithium ions

reversibly. Though TiS2 shows excellent cyclability, its average potential is rather low

(2.1 V vs Li+/Li), consequently the energy density is not sufficient enough for high

energy applications (8). Thereafter, study of oxide type compounds gains interest

because they could operate at relatively high potentials (> 4 V). These efforts results in

the discovery of LiCoO2, which was reported by Goodenough (9) et al. and later

commercialized by Sony in lithium- ion cells in 1991. However, a maximum of 0.5 Li

ions could be deinserted from the layered LiCoO2 without destabilizing the structure;

therefore the practical capacity is only half of its theoretical capacity (274 mAh.g-1) (9).

Increasing cobalt prices in the market is also another disadvantage. Having high

theoretical capacity and cheaper price of nickel, LiNiO2 is seen as an alternative

material (10,11). The studies show that the reversibility of LiNiO2 is lower than the one

of LiCoO2 (12), thus manganese substitution is suggested. Thereafter, the best

composition, which is found to be LiNi1/3Co1/3Mn1/3O2 (and called NMC111), is

discovered as a solid solution of LiCoO2, LiNiO2 and LiMnO2 compositions (13–16).

LiNi1/3Co1/3Mn1/3O2 delivers a capacity up to 160 mAh.g-1 at 0.1 C rate in a potential

range of 2.5 V – 4.4 V vs Li+/Li (17). Here, the nickel redox provides high capacity, cobalt

14

improves cycling stability and manganese enhances thermal stability. In parallel to

LiNi1/3Co1/3Mn1/3O2, layered LiNi0.80Co0.15Al0.05O2 is also studied (18–20). This material

delivers 200 mAh.g-1 at 0.1 C rate in a potential range of 3 V – 4.5 V (21) and Al

substitution increases the thermal stability (22).

Moreover, polyanionic LiFePO4 is also studied in the literature (23–25). Though LiFePO4

shows excellent cycling stability, its practical energy density is quite limited.

Having lower cost (<10 $.kg-1) and relatively stable structure compared to layer oxides,

spinel LiMn2O4 attracts research interests and was also commercialized by Moli

Energy. However, its practical capacity is rather small (~140 mAh.g-1). To utilize the

advantages of different chemistries, blend of LiMn2O4 with LiNi1/3Co1/3Mn1/3O2 or

LiNi0.80Co0.15Al0.05O2 have also been suggested (26,27). Addition of small amount of layer

oxides suppresses the Mn dissolution as well as improves the capacity retention of

LiMn2O4 cells (28).

I.1.c. High Energy Positive Materials

To extend the driving ranges of the electrical vehicles, specific energy of current

lithium ion batteries should be further increased. The specific energy is related to both

the capacity and the operating potential. Common approaches are focused on the use

of high capacity materials, for instance Ni-rich NMC (LiNixCoyMnzO2, with x≥0.6) and

Li-rich (or Mn rich) NMC [xLi2MnO3. (1-x)LiMO2] or high potential positive materials,

such as LNMO (LiNi0.5Mn1.5O4).

• Ni-rich NMC (LiNi0.80Co0.10Mn0.10O2)

The capacity of the typical NMC111 (LiNi1/3Co1/3Mn1/3O2) is reported as 160 mAh.g-1,

which is smaller than the theoretical value (275 mAh.g-1) upon cycling at 0.1 C rate in

a potential range between 2.5 V - 4.4 V vs Li+/Li (3). Whereas, its Ni-rich version,

LiNi0.80Co0.10Mn0.10O2 (NMC811) delivers much higher specific capacity (200 mAh.g-1 )

upon cycling at 0.1 C rate in a potential range between 3 V- 4.3 V (29). Despite the

capacity and cost advantages, NMC811 shows less thermal stability, which is related

15

to the structure stability. Upon cycling, oxygen release causes structural variation and

thermodynamically stable rock salt LixNi(1-x)O2 structure forms on electrode surface (30).

That increases cell resistance and lowers the cycling stability. Surface coating

techniques (31), doping (32) and core-shell structural design (33) have been proposed to

solve these problems.

• Li-rich (or Mn-rich) NMC [xLi2MnO3. (1-x)LiMO2]

To improve structural stability and specific capacity of NMC, Thackeray et al. (28)

suggest to prepare the solid solution of LiMO2 with electrochemically inactive Li2MnO3

composition. The new lithium rich composition delivers unusual charge capacity

exceeding 250 mAh.g-1 between 2 V and 4.8 V vs Li+/Li. (34,35). Such high capacity is

explained with the presence of cationic (Ni and Co) redox processes up to 4.4 V and

anionic (O2) redox processes between 4.4 V and 4.8 V. Unfortunately, the first charge

capacity could not be recovered at the end of the discharge and low initial columbic

efficiency is observed. That is explained by irreversible oxygen loss (Li2O) from the

Li2MnO3 structure upon delithiation. Besides, such oxygen loss also causes structural

modifications, layered Li2MnO3 changes into spinel LiMnO2 and voltage decay is

observed due to Mn4+ to Mn3+ transitions (36).

• High Voltage Spinel LNMO (LiNi0.5Mn1.5O4)

Spinel type materials are also considered as high energy materials (37–39) due to their

high operating potentials (4.7 V) even though their practical capacity is rather small

(140 mAh.g-1). Moreover, these materials could be easily produced, they show high

rate capability and superior safety behavior. Those properties have already attracted

many research interests. Despite such advantages, cycling at high potential causes

electrolyte decomposition, that later impedes cycling stability (40). Therefore, new

compatible electrolyte must be discovered for further application of these materials. In

addition to these materials, recently, Li-rich disordered rocksalt oxides and sulfides

have been considered as promising high energy positive materials for lithium-ion

batteries (41,42). Their cation disordered rocksalt structure is so robust that less structural

16

problems are expected. Moreover, rocksalt oxides offer capacities exceeding 250

mAh.g-1 at high average potentials (3.6 V), while rocksalt sulfides offer excellent

capacities (>400 mAh.g-1) at relatively low average potentials (2.3 V). Their specific

energy is considerably higher than current materials. Therefore, many researches have

recently been started to develop new Li-rich cation disordered rocksalt compositions.

I.2. Cation Disordered Rocksalts as Positive Materials

I.2.a. The first cation disordered rocksalts: stoichiometric compositions

In 1998, Obrovac et al. discovered cation disordered rocksalts during the preparation

of layer oxides compositions, LiMO2 (where M= Ni, Co, Mn and Fe) with high energy

ball milling (43). Milling of LiCoO2 longer than 16 hours, results in a α-LiFeO2 type

(cation-disordered rocksalt) structure rather than a α-NaFeO2 type (cation-ordered

rocksalt) structure (Figure I - 2). The same trend is observed for all other compositions

(LiNiO2, LiMnO2, LiFeO2 and LiTiO2); longer milling always results in formation of

disordered rocksalt phases. They also observe that disordered structures leads to

smaller discharge capacities (< 70 mAh.g-1) and large irreversibility at the first cycle

(Figure I - 2). The reason of such low capacities is attributed to poor lithium ion

conduction in disordered lattice.

Figure I - 2: X-Ray diffractogram of LiCoO2 and electrochemical performance of LiCoO2 that

is prepared by ball milling (43) at different milling durations

17

Thereafter, Li2MTiO4 (M: Ni, Co, Fe, Mn, V) type disordered rocksalts are reported at

the beginning of 2004 (44-49). To eliminate diffusion limitations in disordered rocksalts,

new synthesis approaches concentrate on low temperature synthesis methods, such as

citric acid assisted sol-gel method (44–46), acid base reactions (47), hydrothermal synthesis

(48) and molten salt synthesis (48,49). With these soft chemistry methods coupled with

carbon coating strategies, average particle size becomes less than 100 nm and electronic

conductivity increases. Thus, much higher practical capacities are reached. For

instance, among Li2MTiO4 compositions, Li2NiTiO4 delivers a charge capacity of 181

mAh.g-1, upon cycling between 2.5 V- 4.8 V (Figure I - 3). However, 50% of charge

capacity is lost at the end of the discharge and only 25% of discharge capacity is

obtained at the end of the 3rd cycle (50). That is explained by irreversible structural

modifications upon cycling.

Figure I - 3: Electrochemical performance of Li2NiTiO4 electrodes that are cycled at a rate of

C/50 between 4.8 V and 2.5 V vs Li+/Li (50) at the first three cycles

Besides, Li2CoTiO4 shows high reversible capacity of 219 mAh.g-1 at relatively low

cycling rate (0.05C), but the large polarization at the first cycle could not be reduced

despite the carbon coating strategies (51).

In brief, the practical capacities of stoichiometric rocksalts are still much lower than

the theoretical capacities. The reason is attributed to poor lithium diffusion through

cation disordered structure. To overcome lithium conduction challenges in cation

18

disordered rocksalts and to understand lithium ion conduction mechanism in these

materials, theoretical calculations are required.

I.2.b. Lithium diffusion mechanism through cation disordered rocksalts

Cation disordered rocksalt crystallises in a face centered cubic (fcc) structure with a

space group of Fm3̅𝑚. In this fcc structure, cations (lithium or transition metals)

statistically occupy 4a positions (½ ½ ½) of octahedral interstitials, whereas anions

occupy 4b positions (0 0 0) as shown in the Figure I - 4.

Figure I - 4: Cation disordered rocksalt structure: face centered cubic structure that contains

lithium (Li), transition metal (TM) and oxygen (O) ions

Previous studies describes the lithium conduction mechanism into fcc crystal lattice

with a theory of Oh-Th-Oh transition in which lithium ions migrate from one

octahedral site (Oh) to another octahedral site with a relaxation at tetrahedral site (Th)

(52,53). Moreover, height of the tetrahedron is reported as an important parameter.

Tetrahedron height should be tall enough that the energy barrier of transition between

Oh-Th sites becomes smaller, thus lithium ions can migrate easily between the sites.

This energy barrier changes based on cation distribution around tetrahedral sites (54).

19

Figure I - 5: Cation arrangements around tetrahedral sites: 0-TM, 1-TM and 2-TM channels

Figure I - 5 shows different cation arrangements around tetrahedral sites, for instance

0-TM represents the tetrahedral sites that are coordinated by only lithium ions without

any transition metals (TM), whereas 1-TM and 2-TM represent the tetrahedral sites

that are coordinated by one and two transition metals respectively.

In 2014, A. Urban et al. perform density functional theory calculations to describe

lithium conduction mechanism in disordered rocksalts (54,55). They describe that both

1-TM and 2-TM channels are inactive for lithium conduction due to strong repulsion

between the cations, whereas 0-TM channels are active for lithium conduction due to

low repulsion between lithium ions (or lithium-vacancy). Thus, energy barrier of

transition between Oh-Th sites is smaller. Besides, they report that increasing lithium

to metal (Li/Me) ratio might create a percolation network through 0-TM channels.

20

Figure I - 6: 0-TM percolation probability (a) and accessible Li content based on Li/Me ratio

(b) in Li-rich disordered rocksalt Li1.211Mo0.467Cr0.3O2 (56)

In another work, they describe the Li/Me threshold as 1.10 to form a percolating

network along 0-TM channels (56). As the material becomes richer in lithium, the

probability for Li+ ions finding an available space in the lattice increases due to the

percolating network formed through the lithium diffusion channels (Figure I - 6). They

also indicate that a new Li-rich rocksalt Li1.211Mo0.467Cr0.3O2 can deliver a reversible

capacity of 300 mAh.g-1 associated to its Li-rich content.

This result perfectly illustrates that:

1. Only reduced capacities are obtained in stoichiometric rocksalts (Li/M=1.0)

because the amount of lithium is not sufficient enough to create percolating

network through lithium diffusion channels.

2. To achieve higher capacities, cation disordered rocksalts should contain excess

lithium amount, hence sufficient lithium conduction paths can form.

21

I.3. Li-rich Cation Disordered Rocksalt Oxides

I.3.a. Li3NbO4 based cation disordered rocksalt oxides

After theoretical calculations reveal the critical Li/Me ratio as 1.10, recent research

interests focus on new Li-rich cation disordered rocksalts that are prepared from

electrochemical inactive ordered rocksalts such as Li2TiO3, Li3NbO4, Li4MoO5 with

transition metal substitution. These materials are not electrochemically active as

cathode materials because they do not have redox active centers, all transition metals

are stabilized at their highest valence state, thus they cannot be oxidized to upper

valence states. For instance, in the case of cation ordered Li3NbO4, which is white

powder and electrochemically inactive, almost no capacity is observed upon cycling

(Figure I - 7).

Figure I - 7: Electrochemical performance of cation ordered Li3NbO4 (57)

Yabuuchi et al. report different overlithiated rocksalts, Li1.3NbxMO2 (where, M: Ni, Fe,

Mn, Co; x=0.3 or 0.43) prepared in disordered rocksalt structure by transition metal

substitution (57). Such materials are synthesized in micronic size and they deliver

smaller capacities (~140 mAh.g-1) based on single electron redox reactions (Me2+/Me3+

or Me3+/Me4+). To increase electronic conductivity, samples are coated with carbon by

mechanical milling, particle size is reduced to submicronic size. Thanks to these

22

improvements, capacities have been increased to 250-300 mAh.g-1, two times higher

than before (57). The compositions are shown in Table I - 1. Among all, Mn doped

Li3NbO4 (Li1.3Nb0.3Mn0.4O2) shows the lowest polarization with a reversible capacity of

260 mAh.g-1 at room temperature (Figure I - 8). Similar composition

(Li1.25Nb0.25Mn0.50O2) is studied by another group as well (58).

Table I - 1: Summary of cation disordered rocksalt oxides reported in the literature

Type Composition T (°C) rate

Initial

charge

capacity

Reversible

Capacity

(mAh.g-1)

Potential

window

(V)

Reference

Li1.30Nb0.30Mn0.40O2 260 260 [57]

Li1.30Nb0.30Fe0.40O2 300 250 [57]

Li1.30Nb0.43Ni0.27O2 280 280 [57]

Li1.30Nb0.43Co0.27O2 310 310 [57]

Li1.30Nb0.30V0.40O2 180 180 4.8 - 1.5 [59]

Li1.30Nb0.30Mn0.40O2 60 10 mA.g-1 380 270 4.8 - 1.5 [57]

Li1.25Nb0.25Mn0.50O2 55 10 mA.g-1 280 290 4.8 - 1.5 [58]

Li2TiO3 - Li2MnO3 Li1.20Ti0.40Mn0.40O2 5 mA.g-1 360 310 4.8 - 1.5 [41]

Li1.20Ti0.40Fe0.40O2 5 mA.g-1 315 210 4.8 - 1.5 [41]

Li1.20Ti0.40Fe0.40O2 10 mA.g-1 350 250 4.8 - 2.0 [60]

Li6/5Ti2/5Mo2/5O2 10 mA.g-1 130 290 4.0 - 1.0 [61]

Li1.20Ni0.333Ti0.333Mo0.133O2 20 mA.g-1 250 230 4.8 - 1.0 [55]

Li3NbO4 - Li3MoO4 Li9/7Nb2/7Mo3/7O2 25 10 mA.g-1 150 280 4.0 - 1.0 [61]

Li3MoO4 Li4/3Mo6+2/9Mo3+

4/9O2 25 10 mA.g-1 180 330 4.0 - 1.0 [61]

Li3MoO4 - LiCrO2 Li1.211Mo0.467Cr0.3O2 25 C/20 280 280 4.3 - 1.5 [56]

Cation Disordered Rocksalt Oxides

50

4.8 - 1.010 mA.g-1

25

Li3NbO4 based

Substitution with

Me3+ and Me2+

Li2TiO3 - Li3MoO4

25

23

Figure I - 8: Electrochemical performances of cation disordered Li1.3Nb0.3Mn0.4O2 cells cycled at

a rate of 10 mA.g-1 between 4.8 V and 1.5 V vs Li+/Li, at 60 °C (57) with SEM image of

Li1.3Nb0.3Mn0.4O2 powder

Despite its high energy density (950 Wh.kg-1 at positive active material level), poor

cyclability is observed at room temperature. When the cells are cycled at 60 °C, inital

charge capacity reaches 380 mAh.g-1 (Figure I - 8). Origin of such a high capacity in

Li1.3Nb0.3Mn0.4O2 electrodes is investigated by X-ray absorption spectroscopy (XAS) (57).

Results indicate that both transition metal and oxide ions play an active role during

cycling, therefore large reversible capacities result from both cationic and anionic

redox processes.

Figure I - 9: Electrochemical performances of cation disordered Li1.3Nb0.3V0.4O2 cells cycled at a

rate of 10 mA.g-1 between 4.8 V and 1.5 V vs Li+/Li (59)

24

The same authors report another disordered rocksalt Li1.3Nb0.3V0.4O2, which is prepared

by substitution of Li3NbO4 with vanadium metal (59). Differently from other

compositions (Table I - 1), this material delivers smaller reversible capacity (170

mAh.g-1) because only cationic redox (V3+ / V4+) is active during cycling (Figure I - 9).

Such result suggests that anionic redox is not always active for all the materials, thus

capacities can vary based on compositions.

I.3.b. Li2TiO3 based cation disordered rocksalt oxides

B. Li et al. studied disordered rocksalt Li2TiO3, which is electrochemically inactive.

Therefore, almost no capacity is obtained at slow cycling rates though it is synthesized

in nanometric size (60). The reason of poor electrochemical activity is attributed to

thermodynamic limitations. To overcome such limitations, substitution of Li2TiO3 with

Fe and Mn is suggested. Thereafter, Fe-substituted Li2TiO3 (Li1.2Ti0.4Fe0.4O2) and Mn-

substituted Li2TiO3 (Li1.2Ti0.4Mn0.4O2) cation disordered rocksalt compositions are

prepared, as illustrated in Figure I - 10 (41,60).

Figure I - 10: Electrochemical performances of Li1.2Ti0.4Fe0.4O2 and Li1.2Ti0.4Mn0.4O2 cells that

are cycled at a rate of 5mA.g-1 between 4.8 V- 1.5 V vs Li+/Li at 50 °C (41)

25

In Li1.2Ti0.4Fe0.4O2 electrodes, initial charge capacity reaches 350 mAh.g-1 and a potential

plateau is observed at 4 V. During discharge, polarization increases, thus discharge

capacity reduces to 250 mAh.g-1. Similar observation is made on Li1.2Ti0.4Mn0.4O2

electrodes: discharge capacity (320 mAh.g-1) is slightly smaller than the charge capacity

(360 mAh.g-1) at the first cycle though subsequent cycles are fully reversible. The

difference between first and subsequent cycles in Li1.2Ti0.4Mn0.4O2 is much smaller

compared to the one in Li1.2Ti0.4Fe0.4O2. The author explains the reason with the

reversibility of anionic redox process. Anionic redox process is almost reversible for

Li1.2Ti0.4Mn0.4O2, whereas it is irreversible in Li1.2Ti0.4Fe0.4O2 electrodes.

In summary, Li1.2Ti0.4X0.4O2 (X: Fe3+, Mn3+) compositions can deliver large reversible

capacities at high operating potentials. However, it should be noted that those results

are obtained at high temperatures (50 °C), thus electrode kinetics is still slow.

I.3.c. Li4MoO5 based cation disordered rocksalt oxides

Another group prepares cation disordered in the Li4MoO5-LiMoO2 system that has the

potential to deliver high capacity based on multi electron (Mo3+ / Mo6+) redox reactions.

As a comparison, Li3NbO4-LiMoO2 and Li2TiO3-LiMoO2 solid solutions are also

prepared (61).

Figure I - 11: Electrochemical performances and capacity retention of Li6/5Ti2/5Mo2/5O2,

Li9/7Nb2/7Mo3/7O2 and Li4/3Mo6+2/9Mo3+4/9O2 cells cycled at a rate of 10mA.g-1 between 4.8 V-

1.0 V at 55 °C (61)

26

Figure I - 11 shows the electrochemical performance of three samples. Among them,

Li4/3Mo6+2/9Mo3+4/9O2 delivers the highest capacity as its lithium content is greater than

Nb and Ti based compositions (61). This result supports the previous observation that

Li-rich content forms percolating network along diffusion channels, thus lithium

conduction increases as well as the capacities (56). Redox activity of the three

compositions are examined with XAS and they confirm that redox activity is based on

pure cationic redox process (Mo3+/Mo6+). Nevertheless, it should be noted that despite

their large reversible capacities (280 - 340 mAh.g-1), operating potential is quite low (~3

V). In addition, the fact that you need to run a first discharge before getting the full

capacity could seriously impede their application in full cells. However, these

materials have good cycling stability, which is attributed to structural stability (61). Ex

situ XRD results reveal that cation disordered structure is rigid, therefore no change in

the crystalline structure is detected and only 1.4% volumetric change is observed at the

end of the second cycle. In brief, Li4MoO5 based cation disordered rock salts can offer

large reversible capacities based on pure cationic redox reactions at relatively low

operating potentials. Their capacity retention is excellent thanks to rigid rocksalt

structure.

27

I.4. Li-rich Cation Disordered Rocksalt Sulfides

I.4.a. Li2TiS3 based cation disordered rocksalt sulfides

In the literature, cation disordered rocksalt sulfides are also studied. For instance, A.

Sakuda et al. mechanochemically synthesize Li2TiS3 powders at different milling

durations (Figure I - 12). Cation disordered rocksalt phase is obtained after 40 h milling

without any remaining precursors (62).

Figure I - 12: (a) X-Ray diffractograms of Li2TiS3 powders milled for different durations, b)

comparison of experimental and simulated diffraction pattern of Li2TiS3 (62)

Figure I - 13: Electrochemical performance of Li2TiS3 electrode at a rate of C/10 between 3.0 V

and 1.5 V vs Li+/Li (62)

Coin cell tests show that Li2TiS3 cells can provide a charge capacity of 274 mAh.g-1 and

discharge capacity of 425 mAh.g-1 based on 1.6 and 2.5 electrons exchanged upon

28

cycling at a rate of C/10 between 3 V and 1.5 V (Figure I - 13). At the second cycle,

reversible 400 mAh.g-1 charge-discharge capacities are obtained at 2.3 V vs Li+/Li.

Capacity retention of Li2TiS3 is compared in both liquid and solid-state electrolyte

systems. When the cells are cycled in carbonate-based electrolytes, they observe that

cycling stability gradually decreases during cycling and 80% capacity is reached at the

end of the 20th cycle (Figure I - 14). The reason of capacity fading in liquid cells is still

unknown. In contrast, superior cycle life is obtained in solid state cells upon cycling at

a rate of 0.5 mA.cm-2. Thus, 98.9 % capacity remains at the end of the 50th cycle (based

on the difference between cycle 5 and 50). Therefore, such materials are suggested to

be used in solid state batteries (63).

Figure I - 14: Capacity retention of Li2TiS3 electrodes (a) in carbonate-based electrolyte at a

rate of C/10 between 3.0 V and 1.5 V vs Li+/Li and (b) in solid state electrolyte at 50°C (62) at

different rates

To discover structural evolution upon cycling, ex situ X-ray diffraction analysis of the

electrodes cycled at different states of charge is performed. Peak intensities of Li2TiS3

reduced and structure became amorphous at the end of the charge, but it returned to

cubic rocksalt structure at the end of discharge (64). Similar observation is reported for

cation disordered Li8FeS5 electrodes as well (65). Cation disordered structure allows

reversible structural change from disordered rocksalt to amorphous phase.

29

I.4.b. Li3NbS4 based cation disordered rock salt sulfides

Same author also reports a disordered rocksalt Li3NbS4 prepared with the same

synthesis recipe (62,66). Disordered rocksalt Li3NbS4 has slightly bigger lattice

parameters (5.13 Å) than Li2TiS3 (5.06 Å).

When the cells are cycled at a rate of C/10 between 3 V – 1 V, a reversible capacity of

486 mAh.g-1 is achieved corresponding to 4.5 exchanged electrons (Figure I - 15).

Cycling curve is reversible in subsequent cycles and its energy density is reported as

high as 923 Wh.kg-1. Li3NbS4 cells show better cycling stability than Li2TiS3 cells, 97%

capacity is reached at the end of the 50th cycle (66). However, average potential of

Li3NbS4 is smaller (2.20 V) than average potential of Li2TiS3 (2.30 V). At high current

rates (1000 mA.g-1), superior cycling stability is observed.

Figure I - 15: Specific capacity vs cell voltage of Li3NbS4 electrodes and capacity retention vs

number of lithium ions exchanged during cycling (66)

Different sulfides, Li2TiS3-Li3NbS4 and Li2SnS3-Li3NbS4 are also reported by the same

author and all of them are synthesized in cation-disordered rocksalt structure (67,68).

Li(8+x)/3Ti(4-4x)/3NbxS4 present reversible capacities higher than 350 mAh.g-1 corresponding

to three electrons process between 3.0 V and 1.5 V (67). On the contrary, tin-substituted

composition, Li3-x/3Nb1-xSn4x/3S delivers smaller capacities with increasing amount of tin

in the composition (68). The reason is explained with low electronic conductivity of tin

30

(7.4x10-7 S.cm-1). The author concludes that sufficient electronic conductivity is

necessary to activate anionic (sulfur) redox process.

Overall, disordered rocksalt sulfides can provide large reversible capacities, however

sufficient electronic conductivity is necessary to activate anionic (sulfur) redox process.

Moreover, capacity fading in the cells containing liquid electrolytes is a major problem

for all disordered rocksalt sulfides. The reason of such a degradation in liquid cells is

still unknown and needs to be elucidated.

I.5. Redox Activity in Disordered Rocksalts

Previously, large capacities delivered by disordered rocksalts were attributed to

activity of both cationic and anionic redox processes. This implies that such processes

should be understood well to design new high-energy chemistries (41,60,69).

Redox activity in a composition can be outlined with the electronic states of the

materials. This can be explained with the metal-ligand band structure (Figure I - 16).

Density of states of a material contains M (metal) at conduction band and L (ligand) at

valence band below energy of fermi level (EF). When purely cationic redox is active,

metal-ligand bands are well separated from each other, therefore there is no

overlapping with these two bands (69). High ionicity between these two bands allows

the metal acting as a redox center and providing necessary electrons during Li

insertion.

Figure I - 16: Cationic redox (Metal (Mn+ d) ligand (Ln- p) band structure) and anionic redox

(Metal (M(n-1)+ d) ligand (L(n+1)- p) band structure)

31

When anionic redox is active, high covalency of metal-ligand band is present. P orbital

of ligand is overlapping with d orbital of metal atom, so ligand transfers electrons, fills

d orbital of metal atom and leaves behind oxidized species (or holes). That means part

of transition metal is reduced to lower valence state, this is called Reductive Coupling

Effect in the literature (70). Such effect is reported for both Li-rich layered (69–74) and

cation-disordered rocksalts (62,73) when anionic redox is activated.

• Anionic Redox in Cation Disordered Rocksalt Oxides

As previously mentioned, Li1.2Ti0.4Mn0.4O2 electrodes delivered large reversible

capacities. To examine such large reversible capacities in Li1.2Ti0.4Mn0.4O2 electrodes,

XAS study is conducted by Li et al. (60). XAS spectra of Ti shows that valence state of

titanium does not change, therefore Ti does not contribute to charge transfer. Whereas,

XAS spectra of Mn and oxygen progressively changed upon charging (Figure I - 17).

This result suggests that charge capacity is produced by both cationic (Mn3+ / Mn4+) and

anionic (O2- / On-) redox activities.

Figure I - 17: XAS spectra of O K-edge and Mn L-edge of Li1.2Ti0.4Mn0.4O2 electrodes during

charge-discharge process (60)

Same study is conducted to examine redox activities in Li1.2Nb0.3Fe0.4O2 electrodes

(Figure I - 18). At the beginning of charge, oxygen species are partially reduced,

whereas iron species are partially oxidized (Fe3+ / Fe4+). However, such reduced oxygen

species disappear at the end of the charge, O K-edge spectra shows similar results for

32

pristine and charge states. This result indicates that anionic redox is activated at the

beginning of the charge, and responsible for the charge transfer. However, the reverse

process is not observed during discharge, indicating anionic redox is not fully

reversible in Li1.2Nb0.3Fe0.4O2 electrodes. Besides, electrode morphology changes at the

end of the discharge (Figure I - 18). This also indicates that such irreversible anionic

redox leads to oxygen release that causes side reactions on the electrode surface as well

as reorganization of particle morphology.

Figure I - 18: Li1.2Nb0.3Fe0.4O2 electrodes at the end of the charge and discharge: Fe L-edge, O

K-edge spectra (41) and TEM images of Li1.2Nb0.3Fe0.4O2 before and after cycling at 50 oC

33

Overall, it seems that anionic redox activity in cation-disordered rocksalt oxides is not

always reversible for all the compositions and irreversible anionic redox can result in

smaller reversible capacities. The reversibility of anionic redox is still unknown,

therefore, further studies are required to elucidate such phenomenon.

• Anionic Redox in Cation Disordered Rocksalt Sulfides

To utilize anionic redox, it is necessary to increase covalency between metal and

ligand, this requires using high electronegative (or less electropositive) metal atom and

less electronegative ligand. For instance, production of persulfide (S22-) is much easier

compared to oxide ions (O22-) because sulfur p orbitals are relatively higher than oxide

p orbitals (69). This creates high covalency between transition metal-sulfur due to

overlapping metal d band and ligand p band, thus anionic redox process is easily

triggered in rocksalt sulfides (69).

Figure I - 19: Coin cell test (a), ex situ X-ray diffraction analysis (b), pair distribution

function analysis of Li2TiS3 electrodes (c) and Ti-S and S-S bonds obtained from ab initio

molecular dynamics calculations (d) (64)

34

Sakuda et al. conducted ex situ X-ray diffraction analysis, pair distribution function

analysis and ab initio molecular dynamics calculations to explain redox activity in

Li2TiS3 electrodes (64). Pair distribution function analysis of electrodes at different states

of charge indicates that at medium range order (r > 5 Å), peak intensities reduces

during lithium extraction and increases during lithium insertion (Figure I - 19c). This

is aligned with ex situ XRD result in which a reversible structural transition is observed

(Figure I - 19b). At short ranges (r < 5 Å) in Figure I - 19c, they observe that covalent S-

S bonds reversibly form and dissociate during charge and discharge correspondingly.

Formation of S-S bonds is attributed to reversible sulfur redox process, sulfur oxidizes

during charge and reduces during discharge. Formation of S-S bonds during charge

are also observed in ab initio molecular dynamics simulations, S atoms lost more than

one lithium neighbor and make covalent S-S bond. Such simulations are done based

on “random” lithium extraction and insertion mechanisms upon cycling. They observe

that atomic distance of Ti-S does not change, S-S distance of S coordinating around Ti

atoms shortens and lengthens during charge and discharge respectively. Such

arrangement allows flexible structural transition from amorphous to rocksalt

structure.

As a summary, anionic redox process is highly reversible in disordered rocksalt

sulfides, flexible structural transition allows such systems to reversibly exchange the

lithium ions.

35

I.6. Conclusion This chapter describes lithium ion technology and current positive electrode materials.

To further increase the specific energy of lithium ion batteries, new high energy Li-rich

and Ni-rich NMC, high-potential spinel LNMO compositions are developed and

extensively studied. Alternatively, Li-rich cation disordered rocksalts have recently

attracted research interests after they have shown promising electrochemical

performances. In this thesis project, we study two different cation disordered families:

rocksalt oxides and sulfides.

As we previously mentioned, Li-rich cation disordered rocksalt oxides can offer large

capacities (≥ 250 mAh.g-1), however there is always irreversibility after first charge. The

origin of such irreversibility is still unknown. To answer this question, we study a

selected composition, Li-rich disordered rocksalt Li2.2NiTi0.2Nb0.6O4. At first, we apply

various synthesis optimizations to improve its electrochemical performance and the

later, we investigate its structural properties with in situ X-ray diffraction technique.

In the second part of the thesis project, we study Li-rich cation disordered rocksalt

Li2TiS3 that have shown unusual electrochemical performances (425 mAh.g-1 at 2.3 V).

Despite its large reversible capacities, strong capacity fading is observed during

cycling. The reason of poor capacity retention is still unknown, therefore our first aim

is to examine degradation problem of Li2TiS3 electrodes with structural (ex situ and

operando neutron diffraction) and morphological (SEM and EDX) studies. To improve

its capacity retention, we propose substitution of Li2TiS3 with selenium, thus we

prepare new compositions (Li2TiSexS3-x), which have shown better capacity retention

than Li2TiS3.

Furthermore, we still lack crucial information about the nature of redox processes in

Li2TiS3 electrodes. Thus, our second aim is to perform electrochemical

characterizations on both Li2TiS3 and Li2TiSexS3-x electrodes. Meanwhile, we perform

surface characterizations on electrodes through X-ray photoelectron spectroscopy to

detect redox active species and to elucidate active redox processes in those materials.

36

Chapter II: Experimental Procedures

This chapter describes the synthesis methods, structural and electrochemical

characterization techniques that are used throughout the thesis. The principles of each

technique, sample preparations and experimental parameters are described.

II.1. Synthesis Methods

Molten salt Synthesis

In the previous chapter, electrochemical performances of disordered rocksalts showed

size dependent properties, thus low temperature synthesis routes were suggested to

prepare these materials. To prepare rocksalt oxides, we select molten salt synthesis

because it allows us preparing ceramic powders at low temperatures. According to the

synthesis procedure, precursors are mixed with an eutectic salt mixture and the

resulting powder is calcined at a temperature above the melting point (Figure II - 1).

During the calcination, the salt melts and acts as the reaction medium, thus reaction

takes place at much lower firing temperatures due to high solubility and diffusivity of

the active species inside the salt (75). After the synthesis, the salt is removed by washing

of the powder with a solvent, such as water.

Figure II - 1: Molten salt synthesis of Li2.2NiTi0.2Nb0.6O4 powder

37

Here are the important parameters for low temperature molten salt synthesis (75):

• Molten salt should be stable, shouldn’t react with precursors during the

synthesis

• Solubility of species in molten salt should be high enough to obtain

targeted phases.

• For low temperature synthesis, molten salt should be selected as it has

low melting point

Mechanochemical Synthesis

Mechanochemical synthesis allows preparing powders in crystalline, meta-stable and

amorphous phases. Among different millers, SPEX and planetary millers are

commonly used in laboratory scale production. In this thesis study, planetary ball

milling is used to prepare disordered rocksalt sulfur-selenium compositions

(Li2TiSexS3-x) that are prepared with a 62:1 (wt%) ball to powder ratio and milled at 510

rpm for 15h, 20h, 25h and 40 h (Figure II - 2). All the samples are prepared in a glovebox

under inert argon atmosphere to avoid air contamination. Appropriate amount of

precursors is weighed and placed into the jar that is then transferred from glovebox to

planetary miller (PM 100) operating under room temperature (Figure II - 2). After the

synthesis, powder is recovered in the glovebox.

Figure II - 2: Mechanochemical synthesis of Li2TiSexS3-x compositions

Rotating disc of the planetary miller induces planetary-earth like movement of the jar

(Figure II - 3). In such a configuration, the jar rotates at its own axes with an inverse

direction of the one of the rotating disc. This creates a centrifugal force that allows the

38

balls to travel freely inside the jar, hit the jar wall as well as each other. As a result,

collision energy is produced. Meanwhile powder particles are trapped between the

balls, consequently they are grinded and rewelded due to mechanically activated

solid-state reactions (76).

Figure II - 3: Mechanochemical synthesis experimental set-up: stainless steel jar, zirconia

balls and planetary miller (PM 100), scheme of grinding medium in PM100

The parameters that has influence on phase and the morphology of the final product

are described below (76):

Grinding medium: Type and size of the balls affect the energy inside the jar, thus phase

of the powder can change. In general, bigger balls create stronger shocks and the

impact energy on the particles becomes greater. Besides, hard materials, such as

stainless steel, hardened steels, tempered steel are frequently used to increase weight

of the grinding medium as they create enough impact energy. In our experiments, we

use zirconium jar with zirconium balls to avoid any reaction between sulfur and steel.

Ball to powder mass ratio: Depending on the ball to powder ratio, energy input can

change. High ball to powder ratio induces high energy inside the jar.

Milling speed: High milling speeds creates high energy input into the powders,

consequently, temperature inside the cell increases. High temperature can be beneficial

to allow the ingredients diffusing enough to trigger chemical reactions. Whereas it

might also cause undesired reactions as well, such as the formation of impurities in the

39

final product if milling takes longer durations. Therefore, powders should be milled at

optimum milling speed.

Milling duration: Milling duration often depends on the milling speed and powder

load (ball to powder ratio). In general, reactions are completed at shorter durations

when the powders are milled at high speeds or at high loads (high ball to powder

ratios) as the input energy is sufficient. It is also noted that milling powders for longer

durations increase the contamination in the final product as the balls often wear out.

II.2. Structural Characterizations

Powder X-ray Diffraction (XRD)

The crystal structure of the synthesized powders is determined by X-ray diffraction

using a BRUKER AXS D8 diffractometer using Cu Kα (λ=1.540 Å) anticathode, the

sample is scanned in the range of 8-80° using a step size of 0.02°. As the sulfides are air

sensitive, all the samples are prepared in the glovebox and the sample surface is

protected with Kapton®, which is a polyimide film, shown in Figure II - 4. Kapton® has

high transmittance to X-rays, however its signal still can be detected at small 2𝜃 angles

(0-28°) in the diffraction patterns of the powders. Besides, its tightness to air, and more

specifically moisture, is limited to couple of hours. Therefore, duration of the X-ray

diffraction is limited to a maximum of three hours.

Figure II - 4: X-ray diffraction analysis: sample holder for powder and electrode

40

X-ray diffraction is a useful technique to determine the crystal structure of the

crystalline powders. Each powder has a unique crystal structure that contains parallel

crystalline planes designated by Miller indexes (hkl) that are separated by “dhkl”

distance called interreticular distance.

Figure II - 5: Scheme of X-ray diffraction technique and Bragg’s law

As the incident X-ray wave interacts with the powder, each plane reflects X-rays

differently, hence this creates a constructive interference. For instance, two different X-

ray waves with a wavelength of λ hit a crystal structure at incident angle θ, first wave

will reflect from top atomic plane, the second wave will reflect from bottom plane

(Figure II - 5). Since the second wave must cross deeper into the bottom plane, it travels

further than the second wave, its distance is much bigger than the first wave. If this

distance is equal to nλ then these two waves are in phase, therefore they constructively

interfere. This verifies Bragg’s law (Eq. II - 1).

Bragg’s law: nλ = 2 dhkl sinθ (Eq. II - 1)

where n: number of X-ray waves, λ is X-ray wavelength, dhkl is distance between the

crystalline planes and θ is the reflected angle.

The most important use of X-ray diffraction method is to identify the materials. Each

material has a characteristic XRD pattern that can be expressed as a fingerprint for

phase identification. There are two important parameters for the phase identification:

peak positions (d-spacing) and peak intensity. Peak positions allow determining the

41

symmetry and cell parameters, whereas peak integrated intensity provides

information about the atomic positions and the occupancies.

X-ray diffractogram provides various information about powders, such as peak

positions, peak profile, peak height and peak width, as illustrated in Figure II - 6.

Figure II - 6: X-ray diffractogram that provides various information about the crystalline

materials

Peak intensities (Ihkl) are proportional to structural factor (Fhkl), it is expressed below

Ihkl = S. jhkl.LPhkl.|Fhkl|2 (Eq. II - 2)

where S: scale factor, jhkl: multiplicity of the reflection, LPhkl: Lorenzian factor and

polarization corresponding diffractometer configuration, Fhkl: structural factor

Structural factor can be defined as

(Eq. II - 3)

Structural factor provides information about the type of the atoms (fj) and the atomic

positions (u, v and w) within the unit cell. Therefore, the target of the structural

analysis is to compare observed and calculated structural factors.

42

The size of the crystalline domains (or crystallite size) can be calculated by using

Scherrer equation (Eq. II - 4). In the equation, β(θ) corresponds to peak widths obtained

from the full-width at half maximum (FWHM) corrected from the instrumental

broadening at corresponding angle (2θ), K is Scherrer constant that varies between

0.62 to 2.08 (0.94 for cubic systems) and λ is X-ray wavelength. The crystallite sizes (L)

can be calculated at each angle, but its mean value is frequently used for a convenience.

𝛽(𝜃) =𝐾𝜆

𝐿𝑐𝑜𝑠𝜃 (Eq. II - 4)

In this thesis project, Fullprof program has been mainly used (77). The peak positions

and peak profiles depend on the crystal structure and the space group

correspondingly. At first, XRD patterns are fitted using Le Bail method in which peak

shape and positions are refined iteratively until minimum Chi-square value (< 5) is

reached. In this way, peak positions are fixed and the unit cell parameters are

calculated with high accuracy. To perform Rietveld refinement and simulated XRD

pattern, structure models (crystal information file - cif) are taken from the databases

(Crystallography Open Database). Rietveld refinement is performed to minimize the

difference between the observed and the calculated data by refining the positions and

occupancies of the atoms in the crystal.

In the case of the in situ analysis, the data are refined using Jana2006 to enable the

correction of the effect of the variable divergence slit (78). The refinement has been

carried using the Fundamental Parameter Approach, in which the instrumental

resolution is calculated from the description of the main parameter of the optical set-

up. This allows a direct correction of the peak profile.

• Ex situ XRD analysis

Ex situ XRD analysis is another way to detect structural evolution in disordered

rocksalts. Coin cells are cycled in the standard procedure that is described in Section

2.3 and stopped at different states of charge. After cycling, cells are dismounted in the

glovebox, electrodes are recuperated from coin cells, they are washed with dimethyl

43

carbonate (DMC) and dried inside the glovebox. After drying, electrodes are placed on

the surface of a glass plate and covered with Kapton® to protect from air atmosphere

(Figure II - 4).

• In situ XRD analysis

To observe structural evolution in disordered Li2.2NiTi0.2Nb0.6O4 during cycling, in situ

XRD analysis is performed using Biologic Instrument and BRUKER diffractometer.

The cell is cycled in GITT (Galvanostatic Intermittent Titration Technique) mode.

Figure II - 7: Scheme of in situ XRD analysis with GITT mode of cycling

The cell is charged at a rate of C/50 for 30 min, then the current is stopped and it follows

a relaxation for 210 min without current pulses. During cell relaxation, XRD analysis

is performed (Figure II - 7). Thereafter, same procedure is applied for discharging the

cell. The loop is arranged until the cell fully charge-discharged.

Figure II - 8: In situ cell assembly

The in situ cell contains 30-40 mg of active material (70wt% Li2.2NiTi0.2Nb0.6O4 and 30%

Carbon black) as positive electrode, glass fiber (Whatman®) disk as separator, LP100 (1

M LiPF6 in EC: PC: DMC= 1:1:3) as non-aqueous electrolyte, Li foil as negative

44

electrode (Figure II - 8). Beryllium, which is transparent to X-rays, is used as current

collector and a piece of aluminum foil (2 µm) is used to protect beryllium from

oxidation at the potentials higher than 4 V. The sample powder is prepared by mixing

70% active materials with 30% carbon, and then dried in a Buchi® oven at 60 °C under

vacuum for 24 hours.

Operando Neutron diffraction

To observe the structural evolutions in the composition of Li2TiS3 and to detect where

extra Li+ ions go into the structure, we perform operando neutron diffraction at Institute

Laue Langevin (ILL) in Grenoble. The measurements are performed on the D20

diffractometer (high resolution configuration) at a wavelength of 1.594 Å between 1°

and 170° with a step of 0.05°. The cell is cycled at a rate of C/10 in a range from 3.5 V

to 1.0 V.

Neutron diffraction is a complementary technique to X-ray diffraction. Differently

from X-rays, neutrons interact with the nucleus of the atoms, instead of electrons. Each

atom has a specific coherent elastic scattering length that can be positive or negative

value. This property allows determining light elements (Li, B, H, C, O and F),

differentiating between the elements that have close atomic numbers as well as the

isotopes in the crystalline structure. Moreover, it is a useful technique to characterize

the materials in operando diffraction mode, both electrochemical and structural

characterizations are simultaneously performed. However, large amount of active

material should be subjected to the neutron beams to get sufficient intense diffraction

peaks. In general, working with the cells containing large amount of powders is often

challenging, it often results in large cell polarization and smaller capacities. Therefore,

electrochemical performance of neutron cell must be optimized beforehand. Besides,

separator and electrolyte in the neutron cell contain hydrogen that presents high

inelastic coherent scattering cross section and thus create a large diffraction

background. To avoid such large background, electrolytes are usually prepared from

deuterated solvents that does not contain any 1H.

45

Figure II - 9: Operando neutron diffraction cell assembly79 and experiment in ILL

The cell used for neutron diffraction is designed by M. Bianchini79 (Figure II - 9). The

bottom part of the cell is made of Ti-Zr alloy. The neutron scattering lengths of titanium

(bTi= -3.37x10-12) and zirconium (bZr= 7.16x10-12) are opposite in sign and almost

related to 2:1. Such alloy is prepared as Ti2.08Zr that becomes transparent to neutrons,

hence it does not contribute to the background of diffraction pattern.

• Cell preparation for operando neutron diffraction

The cell contains 250 mg of sample powder (Li2TiS3), positive electrode, glass fiber

(Whatman®) disk as separator, deuterated electrolyte (1 M LiPF6, d-PC: d-DMC= 1:3,

Eurisotop), which is dried on molecular sieves, and lithium metal as negative electrode

(Figure II - 9). The positive electrode is prepared containing 70% active materials and

30% carbon and it is dried in Buchi® at 60 °C vacuum for 24 hours.

Before the experiment in ILL, we studied and optimized the electrochemistry of Li2TiS3.

At the beginning, we observed a large polarization and poor reversibility when the

cells were cycled in a standard schedule (3 V – 1.5 V). Thereafter, we optimized the

amount of the powder and found an adapted cycling schedule (3.5 V – 1 V).

46

Scanning Electron Microscopy (SEM)

Morphology of synthesized powders are investigated by Scanning Electron

Microscopy equipped with a field electron gun (MEB LEO 1530 Gemini). The surface

of the sample is scanned through focused beam of electrons. As electrons interact with

the atoms, electrons are removed from the inner shell and empty place is filled by

electrons at higher shell energy level, resulting in an energy release. Electrons emitted

by the atoms are collected by detectors and the detected signal is transferred into

information about surface topography. Two different types of electrons are emitted

from the surface: secondary and backscattered electrons. Secondary electrons are

collected close to the specimen surface; hence they create high resolution images.

Backscattered electrons emerge from deeper locations within the specimen; therefore,

they produce low resolution images compared to secondary electrons. As the intensity

of backscattered electron signal depends on the atomic number (Z), they provide

information about the element distribution in the sample.

Energy-Dispersive X-ray Spectroscopy (EDS)

To perform the chemical mapping and determine elements in the compositions, we

used Energy-Dispersive X-ray Spectroscopy using 15 kV in MEB LEO 1530 Gemini.

The surface of the sample is bombarded with focused beam of electrons and emitted

X-rays are collected by the detectors. This provides the information about the element

distribution in a specific region of the sample. Although this technique is easy and

practical to determine the elements in a composition, most of the instruments cannot

detect light elements (Z<10) precisely, thus quantitative analysis of powders is often

difficult.

• Sample preparation for SEM and EDS Analyses

A specific sealed sample holder is used to transfer air sensitive samples from glovebox

to the SEM vacuum chamber. The samples are placed on the surface of adhesive carbon

tape that has been already sticked onto SEM stub (Figure II - 10). Data is recorded for

47

various magnifications (2000, 5000, 10000 and 30000) by using 5 kV accelerating

voltage.

Figure II - 10: Sealed sample holder for SEM of sulfide electrodes

Raman Spectroscopy

To observe chemical bonds in different disordered rocksalt sulfur-selenium

compositions, Raman measurements are carried out on a LabRam HR instrument

(Horiba Jobin-Yvon) in backscattering geometry. 632 nm laser excitation wavelength

on a laser power of 10% filter (≈ 0.17 mW) is used. To prevent heating of the materials

by the laser beam and to obtain a sufficient signal to noise ratio, short acquisition times

(10 s) are used with multiple accumulations (30).

Working mechanism of Raman Spectroscopy is based on inelastic scattering, so called

“Raman Effect”. As the sample is irradiated by photons with a single frequency, the

atoms vibrate and rotate as a response of electromagnetic field that is formed upon

photon-sample interaction. As a result, photons are scattered at different frequencies.

In Raman spectroscopy, one can know the characteristics of each material from

• Raman frequencies, which allows distinguishing the species from each other.

• Width of Raman peaks, which gives information about the quality of crystals

and amount of plastic deformations

• Polarization of Raman peak, which is more related to crystal symmetry and

orientation.

48

Different laser can be used for Raman light source. As the wavelength of the laser gets

smaller, signal quality increases, however it also causes heating of the sample, thus the

sample is more easily damaged.

• Sample preparation for Raman Spectroscopy

Powders are prepared in the glovebox and a sealed sample holder is used for Raman

measurements. The powder is put in the center of the sample holder that is closed by

a cover containing a transparent glass that is invisible to the laser. Sample holder is

transferred from glovebox to the LabRam HR instrument coupled with Atomic Force

Microscopy (AFM) machine. AFM probe is used to select the particle region, and then

laser beam is focused on selected particles.

X-ray Photoelectron Spectroscopy (XPS)

To determine redox active species and to elucidate the redox mechanism in disordered

rocksalt sulfur-selenium compositions (Li2TiSexS3-x), X-ray Photoelectron Spectroscopy

(XPS) is performed using Versaprobe II PHI 5000 (ULVAC-PHI) spectrometer

equipped with a 100 μm focused monochromatic Al Kα X-ray source (1486.6 eV). At

the beginning, powders are analyzed and fitting models are developed. The same

fitting models are applied to the electrodes at different states of the charge.

Principle of XPS is based on the photoelectronic effect in which sample is irradiated by

photons with an energy of hυ and electrons are ejected from the core levels with a

kinetic energy of Ek. The kinetic energies of the ejected electrons are measured, binding

energies are calculated by the spectrometer (Figure II - 11).

Binding energy of core electron (Eb) is expressed by Einstein relationship (Eq. II - 5):

Eb = hѵ –E

k - Ф

sp Eq. II - 5

where Ek: kinetic energy of ejected electron, which is measured by spectrometer, hυ:

energy of photons, Фsp: work function induced by the analyzer.

49

Figure II - 11: Principle of X-ray photoelectron spectroscopy

XPS provides information about

• Chemical composition: Each element has unique core electrons energies. By

measuring the binding energies of core electrons (at K, L or M core levels),

elements present on the sample can be identified. Such elemental analysis is

performed by the “Survey” mode of XPS technique and binding energies can be

found in XPS database.

• Chemical state: The binding energy of a core electron is sensitive to the chemical

environments of the elements. If the same element is bonded to a different

chemical species, the binding energy of its core electron can be different. This

arises from different electronegativities of the atoms in a bond. As the electron

density of valence is shifted to more electronegative (electron-withdrawing)

atom, less energy is required to extract an electron from such electron-enriched

environment. Thus, the kinetic energy of the ejected electrons from the more

electronegative atom increases, but the binding energy of core electron

decreases. Such change in the binding energy results in a shift of core peak at

XPS spectrum. From this observation, the change in the valence states of the

species can be detected by High Resolution XPS technique.

50

• Quantification: As the number of photoelectrons depends on the atomic

concentration of the elements present in the sample, XPS also allows

quantifying the chemical composition. Atomic concentration (Ci) can be

expressed as below, where Ii: peak intensity (peak area after background

removal) and Si: sensivity factor

(Eq. II - 6)

• Sample preparation for XPS analysis

Electrodes after cycling are dismounted in the glovebox, washed with DMC (dimethyl

carbonate) and let to dry overnight. Then electrodes are cut into small pieces and

placed into sample holder. Later, it is transferred into vacuum chamber using a transfer

vessel to avoid any air contamination (Figure II - 12).

Figure II - 12: Sample holder for XPS analysis and the XPS machine

The first aim is to identify the elements present in the samples, therefore survey spectra

is recorded over a spectral range of 0−1200 eV using a pass energy of 117 eV,

corresponding to an energy resolution of 1.6 eV and an acquisition time of 0.5 s.eV-1.

The high-resolution spectral analysis is carried out using a pass energy of 23 eV

allowing an energy resolution of 0.5 eV.

51

II.3. Electrochemical Characterizations

Electrochemical Cell Descriptions

• Preparation of oxide electrodes

The electrodes are prepared according to 80:10:10 formulation: 80% active mass of

material (Li2.2NiTi0.2Nb0.6O4) is mixed with 10% Carbon SP (5321) in a mortar. To have

a bulk electrode and to adjust the viscosity of the slurry, 10% of PVDF (Polyvinylidene

fluoride) binder is added with few drops of NMP (N-methyl-2-pyrrolidone) solvent.

Then, slurry is coated on an Al current collector (20 μm) by using 100 μm doctor blade

and left to dry in the oven at 55 °C for 12 hours. Later, electrodes are cut (14 mm disk

diameter), pressed (10 tons), weighed (loading > 4 mg) and dried in the Buchi® oven at

80 °C for 48 hours under vacuum.

• Preparation of sulfide-selenide based electrodes

Unlike oxide-based electrodes, sulfide and selenide-based electrodes are prepared in a

glovebox under inert argon atmosphere to avoid air contamination. Electrodes are

prepared based on the same formulation (80:10:10) as described previously for

preparation of oxide electrodes. Slurry is coated manually on the Al foil by using 100

µm doctor blade and left to dry in the glovebox for three days. Later, electrodes are cut

(14 mm), pressed (10 tons), weighed (loaded > 2 mg) in the glovebox and dried in the

Buchi® oven at 60 °C for 12h under vacuum.

• Coin cell preparation

The electrochemical characterizations of synthesized powders are performed in coin

cell, which is shown in Figure II - 13. The cells are prepared by using 14 mm diameter

positive electrodes, 16 mm Li foil as a negative electrode and 150 μl electrolyte of LP100

containing 1 M LiPF6 in EC:PC:DMC, where EC: ethylene carbonate; PC: propylene

carbonate; DMC: dimethyl carbonate, in volumetric ratio of 1:1:3 respectively. Two

52

types of separators (Φ = 16.5 mm) are used; polypropylene separator (Celgard® 2400)

and polyolefin separator (Viledon®).

Figure II - 13: Coin cell assembly

Electrochemical Techniques

• Galvanostatic Cycling (GC)

To evaluate the specific capacity, average potential and cyclability of rocksalt oxides

and sulfides, galvanostatic cycling technique is used. This technique also gives

information about the potential-composition relations, such as number of lithium

intercalated, system reversibility and diffusion properties (Figure II - 14). To observe

such properties, common approach is to cycle the cells into fully charged and

discharged states with a constant current rate of C/n, which is related to the

intercalation of lithium ions through the electrodes in n hours. The working electrode

undergoes an oxidation process by applying a positive current to the electrochemical

cell until it reaches the high cut off potential, then it goes through a reduction process

by applying a negative current until it reaches the low cut off potential. Difference

between charged and discharged states gives information about the (ir)reversibility of

the system.

53

Figure II - 14: Galvanostatic cycling

• Galvanostatic cycling schedule

Disordered Li2TiS3 is cycled at a rate of C/10 in a potential range from 3 V to 1.5 V vs

Li+/Li, whereas disordered Li2.2NiTi0.2Nb0.6O4 is cycled at a rate of C/50 in a potential

range of 4.8 V and 2 V. ARBIN cycling instrumentation that operates at 22 °C is used

to cycle the cells. To detect if the materials have diffusion limitations, high temperature

cycling schedule is carried out. In that case, coin cells are placed into an oven that

operates at 55 °C and is connected to the ARBIN instrumentation.

• Galvanostatic Intermittent Titration Technique (GITT)

To study the thermodynamics and kinetics of disordered Li2.2NiTi0.2Nb0.6O4,

Galvanostatic Intermittent Titration Technique (GITT) is used. This technique allows

obtaining galvanostatic profile at equilibrium potential.

In GITT technique, the cell is charged (or discharge) with a small constant current for

a certain time, then the current is interrupted and the cell is given a rest until the

potential is stabilized (Figure II - 15). This procedure is repeated until the cell reached

fully charged or discharged states. With imposed current, lithium is deinserted (or

54

inserted) into the active material, thus potential changes based on concentration

gradient. Without no imposed current, potential reaches its thermodynamically

equilibrium condition with ion transport that occurs from the bulk to the electrodes.

Figure II - 15: GITT mode of cycling

The equilibrium potential is key for various investigations in electrodes. Most of the

active materials undergoes structural modifications upon cycling that adversely affect

the electrochemical performances. With galvanostatic mode of cycling, it is quite

difficult to detect the actual start of such modifications, one can only detect the

modifications in the electrodes at the end of the charge and discharge states. On the

contrary, GITT allows performing X-ray diffraction analysis at the equilibrium

potentials, thus it makes possible to detect step by step the structural modifications

based on lithium (de)insertion process. It also allows determining diffusion and kinetic

properties of electrodes at the equilibrium potentials.

55

• Cycling Voltammetry (CV)

To determine redox potentials, electrode kinetics and reversibility of sulfur-selenium

compositions, cycling voltammetry technique is applied. According to this technique,

cell is submitted to a potential evolving linearly at a specific scan rate (V.s-1) and the

current evolves as a response of potential change. This gives information about the

different redox processes that take place at the electrodes.

Figure II - 16: Cyclic Voltammetry test, parameters obtained in a voltammogram

When an electrochemical cell is scanned with a scan rate (µV.s-1), positive current (iox)

is measured for an oxidation reaction, whereas a negative current (ired) is measured for

reduction reaction in current-potential curve (cyclic voltammogram). Besides,

oxidation (Eox) and reduction (Ered) redox potentials can be identified in cyclic

voltammogram (Figure II - 16). The difference between these two potentials (∆E) gives

indications on the reversibility of the redox reactions.

For a Nernstian system, redox reactions are considered to be reversible, thus the

potential difference is expressed as below

∆𝐸 = 𝐸𝑟𝑒𝑑 − 𝐸𝑜𝑥 +56.5

𝑛 𝑚𝑉 (Eq. II - 7)

where n is the number of electrons exchanged.

56

Figure II - 17: Voltammogram of the electrodes scanned from slow scan rates (5 µV.s-1) to

higher scan rates (100 µV.s-1) with gradual increments, calculation of potential difference for

each scan rates to determine electrode kinetics at slower and higher scan rates

In our experiments, two different cells are cycled with a slow scan rate (5µV.s-1) to

observe lithium (de)insertion process in the sulfur-selenium compositions. To

investigate the electrode kinetics at faster scan rates and reversibility of electrode

kinetics, the cells are scanned from slow scan rates (5 µV.s-1) to higher scan rates (100

µV.s-1) with gradual increments (Figure II - 17). The amplitude of the current increases

with scan rates, indicating the relation between electrode kinetics of the redox

reactions with applied scan rates that is expressed in Randles Sevcik equation for a

reversible process:

Ip= (2.687x105)n3/2v1/2D1/2AC (Eq. II - 8)

Ip is peak current, n is number of transferred electrons, v is scan rate (V.s-1), D is

diffusion coefficient (cm2.s-1), A is electrode area (cm2) and C is concentration (mol.L-1)

By calculating potential difference (∆E) for each scan rate, diffusion coefficient of the

species can be calculated using Randles-Sevcik equation (Eq. II - 8). Thus, the electrode

kinetics of different species can be distinguished from each other.

57

Chapter III: Development of New

Li2.2NiTi0.2Nb0.6O4 Positive Material in

Disordered Rocksalt Structure

This chapter describes molten salt synthesis of Li2.2NiTi0.2Nb0.6O4 as well as structural

and electrochemical characterizations. To improve the capacity and develop a model

material, various synthesis optimizations are carried out. Thereafter, understanding

studies on model material are performed.

Introduction

In the first chapter, we mention that Li-rich disordered rocksalt oxides can provide

large reversible capacities at relatively high operating potentials. After those results,

the study of disordered rocksalts has been started in CEA LITEN. Several Li-rich

disordered rocksalt Li2+1/3*zNiTi1-4/3*zNbzO4 compositions have been prepared with

molten salt synthesis. Such compositions are a derivative of Li2MTiO4 composition,

where M: Ni, Co, Fe, Mn (80–84). Among them, Li2.2NiTi0.2Nb0.6O4 is selected in this thesis

project. The first target of the project is to enhance the electrochemical properties of

Li2.2NiTi0.2Nb0.6O4 with optimizing synthesis parameters to prepare a model material.

The next target is to perform in situ X-ray diffraction analysis on model material to

examine a major problem in cation disordered rocksalt oxides, “irreversibility at the

first cycle”.

Rocksalt Li2.2NiTi0.2Nb0.6O4 is synthesized based on the same recipe as in the literature

(48). Li(CH3COO).4H2O (99.0% sigma), Ni(CH3COO)2.4H2O (99% Acros), TiO2 (99%

Sigma), Nb2O5 (99.0% Acros) are weighed in molar ratio of 2.2:1:0.2:0.3 and put into 50

ml stainless steel jar that contains zirconia balls and mixed at 350 rpm for two hours.

After two hours, eutectic NaCl-KCl salt in a salt ratio to precursor ratio of 4:1 is added

into the mixture and then mixing is repeated for two hours. As the mixing is

58

completed, mixture is heated to 300 °C for 2 hours to remove acetate from the

precursors and later it is calcined at 670 °C for 3 hours. Between two heating steps, the

powder is taken from the oven and grinded a few minutes in mortar to have a

homogeneous mixture. The calcination temperature is slightly above the melting point

of eutectic salt (667 °C). After the synthesis, powder is washed with hot water (60 °C)

to remove the remaining salt, filtrated and dried in the oven.

Figure III - 1: Powder XRD patterns of disordered rocksalt Li2.2NiTi0.2Nb0.6O4 powder: before

(1) and after (2) washing with water

Figure III - 1 shows XRD patterns of Li2.2NiTi0.2Nb0.6O4 powder before and after

washing. In unwashed powder, diffraction peaks of disordered rocksalt

Li2.2NiTi0.2Nb0.6O4, NaCl and KCl are detected. By washing the powder with water,

remaining salt is totally removed and only rocksalt cubic phase is observed. Besides,

sharp diffraction peaks indicate good crystallinity of Li2.2NiTi0.2Nb0.6O4 powder.

Figure III - 2 shows XRD pattern of Li2.2NiTi0.2Nb0.6O4 powder in Fullprof program

profile matching. Li2.2NiTi0.2Nb0.6O4 powder is indexed in a rocksalt (cubic) phase

(space group, Fm3̅m) with lattice parameter, a= 4.1842 Å.

59

Figure III - 2: XRD pattern of disordered rocksalt Li2.2NiTi0.2Nb0.6O4 powder in Fullprof

program profile matching

To investigate the electrochemical properties of rocksalt Li2.2NiTi0.2Nb0.6O4, electrodes

are prepared with the procedure explained in Chapter II.3. Three coin cells are cycled

at a rate of C/50 between 4.8 V and 2.0 V vs Li+/Li at room temperature (22 °C).

The electrochemical redox process of Li2.2NiTi0.2Nb0.6O4 can be described as below:

Li2.2NiTi0.2Nb0.6O4 ⇄ Li2.2-xNiTi0.2Nb0.6O4 + xLi+ + xe- (Eq. III - 1)

Considering only Ni2+/Ni4+ redox is active during cycling, the theoretical capacity based

on 2 moles of Li+ exchanged per formula is 264 mAh.g-1.

In Figure III - 3, disordered rocksalt Li2.2NiTi0.2Nb0.6O4 delivers 120 mAh.g-1 charge and

74 mAh.g-1 discharge capacities, corresponding respectively to 0.90 Li+ and 0.56 Li+

exchanged upon cycling. The irreversible capacity at the first cycle is 38%. On the

contrary, second cycle is fully reversible and less polarized, but a much smaller

capacity of 69 mAh.g-1 is delivered. Moreover, the cycle life at room temperature is

rather poor (Figure III - 4). Those results indicate that synthesis improvement is

required to achieve better specific capacities.

60

Figure III - 3: Cell performances of disordered rocksalt Li2.2NiTi0.2Nb0.6O4 electrodes cycled at a

rate of C/50 between 4.8 V-2.0 V vs Li+/Li at 22 °C

Figure III - 4: Cycling life of disordered rocksalt Li2.2NiTi0.2Nb0.6O4 electrodes cycled at a rate

of C/50 between 4.8 V-2.0 V vs Li+/Li at 22 °C

61

III.1. Synthesis Improvement

To improve electrochemical performance of disordered rocksalt Li2.2NiTi0.2Nb0.6O4

material, different synthesis parameters are considered (Figure III - 5).

Figure III - 5: Summary of the parameters considered for synthesis improvement

• Simple tempering (long firing): Rocksalt Li2.2NiTi0.2Nb0.6O4 is prepared using

four different metal oxides reactants. To prepare homogeneous and well-

crystalline ceramic powder, simple tempering, in another words long firing, can

be applied. By letting the ingredients well diffuse during the synthesis, pure

rocksalt phase can be obtained.

• Quenching: Disordered rocksalts have shown electrochemical performances

dependent on particle size; especially diffusion is more problematic in powders

that contain big particles. With fast cooling approach, particle size growth can

be controlled, thus lithium diffusion length becomes shorter, ionic limitations

might be reduced. Consequently, much greater capacities could be achieved.

• Effect of washing (type of solvent): Rocksalt Li2.2NiTi0.2Nb0.6O4 is prepared by

molten salt synthesis, thus washing step is mandatory to remove remaining

salts in the composition. Water is a stable solvent and solubility of various salts

62

in water is quite high, thus it can dissolve many different salts. However,

washing can cause loss of lithium in the composition, for instance lithium ions

in the powders can exchange with protons from the water. This might reduce

the specific capacity; therefore, new solvents should be searched to eliminate

that problem.

• Effect of molten salt to precursor ratio: Molten salts have strong solving

capability; thus, they act as a medium in which diffusion rate of the precursors

is much higher. Ratio of salt to precursor can affect the precursors solubility;

therefore, the effective ratio should be determined.

• Type of molten salt: Each eutectic salt has a different melting point and

solubility. Use of eutectic salt that has a low melting point can reduce the

synthesis temperature and allows preparing powders in smaller particles size.

However, salt should be stable against reactants, should not react with the

precursors throughout the synthesis and it should have enough solubility

power to dissolve the reactants in order to prepare the desired phases.

• Effect of calcination temperature and duration: Low temperature synthesis

methods are useful to synthesize powders in smaller particles size. Finding

optimum calcination temperature and duration is important to optimize the

capacity of Rocksalt Li2.2NiTi0.2Nb0.6O4.

63

III.1.a. Simple tempering (long firing) and quenching strategies

To improve our synthesis recipe, we first consider quench strategy. For a comparison,

we also prepare a reference material prepared by classical approach. Three samples

are prepared with the same recipe, however different firing and cooling steps are

applied.

• The first sample (classical/reference) is heated at 670 °C for 3 hours and let to

cool down in the oven

• The second sample (quench) is heated at 670 °C for 3 hours and it is taken from

the oven and let to cool down outside the oven (no cooling agent is used)

• The third sample (long calcination and quench) is heated at 670 °C for 6 hours

and it is taken from the oven and let to cool down outside the oven.

Figure III - 6: XRD patterns of Li2.2NiTi0.2Nb0.6O4 powders with different strategies: classical

calcination (3h), calcination (3h) with quench and long calcination (6 h) with quench

After calcination, the three samples are washed with water and X-ray diffraction

patterns are compared in Figure III - 6. All samples present the targeted phase with

small amount of impurities. NaNbO3 is present in every sample and is not affected by

the different thermal treatments. Small amount of Li2CO3 is only found in the slowly

64

cooled samples indicating that the carbonation may occur during cooling. Surprisingly

a peak coming from an ordered rocksalt, isostructural to Li3Ni2NbO6 is found for the

material quenched after a long calcination, although it was thought that this would

favor the disordered phase.

Figure III - 7: Cell performances of Li2.2NiTi0.2Nb0.6O4 electrodes cycled at a rate of C/50

between 4.8 V-2.0 V vs Li+/Li at 22 °C

Figure III - 7 shows cell performances of the samples prepared by different strategies

at the first cycle. The “classical” samples show the worst behavior with a very high

polarization that may rise from the formation of Li2CO3 during the cooling. On the

contrary, quenched samples shows lower polarization. Between the two quenched

samples, the samples that was submitted to the shorter thermal treatment exhibits

much higher capacities with a charge capacity of 144 mAh.g-1 and a discharge capacity

of 105 mAh.g-1 upon cycling at a rate of C/50 between 4.8 V-2.0 V vs Li+/Li. Charge

capacity is almost 15% higher than the charge capacity of classical material, whereas

discharge capacity is 30% higher than discharge capacity of classical material.

Moreover, we observe irreversibility at the first cycle for quenched sample is much

smaller (27%) (Figure III - 7). The lower performances of the sample quenched after 6h

may be attributed to the presence of ordered phase that is known to present much

poorer ionic conductivity.

65

Afterwards, we investigate the electrochemical performances of the best quenched

sample at different temperatures (Figure III - 8). At 55°C, both charge and discharge

capacities improve, a charge capacity of 193 mAh.g-1 and a discharge capacity of 144

mAh.g-1 are delivered. This result indicates that ionic and electronic conductivities of

disordered rocksalt Li2.2NiTi0.2Nb0.6O4 might improve at high temperature,

subsequently higher capacities can be reached. Besides, the irreversibility at first cycle

is 27% at 22 °C and 25% at 55 °C. Cycling at high temperature does not eliminate the

irreversibility at the first cycle.

Figure III - 8: Comparison of cell performances of Li2.2NiTi0.2Nb0.6O4 cells (powder fired 3h and

quenched) at 22 °C and 55°C, electrodes cycled at a rate of C/50 between 4.8 V - 2.0 V vs

Li+/Li

Figure III - 9 shows cycling stability at room (22 °C) and high (55 °C) temperatures.

Cycling stability is rather good at room temperature, but discharge capacity is quite

small (96 mAh.g-1). Whereas, cycling stability at 55 °C is reduced and capacity fades

quickly. Poor cycling stability at high temperatures has been already reported in

different positive materials in the literature and it was attributed to the degradation of

active materials (81,85).

66

Figure III - 9: Evolution of discharge capacities of Li2.2NiTi0.2Nb0.6O4 (quenched) at 22 °C,

electrodes cycled at a rate of C/50 between 4.8 V-2.0 V vs Li+/Li

III.1.b. Effect of washing (type of solvent)

In our experiments, we always observe asymmetric X-ray diffraction peaks after

washing our powders. Therefore, we have another assumption that during washing,

some of the lithium ions might have exchanged with protons. Such protons might

interfere with lithium upon insertion/disinsertion and can produce HF that causes

accelerated degradation of the cell. Therefore, our next target is to change our washing

strategy, finding another solvent. To determine effect of washing to the synthesis, two

powders are prepared separately with the same recipe and they are washed either with

ethylene glycol or with water.

67

Figure III - 10: XRD patterns of Li2.2NiTi0.2Nb0.6O4 powders prepared with the same recipe

separately and washed with different solvents: water and ethylene glycol. Kα2 contributions

have been stripped with EVA software for clarity reason. Arrows indicates possible

contributions from H+ exchanged phase.

Figure III - 10 shows the XRD patterns of two different batches after washing. No

remaining salt is detected in diffraction patterns of washed powders, both ethylene

glycol and water effectively remove eutectic salt in powders. Reflections from the EG-

washed samples are broader and may cover small impurities, but they present a

symmetrical profile.

In Figure III - 11, we observe that charge capacity of powder washed with ethylene

glycol is slightly greater, but its discharge capacity is smaller than the powder washed

with water. Although discharge capacity reduces with ethylene glycol washing, we are

now sure that there is no proton-lithium exchange in our system. Therefore, we decide

to wash the powders with ethylene glycol in further synthesis.

68

Figure III - 11: Half-cell performances of Li2.2NiTi0.2Nb0.6O4 washed with different solvents:

water (1) vs ethylene glycol (2), electrodes cycled at a rate of C/50 between 4.8 V and 2.0 V vs

Li+/Li at 22 °C, the first two cycles

III.1.c. Effect of molten salt to precursor ratio

Besides finding a new washing solvent, we examine the effect of salt to precursor ratio

to the synthesis at the same time. According to our classical approach, salt to precursor

ratio is 4:1 (molar basis), whereas new samples are prepared in lower (2:1) and higher

(10:1) salt to precursor ratios. The prepared mixtures are heated at 300 °C for 2 hours,

calcinated at 670 °C for 3 hours and quenched. The powders are washed with water,

filtered and dried.

Figure III - 12 shows the XRD patterns of the samples that are prepared in different

salt to precursor ratios. In each case, disordered rocksalt Li2.2NiTi0.2Nb0.6O4 phases are

successfully obtained. Such result indicates that even smaller salt to precursor ratio

(2:1) is enough to prepare those phases. However, an impurity phase indexed as

NaNbO3 is observed in diffraction patterns of salt to powder ratios (4:1) and (10:1). In

contrast, we do not observe NaNbO3 impurity in powder synthesized in low salt to

69

precursor ratio (2:1). This might be because smaller amount of salt is used for the

synthesis, such low amount might not be seen in the diffraction patterns.

Figure III - 12: XRD patterns of Li2.2NiTi0.2Nb0.6O4 powders prepared by different salt

(NaCl:KCl) to precursor ratios: 2:1 (1), 4:1 (2) and 10:1 (3)

Figure III - 13: Half-cell performances of Li2.2NiTi0.2Nb0.6O4 prepared by different precursor to

salt ratios: 2:1, 4:1 and 10:1, electrodes cycled at a rate of C/50 between 4.8 V and 2.0 V vs

Li+/Li at 22 °C, at the first cycle

70

As illustrated in Figure III - 13, capacities of Li2.2NiTi0.2Nb0.6O4 powders prepared with

salt to precursor ratios of 4:1 and 2:1 are almost the same, whereas much smaller

charge and discharge capacities are obtained from the powder prepared with the

higher salt to precursor ratio (10:1). Those results support our assumption that eutectic

NaCl-KCl salt reacts with the precursors and produce an impurity phase that blocks

the lithium diffusion and leads to smaller capacities. Therefore, as a next step for the

synthesis improvement, we decide to change the eutectic NaCl-KCl with another

eutectic salt that is more stable and does not react with the reactants during the

synthesis.

III.1.d. Molten salt synthesis with a different eutectic salt (LiCl-KCl)

In the literature, different eutectic salts can be found (Table III - 1). Among them, we

select 0.59LiCl-0.41KCl eutectic salt that has a low melting point (354 °C.) We apply

the same synthesis procedure that was described before, but we just change the type

of molten salt. Instead of using eutectic NaCl-KCl salt, 0.59LiCl-0.41KCl eutectic salt in

a salt to precursor ratio of 2:1 is used to synthesize Li2.2NiTi0.2Nb0.6O4 powders.

Table III - 1: Melting point of salts and eutectic salts

At the beginning, firing at low temperature is considered because the melting point of

the new salt is 354 °C. Two different powders are prepared based on the same recipe.

One sample is heated at 450 °C for 12 hours, thereafter it is quenched. Another is

SaltEutectic

molar ratio

melting

point/°C

LiNO3 255

LiOH 477

LiCl 610

LiBr 552

LiOH-LiCl 0.65 : 0.35 277

LiCl-KCl 0.59 : 0.41 354

LiOH-LiNO3 0.40 : 0.60 243

NaCl 801

KCl 771

NaCl-KCl 1.0 : 1.0 667

71

synthesized with the classical method; it is heated at 300 °C for 2 hours and calcined at

670 °C for 3 hours and quenched. The powders are washed with ethylene glycol,

filtered and dried.

Figure III - 14: XRD patterns of Li2.2NiTi0.2Nb0.6O4 powders prepared with different salts:

eutectic NaCl-KCl and 0.59LiCl-0.41KCl at 450 °C and 670 °C

Figure III - 14 shows XRD patterns of three different powders prepared by eutectic

NaCl-KCl and 0.59LiCl-0.41KCl salts. When the powder is prepared by 0.59LiCl-

0.41KCl salt and heated at 450 °C for 12 hours, no disordered rocksalt phase is formed.

The small peaks detected in diffraction patterns belongs to TiO2 and Nb2O5 precursors.

Such a result indicates that diffusion of titanium and niobium oxide precursors are still

low in the 0.59LiCl-0.41KCl salt, even though melting point of 0.59LiCl-0.41KCl is

lower (354 °C). On the contrary, disordered rocksalt phase is observed in the powder

prepared at 670 °C. That result concludes that much higher temperatures are required

to achieve disordered rocksalt phase.

In Figure III - 15, new sample delivers similar charge capacity (145 mAh.g-1) as classical

material, however the discharge capacity is slightly smaller (95 mAh.g-1).

72

Figure III - 15: Half-cell performance of Li2.2NiTi0.2Nb0.6O4 prepared by different eutectic salts:

NaCl-KCl and 0.59LiCl-0.41KCl, electrodes cycled at a rate of C/50 between 4.8 V and 2.0 V

vs Li+/Li at 22 °C, at the first cycle

Figure III - 16: Morphology of Li2.2NiTi0.2Nb0.6O4 powders synthesized with: NaCl-KCl (1:1)

and 0.59LiCl-0.41KCl salts

73

In Figure III - 16, both powders contain agglomerates of spherical particles. For a

synthesis at the same temperature (670 °C), particles become slightly bigger using

eutectic LiCl-KCl salt for the synthesis. This could be related to the diffusion of species;

ion diffusion rate of reactants could be higher in LiCl-KCl salt melt.

In overall, we successfully synthesize Li2.2NiTi0.2Nb0.6O4 in disordered rocksalt phase

using a different eutectic salt (0.59LiCl-0.41KCl) and the electrochemical performances

are almost similar. Therefore, we decide to continue synthesizing Li2.2NiTi0.2Nb0.6O4

powders with 0.59LiCl-0.41KCl salt.

III.1.e. Effect of calcination temperature and duration

Another approach to improve the capacity of disordered rocksalt Li2.2NiTi0.2Nb0.6O4 is

to synthesize the powders at low temperatures. In classical method, the mixtures are

heated at 300 °C to remove the acetate and decomposed at 670 °C. To investigate the

effect of temperature to the synthesis, two samples are prepared and decomposed at

570 °C and 670 °C separately. To discover the effect of firing duration, three samples

are heated at 670 °C for different durations: standard 3 hours (standard), 1 hour and

0.5 h (30 mins).

Figure III - 17 shows samples calcined at different temperatures and durations. In XRD

pattern of the sample calcined at 570 °C, dissymmetrical diffraction peaks are

observed, indicating that new sample contains another phase. The reason might be

slow diffusion of the reactants at 570 °C. For three powders calcined at 670 °C at

different durations, disordered rocksalt phases are successfully obtained. Such result

indicates that diffusion of reactants is high enough at 670 °C, disordered rocksalt phase

can be achieved even at short firing durations (< 3 hours).

74

Figure III - 17: XRD patterns of Li2.2NiTi0.2Nb0.6O4 powders synthesized different synthesis

temperatures and durations: powder synthesized at 670 °C for 3 hours and at 670 °C for 1

hour, 570 °C 3 hours and 670 °C 30 mins (0.5 hour)

Figure III - 18: Cell performances of Li2.2NiTi0.2Nb0.6O4 synthesized at 670 °C for different

durations: 1 h (a), 3 h (b) and 0.5 h (c)

Figure III - 18 shows cell performances of Li2.2NiTi0.2Nb0.6O4 synthesized at 670 °C for

different durations. For Li2.2NiTi0.2Nb0.6O4 powders synthesized for 1 h, reversible

capacities improve, a charge capacity of 166 mAh.g-1 and a discharge capacity of 108

75

mAh.g-1 are obtained. This might be related to the size of the particles; smaller particles

should have been obtained at shorter firing duration. As the lithium ion diffusion

length is reduced, thus specific capacities are enhanced. However, this is not valid for

Li2.2NiTi0.2Nb0.6O4 powders synthesized for 0.5 h, as it is explained with non-

homogeneous phase formation previously.

Figure III - 19: Cycling life of Li2.2NiTi0.2Nb0.6O4 synthesized at 670 °C for different durations:

3 h, 1 h and 0.5 h

In Figure III - 19, electrodes show similar cycling life, except sample synthesized for

0.5 h. Here, we conclude that cell performance of Li2.2NiTi0.2Nb0.6O4 is improved by

synthesizing the powders at shorter firing duration (1 h).

76

III.2. Understanding Studies

After optimizing the electrochemical performance of Li2.2NiTi0.2Nb0.6O4, structural

studies are conducted to understand why the capacity of this material is limited.

Li2.2NiTi0.2Nb0.6O4 shows high irreversibility at the first cycle, but a good reversibility

for the following cycles. The reason of the irreversibility at the first cycle is unknown.

As previously shown, lithium diffusion in Li2.2NiTi0.2Nb0.6O4 is better at high

temperatures (55 °C) compared to room temperatures (22 °C), we draw the hypothesis

that Li2.2NiTi0.2Nb0.6O4 might have lithium diffusion limitations. At the end of the

charge, Li ions are extracted from the shell of the particles and material’s core remains

richer in lithium (Figure III - 20). During discharge, lithium ions insert into the

structure and fill empty spaces in the material surface. As the Li diffusion is limited,

Li accumulates faster at the surface than in the bulk and may create a diffusion barrier.

Figure III - 20: Hypothesis on lithium diffusion mechanism through Li2.2NiTi0.2Nb0.6O4

To test this hypothesis, the cell is cycled at variable charge potentials beginning from

a low potential of 3.9 V up to 4.8 V with a fixed discharge potential (2 V), as illustrated

in Figure III - 21.

77

Figure III - 21: Cell performance of Li2.2NiTi0.2Nb0.6O4 that was cycled at variable charge

potentials beginning from a low potential of 3.9 V up to 4.8 V with a fixed discharge potential

(2 V) at a rate of C/50 vs Li+/Li at 22 °C

Interestingly, we observe that the irreversibility arises at low state of charge (SOC) and

almost no charge and discharge capacities are obtained upon cycling between 3.9 V -

2 V. At potentials greater than 4.2 V, lithium exchange process becomes more

reversible upon cycling. However, following our hypothesis, we would expect to see

a good reversible behavior at low state of charge and an increase in the irreversible

capacities with increasing SOC. After these findings, we conclude that the result does

not align with our hypothesis. Therefore, the irreversibility might arise from the

structural modifications instead of lithium diffusion limitations. Now, we have

another hypothesis in which there might be a structural change at the end of the

charge. Thus, new structure allows only few lithium ion exchange. This can be further

investigated with ex situ and in situ X-ray diffraction techniques.

78

III.2.a. Ex Situ X-ray Diffraction Analysis

To perform ex situ X-ray diffraction, three electrodes are prepared and one electrode

is charged to 4.8 V while another is discharged to 2 V. Detailed procedure can be

found in Chapter II.2.

Figure III - 22: Ex situ XRD analysis of electrodes at different states of charge: pristine

electrode, electrodes at end of charge (4.8 V) and discharge (2 V)

Figure III - 22 shows ex situ XRD analysis of electrodes at different states of charge.

After aligning aluminum peaks, we observe that both peak intensities and lattice

parameters reduce at the end of charge (4.147 Å). Besides, shape of the diffraction peak

at 63.5° becomes much broader, indicating that charged electrode might contain

another phase. On the contrary, lattice parameters of discharged electrode are slightly

bigger than the pristine electrode (4.187 Å). Additionally, the diffraction pattern of

discharged electrode is quite different from the pristine electrode. This result indicates

that structural change occurring during cycling is not completely reversible. It seems

that final composition might not have the same structure as the original composition.

Thereafter we decide to investigate the change of lattice parameters step by step with

in situ X-ray diffraction.

79

III.2.b. In Situ X-ray Diffraction Analysis

Figure III - 23 shows cycling curve of in situ cell containing Li2.2NiTi0.2Nb0.6O4 powder

that is cycled with GITT mode of cycling (for cell preparation, see Chapter II.2). The

cell is cycled for 0.5 hour, and then current is interrupted. Afterwards, the cell is rested

at 3.5 hours and X-ray diffraction is conducted during relaxation.

Figure III - 23: Cycling curve of in situ cell containing Li2.2NiTi0.2Nb0.6O4 powder, GITT mode

of cycling

Figure III - 24 shows global view in situ XRD diffraction pattern of Li2.2NiTi0.2Nb0.6O4

powder. In diffraction patterns, aluminum and beryllium peaks are identified as well

as the disordered rocksalt phase. For clarity sake, some diffraction patterns (pristine,

end of charge and discharge) are selected and shown in Figure III - 25.

80

Figure III - 24: Global view of in situ XRD diffraction pattern of Li2.2NiTi0.2Nb0.6O4 powder

Focusing on the high 2θ region (62°- 64°) in Figure III - 25, we have similar observations

previously described in ex situ X-ray diffraction section. The peak intensities reduce

and the diffraction pattern of charged powder seems to contain another disordered

rocksalt besides disordered rocksalt Li2.2NiTi0.2Nb0.6O4. To be sure about the second

phase, we decide to refine lattice parameters using Fullprof program profile matching.

81

Figure III - 25: In situ X-ray diffraction pattern of Li2.2NiTi0.2Nb0.6O4 powder at different

states of charge: pristine, end of charge and discharge

At first, we start to do fitting the pattern according to our previous assumption that

second disordered rocksalt phase is progressively forming during charging. However,

we detected a discrepancy in the fitting results, thus we were not able to find a good

fit after few steps of lithiation. Afterwards, we assume that the second phase should

have been at the beginning, in other words initial powder (pristine state) might have

already contained two phases before cycling. The refinement of the initial powder is

then carried with both hypotheses, showing a real improvement when introducing a

second phase (Figure III - 26).

82

Figure III - 26: Results of the refinements of the initial powder. Top: with 1 phase, bottom:

with 2 phases

We find that the two phases have very close lattice parameters (Phase 1: 4.1824 Å,

Phase 2: 4.1848 Å), explaining why this phase has not been detected before. During

charge, the lattice parameters are decreasing in a different way and we observe

different slopes (Figure III - 27a).

83

Figure III - 27: Change of (a) lattice parameters and (b) size of the crystalline domains of

Li2.2NiTi0.2Nb0.6O4 powder during in situ XRD analysis

Lattice parameters increase during discharge, but they are slightly bigger than those

in the pristine states (Figure III - 27a). This result indicates that the final phases are

different from the pristine states. The irreversible structural change may thus be one

of the causes of the irreversible capacity observed during cycling.

Meanwhile, we calculate the size of the crystalline domains in Li2.2NiTi0.2Nb0.6O4

powder using Scherrer equation described in Chapter II.2. In Figure III - 27b,

crystalline sizes are decreasing, indicating that powder pulverizes upon charging. It

seems that disordering, such as dislocations or defects happen in the material. During

discharge, we observe abnormal change, a parabolic distribution of crystalline sizes.

The change in crystalline sizes is irreversible as the final crystalline sizes are much

smaller than those in pristine states. This apparition of defects (either dislocation or

long-range disordering) may also be a cause of the capacity fading of this material.

84

III.3. Conclusion

The objective of this chapter was to develop overlithiated disordered rocksalt

Li2.2NiTi0.2Nb0.6O4 with high capacity. Our initial synthesis has resulted in smaller

charge (120 mAh.g-1) and discharge (74 mAh.g-1) capacities than the theoretical

capacity (264 mAh.g-1). To reach desirable capacities, we optimize our synthesis

procedure considering various synthesis parameters.

Table III - 2: Comparison of initial and final synthesis parameters

After optimization of the synthesis parameters (Table III - 2), we finally succeed to

obtain charge and discharge capacities of 166 mAh.g-1 and 108 mAh.g-1 at room

temperature respectively. However, we observe that there is always a high

irreversibility at the end of the first cycle. At first, we suppose that irreversibility arises

from lithium diffusion problem. After testing our hypothesis, we obtain a result that

does not align with our first hypothesis. Thereafter, we test our second hypothesis that

we consider a possible structural change during charge-discharge.

We observe from ex situ XRD result that there should be a second disordered phase

forming during charge process. To validate this, we also perform in situ XRD analysis

of electrodes at different SOC. Refinement of the initial powder indicates that second

disordered rocksalt phase has already been present in the pristine state. In addition,

we find out that the final phases at the end of the discharge are different from those in

the pristine states. From the study of size in the crystalline domains, we conclude that

disordering happening during cycling causes irreversible structural modifications.

85

In general, Li-rich disordered rocksalt oxides discussed in the literature offer large

reversible capacities due to their high amount of lithium content. Our findings show

that this could be possible if the material has pure disordered structure that can allow

lithium ions exchange reversibly during cycling. It seems that diffusion limitations in

disordered rocksalt oxides have been partly solved with overlithiation, nevertheless

preparing those materials in pure disordered structures might gain importance for

future studies.

86

Chapter IV: Developing Li-rich

Rocksalt Metal Sulfides-Selenides:

Li2TiSexS3-x

The first part of this chapter describes mechanochemical synthesis of Li2TiS3

compound as well as structural, morphological and electrochemical characterizations

of this material. To improve the cycle life of Li2TiS3, we propose selenium substitution

with mechanochemical synthesis. Therefore, second part of the chapter describes

synthesis optimization of sulfides-selenides solid solutions, structural and

morphological characterizations and the electrochemical performance comparison of

new materials with Li2TiS3 material.

Introduction

Metal sulfide family is extensively studied in the literature, especially after layered TiS2

showed unique lithium ion intercalating property (86–88). In parallel to TiS2, layered TiS3

was also studied. Upon discharging the cells, three lithium ions are intercalated into

TiS3 (Eq. IV – 1 and Eq. IV - 2), the higher capacity of 80mAh was reached without

precision on the mass of active material (89–91). Unfortunately, this capacity was not

recovered upon charging the cells because two Li ions were not reversible, thus rapid

capacity fading was observed. This was attributed to the structural instability of

layered TiS3 during cycling (62).

TiS3 + 2Li+ +2e- → Li2TiS3 (Eq. IV - 1)

Li2TiS3 + Li+ + e- →Li3TiS3 (Eq. IV - 2)

In 2014, A. Sakuda et al. prepared Li2TiS3 as a disordered rocksalt phase by

mechanochemical synthesis and they observed that this material could provide 425

mAh.g-1 capacity based on 2.5 Li+ ion reversible exchange process. Beyond its

interesting electrochemical performances, authors reported that the cycling life of this

87

material was poor. Rapid capacity fading was observed in the liquid electrolytes,

whereas 92% capacity was maintained after 80 cycles in solid-state electrolytes.

Therefore, this material was suggested for solid-state batteries rather than for

conventional batteries. In this chapter, our first aim is to develop disordered rocksalt

Li2TiS3 with mechanochemical synthesis. Various structural, morphological and

electrochemical characterizations are performed to understand the properties of

Li2TiS3 compound. To solve cycle life problem, we propose substitution of Li2TiS3 with

selenium. Therefore, this chapter also describes the development of a new synthesis

process and new selenium substituted compositions (Li2TiSexS3-x).

IV.1. Disordered Li2TiS3 as a Positive Electrode

IV.1.a. Mechanochemical Synthesis of Li2TiS3 Powders

Li2TiS3 samples are prepared by using the same synthesis recipe as in the literature (62).

Sulfides are known as air sensitive materials; therefore, all samples are prepared in

argon filled glovebox.

In the literature, Li2TiS3 powders are synthesized with planetary miller at milling speed

of 510 rpm for a range of milling durations (from 20 h to 100 h) (62). Pure disordered

rocksalt phase is obtained for milling durations longer than 40 hours though

considerable amount of zirconium contamination is seen in the X-ray diffractogram.

At the beginning of the thesis, our target is to synthesize Li2TiS3 with the same

synthesis recipe, but we would also like to improve the purity of the powders. As

already mentioned in Chapter II.1, powder milling at longer durations causes

undesired reactions and contaminations in the final product as the balls often wear

out. Thus, we consider synthesizing the powders at shorter durations. This can be done

using slightly bigger zirconium balls (5 mm) and jar (50 ml) than those in the literature

(4 mm and 40 ml) with a fix ball-to jar ratio (62:1, in weight). Our synthesis recipe to

prepare Li2TiS3 powders is the following one: 0.4363 g lithium sulfide (Li2S, Sigma

Aldrich, 98%) and 1.0635 g titanium sulfide (TiS2, Sigma Aldrich, 99.8%) are weighed

88

and placed into 50 ml stainless steel jar that contains 285 zirconia balls (5 mm in

diameter). Then, the jar is closed tightly and transferred from an argon-filled glovebox

to the planetary ball-milling machine. To find optimum milling conditions, Li2TiS3

powders are prepared using various milling durations: 15h, 20h, 25h and 40h at 510

rpm. After the milling, the jar is transferred from planetary ball-milling machine to the

glovebox to avoid the air contamination during powder recovery. Thereafter, samples

are analyzed with XRD to identify the phase of the powders. Sample preparation for

XRD is done in the glovebox and Kapton film is used to protect the sample from air

contaminations. The procedure is described in Chapter II.2.

Figure IV - 1: X-ray diffraction patterns of disordered rocksalt Li2TiS3 powders prepared with

planetary miller at 510 rpm for 15h, 20h, 25h and 40h durations and SEM picture of powder

milled for 20 hours

In Figure IV - 1, XRD patterns of the four different Li2TiS3 powders are quite similar,

almost the same peak positions and intensities are observed. SEM images of Li2TiS3

powder (milled for 20h) are also displayed in Figure IV - 1. As seen in the images, the

powder contains agglomerates that consist of nanometer sized spherical particles.

89

Figure IV - 2: X-ray diffraction pattern of disordered rocksalt Li2TiS3 powder (milled for 20 h)

profile matching in Fullprof program

To find the lattice parameters, X-ray diffraction data of Li2TiS3 powder milled at 20

hours is selected to be used in Fullprof profile matching program (Figure IV - 2). At

first, XRD pattern is fitted to optimize the peak positions and intensities iteratively

until the high accuracy fitting (Chi2<5) is reached. The lattice parameters of Li2TiS3 is

found to be 5.0855 Å, which is very close to the one reported in the literature (62). To

optimize the global peak shape and intensities as well as to refine the lattice

parameters, Rietveld refinement is considered. After searching the structure model

(cif-crystal information file) of Li2TiS3 in databases, we could not find corresponding

cif of Li2TiS3. Afterwards, we decide to use the same Wyckoff positions (WP) of

disordered rocksalt (Fm3̅m) Li2TiO3 for Rietveld refinement. The first refinement is

performed with sof (site occupancy factor) modified using stoichiometry of the

composition (66.6% Li and 33.4% Ti metals and 100% S ligand in Li2TiS3 composition).

Accurate occupancies are obtained after high accuracy fitting (Chi2= 1.479) is reached

(Figure IV - 2). The summary data of Rietveld refinement is illustrated in Table IV - 1.

90

Table IV - 1: Rietveld refinement parameters of Li2TiS3 (WP: Wyckoff positions, sof: site

occupation factor, Biso: atomic displacement parameter), global Chi2 = 1.902

Afterwards, electrochemical performances of the four different samples are tested. The

detailed procedure of slurry coating and coin cell preparation can be found in Chapter

2.3. To have reproducible results, three identical cells are cycled at a rate of C/10

between 3 V and 1.5 V vs Li+/Li.

Figure IV - 3: Half-cell performances of Li2TiS3 cells (powders synthesized at 510 rpm for

different durations) at the first cycle, coin cells cycled at a rate of C/10 between 3.0 V – 1.5 V

Li+/Li at 22 °C

Figure IV - 3 shows the electrochemical performances of the different Li2TiS3 electrodes

at the first cycle and the results are summarized in Table IV - 2. For each material,

discharge capacities are found greater than charge capacities, thus there is always an

extra capacity in Li2TiS3 samples. Discharge average potentials are quite similar.

91

Overall, we select the 20 hours milling condition as optimum milling duration for our

future synthesis though 25 hours milling in Table IV - 2 seems more effective. This

discrepancy arises from the fact that the first synthesis is made with fresh balls at 20

hours milling duration. We later find out that there is not a big difference in terms of

cell performance and capacity retention between milling 20h and 25h when the milling

is operated with used balls.

Table IV - 2: Coin cell results of Li2TiS3 compositions synthesized by ball milling for 15h, 20h,

25h and 40h , where CQ1: first charge capacity, DQ1:first discharge capacity, Vav: average

discharge potential at the first cycle

Table IV - 3: ICP-AES (Inductively coupled plasma atomic emission spectroscopy) of Li2TiS3

powders mechanochemically synthesized at different milling durations. Data is normalized

based on fixed Ti amount (Ti=1.00 mol) in the structure.

To investigate the real stoichiometry and the purity of the final products, Li2TiS3

powders are analyzed with ICP-AES technique and results are shown in Table IV - 3.

After data is normalized based on fixed Ti amount (Ti=1.00 mol), quite diverse atomic

ratios are observed in the powders. We find that powders contain a small trace of

zirconium contamination (0.4 - 4%). The atomic ratios of lithium and sulfur in the

powders increase with milling durations (from 15 hours to 25 hours). However, we

observe that atomic ratios obtained from the sample milled at 40 hours are rather out

15 h 20h 25h 40h

CQ1 (mAh.g-1) 271 300 308 99

DQ1 (mAh.g-1) 334 340 354 267

Vav_discharge (V) 2.23 2.22 2.20 2.10

Li2TiS3

92

the error bar. This may arise from the analysis. ICP samples are dissolved in acidic

aqueous solution, so a part of sulfur may have evolve as H2S and leave the sample

before the analysis.

In summary, four different Li2TiS3 powders are mechanochemically synthesized as

disordered rocksalt phase without any remaining precursors. With our optimized

recipe, we can achieve desired phases at shorter milling durations (< 40 hours) than

those in the literature (62). This can be explained by different size of balls, as slightly

bigger balls are used for the mechanochemical synthesis. As bigger balls create

stronger shocks, the impact energy on the particles increases. This allows reaching

desired phases at shorter milling durations. When the electrochemical performances

of powders are investigated, no direct relation between milling durations and

zirconium contamination as well as the performances is observed. Among all milling

durations, we decide to further synthesize our powders milling for 20 hours.

IV.1.b. Electrochemical Performance of Li2TiS3

Now, we investigate the electrochemical performance of Li2TiS3 electrodes in detail

(Figure IV - 4). At the end of the first charge, 1.50 Li+ ions are extracted from Li2TiS3

and the capacity reaches 250 mAh.g-1.

Figure IV - 4: Half-cell performances and cycling life of Li2TiS3 (unpressed) electrodes cycled

at a rate of C/10 between 3 V - 1.5 V vs Li+/Li: the first (black color), second (red color) and

third (blue color) cycles

93

During the discharge, 2 Li+ ions are inserted into the structure, the composition

changes into Li2.5TiS3 and 338 mAh.g-1 of discharge capacity are delivered. The second

and third cycling curves are fully reversible and extra capacity is observed at the end

of the discharge. That is also observed in different rocksalt sulfide compositions

(Li3NbS4 and Li2SnS3) that show larger discharge capacity than the charge capacity

(92,93). However, the origin of this extra capacity is still unknown. Moreover, Li2TiS3

electrodes show poor cycling stability, rapid capacity fading is observed after three

cycles (Figure IV - 4). Poor cycle life of Li2TiS3 cells is already mentioned in the

literature (62). One approach to improve cycle life of Li2TiS3 electrodes is to study the

effect of electrode pressing. To do so, the electrodes are pressed under 10 tons in the

glovebox and they are tested in coin cells.

Figure IV - 5: Effect of pressing to the cycling life: Half-cell performances of Li2TiS3 electrodes

pressed electrodes and cycling life of pressed and unpressed Li2TiS3 electrodes cycled at a rate

of C/10 between 3 V - 1.5 V vs Li+/Li: the first (black color), second (red color) and third (blue

color) cycles

Figure IV - 5 shows electrochemical performances of pressed Li2TiS3 electrodes under

10 tons in a cold-press machine in the glovebox. Here, pressed electrodes deliver much

higher charge capacity (300 mAh.g-1) compared to the unpressed electrodes (250

mAh.g-1). In addition, reversible 340 mAh.g-1 charge and discharge capacities are

obtained at the end of the second and third cycles. Reduction of electrode porosity by

pressing should increase electronic conductivity, thus reversible capacities improve.

94

Besides, we observe that pressing the electrodes also have a positive impact on cycle

life, which is improved considerably. As the pressed electrodes have improved

electronic conductivity, we expected that their rate capability should improve as well.

To investigate this, two different cells are charged to 3 V at C/10 rate. (Figure IV - 6).

To determine cell behavior at high rates, a signature test is used: the cells are

discharged directly at a rate of 10D, when the discharge cut-off voltage is reached, the

current is then reduced to 5D, and this protocol is then continued with 2D, D, D/2, D/3,

D/5 and finally D/10 consecutive rates to reach a full discharge of the material. Between

two discharge phases, electrodes are rested for two hours. In Figure IV - 6, we observe

that Li2TiS3 electrodes retain 74% of discharge capacity at D rate, 64% capacity at 5D

rate and 51% capacity at 10D rate.

Figure IV - 6: Cycling curve of signature test: Current (mA) vs Test time (h) and rate

capability of pressed Li2TiS3 electrodes that are charged at a rate of C/10 to 3 V and discharged

by different discharge (D) rates to 1.5 V (duration of rest time is two hours between two

discharge phases)

The reason of poor cycle life Li2TiS3 cells may be due to sulfur reaction with the

electrolyte. In the literature, it is reported that lithium sulfur batteries also suffer from

poor cycle life (94–96). That is attributed to reaction of sulfur with carbonate-based

electrolytes, therefore tetraglyme-dioxolane-based electrolytes are suggested (97). In

our experiments, we decide to test Li2TiS3 electrodes in tetraglyme-dioxolane based

electrolytes as well.

95

Figure IV - 7: Capacity and average potential of Li2TiS3 electrodes cycled at a rate of C/10

between 3 V - 1.5 V in two different electrolytes: carbonate-based electrolyte (1 M LiPF6 in

EC:PC:DMC 1:1:3 in volume) and tetraglyme-dioxolane based electrolyte (1M LiTFSI +

0.1M LiNO3 in TEGDME:DIOX 1:1 in volume)

Figure IV - 7 shows Li2TiS3 electrodes cycled in two different electrolytes, such as

carbonate-based electrolyte (1 M LiPF6 in EC:PC:DMC 1:1:3 in volume) and

tetraglyme-dioxolane based electrolyte (1M LiTFSI + 0.1M LiNO3 in TEGDME:DIOX

1:1 in volume). Here we observe that cycle life of Li2TiS3 electrodes in carbonate-based

electrolyte is better than the one in tetraglyme-dioxolane-based electrolyte. Besides,

same charge average (2.46 V) and discharge average (2.23 V) potentials are observed

in the two systems. Thus, we decided to keep testing the electrodes with carbonate-

based electrolyte.

IV.1.c. Air Sensitive Property of Li2TiS3

Li2TiS3 is an air-sensitive material, thus this property slows down the rate of process

from material synthesis to coin-cell preparations because it requires always working

under inert atmosphere, often in an Ar-filled glovebox. For industrial perspective,

material stability is important, especially for upscaling process.

In this part, the air-sensitivity of Li2TiS3 is examined. Five different electrodes are

prepared for XRD analysis in the glovebox, later they are left in a fume hood and a dry

room for one day and one week durations. Here, working under fume hood represents

working under air condition. As the dry room has less humidity (dew point: -40oC)

96

compared to air atmosphere, working in dry room conditions will show oxygen

sensitivity rather than humidity.

Figure IV - 8: X-ray diffraction patterns of Li2TiS3 electrodes at different environments:

pristine electrode, after 1 day in dry room, after 1 week in dry room, after 1 day in fume hood,

after 1 week in fume hood

Figure IV - 8 shows XRD pattern of pristine Li2TiS3 electrodes exposed under different

durations and atmospheres. When the electrodes remain for 1 day in the dry room,

peak intensity at 2θ = 36° decreases, but the crystallinity is conserved. The same is

observed at the electrodes remained for 1 week in the dry room. In contrast, almost no

peaks are detected in electrodes remained under fume hood. We detect various TixOy

compounds in electrodes that remained for 1 week under fume hood, indicating that

sulfur and lithium atoms react with air and they may produce H2S gas and amorphous

Li2CO3 compounds that are not observed in the X-ray diffraction patterns.

Figure IV - 9 and Figure IV - 10 show coin cell test results of Li2TiS3 electrodes exposed

under different durations and atmospheres. Electrodes remaining in dry room deliver

almost similar charge and discharge capacities for the first 8 cycles. Thereafter, small

capacity fading is observed in electrodes remained in the dry room. In contrary,

97

smaller capacities at the first cycle are observed in electrodes remaining in fume hood.

Such results indicate that Li2TiS3 material is more sensitive to humidity and it is rather

stable under oxygen environment. This also indicates the possibility of production and

slurry coating of this material in dry room.

Figure IV - 9: (a) Coin cells tests results of standard Li2TiS3 electrodes (black) in a comparison

of electrodes after 1 day (green) and 1 week (red) in dry room

Figure IV - 10: Coin cells tests results of Li2TiS3 electrodes after 1 day (blue) and 1 week

(black) in fume hood

98

• Lithiation of Li2TiS3 with Butyl-lithium Reaction

As previously mentioned regarding Li2TiS3 electrodes, there is always extra capacity

at the first cycle, thus the final composition is different from the initial composition.

Such compositional deviation disappears at the end of the second cycle, fully reversible

lithium intercalation is observed. To eliminate the irreversibility at the first cycle and

increase charge-discharge capacities, we aim at preparing Li2TiS3 in its more lithiated

version, (Li3TiS3). In early studies on TiS2 and TiS3 systems, Dines et al. reported that

butyl-lithium reaction was an effective way for lithiation of the positive materials (6).

Therefore, lithiation of Li2TiS3 is made by butyl-lithium experiment using the same

procedure as in the literature.

• Experimental procedure of Butyl-lithium reaction

Based on the reaction in Eq. IV - 3, 0.025 mol Li2TiS3 powder, 14.60 ml cyclohexane and

magnetic stirrer are put into a small reactor (20 ml). 0.0375 mol butyl lithium (50%

excess) is added slowly into the reactor, the reactor is closed with its cover and left to

stir at 450 rpm in the glovebox (Figure IV - 11a).

Butyl-lithium reaction:

nC4H9Li + Li2TiS3 → Li3TiS3 + ½ C8H18 (Eq. IV - 3)

Mechanochemical reaction:

3/2Li2S + Ti + 3/2S → Li3TiS3 (Eq. IV - 4)

After stirring for a week, the mixture is transferred into small flasks (2 ml) and put into

a centrifuge (Figure IV - 11b and Figure IV - 11c). They are stirred under 14000 rpm for

20 min, then the powders settle in the bottom, while excess cyclohexane solution is

removed from the flask. To remove the organic impurities, the powder is washed with

anhydrous cyclohexane three times (Figure IV - 11d). Thereafter, it is left to dry for two

days in the glovebox. As a comparison, we also prepare Li3TiS3 samples with

mechanochemical synthesis.

99

Based on reaction in Eq. IV - 4, 1.5 mol of Li2S, 1 mol of Ti and 1.5 mol of S are weighed

and placed into the jar. The jar is transferred into planetary ball milling machine and

milling conditions are set to 510 rpm for 20 hours.

Figure IV - 11: Butyl lithium experimental set-up: (a) stirring for a week, (b) taking the

mixture into the small flask, (c) centrifuge of the samples, (d) washing the powder with

cyclohexane

Figure IV - 12 shows the XRD patterns of Li2+xTiS3 powders prepared by butyl lithium

experiment and ball milling. We observe Li2S precursor in X-ray diffraction patterns of

powder prepared by mechanochemistry. It seems that powder prepared by butyl-

lithium also contain small amount of Li2S, which is a side product. The presence of Li2S

impurity indicates that it is more difficult to stabilize Ti3+ in the disordered rocksalt

structure.

Figure IV - 12: XRD pattern of Li2+xTiS3 comparison of butyl lithium reaction vs ball milling

100

Figure IV - 13: Half-cell performances of LixTiS3 that is lithiated by butyl lithium experiment,

electrodes cycled at a rate of C/10 between 3 V - 1.5 V vs Li+/Li; the first cycle (black color),

second cycle (red color) and third cycle (blue color)

In Figure IV - 13, we observe that Li2+xTiS3 cells show almost fully reversible charge-

discharge curve though capacities are rather smaller (~110 mAh.g-1) than expected.

This may be caused by the unsuccessful coating, because slurry gelified during

preparation. It seems that unreacted Li2S stays on the particle surface and reacts with

PVdF binder.

IV.2. Selenium Substituted Li2TiS3

As already mentioned, the cycling stability of Li2TiS3 is rather poor and therefore it

should be improved. We consider that substitution can be a promising solution. For

instance, sulfur anions can be replaced with an alternative anion. Looking at the

periodic table, selenium comes into prominence as a novel prospected material since

it belongs to the same periodic group (6A), hence it is expected to show similar

elemental activity as sulfur. In the literature, it is also well-known that selenium

substitution improves cycling stability of Li2S cells (98,99). However, before the

synthesis, one must consider the theoretical capacity of new compositions. Theoretical

capacities based on two lithium ions exchanged during cycling are reduced by

101

selenium substitution since atomic mass of selenium (78.97 g.mol-1) is much bigger

than atomic mass of the sulfur (32.06 g.mol-1) (Table IV - 4). Therefore, we first consider

the following compositions Li2TiSeS2 and Li2TiSe2S.

Table IV - 4: Theoretical capacities of Se substituted compositions based on two Li+ ions

exchange process

To synthesize selenium substituted Li2TiS3 samples, necessary precursors are searched

from the producers, however neither Li2Se nor TiSe2 are commercially available.

Therefore, alternative synthesis routes are searched. In the literature, the same

precursor problem is mentioned in the development of Li2Se batteries and powders

are synthesized by liquid synthesis routes that are quite challenging and require more

than one step (98). To make one pot and much faster synthesis, we aim at preparing

new compositions with mechanochemical synthesis using metallic precursors, such as

metallic Ti and Se precursors rather than molecular TiSe2 precursor.

IV.2.a. Synthesis Improvements

Before preparing new materials, it is essential to adapt our previous synthesis of Li2TiS3

to the use of metallic Ti and S precursors instead of TiS2. In this way, we can compare

this new synthesis route with the standard route where there are already enough

structural and electrochemical characterizations. In fact, TiS2 precursor is quite

expensive (22 €.g-1) compared to metallic Ti and S precursors (Ti: 6 €.g-1, S: 2 €.g-1),

therefore using metallic precursors would also bring benefits in terms of

manufacturing costs. However, it should be noted that these prices are obtained from

102

the Sigma Aldrich for lab-scale powders, therefore they may not be representative for

industrial powders.

• Wet Mechanochemical Synthesis of Li2TiS3

At the beginning, we aimed at synthesizing Li2TiS3 powder with Li2S powder, metallic

Ti and S powders. Our four different synthesis trials show that metallic precursors do

not react with Li2S precursor, they are rather deposited on the jar surface. To eliminate

this, we consider milling of the precursors with anhydrous solvent. In our first trial,

we try to synthesize Li2TiS3 through a wet mechanochemical synthesis for 20 h and 25

h milling durations. However, we always detect traces of remaining Li2S precursor

with rocksalt phase in XRD patterns. Pure rocksalt Li2TiS3 phase is finally obtained

when the powder is wet milled for 32 hours with already used balls (100).

Synthesis Procedure in detail: 0.436 g Li2S (98%, Sigma Aldrich), 0.435 g Ti (99.98%,

Alfa Aesar) and 0.609 g S (99.99%, Alfa Aesar) precursors are placed into 50 ml stainless

steel jar that contains 285 zirconia balls (5 mm in diameter). Thereafter, anhydrous

hexane (99.9%, Sigma Aldrich) is poured into the jar until covering the entire ball

surface. Milling is operated at 510 rpm for 32 hours. The milling run time is set as 1

min, but pause time is set as 4 min. After milling, jar is transferred from planetary ball

milling machine to glovebox. To evaporate hexane, jar remains open for 3 days in the

glovebox. Then, powder is recovered in the glovebox.

Figure IV - 14 shows the XRD pattern of Li2TiS3 prepared by wet and dry

mechanochemical syntheses. In the two different synthesis processes, Li2TiS3 powders

are successfully achieved in disordered rocksalt phase. Both powders have similar cell

parameters (awet: 5.0853 Å and adry: 5.0855 Å). Moreover, coin cell test results in Figure

IV - 15 show similar charge and discharge capacities of Li2TiS3 prepared by wet and

dry mechanochemical syntheses. Those results confirm that disordered rocksalt Li2TiS3

can be prepared by a new synthesis process that is more flexible and less expensive.

With this new process, different metal sulfides, Li2MS3 or selenides, Li2MSe3 (M: Fe, Ni,

103

Co, Nb, Mo, etc) and different chalcogenides, Li2TiX3, (X: S, Se, Te, etc.) can be prepared

(100,101).

Figure IV - 14: Comparison of wet and dry mechanochemical syntheses: XRD pattern of

Li2TiS3 powders prepared by wet and dry mechanochemical syntheses

Figure IV - 15: Comparison of half-cell performances of Li2TiS3 cells (powders prepared by wet

and dry mechanochemical syntheses)

104

• Wet Mechanochemical Synthesis of Li2TiSeS2

0.336 g Li2S (98%, Sigma Aldrich), 0.410 g TiS2, 0.175 g Ti (99.98%, Alfa Aesar) and 0.578

g Se (99.9%, Alfa Aesar) precursors are placed into 50 ml stainless steel jar that contains

285 zirconia balls (5 mm in diameter). Then, appropriate amount of anhydrous hexane

(99.9 %, Sigma Aldrich) is added until covering the entire ball surface. Milling is

operated at 510 rpm for 20 hours.

• Wet Mechanochemical Synthesis of Li2TiSe2S

0.274 g Li2S, 0.285 g Ti (99.8%, Alfa Aesar) and 0.941 g Se (99.9%, Alfa Aesar) precursors

are placed into 50 ml stainless steel jar that contains 285 zirconia balls (5 mm in

diameter). Then, appropriate amount of anhydrous hexane (99.9 %, Sigma Aldrich) is

added until covering the entire ball surface. Milling is operated at 510 rpm for 20 hours.

After milling, jar is transferred from planetary ball milling machine to glovebox. To

evaporate hexane, jar remains open for 3 days in the glovebox. Then, powder is

recovered in the glovebox.

Figure IV - 16: XRD patterns of Li2TiSexS3-x powders prepared with wet mechanochemical

synthesis

105

XRD patterns of sulfur selenium solid solutions (Li2TiSexS3-x) in Figure IV - 16 show

that selenium substituted samples are successfully obtained in disordered rocksalt

phase by wet mechanochemical synthesis process. XRD patterns of Li2TiSeS2 and

Li2TiSe2S compositions shift to small 2Ѳ angles, indicating that lattice parameters

expand with selenium substitution since the ionic radius of selenium (Se2-, 198 pm) is

larger than the ionic radius of sulfur (S2-, 184 pm). To confirm if selenium is successfully

substituted into Li2TiS3 material, diffraction patterns of Li2TiSeS2 and Li2TiSe2S

powders should be simulated. Before simulating the diffraction patterns, lattice

parameters are found with Profile Matching in Fullprof program. Patterns are fitted

using the same .pcr file generated for Li2TiS3 powder and the lattice parameters are

found after high accuracy fitting (Chi2<5) is reached (Figure IV - 17 and Figure IV - 18).

Afterwards, Rietveld refinement is done using the modified occupancies based on the

stoichiometry of each compositions. The obtained lattice parameters and occupancies

are summarized in Table IV - 5 and Table IV - 6.

Figure IV - 17: XRD pattern of Li2TiSeS2 profile matching in Fullprof program

106

Table IV - 5: Rietveld refinement parameters of Li2TiSeS2 (WP: Wyckoff positions, sof: site

occupation factor, Biso: atomic displacement parameter)

Figure IV - 18: XRD pattern of Li2TiSe2S profile matching in Fullprof program

Table IV - 6: Rietveld refinement parameters of Li2TiSe2S (WP: Wyckoff positions, sof: site

occupation factor, Biso: atomic displacement parameter)

107

To confirm if selenium is incorporated into the crystal lattice, diffraction patterns are

further simulated. Previously found lattice parameters and occupancies are used to

generate simulated diffraction patterns of Li2TiSeS2 and Li2TiSe2S (Figure IV - 19).

Figure IV - 19: Simulated XRD patterns of Li2TiSexS3-x powders

Figure IV - 20: Experimental and simulated peak intensity ratio of (311) and (222) planes,

I(311) / I(222) for Li2TiS3, Li2TiSeS2 and Li2TiSe2S powders

108

Simulated diffraction patterns in Figure IV - 19 look like the experimental diffraction

patterns (Figure IV - 16). We observe again the shift of diffraction patterns into small

2Ѳ angles as well as the increase in peak intensities with selenium substitution. This is

even more characteristic for doublet at 61°, where peak intensity of (311) plane

increases. To confirm selenium incorporation into crystal lattice, the ratio of peak

intensities at (311) and (222) planes, I(311)/I(222), seen in experimental and simulated

diffraction patterns should be calculated for three samples as the (111) plane is biased

by Kapton®. In Figure IV - 20, we observe that the ratio of I(311)/I(222) is almost the same

for Li2TiSeS2 and Li2TiSe2S powders though it is slightly different for Li2TiS3 powder.

This result supports successful selenium substitution into Li2TiS3 material.

SEM images of Li2TiSe2S and Li2TiSeS2 powders are represented in Figure IV - 21. Like

Li2TiS3 powder, selenium substituted samples contain agglomerates that consist of

spherical particles. It seems that the agglomerates tend to have a plate-like morphology

with selenium substitution.

Figure IV - 21: SEM images of Li2TiSeS2 (left) and Li2TiSe2S (right) powders

109

IV.2.b. Electrochemical performances of Li2TiSexS3-x

• Galvanostatic Tests of Li2TiSexS3-x

The electrochemical performances of Li2TiS3, Li2TiSeS2 and Li2TiSe2S electrodes are

investigated considering the possible redox processes (Figure IV - 22).

Figure IV - 22: Half-cell performances and cycle life of Li2TiS3, Li2TiSeS2 and Li2TiSe2S

electrodes cycled at a rate of C/10 between 3 V – 1.5 V vs Li+/Li at 22 oC

110

Figure IV - 22 shows coin cell tests results at the first three cycles of Li2TiS3 upon cycling

at a rate of C/10 between 3 V – 1.5 V vs Li+/Li. As previously described, 1.77 Li+ ions

are extracted from Li2TiS3 at end of initial charge and 300 mAh.g-1 of capacity is

delivered. Here, we explain the possible redox mechanism based on the assumptions

that all the atoms are stabilized at their valence states (Ti4+ and S2-) in pristine electrode

and there is no loss of sulfur atoms during charge-discharge. During charge, sulfur

redox should be active since Ti is already at his maximum valence of Ti4+. Therefore,

charge capacity should have been produced by anionic (sulfur) redox, some part of the

S2- atoms oxidizes into S22- (Eq. IV - 5). During discharge, 2.0 Li+ ions are inserted into

the structure and the composition changed into Li2.23TiS3. Discharge capacity should

have been either provided by anionic redox or both anionic (S2-/S22-) and cationic

(Ti4+/Ti3+) redox processes (Eq. IV - 6). After the first cycle, cycling curve becomes

reversible and 338 mAh.g-1 of capacity is delivered. As previously mentioned, 80% of

capacity is reached at the end of the 7th cycle.

Charge: Li2Ti4+S2-3 → Li0.23Ti4+ (S2-)1.23(S22-)0.885 + 1.77 Li+ + 1.77 e- (Eq. IV - 5)

Discharge: Li0.23Ti4+ (S2-)1.23(S22-)0.885 + 2 Li+ + 2 e-→ Li2.23 (Ti4+ )x(Ti3+)1-x (S2-)3-y(S22-)y

with 2.(0.885-y)+x =2 (Eq. IV - 6)

Figure IV - 22 also shows coin cells tests results of Li2TiSeS2 upon cycling at a rate of

C/10 between 3 V – 1.5 V vs Li+/Li. Li2TiSeS2 electrodes show reversible charge and

discharge capacities similar to Li2TiS3 electrodes. During charge, 1.39 Li+ ions are

extracted from the structure and capacity is recorded as 179 mAh.g-1. If one holds the

same hypothesis to explain redox process of Li2TiS3, the charge capacity should result

from anionic redox that is either sulfur (S2- / S22-) or selenium (Se2-/ Se22-) redox process.

During discharge, more than 2.0 Li+ ions are inserted into the structure, the

composition changed into Li3.08TiSeS2 and 310 mAh.g-1 is delivered. Discharge capacity

should be provided by either anionic (S2-/S22- or Se2-/Se22-) or both anionic and cationic

(Ti4+/Ti3+) redox process. The second and third cycles are fully reversible similar to

Li2TiS3 and cycling life is relatively better than Li2TiS3, 80% of capacity is reached at the

end of 10th cycle.

111

Coin cells tests results at the first three cycles of Li2TiSe2S upon cycling at a rate of C/10

between 3 V – 1.5 V vs Li+/Li show the same trend that is observed for Li2TiSeS2

electrodes. During charge, 1.32 Li+ ions are extracted from the structure and capacity

is recorded as 149 mAh.g-1. In contrary to Li2TiSeS2, much more lithium ions, 3.44 Li+

ions, are inserted into the cubic rocksalt structure and a much higher discharge

capacity (379 mAh.g-1) is obtained. The final composition becomes as Li(4.17)TiSe2S,

which can be regarded as the average (theoretical) composition. More than 3 Li+ uptake

cannot be explained with the same hypothesis, Ti3+/Ti4+ redox process. Therefore, this

should be investigated with further characterizations. Besides, we observe rapid

capacity fading in Li2TiSe2S cells, 80% of capacity is reached at the end of 2nd cycle.

Table IV - 7: Summary of structural and coin cell test results of Li2TiS3, Li2TiSeS2 and

Li2TiSe2S compositions: lattice parameters and cells cycled in galvanostatic mode at a rate of

C/10 between 3 V – 1.5 V vs Li+/ Li

Table IV - 7 shows a summary of structural and coin cell test results of for Li2TiS3,

Li2TiSeS2 and Li2TiSe2S compositions. For Li2TiS3, charge capacity is smaller, whereas

discharge capacity is equal to theoretical capacity (339 mAh.g-1) since rocksalt structure

allows two Li+ reversible exchange. For Se substituted samples, smaller charge

capacities are obtained at the first cycle. On the contrary, discharge capacities are

greater than the theoretical capacities (261 mAh.g-1 for Li2TiSeS2, 213 mAh.g-1 for

Li2TiSe2S). This indicates the presence of extra capacity at the end of the discharge.

Besides, there is a correlation between degree of Se substitution and the formation of

extra capacity at the end of the discharge. Li2TiSe2S shows the highest discharge

capacity. Given average potentials represent an average of charge and discharge

112

potentials during 10 cycles. Average potentials decrease with selenium substitution,

for instance Li2TiSe2S showed the lowest charge and discharge average potentials

(2.24V and 1.98 V respectively).

• Cyclic Voltammetry Tests of Li2TiSexS3-x

To describe electrochemical processes and identify the electron transfer kinetics, cyclic

voltammetry tests of Li2TiS3, Li2TiSeS2 and Li2TiSe2S cells are conducted at various scan

rates (5 µV.s-1, 10 µV.s-1, 20 µV.s-1, 50 µV.s-1 and 100 µV.s-1) between 3 V – 1.5 V vs Li+/

Li.

Figure IV - 23: Redox potentials comparison of Li2TiS3, Li2TiSeS2 and Li2TiSe2S electrodes at

the first (5µV.s-1) and the second (10 µV.s-1) cycles between 3 V -1.5 V vs Li+/Li

Figure IV - 23 shows the cyclic voltammograms of Li2TiS3, Li2TiSeS2 and Li2TiSe2S cells

at the first cycle at 5 µV.s-1 and the second cycle at 10 µV.s-1. During the first cycle, the

Li2TiS3 cell shows one oxidative and one reductive peak from sulfur (S2-/S22-) redox

reaction at 2.69 V and 2.28 V respectively. Potentials are decreased by selenium

substitution, 2.51 V charge, 2.20 V discharge potentials are observed for Li2TiSeS2,

whereas 2.43 V charge and 2.12 V discharge potentials are observed for Li2TiSe2S. These

results indicate that new compositions have different redox activities. During

discharge, we also observe a second phenomenon that could correspond to the extra

capacity. New reductive peaks at 1.88 V appear during discharge and are more visible

at the second cycle. Besides, we observe that an oxidative peak appears at 2.14 V at the

113

second cycle, indicating that the second redox process is reversible. This new peak

becomes much more defined in Li2TiSe2S cell, it is observed at 2.12 V and 1.88 V.

Figure IV - 24 : Cyclic voltammogram of (a) Li2TiS3, (b) Li2TiSeS2 and (c)Li2TiSe2S electrodes

cycled by various scan rates between 3 V – 1.5 V vs Li+/ Li; peak separation (mV) at each scan

rate (µV.s-1) to determine charge transfer kinetics of sulfur redox (d) and selenium redox (e)

processes

To understand the nature of the electrochemical process, and the electron transfer

kinetics, Li2TiS3, Li2TiSeS2 and Li2TiSe2S cells are scanned from 5 µV.s-1 to 100 µV.s-1 and

the current responses are measured and shown in Figure IV - 24a, Figure IV - 24b and

114

Figure IV - 24c. Here, we observe that peak currents increase by increasing scan rates

corresponding to the relationship between peak current and scan rate expressed in

Randles-Sevcik equation (102,103). In addition, peak potentials shift to lower potentials

with increasing scan rates. This may be related to reversibility or irreversibility of the

electrochemical processes. If we describe a reversible process, peak separation (∆E) is

independent of scan rate (v) and anodic and cathodic peak currents are equal (Ipa = Ipc)

(103). Unlike reversible process, electron transfer kinetics are limited in a quasi-

reversible system, thus the peak potential is a function of scan rate (103). To distinguish

between these two processes (reversible or quasi-reversible), peak potential

separations are measured for the first (sulfide) and the second (selenium) redox

processes and the values are shown in Figure IV - 24d and Figure IV - 24e. In Figure

IV - 24d, we observe that peak separations (∆Esulfur) between 20 µV.s-1 to 100 µV.s-1 are

almost stable for Li2TiSexS3-x cells, it seems that sulfur redox process is reversible. To

confirm this, theoretical peak separations are calculated using Nernst equation that is

shown below (Eq. IV - 7) considering all the electrochemical processes are reversible

(102–104). Here, we calculate the peak potential separations based on two Li+ exchange

process for Li2TiS3, three Li+ exchange process for Li2TiSeS2 and four Li+ exchange

process for Li2TiSe2S.

∆𝐸 = 𝐸𝑝𝑎 − 𝐸𝑝𝑐 +56.5

𝑛 𝑚𝑉 at 298 K (Eq. IV - 7)

As we compare the theoretical and experimental peak potential separations, we find

that experimental peak separation values of both sulfur (∆Esulfur) and selenium

(∆Eselenium) redox processes are much higher than theoretical values (28.2 mV for

Li2TiS3, 18.8 mV for Li2TiSeS2 and 14.1 mV for Li2TiSe2S), indicating that sulfur and

selenium redox processes are not completely reversible. In Figure IV - 24e, we observe

that peak potential separations (∆Eselenium) corresponding to selenium redox process

in Li2TiSexS3-x cells increases with increasing scan rate. On the contrary, peak potential

separations (∆Esulfur) corresponding to sulfur redox process is almost stable at higher

115

scan rates (from 50 µV.s-1 to 100 µV.s-1) in Figure IV - 24d. By comparing the two

different redox processes in (Li2TiSexS3-x) cells, we can conclude that electron transfer

kinetics in selenium process is much more sensitive to the scan rate.

Cyclic voltammetry can also allow determining the diffusion property of Li2TiSexS3-x

cells. The peak potentials at each scan rate can be used to calculate the lithium ion

diffusion coefficients using the relationship between measured current and scan rate,

which is expressed as i=avb (102,103). Here a and b parameters are variables and b

parameter is expressed as a sum of faradaic (diffusion controlled) and non-faradaic

(capacitive/surface controlled) currents (103). By plotting logi vs logv, b parameter can

be obtained from the corresponding slope. According to b parameter, the reaction

kinetics are either surface controlled (b=1) or diffusion-controlled process (b=0.5)

(103,105).

Figure IV - 25: a) Logarithmic and b) linear relation between peak current and scan rate of

Li2TiS3, Li2TiSeS2 and Li2TiSe2S electrodes

In Figure IV - 25a, only Li2TiS3 shows a linear relation (diffusion controlled process)

between peak current and square root of scan rate (slope= 0.5, R² ≥ 0.99) among

Li2TiSexS3-x cells. The relation between peak current and scan rate in Li2TiSeS2 and

Li2TiSe2S cells goes in a nonlinear direction (slope> 0.5). For a diffusion controlled

process, peak current is related to the square root of scan rate by i=v1/2, thus it allows

applying Randles-Sevcik equation102,103 (Eq. IV - 8)

116

𝐼𝑝 = (2.69 𝑥 105)𝑛3/2𝐴𝐷01/2

𝐶0𝑉1/2 (Eq. IV - 8)

where Ip is peak current in A, n the number of exchanged electrons, A the active surface

in cm2, Co the lithium concentration in mol.cm3- and V the scan rate in V.s-1.

Figure IV - 25b also shows the linear relation between peak current and scan rate with

the corresponding regression line (R²) for Li2TiS3. The concentration (Co) is calculated

as 0.06 Li+ mol.cm-3 from the unit cell volume and we consider two Li+ ions are

exchanged during charge-discharge. Using these variables in Eq. IV - 8, anodic

diffusion coefficient (Do) of Li2TiS3 is calculated as 1.99x10-9cm2.s-1. Lithium diffusion

kinetics of different sulfurous compounds are studied in the literature and the

diffusion coefficients are reported between 10-9 - 10-8 (106–108). Therefore, DLi+ that we

found is close to the reported values.

We further investigate capacity retention of Li2TiS3, Li2TiSeS2 and Li2TiSe2S cells cycled

at different scan rates (Table IV - 8). By ignoring the fact that the scan rates are quite

slow (µV.s-1), Li2TiSeS2 and Li2TiSe2S electrodes retain the capacities in a better way

than the Li2TiS3 cells at high rates. Although such a result is quite different from what

is observed in galvanostatic tests, it also supports our previous observation in which

electron transfer kinetics in Li2TiSeS2 and Li2TiSe2S compositions are much more

sensitive to scan rate.

Table IV - 8: Cyclic voltammetry test results of Li2TiS3, Li2TiSeS2 and Li2TiSe2S cells cycled at

different cycling rates between 3 V – 1.5 V vs Li+/ Li

117

• Cycle life improvement with different cycling schedules

The goal of the selenium substitution was to improve the cycling stability of Li2TiS3,

therefore, cycling stabilities of Li2TiSeS2 and Li2TiSe2S should be compared with Li2TiS3.

Figure IV - 26 shows capacity retention of Li2TiS3, Li2TiSeS2 and Li2TiSe2S electrodes

cycled in galvanostatic mode at different discharge cut-off potentials. Cycling the cells

between 3 V - 1.5 V vs Li+/Li, Li2TiSeS2 cells show better cycling stability than Li2TiS3

cells, whereas Li2TiSe2S cells show rapid capacity fading. At higher cut-off discharge

potentials (3 V - 1.8 V) and (3 V - 2 V), all selenium substituted cells (Li2TiSeS2 and

Li2TiSe2S) show better cycling stability than Li2TiS3. To discover why Li2TiSe2S cells fail

upon cycling between (3 V - 1.5 V), we perform ex situ XRD and SEM analyses of

electrodes at different states of charge.

Figure IV - 26: Capacity retention (%) of Li2TiS3, Li2TiSeS2 and Li2TiSe2S electrodes cycled in

galvanostatic mode at different cycling discharge cut-off potentials: 3 V - 1.5 V, 3 V - 1.8 V

and 3 V - 2.0 V vs Li+/Li.

118

Figure IV - 27: SEM images of Li2TiSe2S electrodes cycled at a rate of C/10 between (3 V - 1.5

V) and (3 V - 1.8 V)

Figure IV - 28: X-ray diffraction pattern of Li2TiSe2S electrodes cycled at a rate of C/10

between (3 V - 1.5 V) and (3 V - 1.8 V)

119

In Figure IV - 27, we observe porous morphology in the electrodes cycled at 3 V - 1.5

V. The same porous morphology is also seen in the electrodes cycled at 3 V - 1.8 V, but

it is slightly eliminated by increasing the cut-off discharge potential.

In Figure IV - 28, we detect disordered rocksalt phase containing Tix(Se-S)y type of

impurities at two different electrodes. In summary, the change in electrode

morphology and presence of Tix(Se-S)y type of impurities indicate that parasitic

reactions possibly occur during discharging the cell.

We also test a different cycling schedule in which electrodes are first discharged to 1.5

V and then charged to 3 V (Figure IV - 29). Only limited capacities are obtained upon

discharging Li2TiS3 cells, while large discharge capacities are obtained in both Li2TiSeS2

and Li2TiSe2S electrodes upon discharge. Such result indicates that Se substituted

compositions might not be pure, thus they may contain some amount of metallic

selenium that was used during the mechanochemical synthesis. This is investigated

further in the next chapter.

Figure IV - 29: Half-cell performances of Li2TiS3, Li2TiSeS2 and Li2TiSe2S electrodes cycled at

a rate of C/10 between 1.5 V- 3.0 V vs Li+/Li

120

IV.3. Conclusion

In this chapter, Li2TiS3 powder is synthesized mechanochemically in rocksalt phase

and it shows reversible charge and discharge capacities of 339 mAh.g-1 based on two

lithium ions exchange process. To understand degradation process of Li2TiS3 in liquid

cells, various electrochemical and structural characterizations are performed. To

improve cycling stability of Li2TiS3 cells, selenium substitution is considered. Applying

the same synthesis recipe, selenium-substituted samples could not be synthesized,

thus the synthesis recipe had to be improved. After various synthesis strategies, a new

synthesis process is successfully developed and patented (100,101).

Selenium-substituted samples (Li2TiSeS2 and Li2TiSe2S) are synthesized in disordered

rocksalt phase with new synthesis process. To understand their cell performance,

cycling stability and electron transfer kinetics, galvanostatic and voltammetry tests are

conducted. Results show that Li2TiSeS2 and Li2TiSe2S provide larger discharge

capacities than Li2TiS3 cells. Cycling voltammetry test of selenium-substituted samples

at various scan rates show that oxidation and reduction potentials decrease by

selenium substitution. At high scan rates in voltammetry, selenium-substituted

samples show better capacity retention than Li2TiS3 cells. We also find that electron

transfer kinetics of selenium-substituted samples are more sensitive to the scan rate.

Afterwards, electrochemical performances of Li2TiSexS3-x electrodes are investigated at

different cycling schedules. Capacity retention of Li2TiSe2S cells is poor upon cycling

between 3 V - 1.5 V; however, we observe that it improves upon cycling between 3 V -

1.8 V and 3 V - 2.0 V. By examining the cycled electrodes with ex situ XRD and SEM

analyses, we find that Li2TiSe2S electrodes degrade during discharging the cells.

Nevertheless, increasing discharge cut-off potential might limit the parasitic reactions

that forms upon discharge.

Finally, Li2TiSexS3-x electrodes are also tested between 1.5 V – 3 V. Only limited

discharge capacities are detected in Li2TiS3 cells, whereas large discharge capacities are

121

obtained in Li2TiSeS2 and Li2TiSe2S cells. Such results indicate that Se substituted

samples may not be pure, thus may contain considerable amount of metallic selenium

precursor. This is investigated further in the next chapter.

Despite detailed characterization of Li2TiSexS3-x compounds, we still do not know the

redox active species and possible redox mechanisms. To discover such properties, we

further characterize these compounds with X-ray Photoelectron Spectroscopy (XPS),

Raman Spectroscopy and neutron diffraction in the following chapter.

122

Chapter V: Characterization of

Li2TiSexS3-x Compositions

In previous chapter, our new rocksalt sulfide-selenide compositions show unusual

electrochemical performances. We observe large discharge capacities in Se substituted

samples at the first cycle. This could be due to extra Li uptake at the end of the

discharge. Cubic rocksalt structure is dense and all the octahedral sites are already

filled with both Ti4+ and Li+ cations, however it is still questionable where the extra

lithium ions go in the structure. We also observe that cycling stability of Se substituted

samples is rather poor during charge-discharge. Therefore, our next target is to find

what impedes the capacity during cycling.

This chapter describes electrochemical and structural characterization of sulfur-

selenium compositions (Li2TiSexS3-x). Operando neutron diffraction and ex situ X-ray

diffraction are described to detect Li ions in the rocksalt structure. Afterwards, results

of XPS analysis are presented to elucidate the electrochemical activity in those

materials.

V.1. Operando Neutron Diffraction of Li2TiS3

Operando neutron diffraction is performed at ILL, Grenoble. We use a special cell,

“neutron cell” made of Ti2.08Zr alloy, which is transparent to neutron diffraction. The

experimental procedure is described in Chapter 2.2.b.

Before performing operando diffraction experiment, Li2TiS3 powder (≈ 120 mg) is

analyzed with ex situ neutron diffraction in ILL. The target is to collect the diffraction

patterns and to determine the quality of data before the real experiment. In ex situ

neutron diffraction pattern, we observe a large background with very small diffraction

peaks. We conclude that the powder amount is too low to obtain sufficient diffraction

peaks. Afterwards, we decide to increase powder amount (200 mg) to get sufficient

123

intense diffraction peaks during the real experiment. However, we also know that

working with the cells containing large amount of powders is quite challenging, it

often results in large cell polarization with smaller capacities. Therefore, our first goal

is to optimize the electrochemical performances of the neutron cell by applying

different cell testing schedules (Figure V - 1).

Figure V - 1: Cycling of operando neutron cell cycled at a rate of C/10: a) Cell is charged to 3

V and discharged to 1.5 V b) Cell is discharged to 1.5 V and charged to 3 V, c) cell is

discharged to 1.5 V and charged to 3.5 V d) cell is charged to 3.5 V and discharged to 1.0 V

At the beginning, we prepare neutron cell as it is described in Chapter 2.b., for instance

the cell is tested at a rate of C/10 between 3 V and 1.5 V. However, it cannot be charged

to 3 V and a wide potential plateau at 2.6 V with a large polarization (0.68 V) is

observed (Figure V - 1a). We suspect that the parasitic plateau at 2.6 V might be due to

the moisture and a parafilm that is added around the nut to improve the cell tightness.

124

As our main interest in this experiment is to detect extra capacity and to find out where

the extra lithium locates in the structure, we are more interested in discharge state. To

produce similar results, the neutron cell is first discharged to 1.5 V and charged to 3 V

(Figure V - 1b). Thereafter, the high cut-off potential is increased to 3.5 V and we finally

obtain a reversible cycling curve, as shown in Figure V - 1c. Nevertheless, we do not

observe the extra capacity during discharge. We further decrease the low cut-off

potential and the cell is charged first to 3.5 V and discharged to 1.0 V. After such

optimizations, we succeed to obtain reversible cycling curve and observe the extra

capacity during discharge.

Figure V - 2: Operando neutron diffraction pattern of Li2TiS3 powder: pristine powder,

discharged powder and empty neutron diffraction cell (neutron wavelength: 1.594 Å)

Figure V - 3: Simulated neutron diffraction pattern of Li2TiS3

125

Figure V - 2 shows experimental and Figure V - 3 shows simulated operando neutron

diffraction patterns of Li2TiS3 powders. Based on simulated diffraction pattern, we

should see a strong peak at 31°. However, in Figure V - 2, we observe that the first peak

at 31° is lost in the background coming from the cell itself. We could only see a small

peak at 62°, but it is difficult to make a conclusion based on this result. Differently from

the simulated diffraction pattern, two additional peaks at 47 Å and 56 Å in both

pristine and discharge states are seen in Figure V - 2. Those peaks should arise from

the separator within the cell.

In conclusion, we cannot not successfully detect Li+ in the cubic rocksalt structure

because the reflections coming from the material are much weaker than the

background produced by the neutron cell, although 200 mg of powder is loaded in the

cell. Nevertheless, we decide to investigate our another question concerning what

impedes the capacity during cycling and to do so, we perform ex situ XRD study on

electrodes at different states of charge.

126

V.2. Understanding Degradation: Study of the Structural

Property by ex situ XRD

V.2.a. Ex situ XRD and SEM Analyses of Li2TiS3 electrodes

Figure V - 4 shows cycling curve corresponding to the studied electrodes and Figure

V - 5 shows ex situ XRD of Li2TiS3 electrodes at the different states of charge.

Figure V - 4: Cell performances of Li2TiS3 electrodes at different states of charge: electrode at

the end of charge (3 V) and electrode at the end of discharge (1.6 V)

Figure V - 5: Ex situ X-Ray diffractograms of Li2TiS3 at different states of charge: pristine, at

the end of the charge (3 V) and at the end of the discharge (1.6 V)

127

Pristine Li2TiS3 electrode has a rocksalt (cubic) structure. At the end of the charge,

lithium ions deinsert from Li2TiS3 electrode and peak intensities reduce, rocksalt

structure becomes amorphous. The rocksalt structure is recovered after lithium

uptakes at the end of the discharge. Those results are consistent with the literature

where reversible structural change has been observed (62). In addition, authors indicate

that S-S bonds lengthen and shorten during charge and discharge, whereas Ti-S bonds

do not change at all (64). Such rearrangements allow flexible structural change from

crystalline to amorphous and crystalline structure.

Figure V - 6 shows SEM images of Li2TiS3 electrodes cycled at different states of charge.

In Li2TiS3 electrode at the end of the charge, expansion is obvious and some degree of

porosities are observed. However, the morphology of Li2TiS3 electrode at the end of the

discharge is quite similar to Li2TiS3 pristine electrode. Overall, we observe reversible

structural change and detect similar electrode morphologies after charge and

discharge of Li2TiS3 electrodes.

Figure V - 6: SEM images of Li2TiS3 electrodes at different states of charge: pristine electrode,

charged (3 V) and discharged (1.6 V) electrode

128

V.2.b. Ex situ XRD and SEM Analyses of Li2TiSexS3-x electrodes

Figure V - 7 cycling curve corresponding to the studied electrodes and Figure V - 8

shows ex situ XRD of Li2TiSeS2 electrodes at the different states of charge.

Figure V - 7: Cell performances of Li2TiSeS2 electrodes at different states of charge: electrode at

the end of charge (3 V) and electrode at the end of discharge (1.6 V)

Figure V - 8: Ex situ X-Ray diffractograms of Li2TiSeS2 electrodes at different states of charge:

pristine, at the end of charge (3 V) and at the end of discharge (1.6 V)

As shown in Figure V - 8, reversible structural change is also detected in Li2TiSeS2

electrodes. Differently from Li2TiS3, metallic selenium is observed in Li2TiSeS2

129

electrodes at the end of the charge. Thereafter, structure recrystallizes in a cubic

structure at the end of the discharge and metallic Se disappears in Li2TiSeS2 electrode.

Li2TiSe2S electrodes are also tested at different states of charge (Figure V - 9). The same

phenomenon is also observed in Li2TiSe2S electrodes, metallic selenium is detected in

charged electrodes and it disappears in discharged electrodes (Figure V - 10).

Figure V - 9: Cell performances of Li2TiSe2S electrodes at different states of charge: electrode at

the end of charge (3 V) and electrode at the end of discharge (1.6 V)

Figure V - 10: Ex situ X-Ray diffractograms of Li2TiSe2S electrodes at different states of

charge: pristine, at the end of charge (3 V) and at the end of discharge (1.6 V)

130

When Li2TiSe2S electrode at the end of the charge is studied by SEM, (Figure V - 11), a

columnar phase is observed, whereas only one phase is observed in Li2TiSe2S

electrodes at the end of the discharge. This is in accordance to the results of ex situ XRD

analysis, metallic selenium (Se0) could be formed at the end of the charge and

converted into soluble polyselenides containing Se2- at the end of the discharge.

Therefore, only disordered rocksalt structure is observed in the XRD pattern. Li2TiSe2S

electrode at the end of the discharge shows a porous structure that is different from

Li2TiS3 and Li2TiSeS2 electrodes. That might be because selenium has dissolved into the

electrolyte and left empty space in the electrode.

Figure V - 11: SEM images of Li2TiS3, Li2TiSeS2 and Li2TiSe2S electrodes at different states of

charge: pristine, at the end of charge (3 V) and at the end of discharge (1.6 V)

131

Figure V - 12 shows SEM and EDX images of Li2TiSe2S electrodes. After charging the

electrodes at 3 V for 24 hours, we observe a columnar phase that has crystallized in a

different shape. EDX analysis shows that 92% of Se, 4% Ti and 4% S elements are

present in the columnar phase. This result also supports previous hypothesis

concerning formation of metallic selenium during charge process.

Figure V - 12: a) SEM images and b) EDX images of Li2TiSe2S electrodes cycled to 3V and

kept for 24 h

After those results, we have two different hypotheses:

1. Is there Ostwald ripening? We can imagine that some unreacted Se remains

at the surface of the material and it has a very small size, therefore it is not

seen by XRD. Then, when put in the electrolyte, it would dissolve and

precipitate into bigger particles following an Ostwald ripening process (109).

Therefore, diffraction peak of unreacted Se would become visible due to the

bigger size of the particles.

2. Is there a selenium redox activity? If selenium is redox active, the possible

redox process should be Se2- / Se0 that is different from sulfur redox process

(S2-/S22-). Therefore, Se0 could be produced at the end of the charge and

converted into soluble polyselenides containing Se2- at the end of the

discharge.

132

To test these hypotheses, we decide to perform a high-resolution XPS study on rocksalt

sulfide-selenide pristine powders and electrodes at different states of charge.

V.3. XPS Analysis of Li2TiSexS3-x Compositions

Elucidating the electrochemical activity in those materials requires an understanding

of: what species are present in the electrodes, which species are responsible for redox

reactions and how those species change upon charge-discharge. Moreover,

electrochemical and structural characterizations of Li2TiSexS3-x show that not only

sulfur but also selenium should be redox active. To test this hypothesis, we use a

complementary technique, a high-resolution X-ray Photoelectron Spectroscopy (XPS,

introduced in Chapter 2.2.e), which is a surface sensitive technique that gives

information about both chemical states in the compositions and does semi-quantitative

analysis of the compositions. In this chapter, we present the results of a high-resolution

XPS study on disordered rocksalt Li2TiSexS3-x pristine powders and electrodes at

different states of charge. We also prepare tables that contain optimized parameters,

such as peak positions and widths at different core levels. We believe that such results

can help scientists to understand the redox processes in metal sulfides or selenides

since only few publications are present in the literature.

V.3.a. Methodology: Analysis of XPS Spectra

Peak identification/indexation: In general, the peaks obtained from the XPS analysis

are indexed based on the information in XPS Handbook110 and the online XPS

database111 that contains binding energy of species in various compositions. As we are

the first ones who designed sulfur-selenium compositions (Li2TiSexS3-x) and the

binding energy of elements differs based on the electronic environments, we not only

use the data from XPS handbook and database but we also consider similar studies in

the literature to prepare fitting models.

Strategy for peak deconvolution: To extract the hidden information from the peaks,

firstly, we take the peak fittings parameters, such as peak positions, peak ratios and

133

peak widths from the XPS handbook and literature. The latter, we fix these parameters

in the Multitap program that is used to deconvolute the peaks obtained from the High-

Resolution XPS spectra. Afterwards, we build a fitting model that is iteratively

improved until fitting parameters do not change within a specified precision (Chi<5).

The curves are fitted using 80% Gaussian, 20% Lorenzian conditions in Multitap

Program.

• Single peaks: deconvoluting Li 1s, Se 3d and Ti 3s: The binding energy of these

three core peaks are quite close (Li 1s:54-55 eV, Se 3d: 52-53 eV Ti:3s 60-61 eV)

and therefore they can be detected together at the same core level110. For

instance, Ti 3s and Se 3d can be detected at Li 1s core. Here, each core peak is

fitted and evaluated individually for quantitative elemental analysis.

• Overlapping peaks: deconvoluting Ti 2p and S 2p orbitals: For overlapping

peaks, fitting is employed based on the contributions (two peaks) originated

from spin orbit splitting of 2p cores (3/2p and 1/2p). Each peak is fitted within

the same model considering fixed peak separations110 and the ratio between two

peaks are calculated considering these two contributions at the same core level.

We follow the same procedure for fitting of S 2p core peaks.

Quantitative elemental analysis is performed by accounting peak intensities (peak area

after background removal) and the sensivity factors (Chapter 2.2.f). It gives atomic

concentrations in the analyzed samples. For elemental analysis, we account of Li+ in Li

1s core, Ti4+ in Ti 2p core, S2- (could be S22-) in S 2p core for Li2TiS3. It should be noted

that Ti4+ species can be seen at both Ti 2p and Ti 3s core levels, but the amount of

titanium species at Ti 2p core are different from those in Ti 3s core. The difference arises

from the cross-section of each elements. The number of photons obtained from Ti 2p

core is much higher than those coming from Ti 3s core; thus, we only consider one

contribution (Ti4+ in Ti in 2p core) for elemental analysis. Similarly, we account Se2-

(could be Se22-) at Se 3d core instead of Se 3p core for selenium compositions.

134

V.3.b. Survey analysis of Li2TiSexS3-x samples

A survey analysis is performed to identify the elements contained in the composition

with a survey spectrum that is recorded over a spectral range of 0−1200eV. This quick

analysis allows us determining which core of the elements should be analyzed through

XPS and what parameters should be used. Thereafter, we can optimize the quality of

signal/noise ratio for High-Resolution XPS analysis. Nevertheless, it should be noted

that such survey analysis gives general information about species in the compositions.

For detailed analysis, the samples are analyzed further with High-Resolution XPS

(penetration length of X-rays up to 5 nm).

As illustrated in Figure V - 13, O 1s (530 eV) and C 1s (284.8 eV) core peaks are detected

in powders due to sample contamination from glovebox to XPS vacuum chamber. In

Figure V - 14, similar elements are detected in electrodes as well, but differently from

the powders, F 1s core peak (687.7 eV) is detected in electrodes. This originates from

PVdF binder used during electrode preparation.

Figure V - 13: XPS analysis of powders in survey mode

135

Figure V - 14: XPS analysis of electrodes in survey mode

After survey analysis, parameters are optimized based on high signal/noise ratio and

powders are analyzed with high-resolution XPS. We decided to follow the next

procedure during the XPS analysis:

• The first target is to determine active species at specific binding energies and

distinguish them in their different electronic environments.

• Thereafter, fitting models are developed for powders and then same models are

used to analyze the electrodes of the same compositions.

At the beginning, one of our hypotheses was that there might have been a slight trace

of unreacted selenium present in the sulfur-selenium solid solutions, therefore Se

metal would have become visible in X-ray diffractograms after peak intensities

reduced at the end of the charge. To test our hypothesis, we analyze three pristine

powders and Se metal precursor through XPS.

136

Figure V - 15: XPS spectra of powders: Li2TiS3, Li2TiSeS2, Li2TiSe2S and selenium precursor

at Se 3d core

Figure V - 15 shows the Se 3d core peak of Li2TiS3, Li2TiSeS2, Li2TiSe2S and selenium

precursor. The four spectra were calibrated based on C 1s core peak at 284.8 eV. In the

figure, Li 1s and Se 3d binding energies are quite close to each other, Li 1s and Se 3d

core peaks are detected between 53 eV and 54 eV respectively. The intensity of Se2-

peak is much higher compared to intensity of Li+ peak. This could be due to the

elemental sensitivity to XPS analysis, for instance sensitivity factor of Li (0.02) is much

smaller than sensitivity factor of Se (0.67). Besides, binding energy of Se2- is rather

smaller than Se0, core peak of Se0 is detected at 56 eV. Similar binding energy of Se0 is

reported in the literature as well (112). Overall, no signature of selenium precursors is

detected in XPS spectra. Such result is not aligned with our first hypothesis,

consequently we conclude that there is no metallic selenium remaining in the powders

after the synthesis.

137

V.3.c. High Resolution XPS Analysis of Li2TiSexS3-x Powders

Figure V - 16 shows the XPS spectra and peak fittings of Li2TiS3, Li2TiSeS2 and Li2TiSe2S

powders at Li 1s, Ti 2p and S 2p core levels. The curves are fitted using 80% Gaussian,

20% Lorenzian conditions in Multitap program.

Figure V - 16: XPS spectra of Li2TiS3, Li2TiSeS2 and Li2TiSe2S powders at Ti 2p3/2-1/2

(Ti1_orange: Ti 2p3/2, Ti2_yellow: Ti 2p1/2), S 2p3/2-1/2 (S1_light blue: S 2p3/2 and S2_dark blue:

S 2p1/2), Li 1s core (navy blue), Ti 3s (purple) and Se 3d5/2-3/2 (Se1_green:Se 3d5/2, Se2_pink:Se

3d3/2)

• 72-44eV: Li 1s, Ti 3s and Se 3d cores

In Li2TiS3, we observe Li 1s core peak at 54.6 eV with a peak width of 2.1 eV. Previous

studies reported similar binding energy of Li 1s core peak in Li-S systems (113,114). In the

same energy window, we also observe Ti 3s at 60.5 eV with almost the same intensity

as Li 1s core peak (52.4%), but a much wider peak width (3.8). We use the same peak

positions, peak widths and relative intensity of Li 1s and Ti 3s core peaks for fitting

138

selenium-substituted compositions at Li 1s and Se 3d core levels. Table V - 1 shows all

optimized fittings parameters and atomic percentages of Li2TiS3, Li2TiSeS2, Li2TiSe2S

and selenium precursor at Li 1s (or Se 3d), Ti 2p3/2-1/2, and S 2p3/2-1/2.

In Li2TiSeS2, we observe two different selenides (Se1 and Se2) at Se 3d core level and

each selenide contains a doublet, the Se 3d5/2 and Se 3d3/2 peaks, with widths of 0.86 eV.

Binding energy at Se 3d5/2 core of Se1 is detected at 52.7 eV, whereas binding energy of

Se2 is found at 53.8 eV. The parameters can be found in Table V - 1. The difference

between these two energies and the presence of two Se doublets indicate that their

valence states are not the same, Se1 could be regarded as Se2- and the binding energy

of Se2 is quite close to the Se22- (115,116). In the literature, Se 3d core peaks were reported

between 53 and 54 eVs for different metal selenide systems (117,118). By semi-quantitative

analysis, we also observe that Se1 is largely present (72.4%) compared to the amount

of Se2 (27.7%) in Li2TiSeS2. Normally, we expect to detect only one type of selenide at

Se 3d, but it seems that nearly one third of the Se2- is already oxidized to Se22- at least

at the surface of the material. This means that the amount of Li in Li2TiSeS2 is less than

what we calculated (Li2-xTiSeS2).

In Li2TiSe2S, similar peak positions are detected. Besides, we observed that ratio of Se1

to Se2 increased since 81.6% of Se1 is present compared to Se2 (18.4%). The amount of

Se2- is much higher in Li2TiSe2S, as expected. Nevertheless, we observe again that some

part of Se is oxidized, thus the amount of Li in Li2TiSe2S, is less than what we calculated

(Li2-xTiSe2S).

In conclusion, we do not observe unreacted selenium in XPS analysis on powders.

Instead, we observe one type of Li+, but two different selenide types (Se2- and Se22-) in

this energy region. Besides, Ti-S and Ti-Se peaks are not easily distinguished because

their binding energy are very close (9). Therefore, they are considered as one single Ti

peak.

139

• 473eV-449eV: Ti 2p core

In Li2TiS3 powder, two different Ti 2p doublets, such as Ti1 and Ti2 are observed at

455.1 eV and 457.4 eV, indicating Ti4+ present in sulfur (TiS2) and oxygen (TiO2)

environments. Presence of TiO2 peak at Ti 2p core indicates that there should be air

exposure during sample transfer from transfer box to XPS vacuum chamber. Similar

phenomenon was observed in the literature as well (119–121).

In both Li2TiSeS2 and Li2TiSe2S compositions, we observed similar binding energies

and peak widths of Ti 2p core levels. Here, Ti4+ is present in sulfur-selenium, [Ti(S-Se)2]

and oxygen (TiO2) environments. In addition, correlation between Se substitution and

Ti4+ in oxygen (TiO2) environment may exist because the more Se is present in the

composition, the greater intensity of Ti4+ in oxygen (TiO2) environment is detected.

Selenium compounds thus seem more air-sensitive.

• 171eV-150eV: S 2p and Se 3p cores

In Li2TiS3, we observe two different sulfur doublets (S1 and S2) at 159.9 eV and 160.6 eV

with widths of 1.2 eV. According to semi-quantitative analysis, S1 presents 49.8%,

while S2 presents 50.2% in Li2TiS3 composition. In the literature, many metal sulfides

contains two or three sulfur species at S 2p core and they are regarded as terminal and

bridge sulfur species (122–124). Since the energy difference between two different sulfur

species is quite small (< 1 eV), these sulfides should be regarded as S2- (S1) and S22- (S2).

Contrary to Li2TiS3, only one S 2p doublet is observed in both Li2TiSeS2 and Li2TiSe2S

compositions; however, two different selenium doublets, Se1 and Se2, are observed in

Li2TiSeS2 at 158.8 eV and 159.4 eV in percentages of 21.2% and 7.2% respectively

(Sulfur: 71.6%). As already mentioned before, Se2 might be regarded as Se22-. Similar

species are detected in Li2TiSe2S, the amount of selenium species increase in Li2TiSe2S,

for instance Se2 now reaches 20%, much higher than the one in Li2TiSeS2. On the

contrary, the percentage of Se1 found is 38.5% (Sulfur: 41.5%). In conclusion, all the

species in the powders are detected at Li 1s, Ti 2p and S 2p cores and curve fittings

140

models are developed. Such models are used for curve fitting of XPS analysis on

electrodes at different states of charge.

Table V - 1 :Optimized fitting parameters of Li2TiS3, Li2TiSeS2, Li2TiSe2S and selenium

precursor at Li 1s/Se 3d, Ti 2p3/2-1/2, and S 2p3/2-1/2

141

V.3.d. XPS Analysis of Li2TiSexS3-x Pristine Electrodes

After finding the best fitting models through powder analysis, we decide to continue

XPS analysis on various electrodes. Before analyzing the electrodes at different states

of charge, our next target is to investigate the surface and the bulk properties of the

pristine electrodes. One way to reach the bulk of the electrode is to scratch the electrode

with a spatula. The scratched electrodes (pristine state) are later compared with the

non-scratched electrodes and powders, XPS spectra of each compounds are shown in

Figure V - 17 (Li2TiS3), Figure V - 18 (Li2TiSeS2) and Figure V - 19 (Li2TiSe2S).

Figure V - 17: XPS spectra of Li2TiS3 at Ti 2p3/2-1/2, S 2p3/2-1/2 and Li 1s core: Li2TiS3 powder vs

non-scratched Li2TiS3 electrodes and scratched Li2TiS3 electrodes

In Figure V - 17, similar active species are observed in Li2TiS3 powder and electrodes.

We observe LiF peak at Li 1s core. That may come from reaction of sulfur with binder.

It has stronger peak intensity in non-scratched electrode, indicating that LiF is mostly

present on the electrode surface rather than in the bulk. In S 2p core, additional sulfur

species (S-C) are detected in Li2TiS3 electrodes, unlike Li2TiS3 powder. That is attributed

to sulfur-carbon bond in the literature since electrodes contain important amount of

142

carbon (10%) to increase their electronic conductivity (123,124). This explains why the S-

C species was not detected before in sulfur-selenium powders. Unlike powders, we

also observe that the amount of S2- reduces in electrodes (10%), whereas the amount of

S22- increases (80%) with S-C (10%). We explain this in a new hypothesis that sulfur

may react with PVdF binder (production of LiF) and produce more oxidized sulfur

species. That also indicates that less lithium is present in the composition, the actual

formula might be Li2-xTiS3. Additionally, it might be the reason of why S-C bond is only

seen in electrodes, not in powders. Sulfur may react with carbon in the PVdF and

created S-C compound. In Ti 2p core, signal of Ti4+ in TiS2 environment becomes much

lower in electrodes, whereas signal of Ti4+ in TiO2 environment becomes stronger in

electrodes.

Same analyses are conducted for both Li2TiSeS2 and Li2TiSe2S compositions. It should

be noted that all the peak positions and peak separations are fixed at the beginning of

the XPS analysis of scratched and non-scratched electrodes, thus the results obtained

in both electrodes and powders are similar (Figure V - 18 and Figure V - 19). The main

differences between scratched and non-scratched electrodes appear in Li 1s core and S

2p core levels. Much higher selenium peak intensities at Li 1s core level are observed

in non-scratched electrodes, whereas intensities become lower in S 2p core peaks.

Similar phenomenon is observed for Li2TiSeS2 and Li2TiSe2S, LiF and S-C species are

detected in Li 1s core and S 2P core levels respectively. That also strengthens our

hypothesis that sulfur might react with PVdF binder.

Overall, peak intensity of lithium and sulfur species are much stronger in scratched

electrodes, therefore we decide to perform post mortem analysis on scratched

electrodes. Since our next target is to analyze sulfur-selenium redox process, S 2p core

levels of scratched electrodes are investigated through high-resolution XPS.

143

Figure V - 18: XPS spectra of Li2TiSeS2 at Ti 2p3/2-1/2, S 2p3/2-1/2 and Li 1s core: powder vs

scratched electrodes and non-scratched electrodes

Figure V - 19 : XPS spectra of Li2TiSe2S at Ti 2p3/2-1/2, S 2p3/2-1/2 and Li 1s core: powder vs

scratched electrodes and non-scratched electrodes

144

V.3.e. XPS Analysis on Li2TiSexS3-x Electrodes at Different States of Charge

In this part, our target is to find the appropriate answer to our second hypothesis. As

we observed Se metal at the end of charge, our second hypothesis was that there

should have been a selenium redox activity and it occured from the change of Se2- into

Se0, that is why we observed Se metal at the end of the charge. According to our XPS

analysis on powders, we found out that there is no Ostwald ripening since no signature

of selenium precursor is found in XPS pattern of Li2TiSexS3-x compositions. However

we still need to verify the origin of extra capacity and elucidate redox process of Li2TiS3

composition. In this purpose, electrodes are cycled using the standard procedure

described in Chapter 2.3, dismounted and prepared for High resolution XPS as

described in Chapter 2.2.

Li2TiS3 electrodes at different states of charge

At S 2p core, three different sulfur species are observed in pristine electrode, shown in

Figure V - 20. S1 and S2 represents S2- (23.7%) and S22- (70.7) respectively, whereas S3

(S-C) (5.6%) corresponds to sulfur-carbon bond.

Figure V - 20: XPS spectra of Li2TiS3 scratched electrodes at different states of charge:

pristine, charged electrode (3 V) and discharged electrode (1.6 V)

145

At the end of charge, lithium ions are extracted from Li2TiS3 electrode, S2 (S22-) shifts to

higher binding energy and peak intensity reduces to 17%. This indicates oxidation of

S2 species. In the meantime, S1 (S2-) shifts to lower binding energy and peak intensity

increases to 43%. This indicates that sulfur might be partially reduced to maintain the

charge balance around Ti-S coordination. In electrochemical results, we assume that

sulfur oxidizes during charge and reduces during discharge. Therefore, this result is

different from what we expected. Besides, the peak at 168 eV indexed as SOx- (14%)

indicates that either Li2TiS3 might have reacted with carbonate-based electrolyte after

cycling or there is carbonate contamination on electrode surface after the cell is

dismounted. At the end of the discharge, lithium ions are inserted into Li2TiS3

electrode, all the sulfur species are found at the same binding energies as those of the

pristine state. This indicates that sulfur redox process is reversible.

Li2TiSeS2 electrodes at different states of charge

Figure V - 21 shows XPS spectra of Li2TiSeS2 scratched electrodes at different states of

charge. In pristine Li2TiSeS2 electrode Se2-(22%), Se22- (28%), S2- (45%) and S-C (5%)

species are detected at S 2p core level.

At the end of charge, lithium ions are extracted from Li2TiSeS2 electrode, sulfur

oxidizes, S2- shifts to higher binding energy and peak intensity reduces down to 22%.

On the contrary, the changes in binding energies of Se2- and Se22- are quite small and we

observe that peak intensity of Se22- increases (41%) compared to Se2- (12%). These

results seems to confirm an oxidation of the anionic network, however right now, we

are not able to confirm the exact selenium redox process because the deconvolution at

S 2p is quite difficult due to the high number of contributions with very close energies

(See Appendix).

At the end of the discharge, all the sulfur (S2-) and selenium species (Se2- and Se22-) are

detected at the similar binding energies as those of the pristine state of Li2TiSeS2

electrode. It seems that selenium redox is also reversible.

146

Figure V - 21: XPS spectra of Li2TiSeS2 and Li2TiSe2S scratched electrodes at different states

of charge: pristine, charged electrode (3 V) and discharged electrode (1.6 V)

In Figure V - 22, reversible selenium redox in Li2TiSeS2 is also seen at Se 3d core.

Similarly to Figure V - 21, shifts in binding energies are very small, however change in

the intensities is evident. It seems that Se2- converts to Se22- during charge, but it returns

to the original position during discharge. It suggests a reversible redox activity.

147

Figure V - 22 : XPS spectra of Li2TiSeS2 electrodes at different states of charge: Comparison of

Se 3d and S 2p core levels

Li2TiSe2S electrodes at different states of charge

Figure V - 21 also shows Li2TiSe2S electrodes at different states of charge. In pristine

Li2TiSe2S electrode Se2-(25%), Se22- (32.5%), S2- (35%) and S-C (7.5%) species are detected

at S 2p core level. Pristine electrode now contains much more Se22- (32.5%) than in

Li2TiSeS2 (28%). At the end of charge, we observe similar sulfur and selenium redox

activities as those in Li2TiSeS2 electrodes. At the end of the discharge, the peak

positions of sulfur and selenium species are found very close to those in the pristine

state. With semi-quantitative analysis, we detect Se2-(20%), Se22- (44%), S2- (25%) and S-

C (11%) species at S 2p core.

148

V.4. Discussion on the XPS Results

In pristine Li2TiS3 electrodes, we consider S2- as terminal disulfide species, whereas S22-

as central disulfide species that are bonded with Li+ and Ti4+ ions (122). Upon charging,

lithium ions are extracted from structure, central disulfide species (S22-) lose some

lithium ions and only Ti-S bonded sulfur species (S22-) remain. Meanwhile, central

disulfide species that lose neighbor lithium ions prefer creating disulfides (S2-),

therefore we observe an increase of S2- species at the end of the charge. The similar

observation is reported by A. Sakuda et al., their Monte Carlo simulation results

showed that sulfur atoms that lose at least one lithium counter made S-S bonds (S2-)

(64). With formation of disulfides, the structural volume is increased. In this way,

rocksalt Li2TiS3 has undergone smooth structural transition from crystalline to

amorphous to crystalline structure, without direct crystalline to crystalline structural

transition. At the end of the discharge process, we observe the same species at similar

binding energy, this result is consistent with what we observed in ex situ XRD results

previously.

In pristine Se-substituted electrodes (Li2TiSeS2 and LiTiSe2S), we detect central

disulfide species (S22-) and selenium species (Se22-) with diselenides (Se2-) without any

terminal disulfides (115,122,125). Like sulfide electrode (Li2TiS3), central disulfide species

(S22-) and selenium species (Se22-) lose some lithium ions and Ti-S and Ti-Se bonded

sulfur and selenium species (S22-, Se2-) remain at the end of the charge (3 V). Differently

from sulfide electrode (Li2TiS3), diselenides species tend to make bonds with central

disulfide species and produce Se-S bonds, instead of producing diselenides at the end

of the charge. That is why we observe the increase of selenium species (Se22-) instead

of diselenides (Se2-) in the XPS spectra. This also indicates that the possible redox

mechanisms between sulfur and selenium is quite different.

In previous chapter, we show that selenium redox activity appears during first

discharge at 1.88 V vs Li+/Li for both Li2TiSeS2 and Li2TiSe2S electrodes. Therefore, here

149

we only consider sulfur redox process, the shift of S22- species from low binding to high

binding energy creates available charge capacity. At the end of the discharge, we detect

similar species in selenium-subsituted samples (Li2TiSexS3-x), but it seems that more

diselenides are produced. Here, we consider that disulfides and diselenides do not

contribute to charge-discharge capacity, thus they are redox inactive species. If we

assume continuous production of diselenides in Li2TiSe2S cells at the end of the

discharge, this might also explain why we observe huge capacity fading in the

Li2TiSe2S cells previously.

According to our second hypothesis, the possible redox mechanisms of selenium

should be undergone from Se2- to Se0. With the current study, it is quite difficult to

determine this point as the actual redox process of selenium starts at the first discharge

process. Therefore, further ex situ XPS analysis on Li2TiSexS3-x electrodes at different

states of charges (2nd and 3rd cycles) are required. In sulfur 2p core level, there are high

number of contributions coming from both sulfur and selenium species. To make clear

observations on selenium redox process, it might be better to work with only selenium

species. For instance, a new composition Li2TiSe3 might be developed and electrodes

at different states of charges might be analyzed. For a comparison, high performance

pure Se electrodes might also be developed and analyzed with the same approach.

150

V.5. Conclusion

In this chapter, we focused on characterization of Li2TiSexS3-x through electrochemical,

structural and morphological studies. We observed that these materials provide extra

capacity during discharge. To understand the origin of extra capacity and the reason

of degradation process upon cycling, various structural characterizations were done.

Ex situ XRD showed that cubic rocksalt structure reversibly changed upon cycling. In

addition, metallic selenium was observed in diffraction pattern and a columnar phase

was observed in SEM images of selenium-substituted samples. After such results, we

had hypotheses, such as either unreacted selenium remains in the electrodes or

selenium redox process is present (Se2- / Se0).

To investigate our hypotheses, we performed ex situ XPS study. Our first hypothesis

failed because we did not observe unreacted selenium in powders. XPS analysis on

Li2TiS3 showed that sulfur redox process is reversible. In selenium-substituted samples,

it seems that selenium redox is also reversible, however we cannot explain and confirm

the exact Selenium redox process at this point. The deconvolution at S 2p is quite

difficult due to the high number of contributions with very close energies, thus further

analysis is necessary.

Finally, the presence of LiF at Li 1s core and more oxidized sulfur species at S 2p core

in pristine electrodes suggest that sulfur might react with PVdF binder and produce

more oxidized sulfur species. This hypothesis should be investigated by further XPS

study on Li2TiSexS3-x electrodes prepared with different binders. It may also help to

clarify the degradation process upon cycling.

151

General Conclusion and Perspectives

152

The purpose of this thesis study is to develop overlithiated disordered rocksalt

materials that provide high specific energies (> 900Wh.kg-1 at material level) and

present a cobalt-free alternative to conventional materials used in Li-ion batteries. Li-

rich cation disordered rocksalt oxides can offer large capacities (≥ 250 mAh.g-1),

however there is always irreversibility in the first cycle and origin of this irreversibility

is still unknown. Meanwhile, disordered rocksalt Li2TiS3 has attracted research interest

thanks to its excellent reversible capacities (425 mAh.g-1 at 2.3 V vs Li+/Li). However,

detailed electrochemical and structural studies of such a promising material are still

missing, especially redox properties have not been reported yet. In addition, the

question of why Li2TiS3 cells failed in traditional batteries with liquid electrolytes has

not been answered yet. Therefore, we were interested to investigate those properties

and examine the reason of degradation during charge-discharge.

In this thesis project, we have studied overlithiated disordered rocksalt family in two

different axes: disordered rocksalt oxide (Li2.2NiTi0.2Nb0.6O4) and sulfide (Li2TiSexS3-x).

In the first part of the thesis, our aim was to optimize the performance of our patented

composition, disordered rocksalt Li2.2NiTi0.2Nb0.6O4. The disordered rocksalt (cubic)

structures are generally stabilized at high temperatures (1000 oC), however the

performances are limited due to particle growth. Thus, an optimization of the synthesis

parameters of our molten salt synthesis route was carried out. The nature of the

eutectic salt, the washing solvent, as well as the duration and the temperature of the

synthesis process were investigated. An optimum was found using 0.59LiCl-0.41KCl

eutectic salt at 670°C for 1h followed by a quench, and ethylene glycol as washing

media. The material delivered 166 mAh.g-1 of charge capacity with 108 mAh.g-1 of

reversible capacity. In situ XRD was performed to investigate the irreversibility (35%)

at the first cycle. The result showed that Li2.2NiTi0.2Nb0.6O4 has undergone structural

evolutions during charge-discharge. In addition, the particle size analysis showed that

dislocations or defects might have arisen in Li2.2NiTi0.2Nb0.6O4 powder, this

phenomenon may have been the cause of the irreversibility. This result indicated that

153

purity of the disordered rocksalts systems is important to design high energy

materials.

In the second part of the thesis, disordered rocksalt Li2TiS3 powders were prepared via

mechanochemical synthesis for shorter milling duration (15h - 25h) than those in the

literature (40h – 100h). As the Li2TiS3 electrodes were tested in carbonate-based

electrolyte, they delivered a reversible of 250 mAh.g-1, however a capacity decay

starting form the second cycle was observed. Thereafter, electrode pressing was

considered to improve its electronic conductivity and we observed that both reversible

capacities (339 mAh.g-1) and cycling stability were improved. To further improve

cycling stability of Li2TiS3 cells, the partial substitution of Li2TiS3 with selenium was

considered. Our efforts to prepare Li2TiSexS3-x compositions have resulted in the

discovery of a new synthesis method and high capacity compositions (Li2TiSeS2 and

Li2TiSe2S). X-ray diffraction patterns showed that all Li2TiSexS3-x powders that were

prepared in cation disordered rocksalt phase follow a solid solution behavior.

Thereafter, electrochemical properties of new materials have been investigated in

comparison with Li2TiS3. Li2TiSeS2 and Li2TiSe2S cells showed better cycling stability

than Li2TiS3 cells at higher cut-off discharge potentials (1.8V and 2 V) while cycling

stability of Li2TiS3 cells did not improve even at higher cut-off potentials. Furthermore,

Li2TiSe2S cells failed upon cycling between 3 V and 1.5 V. To find out the reason of

degradation during cycling, we decided to investigate Li2TiSexS3-x with deep structural

characterizations.

In the last part of the thesis, we examined the lithium insertion-extraction mechanism

through disordered rocksalt Li2TiS3 with structural studies (operando neutron

diffraction, FTIR and Raman). Unfortunately, poor crystallinity of Li2TiS3 powder

made definitive conclusions difficult even though various characterization tools were

used. Afterwards, Li2TiSeS2 and Li2TiSe2S electrodes were characterized by combining

both ex situ XRD and SEM techniques. Selenium metal was observed in the charged

electrodes, but it could not be detected in both pristine and discharged electrodes.

154

Thereafter, we had a hypothesis that there might have been Oswald ripening or

Selenium redox activity during charge. To tests these hypotheses, Li2TiSexS3-x powders

and Se precursors were further analyzed with X-Ray Photoelectron Spectroscopy (XPS)

analysis. The result showed that there were no remaining selenium precursors in the

Li2TiSexS3-x powders, thus our first hypothesis related to Oswald ripening was not true.

Thereafter, we decided to discover redox active species and elucidate redox processes

in Li2TiSexS3-x by XPS. At first, we identified species (elements) present in Li2TiSexS3-x at

their core levels and developed a fitting model for both powders and electrodes.

Surprisingly, high amounts of oxidized sulfur species at 2p sulfur core and LiF species

at Li 1s core were detected in pristine electrodes, showing the possible reaction of

sulfur in the Li2TiSexS3-x with the PVdF binder during electrode preparations. Ex situ

XPS analyses of Li2TiSexS3-x electrodes at different states of charge have confirmed our

previous observations that sulfur redox process was reversible. However, it is still

quite difficult to identify exact redox process in Li2TiSexS3-x due to the presence of a

high number of contributions of selenium species. We suggest that further XPS

analyses on Se metal or Li2TiSe3 can be useful to elucidate redox process in Li2TiSexS3-x

samples.

Overall, the degradation process of Li2TiS3 in liquid cells is still a question even though

various structural characterization tools were used to understand the underlying

reason. Further XPS analysis of Li2TiS3 electrodes at different states of charge and

during cycling (after 1st, 5th, 10th and 20th cycles for instance) might be useful to clarify

this issue. The degradation might result from the Li metal anode as the lithium metal

is very reactive. Thus, testing the Li2TiS3 electrodes against graphite or silicon type

negative electrodes can be interesting. In this thesis study, we could only investigate

reversible sulfur (anionic) redox process but did not focus on cationic process as the

titanium redox is generally active at lower potentials. To determine titanium redox

activity, various ex situ XPS experiments can be performed at Li2TiS3 electrodes cycled

at lower discharge cut-off potentials. XPS could also be useful to examine the true

155

selenium redox process and explain the main differences between sulfur and selenium

redox processes. To this purpose, a new Li2TiSe3 electrode can be characterized at

different states of charge. Consequently, different high energy compounds, such as Li-

excess disordered sulfides (Li2MS3 M: Ni, Co, Fe, Mn) or chalcogenides (Li2TiX3 X: S,

Se, Te) can be prepared with our patented synthesis process and characterized with

the techniques that were reported in this thesis.

156

Bibliography

(1) Armand, M.; Tarascon, J.-M. Building Better Batteries. nature 2008, 451 (7179),

652.

(2) Placke, T.; Kloepsch, R.; Dühnen, S.; Winter, M. Lithium Ion, Lithium Metal,

and Alternative Rechargeable Battery Technologies: The Odyssey for High

Energy Density. J. Solid State Electrochem. 2017, 21 (7), 1939–1964.

(3) Ding, Y.; Cano, Z. P.; Yu, A.; Lu, J.; Chen, Z. Automotive Li-Ion Batteries:

Current Status and Future Perspectives. Electrochem. Energy Rev. 2019, 2 (1), 1–

28.

(4) Pillot, C. The Rechargeable Battery Market 2017-2025, 2018.

(5) Myung, S.-T.; Maglia, F.; Park, K.-J.; Yoon, C. S.; Lamp, P.; Kim, S.-J.; Sun, Y.-K.

Nickel-Rich Layered Cathode Materials for Automotive Lithium-Ion Batteries:

Achievements and Perspectives. ACS Energy Lett. 2017, 2 (1), 196–223.

(6) Claus Daniel, B. J. Handbook of Battery Materials; Wiley Online Library.

(7) Whittingham, M. S.; Gamble, F. R. The Lithium Intercalates of the Transition

Metal Dichalcogenides. Mater. Res. Bull. 1975, 10 (5), 363–371.

(8) Whittingham, M. S. Electrical Energy Storage and Intercalation Chemistry.

Science 1976, 192 (4244), 1126–1127.

(9) Mizushima, K.; Jones, P. C.; Wiseman, P. J.; Goodenough, J. B. LixCoO2 (0. Mater.

Res. Bull. 1980, 15 (6), 783–789.

(10) Ohzuku, T.; Ueda, A.; Nagayama, M. Electrochemistry and Structural

Chemistry of LiNiO2 (R3̅m) for 4 Volt Secondary Lithium Cells. J. Electrochem.

Soc. 1993, 140 (7), 1862–1870.

(11) Broussely, M.; Perton, F.; Labat, J.; Staniewicz, R. J.; Romero, A. Li/LixNiO2 and

Li/LixCoO2 Rechargeable Systems: Comparative Study and Performance of

Practical Cells. J. Power Sources 1993, 43 (1), 209–216.

(12) Arai, H.; Okada, S.; Sakurai, Y.; Yamaki, J. Reversibility of LiNiO2 Cathode. Solid

State Ion. 1997, 95 (3), 275–282.

157

(13) MacNeil, D. D.; Lu, Z.; Dahn, J. R. Structure and Electrochemistry of Li[NixCo1

− 2xMnx]O2 ( 0 ⩽ x ⩽ 1 / 2 ). J. Electrochem. Soc. 2002, 149 (10), A1332–A1336.

(14) Patoux, S.; Doeff, M. M. Direct Synthesis of LiNi1/3Co1/3Mn1/3O2 from Nitrate

Precursors. Electrochem. Commun. 2004, 6 (8), 767–772.

(15) Yabuuchi, N.; Ohzuku, T. Novel Lithium Insertion Material of

LiCo1/3Ni1/3Mn1/3O2 for Advanced Lithium-Ion Batteries. J. Power Sources 2003,

119–121, 171–174.

(16) Hwang, B. J.; Tsai, Y. W.; Carlier, D.; Ceder, G. A Combined

Computational/Experimental Study on LiNi1/3Co1/3Mn1/3O2. Chem. Mater. 2003,

15 (19), 3676–3682.

(17) Shaju, K. M.; Subba Rao, G. V.; Chowdari, B. V. R. Performance of Layered

Li(Ni1/3Co1/3Mn1/3)O2 as Cathode for Li-Ion Batteries. Electrochimica Acta 2002, 48

(2), 145–151.

(18) Zhuang, G. V.; Chen, G.; Shim, J.; Song, X.; Ross, P. N.; Richardson, T. J. Li2CO3

in LiNi0.8Co0.15Al0.05O2 Cathodes and Its Effects on Capacity and Power. J. Power

Sources 2004, 134 (2), 293–297.

(19) Lee, K. K.; Yoon, W. S.; Kim, K. B.; Lee, K. Y.; Hong, S. T. Characterization of

LiNi0.85Co0.10M0.05O2 (M = Al, Fe) as a Cathode Material for Lithium Secondary

Batteries. J. Power Sources 2001, 97–98, 308–312.

(20) Cho, Y.; Cho, J. Significant Improvement of LiNi0.8Co0.15Al0.05O2 Cathodes at 60 °

C by SiO2 Dry Coating for Li-Ion Batteries. J. Electrochem. Soc. 2010, 157 (6),

A625–A629.

(21) Jo, M.; Noh, M.; Oh, P.; Kim, Y.; Cho, J. A New High Power LiNi0.81Co0.1Al0.09O2

Cathode Material for Lithium-Ion Batteries. Adv. Energy Mater. 2014, 4 (13),

1301583.

(22) Guilmard, M.; Croguennec, L.; Delmas, C. Thermal Stability of Lithium Nickel

Oxide Derivatives. Part II:  LixNi0.70Co0.15Al0.15O2 and LixNi0.90Mn0.10O2 (x = 0.50

and 0.30). Comparison with LixNi1.02O2 and LixNi0.89Al0.16O2. Chem. Mater. 2003,

15 (23), 4484–4493.

158

(23) Yamada, A.; Chung, S. C.; Hinokuma, K. Optimized LiFePO4 for Lithium

Battery Cathodes. J. Electrochem. Soc. 2001, 148 (3), A224–A229.

(24) Huang, H.; Yin, S.-C.; Nazar, L. F. Approaching Theoretical Capacity of LiFePO4

at Room Temperature at High Rates. Electrochem. Solid-State Lett. 2001, 4 (10),

A170–A172.

(25) Prosini, P. P.; Lisi, M.; Zane, D.; Pasquali, M. Determination of the Chemical

Diffusion Coefficient of Lithium in LiFePO4. Solid State Ion. 2002, 148 (1), 45–51.

(26) Kim, D. K.; Muralidharan, P.; Lee, H.-W.; Ruffo, R.; Yang, Y.; Chan, C. K.; Peng,

H.; Huggins, R. A.; Cui, Y. Spinel LiMn2O4 Nanorods as Lithium Ion Battery

Cathodes. Nano Lett. 2008, 8 (11), 3948–3952.

(27) Hosono, E.; Kudo, T.; Honma, I.; Matsuda, H.; Zhou, H. Synthesis of Single

Crystalline Spinel LiMn2O4 Nanowires for a Lithium Ion Battery with High

Power Density. Nano Lett. 2009, 9 (3), 1045–1051.

(28) M. Thackeray, M.; S. Johnson, C.; T. Vaughey, J.; Li, N.; A. Hackney, S.

Advances in Manganese-Oxide ‘Composite’ Electrodes for Lithium-Ion

Batteries. J. Mater. Chem. 2005, 15 (23), 2257–2267.

(29) Zhang, S. S. Understanding of Performance Degradation of

LiNi0.80Co0.10Mn0.10O2 Cathode Material Operating at High Potentials. J. Energy

Chem. 2020, 41, 135–141.

(30) Li, J.; Liu, H.; Xia, J.; Cameron, A. R.; Nie, M.; Botton, G. A.; Dahn, J. R. The

Impact of Electrolyte Additives and Upper Cut-off Voltage on the Formation of

a Rocksalt Surface Layer in LiNi0.8Mn0.1Co0.1O2 Electrodes. J. Electrochem. Soc.

2017, 164 (4), A655–A665.

(31) Zhao, E.; Chen, M.; Hu, Z.; Chen, D.; Yang, L.; Xiao, X. Improved Cycle Stability

of High-Capacity Ni-Rich LiNi0.8Mn0.1Co0.1O2 at High Cut-off Voltage by Li2SiO3

Coating. J. Power Sources 2017, 343, 345–353.

159

(32) Chen, M.; Zhao, E.; Chen, D.; Wu, M.; Han, S.; Huang, Q.; Yang, L.; Xiao, X.;

Hu, Z. Decreasing Li/Ni Disorder and Improving the Electrochemical

Performances of Ni-Rich LiNi0.8Co0.1Mn0.1O2 by Ca Doping. Inorg. Chem. 2017, 56

(14), 8355–8362.

(33) Sun, Y.-K.; Myung, S.-T.; Kim, M.-H.; Prakash, J.; Amine, K. Synthesis and

Characterization of Li[(Ni0.8Co0.1Mn0.1)0.8(Ni0.5Mn0.5)0.2]O2 with the Microscale

Core−Shell Structure as the Positive Electrode Material for Lithium Batteries. J.

Am. Chem. Soc. 2005, 127 (38), 13411–13418.

(34) Zhang, L.; Borong, Wu.; Ning, Li.; Feng, Wu. Hierarchically Porous Micro-Rod

Lithium-Rich Cathode Material Li1.2Ni0.13Mn0.54Co0.13O2 for High Performance

Lithium-Ion Batteries. Electrochimica Acta 2014, 118, 67–74.

(35) He, X.; Wang, J.; Kloepsch, R.; Krueger, S.; Jia, H.; Liu, H.; Vortmann, B.; Li, J.

Enhanced Electrochemical Performance in Lithium Ion Batteries of a Hollow

Spherical Lithium-Rich Cathode Material Synthesized by a Molten Salt Method.

Nano Res. 2014, 7 (1), 110–118.

(36) Song, B.; Liu, Z.; On Lai, M.; Lu, L. Structural Evolution and the Capacity Fade

Mechanism upon Long-Term Cycling in Li-Rich Cathode Material. Phys. Chem.

Chem. Phys. 2012, 14 (37), 12875–12883.

(37) Xiao, J.; Chen, X.; Sushko, P. V.; Sushko, M. L.; Kovarik, L.; Feng, J.; Deng, Z.;

Zheng, J.; Graff, G. L.; Nie, Z.; Choi, D.; Liu, J.; Zhang, J.-G.; Whittingham, M.

S. High-Performance LiNi0.5Mn1.5O4 Spinel Controlled by Mn3+ Concentration

and Site Disorder. Adv. Mater. 2012, 24 (16), 2109–2116.

(38) Kim, J.-H.; Myung, S.-T.; Yoon, C. S.; Kang, S. G.; Sun, Y.-K. Comparative Study

of LiNi0.5Mn1.5O4-δ and LiNi0.5Mn1.5O4 Cathodes Having Two Crystallographic

Structures:  Fd3̄m and P4332. Chem. Mater. 2004, 16 (5), 906–914.

(39) Zhou, L.; Zhao, D.; Lou, X. (David). LiNi0.5Mn1.5O4 Hollow Structures as High-

Performance Cathodes for Lithium-Ion Batteries. Angew. Chem. Int. Ed. 2012, 51

(1), 239–241.

160

(40) Santhanam, R.; Rambabu, B. Research Progress in High Voltage Spinel

LiNi0.5Mn1.5O4 Material. J. Power Sources 2010, 195 (17), 5442–5451.

(41) Yabuuchi, N. Material Design Concept of Lithium-Excess Electrode Materials

with Rocksalt-Related Structures for Rechargeable Non-Aqueous Batteries.

Chem. Rec. 2018.

(42) Pralong, V.; Gopal, V.; Caignaert, V.; Duffort, V.; Raveau, B. Lithium-Rich Rock-

Salt-Type Vanadate as Energy Storage Cathode: Li2–XVO3. Chem. Mater. 2012, 24

(1), 12–14.

(43) Obrovac, M. N.; Mao, O.; Dahn, J. R. Structure and Electrochemistry of LiMO2

(M= Ti, Mn, Fe, Co, Ni) Prepared by Mechanochemical Synthesis. Solid State Ion.

1998, 112 (1–2), 9–19.

(44) Yang, M.; Zhao, X.; Bian, Y.; Ma, L.; Ding, Y.; Shen, X. Cation Disordered Rock

Salt Phase Li2CoTiO4 as a Potential Cathode Material for Li-Ion Batteries. J.

Mater. Chem. 2012, 22 (13), 6200–6205.

(45) Fattakhova, D.; Petrykin, V.; Brus, J.; Kostlánová, T.; Dědeček, J.; Krtil, P.

Solvothermal Synthesis and Electrochemical Behavior of Nanocrystalline Cubic

Li–Ti–O Oxides with Cationic Disorder. Solid State Ion. 2005, 176 (23), 1877–

1885.

(46) Prabaharan, S. R. S.; Michael, M. S.; Ikuta, H.; Uchimoto, Y.; Wakihara, M.

Li2NiTiO4—a New Positive Electrode for Lithium Batteries: Soft-Chemistry

Synthesis and Electrochemical Characterization. Solid State Ion. 2004, 172 (1), 39–

45.

(47) Küzma, M.; Dominko, R.; Meden, A.; Makovec, D.; Bele, M.; Jamnik, J.;

Gaberšček, M. Electrochemical Activity of Li2FeTiO4 and Li2MnTiO4 as Potential

Active Materials for Li Ion Batteries: A Comparison with Li2NiTiO4. J. Power

Sources 2009, 189 (1), 81–88.

161

(48) Zhang, W.; Luo, J.; Li, X.; Liang, J.; Zhu, Y.; Qian, Y. Facile Synthesis and

Electrochemistry of a New Cubic Rocksalt LixVyO2 (x = 0.78, y = 0.75) Electrode

Material. J. Mater. Chem. A 2017, 5 (10), 5148–5155.

(49) Jiang, K.; Hu; Sun; Wang; Jin; Ren; Chen, G. Z. Electrochemical Synthesis of

LiTiO2 and LiTi2O4 in Molten LiCl. Chem. Mater. 2004, 16 (22), 4324–4329.

(50) Trócoli, R.; Cruz-Yusta, M.; Morales, J.; Santos-Peña, J. On the Limited

Electroactivity of Li2NiTiO4 Nanoparticles in Lithium Batteries. Electrochimica

Acta 2013, 100, 93–100.

(51) Yang, M.; Zhao, X.; Bian, Y.; Ma, L.; Ding, Y.; Shen, X. Cation Disordered Rock

Salt Phase Li2CoTiO4 as a Potential Cathode Material for Li-Ion Batteries. J.

Mater. Chem. 2012, 22 (13), 6200.

(52) Van der Ven, A.; Ceder, G. Lithium Diffusion Mechanisms in Layered

Intercalation Compounds. J. Power Sources 2001, 97–98, 529–531.

(53) Van der Ven, A.; Bhattacharya, J.; Belak, A. A. Understanding Li Diffusion in

Li-Intercalation Compounds. Acc. Chem. Res. 2013, 46 (5), 1216–1225.

(54) Urban, A.; Lee, J.; Ceder, G. The Configurational Space of Rocksalt-Type Oxides

for High-Capacity Lithium Battery Electrodes. Adv. Energy Mater. 2014, 4 (13),

1400478.

(55) Lee, J.; Papp, J. K.; Clément, R. J.; Sallis, S.; Kwon, D.-H.; Shi, T.; Yang, W.;

McCloskey, B. D.; Ceder, G. Mitigating Oxygen Loss to Improve the Cycling

Performance of High Capacity Cation-Disordered Cathode Materials. Nat.

Commun. 2017, 8 (1), 981.

(56) Lee, J.; Urban, A.; Li, X.; Su, D.; Hautier, G.; Ceder, G. Unlocking the Potential

of Cation-Disordered Oxides for Rechargeable Lithium Batteries. Science 2014,

343 (6170), 519–522.

162

(57) Yabuuchi, N.; Takeuchi, M.; Nakayama, M.; Shiiba, H.; Ogawa, M.; Nakayama,

K.; Ohta, T.; Endo, D.; Ozaki, T.; Inamasu, T.; Sato, K.; Komaba, S. High-

Capacity Electrode Materials for Rechargeable Lithium Batteries: Li3NbO4-

Based System with Cation-Disordered Rocksalt Structure. Proc. Natl. Acad. Sci.

2015, 112 (25), 7650–7655.

(58) Wang, R.; Li, X.; Liu, L.; Lee, J.; Seo, D.-H.; Bo, S.-H.; Urban, A.; Ceder, G. A

Disordered Rock-Salt Li-Excess Cathode Material with High Capacity and

Substantial Oxygen Redox Activity: Li1.25Nb0.25Mn0.5O2. Electrochem. Commun.

2015, 60, 70–73.

(59) Nakajima, M.; Yabuuchi, N. Lithium-Excess Cation-Disordered Rocksalt-Type

Oxide with Nanoscale Phase Segregation: Li1.25Nb0.25V0.5O2. Chem. Mater. 2017,

29 (16), 6927–6935.

(60) Li, B.; Jiang, N.; Huang, W.; Yan, H.; Zuo, Y.; Xia, D. Thermodynamic Activation

of Charge Transfer in Anionic Redox Process for Li-Ion Batteries. Adv. Funct.

Mater. 2018, 28 (4), 1704864.

(61) Hoshino, S.; Glushenkov, A. M.; Ichikawa, S.; Ozaki, T.; Inamasu, T.; Yabuuchi,

N. Reversible Three-Electron Redox Reaction of Mo3+/Mo6+ for Rechargeable

Lithium Batteries. ACS Energy Lett. 2017, 2 (4), 733–738.

(62) Sakuda, A.; Takeuchi, T.; Okamura, K.; Kobayashi, H.; Sakaebe, H.; Tatsumi, K.;

Ogumi, Z. Rock-Salt-Type Lithium Metal Sulphides as Novel Positive-

Electrode Materials. Sci. Rep. 2015, 4 (1).

(63) Sakuda, A. Favorable Composite Electrodes for All-Solid-State Batteries. J.

Ceram. Soc. Jpn. 2018, 126 (9), 675–683.

(64) Sakuda, A.; Ohara, K.; Kawaguchi, T.; Fukuda, K.; Nakanishi, K.; Arai, H.;

Uchimoto, Y.; Ohta, T.; Matsubara, E.; Ogumi, Z.; Kuratani, K.; Kobayashi, H.;

Shikano, M.; Takeuchi, T.; Sakaebe, H. A Reversible Rocksalt to Amorphous

Phase Transition Involving Anion Redox. Sci. Rep. 2018, 8 (1).

163

(65) Shimoda, K.; Murakami, M.; Takeuchi, T.; Matsunaga, T.; Ukyo, Y.; Sakaebe, H.;

Kobayashi, H.; Matsubara, E. Structural and Dynamic Behavior of Lithium Iron

Polysulfide Li8FeS5 during Charge–Discharge Cycling. J. Power Sources 2018,

398, 67–74.

(66) Sakuda, A.; Takeuchi, T.; Kobayashi, H.; Sakaebe, H.; Tatsumi, K.; Ogumi, Z.

Preparation of Novel Electrode Materials Based on Lithium Niobium Sulfides.

Electrochemistry 2014, 82 (10), 880–883.

(67) Sakuda, A.; Takeuchi, T.; Shikano, M.; Ohara, K.; Fukuda, K.; Uchimoto, Y.;

Ogumi, Z.; Kobayashi, H.; Sakaebe, H. Development of Li2TiS3–Li3NbS4 by a

Mechanochemical Process. J. Ceram. Soc. Jpn. 2017, 125 (4), 268–271.

(68) Sakuda, A.; Kuratani, K.; Takeuchi, T.; Kiuchi, H.; Kawaguchi, T.; Shikano, M.;

Sakaebe, H.; Kobayashi, H. Cubic Rocksalt Li2SnS3 and a Solid Solution with

Li3NbS4 Prepared by Mechanochemical Synthesis. Electrochemistry 2017, 85 (9),

580–584.

(69) Grimaud, A.; Hong, W. T.; Shao-Horn, Y.; Tarascon, J.-M. Anionic Redox

Processes for Electrochemical Devices. Nat. Mater. 2016, 15 (2), 121.

(70) Sathiya, M.; Rousse, G.; Ramesha, K.; Laisa, C. P.; Vezin, H.; Sougrati, M. T.;

Doublet, M.-L.; Foix, D.; Gonbeau, D.; Walker, W.; Prakash, A. S.; Ben Hassine,

M.; Dupont, L.; Tarascon, J.-M. Reversible Anionic Redox Chemistry in High-

Capacity Layered-Oxide Electrodes. Nat. Mater. 2013, 12 (9), 827–835.

(71) Sathiya, M.; Abakumov, A. M.; Foix, D.; Rousse, G.; Ramesha, K.; Saubanère,

M.; Doublet, M. L.; Vezin, H.; Laisa, C. P.; Prakash, A. S.; Gonbeau, D.;

VanTendeloo, G.; Tarascon, J.-M. Origin of Voltage Decay in High-Capacity

Layered Oxide Electrodes. Nat. Mater. 2015, 14 (2), 230–238.

(72) Assat, G.; Tarascon, J.-M. Fundamental Understanding and Practical

Challenges of Anionic Redox Activity in Li-Ion Batteries. Nat. Energy 2018, 3 (5),

373–386.

164

(73) Jacquet, Q.; Perez, A.; Batuk, D.; Van Tendeloo, G.; Rousse, G.; Tarascon, J.-M.

The Li3Ru y Nb1– y O4 (0 ≤ y ≤ 1) System: Structural Diversity and Li Insertion and

Extraction Capabilities. Chem. Mater. 2017, 29 (12), 5331–5343.

(74) McCalla, E.; Sougrati, M. T.; Rousse, G.; Berg, E. J.; Abakumov, A.; Recham, N.;

Ramesha, K.; Sathiya, M.; Dominko, R.; Van Tendeloo, G.; Novák, P.; Tarascon,

J.-M. Understanding the Roles of Anionic Redox and Oxygen Release during

Electrochemical Cycling of Lithium-Rich Layered Li4FeSbO6. J. Am. Chem. Soc.

2015, 137 (14), 4804–4814.

(75) Kimura, T. Molten Salt Synthesis of Ceramic Powders. In Advances in Ceramics

- Synthesis and Characterization, Processing and Specific Applications; 2011.

(76) Suryanarayana, C. Mechanical Alloying and Milling. Prog. Mater. Sci. 2001, 46

(1), 1–184.

(77) Rodriguez-Carvajal, J. Recent Developments of the Program FULLPROF,

Commission on Powder Diffraction. IUCr Newsl 2001, 26.

(78) Petříček, V.; Dušek, M.; Palatinus, L. Crystallographic Computing System

JANA2006: General Features. Z. Für Krist. - Cryst. Mater. 2014, 229 (5), 345–352.

(79) Bianchini, M.; Leriche, J. B.; Laborier, J.-L.; Gendrin, L.; Suard, E.; Croguennec,

L.; Masquelier, C. A New Null Matrix Electrochemical Cell for Rietveld

Refinements of In-Situ or Operando Neutron Powder Diffraction Data. J.

Electrochem. Soc. 2013, 160 (11), A2176–A2183.

(80) Sebastian, L.; Gopalakrishnan, J. Li2MTiO4 (M=Mn, Fe, Co, Ni): New Cation-

Disordered Rocksalt Oxides Exhibiting Oxidative Deintercalation of Lithium.

Synthesis of an Ordered Li2NiTiO4. J. Solid State Chem. 2003, 172 (1), 171–177.

(81) Wang, Y.; Wang, Y.; Wang, F. Facile Molten Salt Synthesis of Li2NiTiO4 Cathode

Material for Li-Ion Batteries. Nanoscale Res. Lett. 2014, 9 (1), 197.

(82) Küzma, M.; Dominko, R.; Meden, A.; Makovec, D.; Bele, M.; Jamnik, J.;

Gaberšček, M. Electrochemical Activity of Li2FeTiO4 and Li2MnTiO4 as Potential

Active Materials for Li Ion Batteries: A Comparison with Li2NiTiO4. J. Power

Sources 2009, 189 (1), 81–88.

165

(83) Yang, M.; Zhao, X.; Bian, Y.; Ma, L.; Ding, Y.; Shen, X. Cation Disordered Rock

Salt Phase Li2CoTiO4 as a Potential Cathode Material for Li-Ion Batteries. J.

Mater. Chem. 2012, 22 (13), 6200–6205.

(84) Kawano, Y.; Kitajou, A.; Okada, S. Synthesis and Cathode Properties of a Cubic

Rocksalt-Type Si-Doped Li2NiTiO4 for Lithium-Ion Batteries. J. Power Sources

2013, 242, 768–774.

(85) Lee, J.; Seo, D.-H.; Balasubramanian, M.; Twu, N.; Li, X.; Ceder, G. A New Class

of High Capacity Cation-Disordered Oxides for Rechargeable Lithium

Batteries: Li–Ni–Ti–Mo Oxides. Energy Environ. Sci. 2015, 8 (11), 3255–3265.

(86) Whittingham, M. S. Electrical Energy Storage and Intercalation Chemistry.

Science 1976, 192 (4244), 1126–1127.

(87) Whittingham, M. S. Chemistry of Intercalation Compounds: Metal Guests in

Chalcogenide Hosts. Prog. Solid State Chem. 1978, 12 (1), 41–99.

(88) Whittingham, M. S. The Role of Ternary Phases in Cathode Reactions. J.

Electrochem. Soc. 1976, 123 (3), 315–320.

(89) Whittingham, M. S. Chemistry of Intercalation Compounds: Metal Guests in

Chalcogenide Hosts. Prog. Solid State Chem. 1978, 12 (1), 41–99.

(90) Holleck, G. L.; Driscoll, J. R. Transition Metal Sulfides as Cathodes for

Secondary Lithium Batteries—II. Titanium Sulfides. Electrochimica Acta 1977, 22

(6), 647–655.

(91) Dines, M. B. Lithium Intercalation via N-Butyllithium of the Layered Transition

Metal Dichalcogenides. Mater. Res. Bull. 1975, 10 (4), 287–291.

(92) Sakuda, A.; Takeuchi, T.; Kobayashi, H.; Sakaebe, H.; Tatsumi, K.; Ogumi, Z.

Preparation of Novel Electrode Materials Based on Lithium Niobium Sulfides.

Electrochemistry 2014, 82 (10), 880–883.

(93) Sakuda, A.; Kuratani, K.; Takeuchi, T.; Kiuchi, H.; Kawaguchi, T.; SHIKANO,

M.; Sakaebe, H.; Kobayashi, H. Cubic Rocksalt Li2SnS3 and a Solid Solution with

Li3NbS4 Prepared by Mechanochemical Synthesis. Electrochemistry 2017, 85,

580–584.

166

(94) Pang, Q.; Liang, X.; Kwok, C. Y.; Nazar, L. F. Advances in Lithium–Sulfur

Batteries Based on Multifunctional Cathodes and Electrolytes. Nat. Energy 2016,

1 (9), 1–11.

(95) Safa, M.; Hao, Y.; Chamaani, A.; Adelowo, E.; Chawla, N.; Wang, C.; El-Zahab,

B. Capacity Fading Mechanism in Lithium-Sulfur Battery Using Poly(Ionic

Liquid) Gel Electrolyte. Electrochimica Acta 2017, 258, 1284–1292.

(96) Yan, J.; Liu, X.; Li, B. Capacity Fade Analysis of Sulfur Cathodes in Lithium–

Sulfur Batteries. Adv. Sci. 2016, 3 (12), 1600101.

(97) Manthiram, A.; Fu, Y.; Chung, S.-H.; Zu, C.; Su, Y.-S. Rechargeable Lithium–

Sulfur Batteries. Chem. Rev. 2014, 114 (23), 11751–11787.

(98) Eftekhari, A. The Rise of Lithium–Selenium Batteries. Sustain. Energy Fuels 2017,

1 (1), 14–29.

(99) Hu, J.; Zhong, H.; Yan, X.; Zhang, L. Confining Selenium Disulfide in 3D Sulfur-

Doped Mesoporous Carbon for Rechargeable Lithium Batteries. Appl. Surf. Sci.

2018, 457, 705–711.

(100) CELASUN, Y.; COLIN, J.-F.; Martinet, S.; Peralta, D. Synthesis Method.

US20200086389A1, March 19, 2020.

(101) CELASUN, Y.; COLIN, J.-F.; Martinet, S.; Peralta, D. Lithiated Material.

US20200087159A1, March 19, 2020.

(102) Huang, X.; Wang, Z.; Knibbe, R.; Luo, B.; Ahad, S. A.; Sun, D.; Wang, L. Cyclic

Voltammetry in Lithium–Sulfur Batteries—Challenges and Opportunities.

Energy Technol. 2019, 7 (8), 1801001.

(103) Kim, T.; Choi, W.; Shin, H.-C.; Choi, J.-Y.; Kim, J. M.; Park, M.-S.; Yoon, W.-S.

Applications of Voltammetry in Lithium Ion Battery Research. J. Electrochem.

Sci. Technol. 2020, 11 (1), 14–25.

(104) Garche, J.; Dyer, C. K.; Moseley, P. T.; Ogumi, Z.; Rand, D. A. J.; Scrosati, B.

Encyclopedia of Electrochemical Power Sources; Newnes, 2013.

167

(105) Opitz, M.; Yue, J.; Wallauer, J.; Smarsly, B.; Roling, B. Mechanisms of Charge

Storage in Nanoparticulate TiO2 and Li4Ti5O12 Anodes: New Insights from Scan

Rate-Dependent Cyclic Voltammetry. Electrochimica Acta 2015, 168, 125–132.

(106) Chung, S.-H.; Chang, C.-H.; Manthiram, A. A Carbon-Cotton Cathode with

Ultrahigh-Loading Capability for Statically and Dynamically Stable Lithium–

Sulfur Batteries. ACS Nano 2016, 10 (11), 10462–10470.

(107) Huang, J.-Q.; Zhang, Q.; Peng, H.-J.; Liu, X.-Y.; Qian, W.-Z.; Wei, F. Ionic Shield

for Polysulfides towards Highly-Stable Lithium–Sulfur Batteries. Energy

Environ. Sci. 2013, 7 (1), 347–353.

(108) Balach, J.; Jaumann, T.; Klose, M.; Oswald, S.; Eckert, J.; Giebeler, L. Functional

Mesoporous Carbon-Coated Separator for Long-Life, High-Energy Lithium–

Sulfur Batteries. Adv. Funct. Mater. 2015, 25 (33), 5285–5291.

(109) Huo, S. H.; Qian, M.; Schaffer, G. B.; Crossin, E. 21 - Aluminium Powder

Metallurgy. In Fundamentals of Aluminium Metallurgy; Lumley, R., Ed.;

Woodhead Publishing Series in Metals and Surface Engineering; Woodhead

Publishing, 2011; 655–701.

(110) Moulder, J. F. Handbook of X-Ray Photoelectron Spectroscopy: A Reference

Book of Standard Spectra for Identification and Interpretation of XPS Data;

Physical Electronics Division, Perkin-Elmer Corporation, 1992.

(111) NIST X-ray Photoelectron Spectroscopy (XPS) Database Main Search Menu

https://srdata.nist.gov/xps/main_search_menu.aspx (accessed Jun 3, 2020).

(112) Cui, Y.; Abouimrane, A.; Lu, J.; Bolin, T.; Ren, Y.; Weng, W.; Sun, C.; Maroni, V.

A.; Heald, S. M.; Amine, K. (De)Lithiation Mechanism of Li/SeSx (x = 0–7)

Batteries Determined by in Situ Synchrotron X-Ray Diffraction and X-Ray

Absorption Spectroscopy. J. Am. Chem. Soc. 2013, 135 (21), 8047–8056.

(113) Wei Seh, Z.; Wang, H.; Hsu, P.-C.; Zhang, Q.; Li, W.; Zheng, G.; Yao, H.; Cui, Y.

Facile Synthesis of Li2S–Polypyrrole Composite Structures for High-

Performance Li2S Cathodes. Energy Environ. Sci. 2014, 7 (2), 672–676.

168

(114) Wu, F.; Lee, J. T.; Fan, F.; Nitta, N.; Kim, H.; Zhu, T.; Yushin, G. A Hierarchical

Particle–Shell Architecture for Long-Term Cycle Stability of Li2S Cathodes. Adv.

Mater. 2015, 27 (37), 5579–5586.

(115) Eftekhari, A. The Rise of Lithium–Selenium Batteries. Sustain. Energy Fuels 2017,

1 (1), 14–29.

(116) Liu, T.; Jia, M.; Zhang, Y.; Han, J.; Li, Y.; Bao, S.; Liu, D.; Jiang, J.; Xu, M.

Confined Selenium within Metal-Organic Frameworks Derived Porous Carbon

Microcubes as Cathode for Rechargeable Lithium–Selenium Batteries. J. Power

Sources 2017, 341, 53–59.

(117) Loor, M.; Bendt, G.; Schaumann, J.; Hagemann, U.; Heidelmann, M.; Wölper,

C.; Schulz, S. Synthesis of Sb2Se3 and Bi2Se3 Nanoparticles in Ionic Liquids at

Low Temperatures and Solid-State Structure of [C4C1Im]3[BiCl6]. Z. Für Anorg.

Allg. Chem. 2017, 643 (1), 60–68.

(118) Shkvarin, A. S.; Yarmoshenko, Yu. M.; Yablonskikh, M. V.; Merentsov, A. I.;

Shkvarina, E. G.; Titov, A. A.; Zhukov, Yu. M.; Titov, A. N. The Electronic

Structure Formation of CuxTiSe2 in a Wide Range (0.04 < x < 0.8) of Copper

Concentration. J. Chem. Phys. 2016, 144 (7), 074702.

(119) Luo, H.; Krizan, J. W.; Seibel, E. M.; Xie, W.; Sahasrabudhe, G. S.; Bergman, S.

L.; Phelan, B. F.; Tao, J.; Wang, Z.; Zhang, J.; Cava, R. J. Cr-Doped TiSe2 – A

Layered Dichalcogenide Spin Glass. Chem. Mater. 2015, 27 (19), 6810–6817.

(120) Hawkins, C. G.; Whittaker-Brooks, L. Controlling Sulfur Vacancies in TiS2–x

Cathode Insertion Hosts via the Conversion of TiS3 Nanobelts for Energy-

Storage Applications. ACS Appl. Nano Mater. 2018, 1 (2), 851–859.

(121) Martinez, H.; Auriel, C.; Gonbeau, D.; Loudet, M.; Pfister-Guillouzo, G. Studies

of 1T TiS2 by STM, AFM and XPS: The Mechanism of Hydrolysis in Air. Appl.

Surf. Sci. 1996, 93 (3), 231–235.

(122) Fantauzzi, M.; Elsener, B.; Atzei, D.; Rigoldi, A.; Rossi, A. Exploiting XPS for the

Identification of Sulfides and Polysulfides. RSC Adv. 2015, 5 (93), 75953–75963..

169

(123) Liang, X.; Hart, C.; Pang, Q.; Garsuch, A.; Weiss, T.; Nazar, L. F. A Highly

Efficient Polysulfide Mediator for Lithium–Sulfur Batteries. Nat. Commun. 2015,

6, 5682.

(124) Pang, Q.; Kundu, D.; Cuisinier, M.; Nazar, L. F. Surface-Enhanced Redox

Chemistry of Polysulphides on a Metallic and Polar Host for Lithium-Sulphur

Batteries. Nat. Commun. 2014, 5, 4759.

(125) Zhang, Z.; Yang, X.; Guo, Z.; Qu, Y.; Li, J.; Lai, Y. Selenium/Carbon-Rich Core–

Shell Composites as Cathode Materials for Rechargeable Lithium–Selenium

Batteries. J. Power Sources 2015, 279, 88–93.

170

Appendix Table - 1: Binding energy of species at different core levels

171

172

Résumé

« Synthèse et caractérisation de nouveaux matériaux d’électrode positive pour des

applications Li-ion à haute énergie »

Cette thèse concerne le développement de rocksalts désordonnés surlithiés pour les systèmes

Li-Ion à haute densité d’énergie. Dans un premier volet, les paramètres de synthèse ont été

optimisés pour améliorer les performances du rocksalt désordonné Li2.2NiTi0.2Nb0.6O4. Pour

comprendre l’origine de sa forte irréversibilité au premier cycle, des analyses in situ

structurales et électrochimiques avancées montrent un changement structural avec

l’apparition d’un désordre durant la première charge. Ensuite, le rocksalt désordonné Li2TiS3

a été préparé selon notre procédé breveté. De nouvelles compositions avec une substitution au

Sélénium, Li2TiSexS3-x, ont permis d’obtenir de fortes capacités de décharge à des potentiels

inférieurs avec une meilleure cyclabilité. L’activité réversible redox du soufre a été confirmée

par électrochimie et par analyses de surface ex situ mais des caractérisations plus poussées sont

nécessaires pour élucider le procédé redox complexe du sélénium.

Abstract

« Synthesis and characterization of new positive electrode materials for Li-ion high

energy applications »

This thesis focuses on the development of overlithiated disordered rocksalts for high-energy

Li-ion systems. Firstly, synthesis parameters have been optimized to improve the

performances of the disordered rocksalt Li2.2NiTi0.2Nb0.6O4. To examine its high irreversibility

(35%) at the first cycle, in situ advanced structural and electrochemical analyses have been

performed. Results show that a structural change and disordering happen during the first

charge. In a second part, the disordered rocksalt Li2TiS3 has been prepared with our patented

process. To improve cycling stability of the cells, Li2TiS3 has been partially substituted with

selenium and new Li2TiSexS3-x compositions have been prepared. Li2TiSexS3-x cells have large

discharge capacities at slightly lower potentials. Reversible sulfur redox activity is confirmed

by electrochemistry and ex situ surface analyses, however further characterizations are

required to elucidate the relatively complex selenium redox process.