structural tuning and spectroscopic characterizations of

91
Western University Western University Scholarship@Western Scholarship@Western Electronic Thesis and Dissertation Repository 10-26-2021 11:00 AM Structural Tuning and Spectroscopic Characterizations of Structural Tuning and Spectroscopic Characterizations of Polysulfide as Battery Materials Polysulfide as Battery Materials Daiqiang Liu, The University of Western Ontario Supervisor: Song, Yang, The University of Western Ontario A thesis submitted in partial fulfillment of the requirements for the Master of Science degree in Chemistry © Daiqiang Liu 2021 Follow this and additional works at: https://ir.lib.uwo.ca/etd Part of the Materials Chemistry Commons Recommended Citation Recommended Citation Liu, Daiqiang, "Structural Tuning and Spectroscopic Characterizations of Polysulfide as Battery Materials" (2021). Electronic Thesis and Dissertation Repository. 8238. https://ir.lib.uwo.ca/etd/8238 This Dissertation/Thesis is brought to you for free and open access by Scholarship@Western. It has been accepted for inclusion in Electronic Thesis and Dissertation Repository by an authorized administrator of Scholarship@Western. For more information, please contact [email protected].

Upload: others

Post on 08-Jun-2022

7 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Structural Tuning and Spectroscopic Characterizations of

Western University Western University

Scholarship@Western Scholarship@Western

Electronic Thesis and Dissertation Repository

10-26-2021 11:00 AM

Structural Tuning and Spectroscopic Characterizations of Structural Tuning and Spectroscopic Characterizations of

Polysulfide as Battery Materials Polysulfide as Battery Materials

Daiqiang Liu, The University of Western Ontario

Supervisor: Song, Yang, The University of Western Ontario

A thesis submitted in partial fulfillment of the requirements for the Master of Science degree in

Chemistry

© Daiqiang Liu 2021

Follow this and additional works at: https://ir.lib.uwo.ca/etd

Part of the Materials Chemistry Commons

Recommended Citation Recommended Citation Liu, Daiqiang, "Structural Tuning and Spectroscopic Characterizations of Polysulfide as Battery Materials" (2021). Electronic Thesis and Dissertation Repository. 8238. https://ir.lib.uwo.ca/etd/8238

This Dissertation/Thesis is brought to you for free and open access by Scholarship@Western. It has been accepted for inclusion in Electronic Thesis and Dissertation Repository by an authorized administrator of Scholarship@Western. For more information, please contact [email protected].

Page 2: Structural Tuning and Spectroscopic Characterizations of

ii

Abstract

Polysulfide materials have drawn extensive attention for next-generation battery development

since the current lithium-ion battery has almost reached its limit regarding energy density and

safety issues. Many phosphorus and sodium sulfide materials have been used in different battery

types, such as solid-state and sodium-sulfur batteries. However, there are still issues that prevent

these techniques from applications. In recent years, there has been increasing attention on

investigations of the structural and phase transformations of electrode and electrolyte materials

under high pressure. Many studies have shown that external pressure can affect structural

properties and influence electrical properties. In this study, three battery-related materials, P4S3,

P4S10, and Na2S4, were investigated under high-pressure conditions for the first time using in-situ

vibrational spectroscopy. Structural reversibility and phase transition upon compression were

studied for all three materials. In terms of P4S3, two phase transitions were identified, one at about

3 GPa and 7 GPa. The amorphous state was at 13 GPa. For P2S5, two phase transitions were

observed upon compression at 3 GPa and 7.5 GPa. When the pressure reached above 12 GPa, the

material became amorphous. For Na2S4, three phase transitions occurred upon compression at

about 1, 6 and 11 GPa. The material would reach the amorphous state at 16 GPa. Overall, our high-

pressure studies of these three materials could further understand how these materials react to

external compression and could promote the improvement of their performance as battery

materials or the development of new battery materials.

Page 3: Structural Tuning and Spectroscopic Characterizations of

iii

Keywords

Polysulfide, phosphorus sulfide, sodium sulfide, battery material, electrode, electrolyte, high

pressure, Raman spectroscopy, phase transition, reversibility, infrared spectroscopy, diamond

anvil cell.

Page 4: Structural Tuning and Spectroscopic Characterizations of

iv

Summary of Lay Audience

In recent years, Lithium-ion battery (LIB) as a mature battery technique has been almost reached

its limitation due to the increasing demand for batteries with better performance. Several battery

techniques have been considered to be the next-generation batteries, such as all-solid-state batteries

and sodium-sulfur batteries. Many sulfur-containing materials have been proposed for these

battery techniques. In Particular, Polysulfide played an important role in determining battery

performance. Therefore, understanding the structures and enabling the structural modifications of

these polysulfides are of great importance in improving the battery performance. Moreover, many

studies have shown that external pressure can affect structural properties and influence the

electrical properties of a material.

In this study, three battery-related polysulfide materials P4S3, P4S10, and Na2S4, were investigated

under high-pressure conditions for the first time using in-situ vibrational spectroscopy (Raman and

IR). This information allows the precise determination of structural stability and reversibility of

these materials. In addition, the pressure-induced modification was also observed for the first time.

This new structural information provides a further understanding of how these materials react to

external compression and could promote the improvement of their performance as battery

materials or the development of new battery materials.

Page 5: Structural Tuning and Spectroscopic Characterizations of

v

Co-Authorship Statement

In Chapter 3, the far-IR region (10-400 cm-1) data of P2S5 were collected by Dr. Zhenxian Liu

from Brookhaven National Laboratory NSLS-II.

Daiqiang Liu is responsible for the sample loading, data collection and data analysis of all the

Raman spectra and wrote the draft of this thesis. Dr. Yang Song edited the content.

Page 6: Structural Tuning and Spectroscopic Characterizations of

vi

Acknowledgements

First and foremost, I would like to express my sincere gratitude from the bottom of my heart to my

supervisor Prof. Yang Song, for offering me the opportunity to do the research and providing

invaluable support throughout the study. I appreciate everything he did for me, especially in

improving my academic writing skills. Without his patient guidance and excellent supervision, I

could not make this far.

I would also like to thank Prof. John F. Corrigan, Prof. Lijia Liu, and Prof. Yang Song for their

excellent graduate courses. I would like to thank the members of my examination committee, Prof.

T. K. Sham and Prof. Lijia Liu, for their insightful comments and valuable time. I would like to

thank Ms Susan England and Dr M. Naeem Shahid in the first-year chemistry Lab, for the excellent

working experience.

I would also like to thank all my group members: Mr Rongfeng Guan, Dr Shan Jiang, Miss Boqing

Li, and Miss Jingyan Liu, for all their helpful discussions and constructive feedbacks. I want to

extend my thanks to all people I have met in the chemistry department. I am glad to have you

during my research experience.

Finally, I would like to say thank you to my parents. Thank you for your unconditional love and

support, even sometimes I make bad choices. Thank you for helping me to shape my life with

positivity and passion. Without you, I’d never been the person I am today. Thank you for

everything.

Page 7: Structural Tuning and Spectroscopic Characterizations of

vii

Table of Contents

Abstract ········································································································· ii

Keywords ······································································································ iii

Summary of Lay Audience ················································································· iv

Co-Authorship Statement ···················································································· v

Acknowledgements ·························································································· vi

Table of Contents ····························································································· vii

List of Tables·································································································· ix

List of Figures ································································································· x

List of Abbreviations ······················································································· xiii

1 Introduction ······························································································· 1

1.1 Polysulfide materials ··············································································· 1

1.1.1 Polysulfide materials in battery applications ············································· 1

1.1.2 Phosphorus sulfides ·········································································· 4

1.1.3 Sodium polysulfides ·········································································· 6

1.2 High-pressure materials science·································································· 7

1.2.1 High-pressure phenomena ··································································· 7

1.2.2 Battery materials at high pressure ························································· 10

1.3 Motivations and outline ·········································································· 13

1.4 Reference ··························································································· 15

2 Methodology ···························································································· 22

2.1 High-pressure apparatus ·········································································· 22

2.1.1 Diamond anvil cell ·········································································· 22

2.1.2 High-pressure gauge········································································· 24

2.1.3 Pressure-transmitting medium ····························································· 25

2.2 In situ characterizations at high pressure ······················································ 26

2.2.1 Principles of vibrational spectroscopy ···················································· 26

2.2.2 Raman spectroscopy········································································· 27

2.2.3 Infrared spectroscopy ······································································· 29

2.3 References ·························································································· 32

3 Investigation on the structural evolution of P4S3 and P2S5 (P4S10) under high-pressure ······ 34

3.1 Introduction ························································································ 34

Page 8: Structural Tuning and Spectroscopic Characterizations of

viii

3.2 Experimental ······················································································· 38

3.2.1 Material ······················································································· 38

3.2.2 High-Pressure Experiment procedures ··················································· 39

3.2.3 Characterizations ············································································ 39

3.3 Result and Discussion ············································································ 40

3.3.1 Raman results of P4S3 upon compression and decompression ························ 40

3.3.2 Pressure dependence of Raman Modes of P4S3 ········································· 42

3.3.3 Computed unit cell parameters and equation of state of P4S3 ························· 45

3.3.4 Raman and IR Spectra of P2S5 at ambient pressure ····································· 48

3.3.5 In-situ high-pressure Raman spectroscopy of P2S5 ····································· 50

3.3.6 In-situ high-pressure IR spectroscopy of P2S5 ··········································· 51

3.3.7 Pressure dependence of IR and Raman modes of P2S5 ································· 52

3.3.8 Discussion ···················································································· 55

3.4 Conclusion ························································································· 56

3.5 Reference ··························································································· 57

4 Investigation of structural evolutions of Na2S4 under high-pressure ···························· 60

4.1 Introduction ························································································ 60

4.2 Experimental ······················································································· 63

4.2.1 Material ······················································································· 63

4.2.2 High-Pressure Experiment procedures ··················································· 63

4.2.3 Characterizations ············································································ 63

4.3 Result and Discussion ············································································ 64

4.3.1 Raman results of Na2S4 upon compression and decompression ······················ 64

4.3.2 Pressure effects on Raman modes of Na2S4 ············································· 67

4.3.3 Discussion ···················································································· 69

4.4 Conclusions ························································································ 70

4.5 Reference ··························································································· 72

5 Summary and future work ············································································· 75

5.1 Summary ··························································································· 75

5.2 Future Work ························································································ 76

Curriculum Vitae ····························································································· 78

Page 9: Structural Tuning and Spectroscopic Characterizations of

ix

List of Tables

Table 1.1 Selected representative high-pressure phenomena ......................................................... 9

Table 2.1 Commonly used PTMs with their pressure limitations 12, 16 ......................................... 26

Table 3.1 P4S3 physical and chemical properties information ...................................................... 35

Table 3.2 P4S3 crystal structure information12 .............................................................................. 36

Table 3.3 P2S5 (P4S10) physical and chemical properties information .......................................... 37

Table 3.4 P2S5 (P4S10) crystal structure information14 .................................................................. 37

Table 3.5 Peak Assignment of Raman Modes of P4S3.................................................................. 41

Table 3.6 Pressure dependence (dν/dP, cm-1/GPa) of vibrational modes of P4S3 upon compression

from Raman spectra ...................................................................................................................... 44

Table 3.7 Unite cell parameters of P4S3 as a function of pressure ................................................ 46

Table 3.8 Peak assignment of Raman and IR modes of P2S5 ....................................................... 49

Table 3.9 Pressure dependence (dv/dP, cm-1/GPa) of vibrational modes of P2S5 upon compression

from Raman and IR spectra. ......................................................................................................... 54

Table 4.1 Na2S4 physical and chemical properties information ................................................... 61

Table 4.2 Na2S4 crystal structure information9 ............................................................................. 61

Table 4.3 Peak Assignment of Raman Modes of Na2S4 ............................................................... 65

Table 4.4 Pressure dependence (dν/dP, cm-1/GPa) of vibrational modes of Na2S4 upon compression

from Raman spectra ...................................................................................................................... 69

Page 10: Structural Tuning and Spectroscopic Characterizations of

x

List of Figures

Figure 1.1 Li-S battery schematic diagram and "shuttle effect" caused by polysulfides.8 ............. 2

Figure 1.2 Structure of conventional battery (left) and all-solid-state battery (right)12 .................. 3

Figure 1.3 Structure of prototype Na-S battery15 ............................................................................ 4

Figure 1.4 Structures of phosphorus and phosphorus sulfide group20 ............................................ 5

Figure 1.5 The assessed Na-S phase diagram35 .............................................................................. 6

Figure 1.6 Examples of high-pressure phenomena45 ...................................................................... 8

Figure 1.7 Pressure-induced amorphization effect on resistance change of Li4Ti5O1265 .............. 10

Figure 1.8 Pressure-induced phase transition effect on the conductivity of Na2C6O6 67 .............. 11

Figure 1.9 Pressure-induced polymerization of anionic iodide chains with the formation of a 3D

network71 ....................................................................................................................................... 12

Figure 2.1 The cylinder part and the piston part of the DAC ....................................................... 23

Figure 2.2 Schematics of a symmetric diamond anvil cell6 .......................................................... 23

Figure 2.3 Origin of ruby R-line excitation and an example of ruby fluorescence shift under

different Pressure7 ......................................................................................................................... 24

Figure 2.4 Illustration of IR absorption, Rayleigh scattering and Raman scattering process ....... 27

Figure 2.5 Schematic illustration of the Raman system19 ............................................................. 29

Figure 2.6 Schematic illustration of the IR system19 .................................................................... 30

Figure 3.1 Molecular structure and crystal structure of P4S3. Orange and yellow spheres represent

phosphorus and sulfur atoms. ....................................................................................................... 35

Page 11: Structural Tuning and Spectroscopic Characterizations of

xi

Figure 3.2 Molecular structure and crystal structure of P4S10. Orange and yellow spheres represent

phosphorus and sulfur atoms. ....................................................................................................... 37

Figure 3.3 Raman Spectrum of P4S3 with peak assignment collected at 0.62 GPa ...................... 40

Figure 3.4 Raman spectra of P4S3 upon compression and decompression. The formation of new

Peaks is marked on the spectra by squares. .................................................................................. 42

Figure 3.5 Pressure dependence of Raman modes of P4S3. The vertical dashed lines indicate the

proposed phase boundaries, and different symbols represent Raman modes with different origins.

....................................................................................................................................................... 43

Figure 3.6 Pressure dependence of lattice parameters and unit cell volume from the calculations

(a). The cell volume data was compared with a third-order Birch-Murnaghan equation of state

(EOS) (b). ...................................................................................................................................... 47

Figure 3.7 Pressure dependence of normalized cell parameters and cell volume of P4S3. The dashed

line is phase boundaries concluded from Raman spectra. ............................................................ 48

Figure 3.8 Raman and IR spectra of P2S5 collected at ambient condition and 0.82 GPa, respectively.

....................................................................................................................................................... 49

Figure 3.9 Raman spectra of P2S5 upon compression and decompression ................................... 50

Figure 3.10 IR spectra of P2S5 upon compression ........................................................................ 52

Figure 3.11 Pressure dependence of Raman (a) and IR (b) modes of P2S5. The vertical dashed lines

indicate the proposed phase boundaries. Different symbols represent Raman modes with different

origins. .......................................................................................................................................... 53

Figure 4.1 a. Crystal structure of Na2S4 with space group. b. Crystal structure of Na2S4 along with

c. c. Structural formula. Purple and yellow spheres represent sodium and sulfur atoms. ............ 61

Figure 4.2 Raman spectrum of Na2S4 with peak assignment collected at 0.38 GPa .................... 64

Page 12: Structural Tuning and Spectroscopic Characterizations of

xii

Figure 4.3 Raman spectra of Na2S4 upon compression and decompression. Formation of new Peaks

and splitting of peaks are marked on the spectra by arrows. ........................................................ 66

Figure 4.4 Pressure dependence of selected Raman modes of Na2S4.The vertical dashed lines

indicate the proposed phase boundaries. Different symbols represent Raman modes with different

origins. .......................................................................................................................................... 67

Page 13: Structural Tuning and Spectroscopic Characterizations of

xiii

List of Abbreviations

ASSB all-solid-state battery

CCD charge-coupled device

DAC diamond anvil cell

EDM electrical discharge machining

EV electronic vehicle

FIS frontier synchrotron infrared spectroscopy

FTIR Fourier-transform infrared

GPa gigapascal

IR infrared

LIB lithium-ion battery

NSLS-II National Synchrotron Light Source II

PTM pressure transmitting medium

SEI solid electrolyte interface

SR synchrotron radiation

Page 14: Structural Tuning and Spectroscopic Characterizations of

1

Chapter 1

1 Introduction

1.1 Polysulfide materials

1.1.1 Polysulfide materials in battery applications

As a mature battery technique, lithium-ion battery (LIB) has been serving in various fields, from

portable electronics to electric vehicles, for over 30 years since its commercialization.1, 2 However,

the technique almost reaches its limitation for multiple parameters in terms of energy density

(theoretical capacity of 200-250 Wh/kg) and cycle life (~1000 cycles).3 Although it is still

dominant for portable electronics applications, it started showing deficiency in larger-scale

applications such as the electronic vehicle (EV) and stationary uses.3, 4 For example, its energy

density is not good enough for the market requirements; the preferred energy density for an EV

should reach a practical specific energy of 550 Wh/kg.4 Also, the cycle life and safety issues are

problematic for larger-scale stationary applications due to liquid electrolytes. The complex

composition of liquid electrolytes would lead to several issues, such as solid electrolyte interphase

(SEI), electrode dissolution/corrosion, and electrolyte flammability.5 All these issues have limited

the development of LIB technology to achieve higher energy density and better safeness to satisfy

the need for a next-generation battery.6 Sulfur, a safe, abundant nonmetal with a high theoretical

energy density (∼2600Wh kg−1), has been introduced as a promising cathode candidate for a

lithium battery to enhance the energy density.7 However, it was reported that the performance of

the Li-S battery was not as good as the expectation due to the "shuttle effect"(Figure 1.1).8 During

the discharging, sulfur at the cathode would react with lithium-ion to form long polysulfide chains.

Then these polysulfides would diffuse to the anode and react with lithium to form short chains of

polysulfide. This "shuttle effect" process will cause battery working materials to lose, leading to

short battery life.9 Therefore, in recent years, many sulfur-containing materials have been proposed

for different types of battery techniques. In particular, binary sulfide materials have drawn

extensive attention due to their relatively high specific capacity and energy density as electrodes10

and high ionic conductivity as solid electrolytes.6, 11

Page 15: Structural Tuning and Spectroscopic Characterizations of

2

Figure 1.1 Li-S battery schematic diagram and "shuttle effect" caused by polysulfides.8

The all-solid-state battery (ASSB) is introduced as one of the most promising energy technologies

for the next-generation battery. Organic liquid electrolytes are usually used in a traditional LIB

due to their high ionic conductivity and good wettability. However, liquid electrolytes would

damage the electrode (SEI) and lead to safety issues (flammability). Therefore, in an ASSB,

replacing the liquid electrolytes with solid electrolytes is the main concept of the technology (see

Figure 1.2).3 Advantages of solid-state electrolytes are the increased safety since most solid-state

electrolytes are non-flammable and have a more comprehensive operating temperature range(-50

to 150°C).3 In addition, sulfide materials, especially in the polysulfide form, are a good candidate

as a solid electrolyte due to their relatively high ionic conductivity (10-4 S/cm) among all solid

electrolyte materials. Typically, the ionic conductivity needs to be at least 10-4 S/cm for a good

solid electrolyte and even better if >10−3 S/cm at room temperature.3 However, the system is still

under development due to the compatibility between electrodes and solid electrolytes.

Page 16: Structural Tuning and Spectroscopic Characterizations of

3

Figure 1.2 Structure of conventional battery (left) and all-solid-state battery (right)12

Sodium-sulfur battery (Na-S battery) is another type of battery system considered for large-scale

applications such as stationary applications. The battery system was first introduced in 1966, and

it features a high energy density (230W/kg), high efficiency (90%) and long cycle life(~25 years).13

The battery system has the basis of molten salt technology, i.e. both electrodes are in a molten

state. Therefore, the battery needs to operate at 300°C to keep sodium and sulfur electrodes in a

molten state.14 The battery system separates the molten sodium negative electrode and molten

sulfur positive electrode by a solid beta alumina electrolyte.15 The structure of the system is shown

in Figure 1.3; Na+ ion can pass through the electrolyte and react with sulfur to form sodium

polysulfides as an intermediate product:

2Na + 4S = Na2S4 (1.1)

During the discharging process, as the reaction described eq. (1.1) Na+ flows through the solid

electrolyte, combines with sulfur, and produces 2.0 V potential.13 The process can be reversed in

the charging process; the sodium ions would be released from sodium polysulfides back to the

molten state through the electrolyte. Although the technology has already been used in Japan16,

there is still an unsolved problem with the corrosion effect caused by sodium polysulfide. The cell

case is corroded by sodium polysulfide and then cause the degradation of the cell performance.17,

18

Page 17: Structural Tuning and Spectroscopic Characterizations of

4

Figure 1.3 Structure of prototype Na-S battery15

As can be seen, in all the above Li-S and Na-S sulfur battery applications, polysulfide materials

played an important role in determining battery performance. Therefore, understanding the

structures and enabling the structural modifications of these polysulfides are of great importance

in improving the battery performance. Polysulfide materials are compounds that contain two or

more element sulfur atoms linked together by covalent bonds. It is foreseeable that many

compounds exist as polysulfide materials since the S-S bond energy is very similar to the bond

energy of C-C bonds19. In this research, three polysulfide materials were investigated. They belong

to two different polysulfide material classes, phosphorus sulfides and sodium polysulfides.

1.1.2 Phosphorus sulfides

Phosphorus sulfides are inorganic compounds with the formula P4Sx where x ≤10.20 The reaction

between phosphorus and sulfur was first reported in 1740 by A.S. Marggraf. 21 Through almost

three centuries of studies, there are four crystalline compounds, P4S3, P4S5, P4S7 and P4S10 (Figure

1.4), in the group that have been confirmed structurally by X-ray diffraction.22 In 2017, sulfur-

rich phosphorus sulfide materials (P4S10+n) were reported and tested for a potential battery

electrode application.23 In recent years, phosphorus sulfides have been considered suitable for

Page 18: Structural Tuning and Spectroscopic Characterizations of

5

electrode materials because of their relatively high specific capacity. In 2020, the electrochemical

behaviour of P4S3 was first reported featuring a high specific capacity as anode material for Li and

Na batteries.24 However, the low cycle life problem still needs to be addressed by future

development. Furthermore, phosphorus sulfides are an important component of solid electrolytes

due to their excellent performance in ASSB. Especially for P2S5 (P4S10), it is an essential

component of the Li2S-P2S5 binary system.25-32 The system is one of the most promising

glass/ceramic type solid electrolytes introduced by M. Tatsumisago group.28, 32 It features good

stability and high conductivities of more than 10−4 S/cm at room temperature without adding any

extra element (Si, Ge, Al), which has good potential as a solid electrolyte for the solid-state

battery.3 Nevertheless, issues still need to be solved, such as the low charging cycle life and

capacity fade caused by the difference in potential between the positive electrode and the

electrolyte.29

Figure 1.4 Structures of phosphorus and phosphorus sulfide group20

P4 P4S3 P4S5

P4S7 P4S10

Page 19: Structural Tuning and Spectroscopic Characterizations of

6

1.1.3 Sodium polysulfides

Since Friedrich reported the first sodium-sulfur phase diagram in 1914, the sodium-sulfur system

has been studied for over a hundred years. The research introduced sodium polysulfide and

speculated the existence of tetrasodium polysulfides, i.e., Na4Sx, where x = 3, 5, 7, 933. However,

in the same year, Rule and Thomas found that only disodium polysulfides could exist as

compounds.34 Later, their findings were confirmed by Pearson and Robinson, and the phase

diagram got improved(see Figure 1.5).35 Therefore, sodium polysulfides are a group of salt

materials with the typical formula Na2Sx with x =2, 3, 4, 5.

Figure 1.5 The assessed Na-S phase diagram35

Page 20: Structural Tuning and Spectroscopic Characterizations of

7

Furthermore, with the rise of the battery industry, in the 1960s, sodium-sulfur batteries were first

reported as a new high-energy-density power source.36 This discovery has gained the sodium-

sulfur system surge of interest again. Consequently, sodium polysulfide materials also received

massive attention due to their essential role in the battery system. In a Na-S battery, sodium

polysulfide would form an intermediate product during its charge/discharge process.37 The

polysulfide intermediates (Na2S4) formed during the charge/discharge process would react with

the solid electrolyte. This corrosion effect would result in several unsatisfactory performances such

as low discharge capacity, fast capacity decay during cycling, deterioration of the sodium anode.

The issue has been preventing further applications of the type of battery for decades. Furthermore,

some sodium polysulfide materials are also used in polysulfide bromide batteries as electrolyte

material.38 The technique is based on the reversible electrochemical reaction between two salt

solution electrolytes, sodium bromide and sodium polysulfide, via the following reaction:

3𝑁𝑎𝐵𝑟 + 𝑁𝑎2𝑆4 ↔ 2𝑁𝑎2𝑆2 + 𝑁𝑎𝐵𝑟3 (1.2)

The reaction proceeds from left to right during the charging process, and the discharging process

goes the other way.39 Two energy storage plants with the technique were built in UK and USA.

However, both locations have never been fully commissioned due to the high fabrication cost issue

caused by low energy density and the potential pollution issue from bromine.40

1.2 High-pressure materials science

1.2.1 High-pressure phenomena

Pressure is one of the primary thermodynamic parameters in the chemical system, as it influences

the equilibrium and rate of transformations of materials and reactions. Compared to other physical

parameters such as temperature and concentration, pressure can vary in a wide range of over 60

orders of magnitude.41 Pressure could be as low as 10-32 atm in the intergalactic spaces; it also

could be ~10-16 atm, which is the lowest pressure laboratory vacuum could reach; at the Earth's

centre, the pressure is about to 106 atm; and at the centre of a neutron star, the pressure could be

up to ~1030 atm.42 However, most of the modern chemical knowledge has been found or developed

from studies conducted at ambient pressure, near 1 atm.43 Therefore, High-pressure studies could

provide valuable information about materials under conditions where it exists and could discover

Page 21: Structural Tuning and Spectroscopic Characterizations of

8

novel structural or materials that are not observed under ambient condition. By definition, when a

system has constant entropy (S) or Temperature (T), the change of internal energy could be

measured by pressure (P) as shown in equation 1.2, where E, F, V are internal energy, Helmholtz

free energy and volume, respectively.44

𝑃 = − (𝜕𝐸

𝜕𝑉)

𝑆= − (

𝜕𝐹

𝜕𝑉)

𝑇 (1.3)

Figure 1.6 Examples of high-pressure phenomena45

When a material is under a high-pressure condition, the volume of the system would be reduced.

The volume reduction would result in the shortening of intermolecular distances and electron

Page 22: Structural Tuning and Spectroscopic Characterizations of

9

orbitals overlap.42 These effects may cause a significant change in the bonding pattern, reactivity

and the structural arrangement of the matter, resulting in a number of novel phenomena, such as

phase transformations, ionization, condensation, polymerization, amorphization, dissociation

atomization and metallization, which could not be found at ambient conditions.42 Many novel

structures and materials could be developed, and there are examples shown in Figure 1.6. For

example, iodine metallizes at 16 GPa46 and lithium forms chains or pairs when the pressure reaches

39 GPa.47 Furthermore, many pressure-induced novel structures and materials have been reported

across miscellaneous topics. Selected representative high-pressure phenomena findings are

summarized in Table 1.1. From the examples summarized in the table, it can be found that many

materials that are well-known under ambient conditions could have novel changes/structures

induced by high pressure. Therefore, high-pressure has been considered an effective tool for

modifying and tuning all chemical, structural, mechanical, electronic, magnetic, and phonon

properties.

Table 1.1 Selected representative high-pressure phenomena

Materials High-pressure Phenomena Reference

Hydrogen (H2)

Seven high-pressure solid phases have been reported so

far up to 425 GPa. probable metallization of hydrogen

above 400 GPa

48-53

CO2 Three phase transitions at 10, 20, and 40 GPa, became a

nonlinear quartz-like solid. 54

Methane-hydrogen

system

four molecular compounds, CH4(H2)2, (CH4)2H2,

CH4(H2)4, and CH4H2, at pressures up to 10 GPa. 55

Nitrogen-helium

system

formation of a stoichiometric, solid van der Waals

compound of composition He(N2)11 when helium-

nitrogen mixture under the high-pressure condition

56

Iron-nickel system Solid-solid phase transition above 200 GPa for Pure iron

and iron alloy would have a bcc structure in Earth's core. 57

Osmium (Os)

The material was compressed up to 750 GPa, unit cell

parameter ratio of Os exhibits anomalies at 150 and 440

GPa, which might be related to an electronic transition

associated with pressure-induced interactions between

core electrons.

58

Page 23: Structural Tuning and Spectroscopic Characterizations of

10

1.2.2 Battery materials at high pressure

As discussed above, external pressure could be an additional driving force for structural evolutions

and formations of new material. Therefore, many battery materials, including electrode and

electrolyte materials, have been studied under high pressure. Some of these studies have shown

that the pressure-induced structural transformation would affect the material's performance in the

battery application.59-64 Thus, the pressure-induced transformation provides a new method to

synthesize novel structures and materials with properties that differ from the ambient condition.

Figure 1.7 Pressure-induced amorphization effect on resistance change of Li4Ti5O1265

As mentioned in 1.1.1, LIB technology has almost reached its theoretical energy density. Therefore,

there are many high-pressure studies on electrode materials seeking novel changes to improve their

performance for LIB. In addition, the structural stability of electrode materials can also be

investigated by high pressure because structural stability is an important factor related to batteries'

capacity retention and cycle life. Moreover, the extreme pressure condition can tune the lattice

strain, an important parameter of structural stability and compressibility.66 For example, the high-

pressure study on an anode material, Li4Ti5O12, shows that the material becomes amorphous when

Page 24: Structural Tuning and Spectroscopic Characterizations of

11

the pressure reaches around 40 GPa. As a result, the pressure-induced amorphization process

enhanced the material's conductivity by one order of magnitude (shown in Figure 1.7).65

In another high-pressure study of electrode material, Na2C6O6 was investigated up to 30 GPa using

Raman, IR, X-ray diffraction and alternating current impedance spectroscopy. The conductivity of

Na2C6O6 was found to increase one order of magnitude after a phase transition occurred at 11 GPa

(shown in Figure 1.8).67 Furthermore, a high-pressure study of lithium acetylide (Li2C2), a cathode

candidate material, shows that the material's conductivity was enhanced 109 times by compression

to 40 GPa. 68 The enhanced electrical properties are associated with the pressure-induced

polymerization where the unsaturated bonds C≡C formed the extended framework that enhances

the conductivity.

Figure 1.8 Pressure-induced phase transition effect on the conductivity of Na2C6O6 67

Page 25: Structural Tuning and Spectroscopic Characterizations of

12

For electrolyte materials, structural phase transformations and electronic changes caused by

increased pressure would result in modulation of the bandgap, charge carrier concentration, and

carrier mobility.69 In particular, with the rapid development of all-solid-state batteries, an

increasing number of high-pressure studies on electrolyte materials were reported recently.69, 70

Several studies show that the structural changes would lead to enhanced ionic conductivity of the

material. For example, a high-pressure study of polyiodides showed that the applied external

pressure induced a phase transition to form a discrete-heptaiodide unit, followed by polymerization

leading to a 3D anionic network (Figure 1.9). This pressure-induced polymerization also

influenced the electrical conductivity by changing the insulating salt into a semiconducting

polymer, making it a potential solid-state electrolyte.71 All these examples discussed above suggest

that pressure is an efficient driving force to tune the structures and thus the associated transport

properties of electrode and electrolyte materials, shedding light on the improvement of battery

performance.

Figure 1.9 Pressure-induced polymerization of anionic iodide chains with the formation of a

3D network71

Page 26: Structural Tuning and Spectroscopic Characterizations of

13

1.3 Motivations and outline

As discussed above, pressure is a powerful tool for tuning the chemical bond, crystal structure,

and properties of materials. It could be a new route to synthesize new structures and materials.

Pressure-induced phase transformation would also influence the properties of materials and trigger

novel changes that enhance the properties of battery applications. Therefore, it is essential to

conduct high-pressure studies on batteries materials; these studies would provide valuable

information on materials’ structural compressibility, structural evolutions, and property changes

under high-pressure conditions. Moreover, as the examples mentioned in the previous section,

different transformations would occur when different materials are under extreme conditions. The

determining factors associated with the different battery performances remain unclear. Therefore,

it is worthy of investigating the structural stability of these polysulfide materials and examining

the possibility of new structure formation.

In particular, the two phosphorus sulfides, P4S10, and P4S3 are highly related to battery application

but still need further investigation. For instance, P4S10 has shown potential as a component of the

Li2S-P2S5 solid-state electrolyte system but with relatively low ionic conductivity. Moreover, P4S3

as an electrode material, the system's cycle life still needs to be improved. Although P4S3 and P4S10

are basic phosphorus sulfide materials and their reactivities and various properties have been

widely studied with extensive practical applications, the high-pressure studies of these materials

have not been reported yet and could relate to their battery applications.

In the Na-S battery system, Na2S4 is the primary sodium polysulfide intermediate product and is

also the cause of the corrosion issue. Even though the Na-S system was introduced in the 1960s

with extensive studies in boding, molecular and crystal structures at ambient conditions, the

material has not been studied under high pressure. Thus, the high-pressure study on Na2S4 may

provide additional information on the material and promote the corrosion issue solution and

associated properties desirable for battery applications.

Herein, the first high-pressure study of these three materials using in situ Raman spectroscopy and

IR spectroscopy are reported from this study. Specifically, P4S3, P4S10, and Na2S4 were compressed

up to 25, 17, and 30 GPa, respectively, followed by decompression and recovery. The Raman and

Page 27: Structural Tuning and Spectroscopic Characterizations of

14

IR spectroscopy revealed the exciting effect of high pressure on these three materials upon both

compression and decompression. It allows the precise determination of structural stability and

reversibility of these materials. In addition, the pressure-induced modification was also observed

for the first time. This new structural information provides a further understanding of how these

materials react to external compression. And may improve their performance as battery materials

or develop new battery materials.

Chapter 2 provides the experimental methodology of this research, from high-pressure generation

devices to structural investigation instrument setup. Chapter 3 will focus on the structural evolution

of two phosphorus sulfide materials under high pressure characterized by Raman and Infrared

spectroscopy. In Chapter 4, Na2S4 from the sodium sulfide group will be investigated under high

pressure. The study focused on the structural evolution and possible electronic property changes

during the compression. Finally, Chapter 5 provides a summary of the thesis and some suggestions

for future work.

Page 28: Structural Tuning and Spectroscopic Characterizations of

15

1.4 Reference

1. Asenbauer, J.; Eisenmann, T.; Kuenzel, M.; Kazzazi, A.; Chen, Z.; Bresser, D., The

Success Story of Graphite as a Lithium-ion Anode Material – Fundamentals, Remaining

Challenges, and Recent Developments Including Silicon (oxide) Composites. Sustainable Energy

& Fuels 2020.

2. Ye, H.; Zhang, Y.; Yin, Y.-X.; Cao, F.-F.; Guo, Y.-G., An Outlook on Low-Volume-

Change Lithium Metal Anodes for Long-Life Batteries. ACS Central Science 2020, 6 (5), 661-671.

3. Kudu, Ö. U.; Famprikis, T.; Fleutot, B.; Braida, M.-D.; Le Mercier, T.; Islam, M. S.;

Masquelier, C., A Review of Structural Properties and Synthesis Methods of Solid Electrolyte

Materials in the Li2S − P2S5 Binary System. Journal of Power Sources 2018, 407, 31-43.

4. Benveniste, G.; Rallo, H.; Canals Casals, L.; Merino, A.; Amante, B., Comparison of the

state of Lithium-Sulphur and lithium-ion batteries applied to electromobility. Journal of

Environmental Management 2018, 226, 1-12.

5. Wang, C.; Fu, K.; Kammampata, S. P.; McOwen, D. W.; Samson, A. J.; Zhang, L.;

Hitz, G. T.; Nolan, A. M.; Wachsman, E. D.; Mo, Y.; Thangadurai, V.; Hu, L., Garnet-Type

Solid-State Electrolytes: Materials, Interfaces, and Batteries. Chemical reviews 2020, 120 (10),

4257-4300.

6. Zhang, Q.; Cao, D.; Ma, Y.; Natan, A.; Aurora, P.; Zhu, H., Solid-State Batteries:

Sulfide-Based Solid-State Electrolytes: Synthesis, Stability, and Potential for All-Solid-State

Batteries (Adv. Mater. 44/2019). Advanced Materials 2019, 31 (44), 1970311.

7. Lithium sulfur batteries / edited by Mark Wild, and Greg Offer. Wiley: Hoboken, New

Jersey, 2019.

8. Xiong, S.; Regula, M.; Wang, D.; Song, J., Toward Better Lithium–Sulfur Batteries:

Functional Non-aqueous Liquid Electrolytes. Electrochemical Energy Reviews 2018, 1 (3), 388-

402.

9. Jo, S.-C.; Kim, M.-J.; Choi, I.-H.; Kim, B. G., Flexible High-Energy-Density Lithium-

Sulfur Batteries Using Nanocarbon-Embedded Fibrous Sulfur Cathodes and Membrane Separators.

NPG Asia Materials 2021, 13 (1).

10. Chuang, C.-C.; Hsieh, Y.-Y.; Chang, W.-C.; Tuan, H.-Y., Phosphorus-Sulfur/Graphene

Composites as Flexible Lithium-Sulfur Battery Cathodes with Super High Volumetric Capacity.

Chemical Engineering Journal 2020, 387, 123904.

Page 29: Structural Tuning and Spectroscopic Characterizations of

16

11. Li, Y.; Zhang, D.; Xu, X.; Wang, Z.; Liu, Z.; Shen, J.; Liu, J.; Zhu, M., Interface

engineering for composite cathodes in sulfide-based all-solid-state lithium batteries. Journal of

Energy Chemistry 2021, 60, 32-60.

12. Safe and Reliable All-Solid-State Battery [photograph].

http://www.spring8.or.jp/en/news_publications/research_highlights/no_49/.

13. Chen, H.; Xu, Y.; Liu, C.; He, F.; Hu, S., Chapter 24 - Storing Energy in China—An

Overview. In Storing Energy, Letcher, T. M., Ed. Elsevier: Oxford, 2016; pp 509-527.

14. Gupta, N.; Kaur, N.; Jain, S. K.; Singh Joshal, K., Chapter 3 - Smart grid power system.

In Advances in Smart Grid Power System, Tomar, A.; Kandari, R., Eds. Academic Press: 2021; pp

47-71.

15. Hussain, F.; Rahman, M. Z.; Sivasengaran, A. N.; Hasanuzzaman, M., Chapter 6 - Energy

storage technologies. In Energy for Sustainable Development, Hasanuzzaman, M. D.; Rahim, N.

A., Eds. Academic Press: 2020; pp 125-165.

16. Gao, D. W., Chapter 1 - Basic Concepts and Control Architecture of Microgrids. In Energy

Storage for Sustainable Microgrid, Gao, D. W., Ed. Academic Press: Oxford, 2015; pp 1-34.

17. Okuyama, R.; Nakashima, H.; Sano, T.; Nomura, E., The Effect of Metal Sulfides in the

Cathode on Na/S Battery Performance. Journal of Power Sources 2001, 93 (1), 50-54.

18. Heusler, K. E.; Grzegorzewski, A.; Knödler, R., Corrosion of Metallic Materials in Sodium

Polysulfide. Journal of The Electrochemical Society 1993, 140 (2), 426-431.

19. David Lee Pringle, The Nature of the Polysulfide Anion. Retrospective Theses and

Dissertations 1967.

20. Cowley, A. H., The Structures and Reactions of the Phosphorus Sulfides. Journal of

Chemical Education 1964, 41 (10), 530.

21. Pernert, J. C.; Brown, J. H., Some Reactions and Properties of the Phosphorus Sulfides.

Chemical & Engineering News Archive 1949, 27 (30), 2143-2145.

22. Jensen, J. O.; Zeroka, D., Theoretical Studies of the Infrared and Raman Spectra of P4S10.

Journal of Molecular Structure: THEOCHEM 1999, 487 (3), 267-274.

23. Li, X.; Liang, J.; Lu, Y.; Hou, Z.; Cheng, Q.; Zhu, Y.; Qian, Y., Sulfur-Rich Phosphorus

Sulfide Molecules for Use in Rechargeable Lithium Batteries. Angew Chem Int Ed Engl 2017, 56

(11), 2937-2941.

Page 30: Structural Tuning and Spectroscopic Characterizations of

17

24. Wei, D.; Yin, J.; Ju, Z.; Zeng, S.; Li, H.; Zhao, W.; Wei, Y.; Li, H., Cage-Like P4S3

Molecule as Promising Anode with High Capacity and Cycling Stability for Li+/Na+/K+ Storage.

Journal of Energy Chemistry 2020, 50, 187-194.

25. Xu, Y.-H.; Li, W.-Z.; Fan, B.; Fan, P.; Luo, Z.-K.; Wang, F.; Zhang, X.-H.; Ma, H.-

L.; Xue, B., Stabilizing electrode/electrolyte interface in Li-S batteries using liquid/solid Li2S-

P2S5 hybrid electrolyte. Applied Surface Science 2021, 546, 149034.

26. Chang, G. H.; Oh, Y. S.; Kang, S.; Park, J.-Y.; Lim, H.-T., Single-Step Prepared Li2S-

P2S5-C Composite Cathode for High Areal Capacity All-Solid-State Lithium-ion Batteries.

Electrochimica Acta 2020, 358, 136884.

27. Kato, A.; Yamamoto, M.; Sakuda, A.; Hayashi, A.; Tatsumisago, M., Mechanical

Properties of Li2S–P2S5 Glasses with Lithium Halides and Application in All-Solid-State Batteries.

ACS Applied Energy Materials 2018, 1 (3), 1002-1007.

28. Agostini, M.; Aihara, Y.; Yamada, T.; Scrosati, B.; Hassoun, J., A Lithium–Sulfur

Battery Using a Solid, Glass-Type P2S5–Li2S Electrolyte. Solid State Ionics 2013, 244, 48-51.

29. Trevey, J. E.; Gilsdorf, J. R.; Miller, S. W.; Lee, S.-H., Li2S–Li2O–P2S5 Solid Electrolyte

for All-Solid-State Lithium Batteries. Solid State Ionics 2012, 214, 25-30.

30. Tsukasaki, H.; Morimoto, H.; Mori, S., Ionic Conductivity and Thermal Stability of Li2O–

Li2S–P2S5 Oxysulfide Glass. Solid State Ionics 2020, 347, 115267.

31. Garcia-Mendez, R.; Smith, J. G.; Neuefeind, J. C.; Siegel, D. J.; Sakamoto, J., Correlating

Macro and Atomic Structure with Elastic Properties and Ionic Transport of Glassy Li2S-P2S5 (LPS)

Solid Electrolyte for Solid-State Li Metal Batteries. Advanced Energy Materials 2020, 10 (19),

2000335.

32. Mizuno, F.; Hama, S.; Hayashi, A.; Tadanaga, K.; Minami, T.; Tatsumisago, M., All

Solid-state Lithium Secondary Batteries Using High Lithium Ion Conducting Li2S–P2S5 Glass-

Ceramics. Chemistry Letters 2002, 31 (12), 1244-1245.

33. Friedrich, K.; Waehlert, M., Investigation of Layer-Forming Systems. Metall und Erz 1914,

160-167.

34. Rule, A.; Thomas, J. S., XXIII.—The Polysulphides of the Alkali Metals. Part I. The

Polysulphides of Sodium. Journal of the Chemical Society, Transactions 1914, 105 (0), 177-189.

35. Pearson, T. G.; Robinson, P. L., CLXXXIX.—The Polysulphides of the Alkali Metals. Part

I. Sodium (i). Journal of the Chemical Society (Resumed) 1930, (0), 1473-1497.

Page 31: Structural Tuning and Spectroscopic Characterizations of

18

36. Oei, D.-G., Sodium-sulfur system. I. Differential thermal analysis. Inorganic Chemistry

1973, 12 (2), 435-437.

37. Kim, I.; Park, J.-Y.; Kim, C.; Park, J.-W.; Ahn, J.-P.; Ahn, J.-H.; Kim, K.-W.; Ahn, H.-

J., Sodium Polysulfides during Charge/Discharge of the Room-Temperature Na/S Battery Using

TEGDME Electrolyte. Journal of The Electrochemical Society 2016, 163 (5), A611-A616.

38. Rydh, C. J.; Sandén, B. A., Energy Analysis of Batteries in Photovoltaic Systems. Part I:

Performance and Energy Requirements. Energy Conversion and Management 2005, 46 (11),

1957-1979.

39. Chen, H.; Cong, T. N.; Yang, W.; Tan, C.; Li, Y.; Ding, Y., Progress in Electrical Energy

Storage System: A Critical Review. Progress in Natural Science 2009, 19 (3), 291-312.

40. Zhang, H., Chapter 9 - Polysulfide-bromine flow batteries (PBBs) for medium- and large-

scale energy storage. In Advances in Batteries for Medium and Large-Scale Energy Storage,

Menictas, C.; Skyllas-Kazacos, M.; Lim, T. M., Eds. Woodhead Publishing: 2015; pp 317-327.

41. Pascarelli, S., Studies of Matter at Extreme Conditions. In Synchrotron Radiation: Basics,

Methods and Applications, Mobilio, S.; Boscherini, F.; Meneghini, C., Eds. Springer Berlin

Heidelberg: Berlin, Heidelberg, 2015; pp 737-760.

42. Schettino, V.; Bini, R., Constraining Molecules at the Closest Approach: Chemistry at High

Pressure. Chemical Society Reviews 2007, 36 (6), 869-880.

43. McMillan, P. F., Chemistry at high pressure. Chemical Society Reviews 2006, 35 (10), 855-

857.

44. Shen, G.; Mao, H. K., High-Pressure Studies with X-Rays Using Diamond Anvil Cells.

Reports on Progress in Physics 2016, 80 (1), 016101.

45. Grochala, W.; Hoffmann, R.; Feng, J.; Ashcroft, N. W., The Chemical Imagination at

Work in Very Tight Places. Angewandte Chemie International Edition 2007, 46 (20), 3620-3642.

46. Chen, L.; Zhang, Z.; Che, R.; Yu, T.; Wang, J., Structural Phase Transitions of Iodine at

High Pressure. Chinese Physics Letters 1994, 11 (2), 99-101.

47. Hanfland, M.; Syassen, K.; Christensen, N. E.; Novikov, D. L., New High-Pressure Phases

of Lithium. Nature 2000, 408 (6809), 174-178.

48. Ji, C.; Li, B.; Liu, W.; Smith, J. S.; Majumdar, A.; Luo, W.; Ahuja, R.; Shu, J.; Wang,

J.; Sinogeikin, S.; Meng, Y.; Prakapenka, V. B.; Greenberg, E.; Xu, R.; Huang, X.; Yang, W.;

Page 32: Structural Tuning and Spectroscopic Characterizations of

19

Shen, G.; Mao, W. L.; Mao, H.-K., Ultrahigh-Pressure Isostructural Electronic Transitions in

Hydrogen. Nature 2019, 573 (7775), 558-562.

49. Dalladay-Simpson, P.; Howie, R. T.; Gregoryanz, E., Evidence for a New Phase of Dense

Hydrogen above 325 Gigapascals. Nature 2016, 529 (7584), 63-67.

50. Loubeyre, P.; Occelli, F.; LeToullec, R., Optical Studies of Solid Hydrogen to 320 GPa

and Evidence for Black Hydrogen. Nature 2002, 416 (6881), 613-617.

51. Loubeyre, P.; LeToullec, R.; Hausermann, D.; Hanfland, M.; Hemley, R. J.; Mao, H.

K.; Finger, L. W., X-Ray Diffraction and Equation of State of Hydrogen at Megabar Pressures.

Nature 1996, 383 (6602), 702-704.

52. Narayana, C.; Luo, H.; Orloff, J.; Ruoff, A. L., Solid Hydrogen at 342 GPa: No Evidence

for an Alkali Metal. Nature 1998, 393 (6680), 46-49.

53. Loubeyre, P.; Occelli, F.; Dumas, P., Synchrotron Infrared Spectroscopic Evidence of the

Probable Transition to Metal Hydrogen. Nature 2020, 577 (7792), 631-635.

54. Iota, V.; Yoo, C. S.; Cynn, H., Quartzlike Carbon Dioxide: An Optically Nonlinear

Extended Solid at High Pressures and Temperatures. Science 1999, 283 (5407), 1510-1513.

55. Somayazulu, M. S.; Finger, L. W.; Hemley, R. J.; Mao, H. K., High-Pressure Compounds

in Methane-Hydrogen Mixtures. Science 1996, 271 (5254), 1400.

56. Vos, W. L.; Finger, L. W.; Hemley, R. J.; Hu, J. Z.; Mao, H. K.; Schouten, J. A., A High-

Pressure Vanderwaals Compound in Solid Nitrogen Helium Mixtures. Nature. 1992, 358 (6381),

46-48.

57. Dubrovinsky, L.; Dubrovinskaia, N.; Narygina, O.; Kantor, I.; Kuznetzov, A.;

Prakapenka, V. B.; Vitos, L.; Johansson, B.; Mikhaylushkin, A. S.; Simak, S. I.; Abrikosov, I.

A., Body-Centered Cubic Iron-Nickel Alloy in Earth Core. Science 2007, 316 (5833), 1880.

58. Dubrovinsky, L.; Dubrovinskaia, N.; Bykova, E.; Bykov, M.; Prakapenka, V.; Prescher,

C.; Glazyrin, K.; Liermann, H. P.; Hanfland, M.; Ekholm, M.; Feng, Q.; Pourovskii, L. V.;

Katsnelson, M. I.; Wills, J. M.; Abrikosov, I. A., The Most Incompressible Metal Osmium at

Static Pressures above 750 Gigapascals. Nature 2015, 525 (7568), 226-229.

59. Gallardo-Amores, J. M.; Biskup, N.; Amador, U.; Persson, K.; Ceder, G.; Morán, E.;

Arroyo y de Dompablo, M. E., Computational and Experimental Investigation of the

Transformation of V2O5 Under Pressure. Chemistry of Materials 2007, 19 (22), 5262-5271.

Page 33: Structural Tuning and Spectroscopic Characterizations of

20

60. García-Moreno, O.; Alvarez-Vega, M.; García-Jaca, J.; Gallardo-Amores, J. M.; Sanjuán,

M. L.; Amador, U., Influence of the Structure on the Electrochemical Performance of Lithium

Transition Metal Phosphates as Cathodic Materials in Rechargeable Lithium Batteries:  A New

High-Pressure Form of LiMPO4 (M = Fe and Ni). Chemistry of Materials 2001, 13 (5), 1570-1576.

61. Amador, U.; Gallardo-Amores, J. M.; Heymann, G.; Huppertz, H.; Morán, E.; Arroyo y

de Dompablo, M. E., High pressure polymorphs of LiCoPO4 and LiCoAsO4. Solid State Sciences

2009, 11 (2), 343-348.

62. Yamaura, K.; Huang, Q.; Zhang, L.; Takada, K.; Baba, Y.; Nagai, T.; Matsui, Y.;

Kosuda, K.; Takayama-Muromachi, E., Spinel-to-CaFe2O4-Type Structural Transformation in

LiMn2O4 under High Pressure. Journal of the American Chemical Society 2006, 128 (29), 9448-

9456.

63. Arroyo-deDompablo, M. E.; Dominko, R.; Gallardo-Amores, J. M.; Dupont, L.; Mali,

G.; Ehrenberg, H.; Jamnik, J.; Morán, E., On the Energetic Stability and Electrochemistry of

Li2MnSiO4 Polymorphs. Chemistry of Materials 2008, 20 (17), 5574-5584.

64. Yoncheva, M.; Stoyanova, R.; Zhecheva, E.; Alcántara, R.; Ortiz, G.; Tirado, J. L., Effect

of the High Pressure on the Structure and Intercalation Properties of Lithium–Nickel–Manganese

Oxides. Journal of Solid State Chemistry 2007, 180 (6), 1816-1825.

65. Huang, Y.; He, Y.; Sheng, H.; Lu, X.; Dong, H.; Samanta, S.; Dong, H.; Li, X.; Kim,

D. Y.; Mao, H.-k.; Liu, Y.; Li, H.; Li, H.; Wang, L., Li-Ion Battery Material under High Pressure:

Amorphization and Enhanced Conductivity of Li4Ti5O12. National Science Review 2018, 6 (2),

239-246.

66. Xiao, F.; Dong, Z.; Mao, H.; Liu, J.; Sun, X.; Song, Y., Morphology- and Lattice

Stability-Dependent Performance of Nanostructured Li4Ti5O12 Probed by in situ High-Pressure

Raman Spectroscopy and Synchrotron X-Ray Diffraction. CrystEngComm 2016, 18 (5), 736-743.

67. Wang, X.; Zhang, P.; Tang, X.; Guan, J.; Lin, X.; Wang, Y.; Dong, X.; Yue, B.; Yan,

J.; Li, K.; Zheng, H.; Mao, H.-k., Structure and Electrical Performance of Na2C6O6 under High

Pressure. The Journal of Physical Chemistry C 2019, 123 (28), 17163-17169.

68. Wang, L.; Dong, X.; Wang, Y.; Zheng, H.; Li, K.; Peng, X.; Mao, H.-k.; Jin, C.; Meng,

Y.; Huang, M.; Zhao, Z., Pressure-Induced Polymerization and Disproportionation of Li2C2

Accompanied with Irreversible Conductivity Enhancement. The Journal of Physical Chemistry

Letters 2017, 8 (17), 4241-4245.

Page 34: Structural Tuning and Spectroscopic Characterizations of

21

69. Wu, L.; Dai, L.; Li, H.; Zhuang, Y.; Liu, K., Pressure-Induced Improvement of Grain

Boundary Properties in Yttrium-Doped BaZrO3. Journal of Physics D: Applied Physics 2016, 49

(34), 345102.

70. Dai, L. D.; Wu, L.; Li, H. P.; Hu, H. Y.; Zhuang, Y. K.; Liu, K. X., Pressure-Induced

Phase-Transition and Improvement of the Microdielectric Properties in Yttrium-Doped SrZrO3.

EPL 2016, 114 (5).

71. Poręba, T.; Ernst, M.; Zimmer, D.; Macchi, P.; Casati, N., Pressure-Induced

Polymerization and Electrical Conductivity of a Polyiodide. Angewandte Chemie International

Edition 2019, 58 (20), 6625-6629.

Page 35: Structural Tuning and Spectroscopic Characterizations of

22

Chapter 2

2 Methodology

2.1 High-pressure apparatus

2.1.1 Diamond anvil cell

Diamond anvil cell (DAC) is commonly used in material science to generate extreme high-pressure

conditions. In 1905, a high-pressure generating mechanical device was invented by Percy

Bridgman1, and it expanded the high-pressure range to two thousand atmospheres for a laboratory

experiment.2 By inspiration of Bridgman's invention, the first DAC was introduced in 1958.3 The

creation of DAC was a significant breakthrough for high-pressure studies. The device allows

scientists and researchers to create up to 400 gigapascals (GPa) pressure by compressing a tiny

amount of sample between two opposing diamond anvils.4 The working principle of the device is

straightforward:

𝑃 =𝐹

𝐴 (2.1)

P is pressure, F is the force applied to the object, and A is the area applied by force, i.e., pressure

is applied force per unit area. The diamond used for DAC is customized in a "brilliant cut" design

to have the culet of the diamond in a very tiny size which could vary from tens to hundreds of

micrometres(μm) depending on the target pressure of the experiment. Therefore, even a moderate

force applying to the diamond would generate very high pressure to the sample between diamonds.

It is noted that diamond, being the hardest material in the world, can resist extreme high pressure

without breaking. More importantly, its transparency allows a broad region of the electromagnetic

to go through, which makes the device an ideal tool for various spectroscopic and diffraction

measurements in situ.5

Page 36: Structural Tuning and Spectroscopic Characterizations of

23

Figure 2.1 The cylinder part and the piston part of the DAC

In this study, the symmetric piston-cylinder type of DAC was used for generating high pressure,

as shown in Figure 2.1. The cell consists of two parts, the cylinder part and the piston part. As the

schematic diagram of DAC shown in Figure 2.2, two brilliant-cut diamonds of the same size were

first glued onto two seats. These seats are usually made of tungsten carbide. The two seats then

were fixed into the two parts of the cell oppositely. Then two diamonds are aligned by adjusting

the position of two seats on each part of the cell. A pre-indented stainless-steel gasket with a drilled

hole that is typically one-third of the culet size in diameter is used as the sample chamber between

the two diamonds. Samples and pressure gauges (ruby etc.) are loaded into the sample chamber

for measurements. Four screws with the same spring washers' pattern are installed to generate

pressure to the sample. By tightening the four screws equally, pressures in the GPa range can be

generated to the sample in the chamber.

Figure 2.2 Schematics of a symmetric diamond anvil cell6

Cylinder Piston

Page 37: Structural Tuning and Spectroscopic Characterizations of

24

For different spectroscopy measurements, different types of diamond would be needed. Typically,

diamonds are classified into two types, type I and type II. The diamond type is distinguished by

the impurity level of nitrogen content, the most common impurity in the diamond crystal structure.

Type I diamonds contain up to 0.1% of nitrogen impurity content, where type II diamond contains

much less nitrogen impurity content. Furthermore, type I diamond is used for Raman spectroscopy

only since type I diamond contains sufficient nitrogen content to be measurable by IR absorption

spectroscopy. On the other hand, type II diamond is ideal for IR spectroscopy. In this study, the

diamonds culet size is 400 µm, and the diamonds are type I for Raman spectroscopy and type II

for IR spectroscopy.

2.1.2 High-pressure gauge

Figure 2.3 Origin of ruby R-line excitation and an example of ruby fluorescence shift under

different Pressure7

To determine the pressure inside the DAC chamber, ruby fluorescence is one of the major methods.

Forman first introduced the method in 1972.8 The technique is a simple, rapid way for measuring

pressure in various laboratories. Ruby is a type of corundum (α-Al2O3) that contains chromium.

The substitution of aluminium by Cr3+ leads to ruby formation.9 Ruby fluorescence spectrum is

very sensitive to pressure change. It is composed of two intense peaks, R1 and R2, at 694.2nm and

692.7 nm under zero pressure at room temperature (298 k).10 As pressure increases, the two peaks

would shift toward the longer wavenumber/lower energy side. As shown in Figure 2.3, the

fluorescence peaks are due to the excitation of Cr3+ by a laser source. The Cr content usually is

Page 38: Structural Tuning and Spectroscopic Characterizations of

25

about 0.5% of the ruby for desired fluorescence. The linear relationship between pressure and the

R1-line shift is calibrated and normalized by Mao and co-workers for pressure measurement.11 It

is good up to 80 GPa for a quasi-hydrostatic environment, and a slight non-linearity was found at

higher pressure.

𝑃 =1904

𝐵[(1 +

∆𝜆

694.24)

𝐵

− 1] (2.2)

In the equation, P represents the current pressure impact on the ruby in GPa, ∆λ represents the

wavelength R1 shifted, and B is a parameter equal to 7.665 for quasi-hydrostatic conditions. The

measurement is accurate to ± 0.05 GPa.

2.1.3 Pressure-transmitting medium

A Pressure-transmitting medium (PTM) is usually a compressible fluid material. Loading a PTM

material into a gasket with the sample can convert the applied stress into a uniform hydrostatic

pressure with no gradient. Hydrostatic pressure is a thermodynamic parameter which means the

pressure is equilibrium at any point in the sample chamber. For the high-pressure study, the

hydrostatic condition is preferred. The experimental result collected under hydrostatic conditions

is believed to have better accuracy and be more comparable with theoretical studies.12 In contrast,

non-hydrostatic stress could distort crystal lattice, lower crystal symmetry, and induce a splitting

of optical spectral lines.13 The use of PTM can help to produce hydrostatic conditions. A proper

PTM usually needs to meet several requirements, such as low thermal conductivity, no reaction

with sample and the ability to keep the pressure under the hydrostatic condition as high as possible.

Some commonly used PTMs with pressure limitations are listed in Table 2.1. In this study, KBr

(Potassium bromide) was used as a solid PTM for IR measurement. KBr is widely used in mid-IR

measurement due to its transmission window from 400 to 4000 cm-1.14 However, KBr is not like

those fluid PTM. The hydrostatic limit of KBr is not well defined. Moreover, for good quality of

IR spectra, the thickness of the sample needs to be adjusted for the desired absorption.15 By filling

the amount of KBr as a spacer into the sample chamber, the thickness of the sample becomes

adjustable and controllable.

Page 39: Structural Tuning and Spectroscopic Characterizations of

26

Table 2.1 Commonly used PTMs with their pressure limitations 12, 16

Pressure transmitting medium Hydrostatic limits (GPa)

1:1 iso-n-pentane 7.4

4:1 methanol-ethanol 10.5

Silicon oil 20

Fluorinert 2.3

Daphne 7474 3.7

Argon 10

Nitrogen 13

Neon 50

Helium 40

2.2 In situ characterizations at high pressure

2.2.1 Principles of vibrational spectroscopy

Vibrational spectroscopy is a technique to determine the molecular structure based on vibrational

energy. When a molecule interacts with electromagnetic radiation, vibration transitions can occur

due to the absorption or emission of the radiation by the bond in the molecule. This type of

transition usually happens between different vibrational levels of the same electronic state. In a

molecule, each chemical bond can vibrate in many ways, resulting in different vibration modes.

Vibration modes can be classified into two types: stretching and bending. Each vibration mode

would have unique vibrational energy. Therefore, measuring the information of vibrational

transition energy could provide information on molecular structures, chemical bonding, and the

physical configuration of the compound. Additionally, vibrational spectroscopy is a powerful

technique to monitor the pressure effects on structural evolution and bonding changes in high-

pressure studies.

Page 40: Structural Tuning and Spectroscopic Characterizations of

27

Figure 2.4 Illustration of IR absorption, Rayleigh scattering and Raman scattering process

Infrared spectroscopy and Raman spectroscopy are two widely used spectroscopic techniques. The

infrared spectrum measures the absorption of infrared light at each vibrational transition, and the

Raman spectroscopy is based on the scattering process. By utilizing the difference between the

incident and scattered radiation, the vibrational transition energy can be determined by the Raman

spectrum. The illustration of IR absorption and Raman scattering process is shown in Figure 2.4.

In a molecule, some vibrational modes are IR-active, and some are Raman-active. Thus, the two

spectroscopic techniques provide complementary information.

2.2.2 Raman spectroscopy

Raman spectroscopy is a non-destructive technique to determine vibrational modes of molecules.

It could provide valuable information on phase transition, intermolecular interactions, and material

structure changes under high pressure.17 The principle of Raman spectroscopy could be explained

by the interaction between light and chemical bonds in the material. When a sample is exposed to

monochromatic light in the visible region, part of the light is scattered by the sample in all

directions, and the scattered light usually could be observed in the direction perpendicular to the

original light.18 There are two types of scattered light based on frequency: one is called Rayleigh

scattering if the scattered light has the same frequency as the incident light. In contrast, the other

Page 41: Structural Tuning and Spectroscopic Characterizations of

28

is called Raman scattering, which is very weak (about 10-5 of the incident beam) and has a different

frequency from the incident light.18 Raman scattering can be separated into stokes lines and anti-

stokes lines. If the frequency of the scattered light is less than that of the incident light, stokes lines

are observed on the Raman spectrum. By contrast, anti-stokes lines will be observed when the

frequency of emitted light is greater than that of the incident light. Usually, the anti-stoke lines

would be weaker than stoke liens since most molecules are at the ground state.18 As mentioned

previously, IR and Raman spectroscopies are based on different selection rules. For vibrational

transitions to be Raman active, a molecule must have a change in its polarizability during vibration

i.e., the derivative of the induced dipole moment with the vibrational displacement must be

nonzero.

(𝑑𝛼

𝑑𝑞)

𝑒 ≠ 0 (2.3)

Where α is the variation of polarizability, q is the normal coordinate, and e is the equilibrium

position. Thus, polarizability is based on the relative tendency of the electron cloud to be distorted

from its original position.

For this study, all the Raman measurements were conducted by a user-customized Raman system

setup. The schematic diagram of the system used is shown in Figure 2.5. The laser used as the

excitation source for the project is a solid-state laser with a wavelength of 532 nm. An Olympus

microscope with a 15x eyepiece and 20x objective focused the laser beam onto the sample. Two

beam splitters are set inside for laser directing, and a built-in CCD camera allows in-situ

observation of the monitor. An adjustable sample stage is used to align the DAC under the

microscope. A CCD detector is used to collect the Raman spectrum. A liquid nitrogen tank is

attached to the detector to keep the detector temperature at about 120 °C. The spectrometer has a

triple grating system with three resolutions; 1200 lines/mm grating is used for all measurements.

WinSpec software is used to control the system and plot spectra. Standard neon lines are used to

calibrate the system and have a possible resolution of ±1 cm-1.

Page 42: Structural Tuning and Spectroscopic Characterizations of

29

Figure 2.5 Schematic illustration of the Raman system19

2.2.3 Infrared spectroscopy

Generally, Infrared (IR) spectroscopy is a non-destructive technique to determine the infrared light

absorption of the analyte as a function of frequency. When an infrared light goes through a

molecule, the molecule absorbs E=hv from the IR source at each vibrational transition. The

intensity of IR absorption is governed by Beer-Lambert law:

𝐼 = 𝐼0𝑒−𝜀𝑐𝑑 (2.4)

I0 is the intensity of the incident light, I is the intensity of the transmitted light, 𝜀 is the molecular

absorption coefficient, c is the concentration of the sample, and d is the cell length.18 Typically,

the IR light source emits all IR frequencies of interest where the far-IR region is 10-400 cm-1, the

mid-IR region is 400-4000 cm-1, and the near-IR region is 4000-14000 cm-1.20 For vibrational

Page 43: Structural Tuning and Spectroscopic Characterizations of

30

transitions to be IR active, the molecule's dipole moment must change in magnitude as the

molecule vibrates, i.e. the molecule must have a permanent dipole moment to be IR-active.

Figure 2.6 Schematic illustration of the IR system19

For this study, all the in-situ mid-IR measurements were conducted in a highly customized IR

spectroscopy system. The schematic diagram of the system is shown in Figure 2.6. The major

component of the IR system is a Fourier transform infrared (FTIR) spectrometer with Model

Vertex 80v and a Glowbar IR light source purchased from Bruker Optics Inc. The system operated

under a vacuum condition of approximately below five mbar in order to avoid the interference of

the impurities from the air. For the optical path of the light source, the collimated IR light source

first gets directed into the relay box through a KBr window. Then the light source is reflected into

a 15x reflective objective lens bypassing an IRIS optic. A designed DAC holder stage is set after

the objective lens. The stage is adjustable for the best sample position. By adjusting the stage, the

beam would focus on the sample chamber of the DAC. Finally, the transmitted IR beam is directed

and collected to a wide-band mercury cadmium telluride (MCT) detector using an identical

Page 44: Structural Tuning and Spectroscopic Characterizations of

31

reflective objective lens as a condenser. The detector can record signals from 4000-400 cm-1

wavenumber. DAC background spectra were collected with the same size as the sample chamber.

All spectra were collected in the local lab by OPUS software with a resolution of 4 cm-1 and 512

scans.

The far-IR data in Chapter 3 were collected at Frontier Synchrotron Infrared Spectroscopy (FIS)

beamline at National Synchrotron Light Source II (NSLS-II), Brookhaven National Laboratory.

Synchrotron radiation (SR) is produced by accelerated electrons travelling at relativistic speeds

along a circular orbit by applied magnetic fields. The radiation is emitted along the tangent of the

orbit.21 SR is extremely intense and covers a broad range of electromagnetic waves, from infrared

to hard x-rays. The FIS beamline features its infrared light with a wide energy range from 2 meV

to 3 eV, which provides unique capabilities for far-IR region (10-400 cm-1) measurements with

low noise and spatial resolution. The sample measurement was conducted using Bruker Vertex

80V and Invenio FTIR spectrometers with an infrared light source generated by the FIS beamline.

All spectra were collected with a resolution of 0.2 cm-1.

Page 45: Structural Tuning and Spectroscopic Characterizations of

32

2.3 References

1. McMillan, P. F., Pressing on: The legacy of Percy W. Bridgman. Nature Materials 2005,

4 (10), 715-718.

2. Miao, M.; Sun, Y.; Zurek, E.; Lin, H., Chemistry Under High Pressure. Nature Reviews

Chemistry 2020, 4 (10), 508-527.

3. Piermarini, G. J., High Pressure X-Ray Crystallography With the Diamond Cell at

NIST/NBS. J Res Natl Inst Stand Technol 2001, 106 (6), 889-920.

4. Shen, G.; Mao, H. K., High-Pressure Studies with X-Rays Using Diamond Anvil Cells.

Reports on Progress in Physics 2016, 80 (1), 016101.

5. Samudrala, G. K.; Moore, S. L.; Vohra, Y. K., Fabrication of Diamond Based Sensors for

Use in Extreme Environments. Materials (Basel) 2015, 8 (5), 2054-2061.

6. Dong, Z.; Song, Y. In Novel Pressure-Induced Structural Transformations of Inorganic

Nanowires, 2011.

7. Ruby Fluorescence Measurement [Photograph].

https://duffy.princeton.edu/laboratory/ruby-fluorescence-measurement.

8. Forman, R. A.; Piermarini, G. J.; Barnett, J. D.; Block, S., Pressure Measurement Made

by the Utilization of Ruby Sharp-Line Luminescence. Science 1972, 176 (4032), 284.

9. Syassen, K., Ruby under pressure. High Pressure Research 2008, 28 (2), 75-126.

10. Goncharov, A. F.; Gregoryanz, E.; Mao, H.-k.; Liu, Z.; Hemley, R. J., Optical Evidence

for a Nonmolecular Phase of Nitrogen above 150 GPa. Physical Review Letters 2000, 85 (6), 1262-

1265.

11. Mao, H. K.; Xu, J.; Bell, P. M., Calibration of the Ruby Pressure Gauge to 800 Kbar under

Quasi-Hydrostatic Conditions. Journal of Geophysical Research: Solid Earth 1986, 91 (B5), 4673-

4676.

12. Klotz, S.; Chervin, J. C.; Munsch, P.; Le Marchand, G., Hydrostatic Limits of 11 Pressure

Transmitting Media. Journal of Physics D: Applied Physics 2009, 42 (7), 075413.

13. Reimann, K.; Syassen, K., Raman Scattering and Photoluminescence in Cu2O under

Hydrostatic Pressure. 1989, 39, 11113.

14. Palik, E. D., - Potassium Bromide (KBr). In Handbook of Optical Constants of Solids, Palik,

E. D., Ed. Academic Press: Burlington, 1997; pp 989-1004.

Page 46: Structural Tuning and Spectroscopic Characterizations of

33

15. Celeste, A.; Borondics, F.; Capitani, F., Hydrostaticity of Pressure-Transmitting Media for

High Pressure Infrared Spectroscopy. High Pressure Research 2019, 39 (4), 608-618.

16. Angel, R. J.; Bujak, M.; Zhao, J.; Gatta, G. D.; Jacobsen, S. D., Effective Hydrostatic

Limits of Pressure Media for High-Pressure Crystallographic Studies. Journal of Applied

Crystallography 2007, 40 (1), 26-32.

17. Goncharov, A. F., Raman Spectroscopy at High Pressures. International Journal of

Spectroscopy 2012, 2012, 617528.

18. Ferraro, J. R.; Nakamoto, K.; Brown, C. W., Introductory Raman Spectroscopy: Second

Edition. Introductory Raman Spectroscopy: Second Edition 2003, 1-434.

19. Guan, J., Effects of High Pressure on Photochemical Reactivity of Organic Molecular

Materials Probed by Vibrational Spectroscopy. PhD Thesis, The University of Western Ontario

2018.

20. Larkin, P. J., Chapter 2 - Basic Principles. In Infrared and Raman Spectroscopy (Second

Edition), Larkin, P. J., Ed. Elsevier: 2018; pp 7-28.

21. González, A., 1.5 X-Ray Crystallography: Data Collection Strategies and Resources. In

Comprehensive Biophysics, Egelman, E. H., Ed. Elsevier: Amsterdam, 2012; pp 64-91.

Page 47: Structural Tuning and Spectroscopic Characterizations of

34

Chapter 3

3 Investigation on the structural evolution of P4S3 and P2S5

(P4S10) under high-pressure

3.1 Introduction

Lithium-ion battery (LIB) has been widely used as a power source in various applications, from

portable consumer electronics to electric vehicles, due to its high specific energy density and

desired cycle life. However, since the theoretical capacity of 200-250 Wh/kg has been almost

reached and some safety issues caused by the traditional liquid electrolyte, the lithium-ion battery

starts becoming straining with its limitations.1-3 In regard to the development of new electrode

materials and battery techniques, many phosphorus sulfide materials have been introduced as

candidates for battery materials in recent years, such as P4S3 and P2S5 (P4S10).1, 2, 4-6 Phosphorus

sulfides is a group of inorganic compounds with the formula P4Sx where x ≤10.7 The reaction

between phosphorus and sulfur to form P-S compounds was first reported in

1740 by A.S. Marggraf.8 Through almost three centuries of studies, only four crystalline

compounds, P4S3, P4S5, P4S7 and P4S10, in the group were confirmed structurally by X-ray

diffraction.9

P4S3 is an inorganic compound known as phosphorus sesquisulfide. P4S3 has been recently

investigated as an anode material because of its high capacity and cycling stability for Li+/Na+/K+

storage.4, 5 The material appears yellow under ambient conditions. The detailed physical and

chemical properties are listed in Table 3.1. Confirmed by crystallographic measurements, the

material has a cage-like molecular structure.10, 11 Three phosphorus atoms form a triangle base.

The fourth phosphorus at the apical connects to each basal phosphorus atom by three bridging

sulfur atoms. The crystal structure of the material is orthorhombic with space group Pnma.12 In

Figure 3.1, the molecular structure and crystal structure are illustrated, and more detailed crystal

structure information is listed in Table 3.2.

Page 48: Structural Tuning and Spectroscopic Characterizations of

35

Figure 3.1 Molecular structure and crystal structure of P4S3. Orange and yellow spheres

represent phosphorus and sulfur atoms.

Table 3.1 P4S3 physical and chemical properties information

IUPAC Name Tetraphosphorus trisulfide CAS Number 1314-85-8

Molecular Formula P4S3 Appearance Yellow-green solid

Molecular Weight 220.093 Density 2.08 g/cm3

Melting Point 172.5°C Boiling point 408°C

P1

P2 P2

P3 S2

S1

S2

Sulfur

Phosphorus

Page 49: Structural Tuning and Spectroscopic Characterizations of

36

Table 3.2 P4S3 crystal structure information12

P4S3 ICSD Code: 417154

Crystal System Orthorhombic Space Group P n m a (No. 62)

Cell Parameter

a = 10.4738 b = 9.5861

c = 13.6715

α = β = γ = 90°

Cell Volume

Z = 8 1372.66 Å3

P1-P2 bond length 2.245 Å P1-S1-P3 bond angle 103.143°

P2-P2 bond length 2.243 Å P2-S2-P3 bond angle 102.883°

P1-S1 bond length 2.098 Å S1-P3-S2 bond angle 99.371°

P1-S2 bond length 2.096 Å S2-P3-S2 bond angle 98.760°

P3-S1 bond length 2.107 Å P1-P2-P2 bond angle 60.032°

P3-S2 bond length 2.113 Å P2-P1-P2 bond angle 59.936°

All-solid-state battery (ASSB) has been considered the most promising next-generation battery.

The core concept of this technique is using solid electrolytes to replace the traditional liquid

electrolyte.2 Compare to the conventional liquid electrolytes, most solid-state electrolytes are non-

flammable and have a more comprehensive operating temperature range.1, 2 Furthermore, the solid-

state electrolyte will not have the solid electrolyte interface (SEI) issue of electrode

dissolution/corrosion issue like the "shuttle effect" (see Chapter One). Sulfide based electrolytes

have attracted great attention in ASSB development due to their high ionic conductivity (over 10-

4 S/cm).2 For example, the Li2S-P2S5 system is one of the most promising solid-state electrolyte

candidates due to its high lithium-ion conductivities and wide electrochemical windows. The

electrolyte could be synthesized via several methods such as ball milling, solid-state route and wet

chemistry method.2 The system's performance can be adjusted by varying the composition of two

components and adding extra compounds. P2S5 is a vital precursor/component of the system. The

material is in yellow solid form under ambient conditions. The formula of phosphorus pentasulfide

could also be written as dimer P4S10, and detailed physical and chemical properties are listed in

Table 3.3. Regarding the molecular structure of P4S10, the phosphorus atoms arrange similarly with

phosphorus (P4) tetrahedral form but have a sulfur atom in between each two phosphorus atoms.

In addition, each atom of phosphorus has an additional sulfur atom connecting to it. The crystal

structure of the material is triclinic with space group P1̅.13, 14 The crystal and molecular structural

Page 50: Structural Tuning and Spectroscopic Characterizations of

37

representations are illustrated in Figure 3.2, and the crystal structure information has been listed

in Table 3.4.

Figure 3.2 Molecular structure and crystal structure of P4S10. Orange and yellow spheres

represent phosphorus and sulfur atoms.

Table 3.3 P2S5 (P4S10) physical and chemical properties information

Table 3.4 P2S5 (P4S10) crystal structure information14

P2S5 ICSD Code: 23843

Crystal System Triclinic Space Group P -1 (No. 2)

Cell Parameter a = 9.072 b = 9.199 c = 9.236

α = 92.58° β = 100.9° γ = 110.18°

Cell Volume

Z = 2 1085.64 Å3

IUPAC Name Diphosphorus pentasulfide CAS Number 1314-80-3

Molecular Formula P2S5 /P4S10 Appearance Yellow powder

Molecular Weight 222.27 Density 12.09 g/cm3

Melting Point 288°C Boiling point 514°C

Sulfur

Phosphorus

Page 51: Structural Tuning and Spectroscopic Characterizations of

38

Pressure is a powerful tool for tuning the chemical bond, crystal structure and properties of

materials. Significantly, the pressure-induced transformation could effectively synthesize novel

systems with properties that differ from those under ambient conditions. Many studies have shown

that external pressure can be an additional driving force for structure evolutions of the material.

Extensive high-pressure studies on battery materials reported that the pressure-induced structural

transformation would affect the material's performance in a battery application.15-22 Although the

two phosphorus sulfides have been well studied for their structures and reactivities under ambient

conditions, the high-pressure study of these materials remains unknown. A thorough investigation

of the structure evolution and phase transformation induced by external pressure could provide

additional information on these materials related to their functions and performance. Therefore, it

is worthy of investigating the structural changes of these two materials and determining the

possibility of new phase formation under high pressure.

Herein, the first high-pressure study of P4S3 and P4S10 using in-situ Raman spectroscopy and IR

spectroscopy are reported from this study. The high-pressure studies of P4S3 and P4S10 were

determined up to 25 and 17 GPa, respectively. The Raman spectra and IR spectroscopy revealed

the exciting effect of high pressure on these two materials. Pressure-induced phase transition was

observed from both materials during the compression process. This new structural information

allows the precise determination of structural stability and compressibility of these materials. The

new findings also could be contributing to the development of new materials/structures for battery

studies.

3.2 Experimental

3.2.1 Material

P2S5 powder is commercially available from Sigma-Aldrich and was obtained from Western

Engineering (purchased from Sigma-Aldrich). P4S3 was purchased from Chem Service Inc. Both

materials were used without further purification. Due to the sensitivity of moisture, both materials

were kept in a nitrogen glove box.

Page 52: Structural Tuning and Spectroscopic Characterizations of

39

3.2.2 High-Pressure Experiment procedures

A gasket was first placed into DAC and pre-indented up to 10 GPa for preparing the sample

chamber. After the pre-indentation, the gasket was removed from the DAC. Then, a needle with a

diameter of 130 microns is used to drill a small hole in the centre of the pre-indentation mark by

the EDM drilling machine. The small hole was used as the sample chamber, and the gasket was

inserted back to the diamond anvil cell. Due to the air sensitivity of samples, DAC was transferred

into a nitrogen glovebox. A ruby was loaded into the cell before moving the DAC into the glovebox.

Next, the sample was loaded into the DAC sample chamber inside the glovebox, and then the cell

was closed and compressed to "finger tight" before taking it out from the glovebox. Ruby

fluorescence was checked for pressure first, and then Raman measurement was taken for near

ambient pressure. For the compression process, the pressure was increased by tightening each bolt

at a certain angle equally. Then the pressure was rechecked with ruby fluorescence following by

Raman/IR measurement. The process was repeated several times until adequate pressure was

reached. Moreover, the pressure was decreased for the decompression process by equally losing

each bolt at a certain angle, followed by pressure and Raman/IR measurement. Similarly, the

process was repeated until the pressure was close to ambient pressure. Before moving the DAC

into the glovebox for IR measurement, a proper amount of KBr was loaded into the cell as a spacer

to adjust the amount of sample loaded.

3.2.3 Characterizations

In-situ Raman spectroscopy (refer to section 2.2.2 in Chapter 2) was applied to monitor pressure-

induced phase transitions of samples. In the Raman system, the scattered light was dispersed using

an imaging spectrograph equipped with a 1200 lines/mm grating achieving a resolution of about

0.2 cm–1. The 532 nm laser was used as the exciting source. The laser beam was focused onto a

spot with a diameter of approximately 5μm by using an objective microscope with a magnification

of 50×.

A customized IR spectroscopy system carried out the in-situ IR absorption experiments with

details described in Chapter 2.

Page 53: Structural Tuning and Spectroscopic Characterizations of

40

The far-IR region data was collected from Frontier Synchrotron Infrared Spectroscopy (FIS)

beamline at National Synchrotron Light Source II (NSLS-II), Brookhaven National Laboratory by

beamline scientist Dr. Zhenxian Liu.

3.3 Result and Discussion

3.3.1 Raman results of P4S3 upon compression and

decompression

Raman spectra of P4S3 were first collected at near ambient conditions (in Figure 3.3). Nine active

Raman modes were observed in the P4S3 sample at 0.62 GPa. The peak assignments are shown in

Table 3.5. The peak assignments of observed Raman modes at 0.62 GPa were done by comparing

the observed Raman modes with references from the previous studies.23, 24 Several peaks

correspond to two modes.

Figure 3.3 Raman Spectrum of P4S3 with peak assignment collected at 0.62 GPa

Page 54: Structural Tuning and Spectroscopic Characterizations of

41

Table 3.5 Peak Assignment of Raman Modes of P4S3

Symmetry Mode Raman Shift (cm-1)

Assignmentb Reference23, 24 Experimentala

A1 ν1 489 489 P-P stretch

ν2 443 441 P-S-P stretch (s)

ν3 414 421 P-S-P stretch (a)

ν4 347 342 P-S-P bend

A2 ν5 187 inactive P-S-P wag

E ν6 489 489 P-S-P stretch (s)

ν7 422 421 P-S-P stretch (a)

ν8 343 342 P-P stretch

ν9 288 288 P-S-P bend

ν10 220 220 P-S-P wag

a Experimental data was collected at 0.62 GPa b (a) and (s) mean asymmetric and symmetric, respectively.

The evolution of the Raman spectra of P4S3 upon compression and decompression is shown in

Figure 3.4. As the pressure increases from the compression plot, most Raman peaks are blue-

shifted, caused by a decrease in the interatomic distance. Upon compression to 3.72 GPa, the

vibrational mode associated with P-S-P wag, ν10, has a discontinuity trend and become vanished

when further compressed. Meanwhile, the peak of modes ν1(ν6), at 489 cm-1, split into two peaks

at 2.58 GPa (marked in the square at 2.58 GPa in Figure 3.4). These observations suggest a phase

transition could occur between 2.5 – 4 GPa. Upon further compression, when the pressure reaches

around 7 GPa, the P-S-P bend mode, ν9, diminishes. At the same pressure, the peak at 343 cm-1

split into two peaks (marked in the square at 7.97 GPa in Figure 3.4). Such observations could be

a sign for another phase transition at 7 GPa. The compression process was stopped at 24.87 GPa

since no further changes were observed. All Raman bands mostly become vanished at about 12

GPa, except a small peak remained even at the highest pressure. These observations indicate a

possible formation of a metastable structure at 12 GPa.

After being compressed to 24.87 GPa, the decompression process was conducted to determine the

reversibility of P4S3. During the decompression, peaks start recovering around 2 GPa, and the

intensity of each peak increases through the process. The pressure was released to 0.03 GPa. All

Page 55: Structural Tuning and Spectroscopic Characterizations of

42

peaks recovered through the decompression process just like they were at ambient pressure,

indicating that all the structural changes of P4S3 induced by the high-pressure condition are

reversible.

Figure 3.4 Raman spectra of P4S3 upon compression and decompression. The formation of

new Peaks is marked on the spectra by squares.

3.3.2 Pressure dependence of Raman Modes of P4S3

To have a quantitative understanding of the Raman shift changes, the vibrational frequencies as a

function of pressure was also plotted, as shown in Figure 3.5. The phase transition boundaries could

be determined by the discontinuity trend of each Raman mode and the appearance or disappearance of

modes. Additionally, the pressure coefficients were also analyzed by a linear fitting and listed in Table

3.6. The phase transitions could be further verified by the significant changes of the pressure

coefficients in different pressure regions.

Page 56: Structural Tuning and Spectroscopic Characterizations of

43

Figure 3.5 Pressure dependence of Raman modes of P4S3. The vertical dashed lines indicate

the proposed phase boundaries, and different symbols represent Raman modes with

different origins.

Page 57: Structural Tuning and Spectroscopic Characterizations of

44

Table 3.6 Pressure dependence (dν/dP, cm-1/GPa) of vibrational modes of P4S3 upon

compression from Raman spectra

Mode Optical mode* Raman Shift (cm-1) Phase

I

Phase

II

Phase

III

ν1 (ν6) P-P stretch + P-S-P stretch (s) 489 3.9 3.2 4.2

ν2 P-S-P stretch (s) 441 2.3 1.3 1.2

ν3 (ν7) P-S-P stretch (a) 421 2.3 1.0 1.4

ν4 (ν8) P-P stretch + P-S-P bend 342 2.4 2.5 1.6

ν9 P-S-P bend 288 0.3 -1.0 -

ν10 P-S-P wag 220 4.6 - -

*(a) and (s) mean asymmetric and symmetric, respectively.

According to the positive value of pressure coefficients, most Raman modes exhibit pressure-induced

blue shift due to the decreasing interatomic distances and the enhancement of the interatomic

interaction. For the first phase transition between 2.5-4 GPa, besides the disappearance of ν10 mode and

the appearance of the new peak around the peak at 480 cm-1, ν9 mode has a discontinuity trend

compared with its phase II pressure dependence value. From the peak assignment, ν10 is associated

with the P-S-P wag mode. Considering P-S-P bend and stretch modes were remaining, the first

structural change is likely caused by the reduced volume from pressure and distortion of the material's

molecular structure. The second phase is observed at about 7 GPa, which is suggested by the formation

of a new peak around the peak of ν4(ν8) modes. The pressure dependence shows no significant

discontinuity for the other modes between phase II and phase III regions, meaning no broken bond.

The structural change probably is still induced by reduced volume and distortion of the cage-like

molecular structure. Above 13 GPa, all Raman modes have vanished except the peak formed around

480 cm-1 after the first transition. The position of the peak suggests that it could be a mode related to

the P-S-P bond or P-P bond, and the broad peak shape and asymmetric profile indicated these bonds

are not equivalent. Moreover, the peak did not shift as the pressure changed further to 25 GPa.

Therefore, upon compression to 13 GPa, the cage-like structure may collapse and a metastable system

with P-S and P-P bonds probably formed.

Page 58: Structural Tuning and Spectroscopic Characterizations of

45

Regarding the structural stability and compressibility of the material, pressure dependence values

in each phase region were also examined. The pressure dependence value of each mode was

compared among all the phases. According to the pressure coefficients at the lower Raman shift

region (200-300 cm-1) in the phase I region, the structure probably significantly changed from the

compression. The compression is more effective on the P-S-P bonds than the P-P bonds in phase

I. In phase II, ν4 (ν8) mode has the largest pressure coefficients, and the pressure dependence value

of the other modes becomes smaller than in phase I. It suggests the compression is more effective

on the P-P than on the P-S-P bonds in the phase two region. In the phase III region, all pressure

coefficients associated with P-S-P modes (ν2 and ν3(ν7)) only slightly changed. The pressure

dependence value associated with P-P modes have more significant differences compared to phase

II. It indicates the P-P bonds are more effectively compressed in phase III.

3.3.3 Computed unit cell parameters and equation of state of P4S3

The unit-cell volume was calculated using the first-principle method based on density functional

theory (DFT). In the calculation, ambient pressure crystal structure featuring orthorhombic lattice

was used as starting point. The unit cell parameters at different pressure upon compression are

listed in Table 3.7 by assuming no change in the crystal structure during the compression process.

In Figure 3.6, the pressure dependence of lattice parameters and unit cell volume from the

calculation was plotted, and the pressure-volume data were fit to a Birch-Murnaghan equation of

state (EOS):25

𝑃 = 3

2𝐾0 [(

𝑉0

𝑉)

7 3⁄

− (𝑉0

𝑉)

5 3⁄

] ∙ {1 −3

4(4 − 𝐾′0) [(

𝑉0

𝑉)

2 3⁄

− 1]} (3.1)

Where P and V are pressure and volumes, K0 and K'0 are the bulk modulus and its first derivative.

V0 is the volume at the ambient pressure condition, which is 1372.66 Å3.

Page 59: Structural Tuning and Spectroscopic Characterizations of

46

Table 3.7 Unite cell parameters of P4S3 as a function of pressure

Pressure a b c cell volume

GPa Å Å Å Å3

0 10.8179 9.9802 14.2924 1543.1

1 10.4397 9.5731 13.6958 1368.8

2 10.2265 9.3673 13.3616 1279.9

4 9.9684 9.0938 12.9673 1175.5

6 9.7862 8.9099 12.7155 1108.7

8 9.6498 8.7597 12.5329 1059.4

10 9.5273 8.6458 12.3829 1019.9

12 9.4372 8.5303 12.2576 986.8

14 9.3511 8.4457 12.1473 959.4

16 9.2662 8.3616 12.0587 934.3

18 9.2069 8.2794 11.9738 912.7

20 9.1462 8.2051 11.9004 893.1

25 9.0183 8.0233 11.7451 849.8

30 8.9312 7.8482 11.6044 813.4

From Figure 3.6, it can be found that all three parameters have the almost same trend as the

pressure increases. All three parameters have a steep trend when the pressure is below about 5 GPa.

This trend is consistent with the previous observation of the largest pressure coefficient at the

lower Raman shift region (150-300 cm-1) in phase I. As the pressure increase further, the high-

pressure effect on the three parameters become less. Same as the cell parameter, the unit cell

volume has the same trend as the pressure increasing. The cell volume decreased dramatically

from 1600 to 1175 Å3 upon compression to 4 GPa and then went to about 800 Å3 at 30 GPa.

Moreover, the trend was fitted with the EOS curve. The fitted bulk modulus (K0) is 9.211 GPa

with a pressure derivative of the modulus (K0') equal to 5.6683, indicating the compressibility of

the material is relatively large (i.e., very compressible).

Page 60: Structural Tuning and Spectroscopic Characterizations of

47

Figure 3.6 Pressure dependence of lattice parameters and unit cell volume from the

calculations (a). The cell volume data was compared with a third-order Birch-Murnaghan

equation of state (EOS) (b).

The calculated cell parameters and unit volume were also normalized with respect to the ambient

condition parameters of P4S3, and the plots are illustrated in Figure 3.7. The figure shows that the

pressure-induced change has the same effect on the three cell parameters when the pressure is

below 3 GPa. Upon compression, the pressure has slightly less effective in the a direction, and

resulting in a steeper trend of the other two parameters. As the pressure further increases toward

13 GPa, the pressure has more effect on the c direction first. After the pressure reaches around 7

GPa, the pressure affects the b and c directions equally. At the 13 GPa, the compression pressure

becomes more effective in the b direction. Therefore, it can be concluded that the high-pressure

effect on the material is isotropic below 3 GPa. As the pressure increases, the compression is more

effective along with the b and c directions. The compression is more effective in the c direction

than the b direction between 3-7 GPa. Then the effect of high pressure become the same on along

b and c a direction. Above 13 GPa, the b axis becomes the most compressible lattice direction

predominantly. Corresponding, the unit cell volume change at the beginning of the compression is

caused by the decrease of all three parameters. At the higher pressure condition, the volume change

is more dominant by b. The phase boundaries concluded from Raman spectra also have the same

trend with the cell parameters, indicating the phase transition could be due to the volume changes.

Page 61: Structural Tuning and Spectroscopic Characterizations of

48

Figure 3.7 Pressure dependence of normalized cell parameters and cell volume of P4S3. The

dashed line is phase boundaries concluded from Raman spectra.

3.3.4 Raman and IR Spectra of P2S5 at ambient pressure

Raman and IR spectra of P2S5 were collected at near ambient conditions (Figure 3.8). Eight active

Raman modes were observed in the P2S5 sample at 0.38 GPa, and four active IR modes were

identified at 0.82 GPa. The peak assignments are shown in Table 3.8. The assignments were done

by comparing the observed Raman and IR modes with references from the previous studies.9

Page 62: Structural Tuning and Spectroscopic Characterizations of

49

Figure 3.8 Raman and IR spectra of P2S5 collected at ambient condition and 0.82 GPa,

respectively.

Table 3.8 Peak assignment of Raman and IR modes of P2S5

Symmetry Mode Reference9 Experimental

Assignmentc

IRa Ramanb

A1 ν1 717 Inactive 713 P=S stretch

ν2 400 - 393 P-S-P bend

ν3 265 264 268 P-S-P stretch (s)

E ν4 546 Inactive - P-S-P stretch (s)

ν5 164 - 164 P-S-P bend

ν6 114 - 125 P=S wag

T1 ν7 413 inactive inactive P-S-P stretch (a)

ν8 206 - - P=S wag

ν9 111 - - P-S-P wag

T2 ν10 692 690 687 P=S stretch

ν11 526 533 - P-S-P stretch (s)

ν12 372 379 382 P-S-P stretch (a)

ν13 272 - - P-S-P wag

ν14 190 - 190 P-S-P bend

ν15 125 - - P=S wag a IR data were collected at 0.82 GPa. bRaman data were collected at ambient conditions. c(a) and (s) mean asymmetric and symmetric, respectively.

Page 63: Structural Tuning and Spectroscopic Characterizations of

50

3.3.5 In-situ high-pressure Raman spectroscopy of P2S5

The evolution of the Raman spectra of P2S5 upon compression and decompression is shown in

Figure 3.9. From the compression plot, it can be observed that most Raman peaks are blue-shifted

as the pressure increases. Upon compression to around 3 GPa, the peaks of modes ν5 and ν10 at 160

and 690 cm-1 become vanished, and the P-S-P stretch (s) mode ν3 and P-S-P bend mode ν14 start

splitting into two peaks at the same pressure. Upon further compression, when the pressure reaches

5 GPa, the P=S wag mode ν6 and the P-S-P stretch (s) mode ν3 disappear. Then at around 7.5 GPa,

the peak split from mode ν14 becomes vanished. The disappearance and splitting of peaks indicate

a transformation involving bond broken/formation probably starts at around 3 GPa and finishes

around 7.5 GPa. During the transformation, the material probably has two or more different

structure co-existence due to the bond breaking and forming process. All Raman bands mostly

vanished at 13 GPa. These observations indicate a possible pressure-induced amorphous state. The

compression process was stopped after the pressure reached 17 GPa.

Figure 3.9 Raman spectra of P2S5 upon compression and decompression

Page 64: Structural Tuning and Spectroscopic Characterizations of

51

The decompression process was also conducted to determine the reversibility of the pressure effect

on P2S5. During the decompression, peaks start recovering around 2 GPa, and the intensity of each

peak increases through the process. The decompression plot shows that several peaks of mode (ν3,

ν7 and ν9) do not recover to their original peak profile, indicating the high-pressure effect on P2S5

is irreversible when decompressing started above 15 GPa. In addition, the decompressed sample

was rechecked after being kept in the cell for 24 hours at ambient pressure. However, the material

seems only recovered partially. For example, the peak of ν7 mode only recovers partially, and the

peak of mode ν10 get broadened, both with different shapes than the initial spectrum at ambient

pressure before compression.

For the reversibility threshold of the material, decompression from different pressure conditions

were also conducted. From the decompression plot, it can be found that the high-pressure effect is

reversible when decompressing from 9, 11 and 12 GPa. The results show that the compression

effect was reversible under 12 GPa. Therefore, when the material is recovered after reaching the

amorphous state, the structure would be irreversible. If the compression pressure is below the

amorphous state, the high-pressure effect on P2S5 would be reversible. According to the

observations, the reversibility threshold should be at around 12 GPa.

3.3.6 In-situ high-pressure IR spectroscopy of P2S5

In-situ IR spectra of P2S5 between 100 and 800 cm−1 upon compression to 12.6 GPa are shown in

Figure 3.10. All modes are blue-shifted from the compression because of the reduced interatomic

distances and the enhancement of the interatomic interaction. Upon compression to about 8 GPa,

the two vibrational modes ν3 and ν13 become vanished, indicating a possible transformation. As

the pressure increases, the remaining peaks become broadened and weakened up to 12.6 GPa. No

other noticeable change was observed from the compression plot. This result also demonstrates

that the amorphous state of the material should be around 12 GPa, consistent with the Raman

results.

Page 65: Structural Tuning and Spectroscopic Characterizations of

52

Figure 3.10 IR spectra of P2S5 upon compression

3.3.7 Pressure dependence of IR and Raman modes of P2S5

To have a quantitative understanding of the Raman and IR modes changes, the vibrational

frequencies as a function of pressure were also plotted, as shown in Figure 3.11. The possible

transformation boundaries could be determined by the discontinuity trend and the appearance or

disappearance of each mode. Additionally, the pressure coefficients were also analyzed by a linear

fitting and listed in Table 3.9. The structural transformation boundaries could be further verified

by the significant changes of the pressure coefficients in different pressure regions.

Page 66: Structural Tuning and Spectroscopic Characterizations of

53

Figure 3.11 Pressure dependence of Raman (a) and IR (b) modes of P2S5. The vertical dashed

lines indicate the proposed phase boundaries. Different symbols represent Raman modes

with different origins.

Overall, regarding the Raman-active modes, the positive pressure coefficients indicate that most

Raman modes exhibit pressure-induced blue shifts due to the reduction of the volume and the

enhancement of the interatomic interaction. For the first transformation boundaries at around 3

GPa, in addition to the appearance/disappearance of peaks, several peaks (ν1, ν3, ν14 and ν6) have

a significant change in their pressure coefficients comparing with phase II. According to the peak

assignment, the disappearance of mode ν10 in phase I may indicate a double bond broken between

P and S. So, the first structural change is probably caused by enhanced intermolecular interactions

and bond broken as a result of the reduced volume by compression. In the second phase region

between 3 and 7.5 GPa, it can be observed that new peaks split from ν3 and ν14 first, then mode ν3

Page 67: Structural Tuning and Spectroscopic Characterizations of

54

becomes vanished upon further compression. The material probably has P=S broken first in this

region, followed by forming a new bonding pattern and completely transforming to a novel

structure at 7.5 GPa. Two or more structures may co-existence in this region during the

transformation to complete. A metastable structure probably forms in the second region since the

pressure coefficients in the phase III region are smaller than the second region. The formed

structure exhibits strong stability until the material reaches to amorphous state at above 12 GPa.

Regarding the pressure dependence and pressure coefficients of IR spectra, two possible phase

transition is observed. The first phase transition is suggested at about 3 GPa by the discontinuity

trend of ν3, ν13 and ν11 modes by comparing the pressure coefficients between phase I and phase II

regions. The observation is consistent with Raman data, where the P=S stretch mode (ν10) at 690

cm-1 was split into two peaks at about 3 GPa. Upon compression to 7.4 GPa, all modes have a

discontinuity trend. All modes have a significant change of pressure coefficients between phase II

and III regions. The observations from IR are consistent with the Raman result, and phase III also

exhibits lower compressibility since all pressure dependence values become smaller than phase II.

The formation of the structure would cause the material to become hard to compress, resulting in

a decreasing pressure coefficient from phase II to phase III.

Table 3.9 Pressure dependence (dv/dP, cm-1/GPa) of vibrational modes of P2S5 upon

compression from Raman and IR spectra.

Modes Raman-active mode* Raman shift (cm-1) Phase I Phase II Phase III

ν1 P=S stretch 713 -0.4 -0.6 -0.2

ν10 P=S stretch 687 1.1 - -

ν2 P-S-P bend 393 1.0 0.8 0.7

ν13 P-S-P stretch (a) 382 1.0 - -

ν3 P-S-P stretch (s) 268 5.0 3.1 -

ν14 P-S-P bend 191 6.1 3.6 3.1

ν5 P-S-P bend 160 1.4 - -

ν6 P=S wag 125 7.5 3.5 -

IR-active mode* Wavenumber (cm-1)

ν10 P=S stretch 690 0.4 0 0

ν11 P-S-P stretch (s) 533 2.0 3.2 0

ν13 P-S-P stretch (a) 377 1.6 3.3 -

ν3 P-S-P stretch (s) 264 1.6 0.5 -

*(a) and (s) represent asymmetric and symmetric, respectively.

Page 68: Structural Tuning and Spectroscopic Characterizations of

55

In regard to structural stability and compressibility of the material, pressure dependence values in

each phase region were also examined. The pressure dependence value of each mode was

compared among all the phases. According to the pressure coefficients in the phase I region, the

compression is more effective on the P-S-P symmetric stretching mode, P-S-P bending mode and

P=S wagging mode in phase I. In phase II, the pressure coefficients of these modes are still larger

than the other modes. However, all pressure coefficients become smaller than the phase I region,

which indicate the material becoming less compressible than phase I. Upon further compression

to phase III, all pressure coefficients become even smaller larger. Only the P-S-P bending mode

had slight changes, and it indicates the P-S-P bonds were effectively compressed in phase III. And

the structure formed after phase II is relatively stable and hard to compress.

3.3.8 Discussion

Comparing the two phosphorus sulfides, each material has a similar structure to the P4 structure.

P4S3 has a P3 triangle structure at the base and has a sulfide insert between base P and apical P. For

P2S5 (P4S10), it has a P4 skeleton structure with sulfur insert in between phosphorus. Both structures

exhibit two possible phase transitions upon compression. The potential phase transitions between

the two materials also had similar pressure boundaries, and it may be due to the similar bond

interaction between P and S. Moreover, there is a possibility that P4S3 did not undergo any phase

transition upon compression since no bond broken was observed. The strong stability may cause

by the cage-like molecular structure and the triangle structure of the P3 base. In contrast with P4S3,

P2S5 (P4S10) features the double bonds between P and S. These double bonds were observed broken

first and forming a metastable structure upon compression process.

The preliminary DFT based first-principle calculations show that at a certain pressure, P2S5 (P4S10),

it can dimerize into a P4S10-P4S10 structure; at higher pressures, it can form a polymeric phase with

a repeating unit of P4S10 as (P4S10)n. These chemical transformations are realized by breaking the

terminal P=S double bonds. The conclusion from the calculation is consistent with the observation

from this study, and it could explain the irreversible changes of the structure from Raman spectra.

The dimer or polymer structure formation would result in the change of Raman spectra for certain

modes. However, in-situ high-pressure X-ray diffraction at the synchrotron radiation facility is still

needed to determine the exact phase boundaries and structures.

Page 69: Structural Tuning and Spectroscopic Characterizations of

56

3.4 Conclusion

In summary, P4S3 were investigated by using Raman spectroscopy under high pressure. P4S3 may

undergo two phase transitions through the compression process up to 24.87 GPa. The first possible

phase transition is between 2.5 – 4 GPa caused by the reduction of the unit cell volume and the

molecular structure distortion. Upon further compression, the second possible phase transition is

at about 7 GPa. The volume change could induce the second phase transition by the further

distortion of the molecular structure since no bond broken observed. The material possibly formed

a metastable structure when the pressure reached about 13 GPa with no further structural changes

up to 24.87 GPa. In addition, pressure-dependent cell parameters and volume were calculated

using the first-principle method. The calculated cell volume has been fitted to the Birch-

Murnaghan equation of state, from which the bulk modulus K0 and its first derivative K0' are

determined to be 9.2110 GPa and 5.6683. The pressure effect is isotropic under 4 GPa affecting

all three parameters. When the pressure passes 5 GPa, the pressure has more effect on along the b

and c directions.

P2S5 (P4S10) were investigated by using Raman and IR spectroscopies under high pressure. P2S5

could possibly have two transformations upon compression. The first transformation is occurred

around 3 GPa, possibly caused by the double bond broken between phosphorus and sulfur. The

second transformation was suggested at about 7.5 GPa and could be due to the formation of a novel

metastable structure. When the pressure reached above 12 GPa, the material became amorphous.

The decompression process suggested that the high-pressure effect is irreversible once the pressure

reaches the amorphous state. High-pressure structure transformations of P4S10 may involve novel

bonding structures, such as a dimer or polymetric structure. However, further high-pressure X-ray

diffraction studies are still required to clarify the two materials' possible phase transitions and high-

pressure phases/structures.

Page 70: Structural Tuning and Spectroscopic Characterizations of

57

3.5 Reference

1. Mizuno, F.; Hama, S.; Hayashi, A.; Tadanaga, K.; Minami, T.; Tatsumisago, M., All

Solid-state Lithium Secondary Batteries Using High Lithium Ion Conducting Li2S–P2S5 Glass-

Ceramics. Chemistry Letters 2002, 31 (12), 1244-1245.

2. Kudu, Ö. U.; Famprikis, T.; Fleutot, B.; Braida, M.-D.; Le Mercier, T.; Islam, M. S.;

Masquelier, C., A Review of Structural Properties and Synthesis Methods of Solid Electrolyte

Materials in the Li2S − P2S5 Binary System. Journal of Power Sources 2018, 407, 31-43.

3. Kato, A.; Yamamoto, M.; Sakuda, A.; Hayashi, A.; Tatsumisago, M., Mechanical

Properties of Li2S–P2S5 Glasses with Lithium Halides and Application in All-Solid-State Batteries.

ACS Applied Energy Materials 2018, 1 (3), 1002-1007.

4. Kulova, T. L.; Skundin, A. M.; Gryzlov, D. Y.; Kudryashova, Y. O.; Chekannikov, A.

A., Phosphorus Sulfide as a Functional Material for Sodium-Ion Batteries. Mendeleev

Communications 2019, 29 (5), 556-557.

5. Wei, D.; Yin, J.; Ju, Z.; Zeng, S.; Li, H.; Zhao, W.; Wei, Y.; Li, H., Cage-Like P4S3

Molecule as Promising Anode with High Capacity and Cycling Stability for Li+/Na+/K+ Storage.

Journal of Energy Chemistry 2020, 50, 187-194.

6. Agostini, M.; Aihara, Y.; Yamada, T.; Scrosati, B.; Hassoun, J., A Lithium–Sulfur

Battery Using a Solid, Glass-Type P2S5–Li2S Electrolyte. Solid State Ionics 2013, 244, 48-51.

7. Cowley, A. H., The Structures and Reactions of the Phosphorus Sulfides. Journal of

Chemical Education 1964, 41 (10), 530.

8. Pernert, J. C.; Brown, J. H., Some Reactions and Properties of the Phosphorus Sulfides.

Chemical & Engineering News Archive 1949, 27 (30), 2143-2145.

9. Jensen, J. O.; Zeroka, D., Theoretical Studies of the Infrared and Raman Spectra of P4S10.

Journal of Molecular Structure: THEOCHEM 1999, 487 (3), 267-274.

10. Van Houten, S.; Vos, A.; Wiegers, G. A., The Crystal Structure of P4S3. Recueil des

Travaux Chimiques des Pays-Bas 1955, 74 (9), 1167-1170.

11. Leung, Y. C.; Waser, J.; Houten, S. v.; Vos, A.; Wiegers, G. A.; Wiebenga, E. H., The

Crystal Structure of P4S3. Acta Crystallographica 1957, 10 (9), 574-582.

12. Raabe, I.; Antonijevic, S.; Krossing, I., Dynamics and Counterion-Dependence of the

Structures of Weakly Bound Ag+–P4S3 Complexes. Chemistry – A European Journal 2007, 13

(26), 7510-7522.

Page 71: Structural Tuning and Spectroscopic Characterizations of

58

13. Vos, A.; Wiebenga, E. H., The Crystal Structures of P4S10 and P4S7. Acta

Crystallographica 1955, 8 (4), 217-223.

14. Vos, A.; Olthof, R.; Van Bolhuis, F.; Botterweg, R., Refinement of the Crystal Structures

of Some Phosphorus Sulphides. Acta Crystallographica 1965, 19 (5), 864-867.

15. Gallardo-Amores, J. M.; Biskup, N.; Amador, U.; Persson, K.; Ceder, G.; Morán, E.;

Arroyo y de Dompablo, M. E., Computational and Experimental Investigation of the

Transformation of V2O5 Under Pressure. Chemistry of Materials 2007, 19 (22), 5262-5271.

16. García-Moreno, O.; Alvarez-Vega, M.; García-Jaca, J.; Gallardo-Amores, J. M.; Sanjuán,

M. L.; Amador, U., Influence of the Structure on the Electrochemical Performance of Lithium

Transition Metal Phosphates as Cathodic Materials in Rechargeable Lithium Batteries:  A New

High-Pressure Form of LiMPO4 (M = Fe and Ni). Chemistry of Materials 2001, 13 (5), 1570-1576.

17. Amador, U.; Gallardo-Amores, J. M.; Heymann, G.; Huppertz, H.; Morán, E.; Arroyo y

de Dompablo, M. E., High pressure polymorphs of LiCoPO4 and LiCoAsO4. Solid State Sciences

2009, 11 (2), 343-348.

18. Yamaura, K.; Huang, Q.; Zhang, L.; Takada, K.; Baba, Y.; Nagai, T.; Matsui, Y.;

Kosuda, K.; Takayama-Muromachi, E., Spinel-to-CaFe2O4-Type Structural Transformation in

LiMn2O4 under High Pressure. Journal of the American Chemical Society 2006, 128 (29), 9448-

9456.

19. Arroyo-deDompablo, M. E.; Dominko, R.; Gallardo-Amores, J. M.; Dupont, L.; Mali,

G.; Ehrenberg, H.; Jamnik, J.; Morán, E., On the Energetic Stability and Electrochemistry of

Li2MnSiO4 Polymorphs. Chemistry of Materials 2008, 20 (17), 5574-5584.

20. Yoncheva, M.; Stoyanova, R.; Zhecheva, E.; Alcántara, R.; Ortiz, G.; Tirado, J. L., Effect

of the High Pressure on the Structure and Intercalation Properties of Lithium–Nickel–Manganese

Oxides. Journal of Solid State Chemistry 2007, 180 (6), 1816-1825.

21. Huang, Y.; He, Y.; Sheng, H.; Lu, X.; Dong, H.; Samanta, S.; Dong, H.; Li, X.; Kim,

D. Y.; Mao, H.-k.; Liu, Y.; Li, H.; Li, H.; Wang, L., Li-Ion Battery Material under High Pressure:

Amorphization and Enhanced Conductivity of Li4Ti5O12. National Science Review 2018, 6 (2),

239-246.

22. Wang, X.; Zhang, P.; Tang, X.; Guan, J.; Lin, X.; Wang, Y.; Dong, X.; Yue, B.; Yan,

J.; Li, K.; Zheng, H.; Mao, H.-k., Structure and Electrical Performance of Na2C6O6 under High

Pressure. The Journal of Physical Chemistry C 2019, 123 (28), 17163-17169.

Page 72: Structural Tuning and Spectroscopic Characterizations of

59

23. Jensen, J. O.; Zeroka, D.; Banerjee, A., Theoretical Studies of the Infrared and Raman

Spectra of P4S3 and P4S7. Journal of Molecular Structure: THEOCHEM 2000, 505 (1), 31-43.

24. Gardner, M., Infrared and Raman Spectra of Some Phosphorus Sulphides. Journal of the

Chemical Society, Dalton Transactions 1973, (6), 691-696.

25. Birch, F., Finite Strain Isotherm and Velocities for Single-Crystal and Polycrystalline NaCl

at High Pressures and 300°K. Journal of Geophysical Research: Solid Earth 1978, 83 (B3), 1257-

1268.

Page 73: Structural Tuning and Spectroscopic Characterizations of

60

Chapter 4 4 Investigation of structural evolutions of Na2S4 under high-

pressure

4.1 Introduction

Sodium sulfur battery (Na-S) is a high-temperature battery technique considered a candidate for

next-generation batteries. It received tremendous attention in recent years because of its low

material cost, energy density (230W/kg), high efficiency (90%) and long cycle life(~25 years).1 In

a Na-S battery, the battery system is based on molten salt technology, i.e. the operation temperature

needs to be over 300 °C to keep both electrodes in molten states.2 The technology of Na–S batteries

have already been used in some countries such as Japan.3 However, there are still issues limiting

its application, such as the safety issue from high operation temperature and corrosion effect from

the formation of polysulfides. During the discharge process, the formation of sodium polysulfides

(mainly Na2S2 and Na2S4) would cause a corrosion effect, leading to cell performance

degradation.4, 5 This corrosion effect is the major issue preventing the technique from further

applications.

The sodium-sulfur system has been studied since 1914; Friedrich reported the first sodium-sulfur

phase diagram.6 Later in the same year, Rule and Thomas updated the diagram and Pearson, and

Robinson confirmed the result in 1930.7, 8 By then, sodium polysulfides have been identified as a

group of salt materials with typical formula Na2Sx with x = 2, 3, 4 and 5. As mentioned above,

sodium tetrasulfide (Na2S4) is an important intermediate product during the charge/discharge

process of the Na-S battery. The material is in yellow solid form under ambient conditions and has

a formula of Na2S4. The detailed physical and chemical properties are listed in Table 4.1. The

material has been well characterized by x-ray diffraction under ambient conditions in previous

crystallographic studies.9 The crystal structure of the material is tetragonal, with eight molecules

per unit cell. The space group of the material is I4̅2𝑑 with the S42- on Ci symmetry sites. In figure

4.1, the structural formula and crystal structure are illustrated, and more detailed crystal structure

information has been listed in Table 4.2.

Page 74: Structural Tuning and Spectroscopic Characterizations of

61

Table 4.1 Na2S4 physical and chemical properties information

Table 4.2 Na2S4 crystal structure information9

Na2S4 ICSD Code: 2586

Crystal System Tetragonal Space Group I -4 2 d (No. 122)

S-S Bond Length 2.074 Å S-S Bond Angle 109.761°

Cell Parameter a = b = 9.5965 c = 11.7885

α = β = γ = 90° Cell Volume 1085.64 Å3

Figure 4.1 a. Crystal structure of Na2S4 with space group. b. Crystal structure of Na2S4 along

with c. c. Structural formula. Purple and yellow spheres represent sodium and sulfur atoms.

IUPAC Name Sodium tetrasulfide CAS Number 12034-39-8

Molecular Formula Na2S4 Appearance Yellow crystalline

powder

Molecular Weight 174.24 Density 1.268 g/cm3

Melting Point 275°C Solubility in Water Soluble in water

Sulfur Sodium

c

a b

Page 75: Structural Tuning and Spectroscopic Characterizations of

62

Pressure is a thermodynamic variable, which can be a powerful tool for tuning the chemical bond,

crystal structure, and properties of materials. Significantly, applying external pressure can

effectively alter the crystal structures of a material. Extensive high-pressure studies on battery

materials reported that the pressure-induced structural transformation would affect the material's

performance in a battery application.10-17 Although the Na-S system was introduced in the 1960s,

the high-pressure study on Na2S4 remains unknown experimentally. Other sodium sulfide

materials such as Na2S was studied under high pressure in 2000 using X-ray diffraction.18 The

result showed that Na2S undergoes two phase transitions upon compression to 22 GPa. The crystal

structure transforms from cubic to orthorhombic at around 7 GPa, then becomes hexagonal at about

15 GPa. In addition, the high-pressure structures of several polysulfide materials in the Na-S

system were predicted through a computational method by Wan and Xu in 2019.19 Their results

showed that several sodium polysulfide materials would have phase transitions at high-pressure

conditions. Therefore, A thorough investigation of the structure evolution and phase

transformation induced by external pressure on Na2S4 could provide additional structural

information and comparison with other materials in the Na-S family. Moreover, it could also

inspire the solution of the corrosion effect and be a way to synthesize novel material structures

with properties that differ from the ambient condition they have.

Herein, the first high-pressure study of Na2S4 by in-situ Raman spectroscopy using a Diamond

Anvil cell is reported from this study. The structural changes of the material were explored upon

compression above 30 GPa. The Raman spectra revealed the exciting effect of high pressure on

the material. Three phase transitions were observed at about 1, 6 and 11 GPa. This new structural

information provides a further understanding of this material, which allows the precise

determination of structural stability and compressibility. It also could be an inspiration for the

development of new materials/structures for battery studies.

Page 76: Structural Tuning and Spectroscopic Characterizations of

63

4.2 Experimental

4.2.1 Material

The Na2S4 material used for the experiment was purchased from Beantown Chemical (purify ~

90%) and used without further purification. Due to the sensitivity of moisture, the material was

kept in a nitrogen glove box.

4.2.2 High-Pressure Experiment procedures

A gasket was first placed into DAC and pre-indented up to 10 GPa for preparing the sample

chamber. After the pre-indentation, the gasket was removed from the DAC. Then, a needle with a

diameter of 130 microns is used to drill a small hole in the centre of the pre-indentation mark by

the EDM drilling machine. The small hole was used as the sample chamber, and the gasket was

inserted back to the diamond anvil cell. Due to the air sensitivity of samples, DAC was transferred

into a nitrogen glovebox. A ruby was loaded into the cell before moving the DAC into the glovebox.

Next, the sample was loaded into the DAC sample chamber inside the glovebox, and then the cell

was closed and compressed to "finger tight" before taking it out from the glovebox. Ruby

fluorescence was checked for pressure first, and then Raman measurement was taken for near

ambient pressure. For the compression process, the pressure was increased by tightening each bolt

at a certain angle equally. Then the pressure was rechecked with ruby fluorescence followed by

Raman. The process was repeated several times until the pressure reached about 32 GPa. Moreover,

the decompression process decreased pressure by equally losing each bolt at a certain angle,

followed by pressure and Raman measurement. The process was repeated until the pressure was

close to the ambient pressure.

4.2.3 Characterizations

In-situ Raman spectroscopy (refer to section 2.2.2 in Chapter 2) was applied to monitor pressure-

induced phase transitions of samples. In the Raman system, the scattered light was dispersed using

an imaging spectrograph equipped with a 1200 lines/mm grating achieving a resolution of about

0.2 cm–1. The 532 nm laser was used as the exciting source. The laser beam was focused onto a

Page 77: Structural Tuning and Spectroscopic Characterizations of

64

spot with a diameter of approximately 5μm by using an objective microscope with a magnification

of 50×.

4.3 Result and Discussion

4.3.1 Raman results of Na2S4 upon compression and

decompression

Raman spectrum of Na2S4 collected at near ambient conditions is shown in Figure 4.2. Nine active

Raman modes were observed in the Na2S4 sample at 0.38 GPa. From the spectrum, the vibrational

modes ν1- ν5 from 200-500 cm-1 correspond to the S42- chain structure. The bands with small

intensity at lower wavenumber below 200 cm-1 are assigned to lattice modes. According to

previous studies, the torsion mode at about 80 cm-1 was not observed from the spectrum. The

detailed peak assignments of each Raman mode of Na2S4 at 0.38 GPa are shown in Table 4.3, and

the assignment is based on previous studies.20-22

Figure 4.2 Raman spectrum of Na2S4 with peak assignment collected at 0.38 GPa

Page 78: Structural Tuning and Spectroscopic Characterizations of

65

Table 4.3 Peak Assignment of Raman Modes of Na2S4

Mode Frequency (cm-1)

Peak Assignment Experiment* Reference20-22

L1 89 88

Lattice mode

L2 100 97

L3 131 128

L4 176 173

ν1 221 206 Symmetric bend

ν2 243 239 Asymmetric bend

ν3 443 445 Symmetric stretching, central S-S

ν4 466 468 Asymmetric stretching

ν5 480 482 Symmetric stretching, terminal S-S

*Experimental data was collected at 0.38 GPa.

The evolution of the Raman spectra of Na2S4 upon compression and decompression is shown in

Figure 4.3. From the compression plot, as the pressure increasing most Raman peaks are blue-

shifted, caused by a decrease in the interatomic distance. Upon compression above 1 GPa, the

lattice mode L1 at around 100 cm-1 becomes diminished. At the same conditions, the lattice mode

L3 at about 130 cm-1 is split into two peaks (see the arrow marked on the spectrum at 1.01 GPa in

Figure 4.3). The observation indicates a possible phase transition at this pressure. With increasing

pressure, a new peak is observed between Lattice mode L1 and L3 at about 6.26 GPa. At the same

pressure condition, Lattice mode L4 and vibrational ν2 become vanished. The appearance and

disappearance of peaks at the lattice mode region could result from the second phase transition.

Upon compression around 11 GPa, most of them vanished except ν3 and ν5, which could signify

the third phase transition. Moreover, when the pressure reaches above 11 GPa, all lattice modes

disappear, indicating the collapse of the crystal structure. All peaks become weaker and broader

as pressure further increases. After the pressure exceeded 20 GPa, all peaks become disappeared,

indicating the material reached an amorphous state. The compression process was stopped at 31.79

Page 79: Structural Tuning and Spectroscopic Characterizations of

66

GPa. No further change is observed from Raman spectra which indicates no structural changes

after the amorphous state.

According to the decompression process, the symmetric stretching mode of central S-S was

recovered first. At about 5 GPa, all other modes start appearing from the Raman spectrum. When

the pressure is released close to ambient pressure at 0.4 GPa, it can be found that all peaks

recovered as they were at ambient pressure. Therefore, it can be concluded that the high-pressure

induced effect on the material is reversible.

Figure 4.3 Raman spectra of Na2S4 upon compression and decompression. Formation of new

Peaks and splitting of peaks are marked on the spectra by arrows.

Page 80: Structural Tuning and Spectroscopic Characterizations of

67

4.3.2 Pressure effects on Raman modes of Na2S4

To have a quantitative understanding of the Raman shift changes, the vibrational frequencies as a

function of pressure were also plotted, as shown in Figure 4.4. The phase transition boundaries

could be determined by the discontinuity trend of each Raman mode and the appearance or

disappearance of modes. Additionally, the pressure coefficients were also analyzed by a linear

fitting and listed in Table 4.4. The phase transitions could be further verified by the significant

changes of the pressure coefficients in different pressure regions.

Figure 4.4 Pressure dependence of selected Raman modes of Na2S4.The vertical dashed lines

indicate the proposed phase boundaries. Different symbols represent Raman modes with

different origins.

Page 81: Structural Tuning and Spectroscopic Characterizations of

68

Overall, the positive pressure coefficients indicate that most Raman modes exhibit pressure-

induced blue shifts due to the reduction of the volume and the enhancement of the interatomic

interaction. For the first phase transition at 1 GPa, in addition to the appearance/disappearance of

peaks, several peaks (L1-L4, ν1 and ν2) have a significant change in their pressure coefficients

compared with phase II. According to the peak assignment, there are no noticeable changes in the

vibrational modes associated with S42- chains, so the first structural change is probably caused by

enhanced intermolecular interactions as a result of the reduced volume by compression. The

second phase transition occurs around 6 GPa can be confirmed by the discontinuity trend of all

Raman modes. The second phase transition may be due to the geometry changes in S42- chains

according to the disappearance of ν2. The third phase at 11 GPa is suggested by the disappearance

of all lattice modes and vibrational modes ν1 and ν4. In addition, the pressure coefficients of each

mode become steep after the pressure past 6 GPa. At 11 GPa, the crystal structure may collapse

since all lattice modes vanish. A metastable sulfur structure may be formed at this stage. Above

20 GPa, all modes have disappeared, and no peak remained in the spectrum. Therefore, the material

probably reaches the amorphous state under the condition since no active Raman mode was

detected.

Regarding structural stability and compressibility of the material, pressure dependence values in

each phase region were also examined. The pressure dependence value of each mode was

compared among all the phases. According to the dependence value of lattice modes in the phase

I region, the phase I structure probably significantly changed from the compression. Phase II has

the largest pressure coefficients of lattice modes, which suggests the compression was more

effective on the lattice and volume reduction than on the S-S bonds. In the phase III region, the

vibrational modes of S-S bonds exhibit the largest values, indicating the intramolecular bonds were

more effectively compressed. In the phase VI region, the material became fairly stable and hard to

compress due to the small pressure dependence values of the remaining modes.

Page 82: Structural Tuning and Spectroscopic Characterizations of

69

Table 4.4 Pressure dependence (dν/dP, cm-1/GPa) of vibrational modes of Na2S4 upon

compression from Raman spectra

Mode Raman mode Raman shift

(cm-1)

Phase

I

Phase

II

Phase

III

Phase

VI

L1

Lattice mode

89 4.0 1.0 0.7 -

L2 101 -2.0 - - -

L3 130 14.0 8.6 4.0 -

L4 176 3.0 6.2 - -

ν1 Symmetric bend 221 -6.0 4.8 8.0 -

ν2 Asymmetric bend 243 1.0 4.8 - -

ν3 Symmetric stretching central S-S 443 2.0 4.0 6.6 1.4

ν4 Asymmetric stretching 466 3.0 3.2 6.8 -

ν5 Symmetric stretching terminal S-S 480 2.0 3.4 5.4 1.5

4.3.3 Discussion

To further interpret the crystal structures of each high-pressure phase inferred from Raman spectra,

the result has been compared by the studies of sulfur-containing materials with similar structures.

Firstly, the result of this study can be understood by comparing it with previous high-pressure

studies on sulfur, considering the S42- chain could have similar changes with S8 structure under

high-pressure conditions. As a well-studied material, sulfur has been investigated under high-

pressure conditions in several studies.23-28 Sulfur has an orthorhombic crystal structure consisting

of eight-membered ring molecules S8 under ambient conditions. Several early studies by Raman

spectroscopic suggest that sulfur have 2 phase transitions upon compression to 25 GPa. The two

phase transitions were according to the discontinuity trend of Raman spectra at 5 GPa and 10 GPa.

However, one study in 1995 by Yoshioka and Nagata has shown that the two phase transitions can

be observed when the energy and power of the exciting laser are sufficiently large.27 In addition,

both phase transitions were not observed through X-ray measurement. Moreover, from the most

recent Sulfur Raman study under high hydrostatic pressure, their result reported the amorphous

state of sulfur is at ~16 GPa where the symmetric bond stretching vibrations of the S8 ring dominate

Page 83: Structural Tuning and Spectroscopic Characterizations of

70

in the Raman spectrum above 10 GPa.28 However, the phase transition at 5 GPa was not identified.

Comparing the observations of our Raman results of Na2S4 from this study, it has also shown two

possible transitions at similar pressures of 6 and 11 GPa inferred from the discontinuity trend of

Raman modes. Furthermore, the Raman spectra were also dominated by symmetric bond

stretching vibrations of S42- above 11 GPa. Therefore, the structural changes in the S4

2- chain might

be similar to those pressure-induced changes in the S8 ring in the pressure region of 6 to 11 GPa.

Secondly, another sodium sulfide compound, Na2S, has been studied under high-pressure

conditions by X-ray diffraction.18 Their result showed that Na2S undergoes 2 phase transitions

upon compression to 22 GPa. Na2S transform from cubic to orthorhombic at 7 GPa and then

become hexagonal at 16 GPa. At ambient pressure, the crystal structure of Na2S4 is tetragonal with

space group I -4 2 d. Although the two materials have different starting crystal structures, similar

transformations in the lattice may occur considering the similar nature of ionic interactions

between the sodium cation and sulphur anion. The tetragonal could transform into an orthorhombic

phase, then followed by a higher symmetry crystal structure before the amorphization state, given

by the splitting observation in one of the lattice modes (L3). Typically, the increase in lattice modes

indicates higher crystal lattice symmetry.

Moreover, a computational study has been made to predict the possible high-pressure phases for

several sodium polysulfides, including Na2S4.19 From the computational research, the researcher

was able to reproduce the phase transition of Na2S. Their result has a good agreement in phase

transitions but with different pressure (2 and 7 GPa). According to their prediction result of Na2S4,

it would become amorphous at about 8 GPa, consistent with our observation but at higher pressures

(i.e., above 11 GPa). Overall, the detailed molecular and crystal structures of Na2S4 need to be

further investigated by in-situ high-pressure X-ray diffraction to be conducted at the synchrotron

radiation facility.

4.4 Conclusions

In summary, a battery-related sodium polysulfide material was investigated under high pressure

by in-situ Raman spectroscopy. The pressure effect on the structural evolution and compressibility

was also studied. Na2S4 may undergo 3 phase transitions at about 1, 6, and 11 GPa according to

Page 84: Structural Tuning and Spectroscopic Characterizations of

71

the discontinuity trend of Raman spectra, and the amorphous state could be above 16 GPa. By

comparison with high-pressure studies of material containing similar structures. The possible high-

pressure phase transformation could be from tetragonal to orthorhombic. Then it would become a

higher symmetry structure before reaching the amorphous state. There is also a possibility that the

starting tetragonal structure remains unchanged until the material reaches its amorphous state.

However, further X-ray diffraction studies are still required for clarification on phase transitions.

These findings provide information on the structural evolution and compressibility of the material

under high-pressure conditions. It also proves that high-pressure technology is a way to tune the

materials' structures. Finally, the phase transition may lead to electrical property changes in the

material, which results in further battery-related research interest.

Page 85: Structural Tuning and Spectroscopic Characterizations of

72

4.5 Reference

1. Chen, H.; Xu, Y.; Liu, C.; He, F.; Hu, S., Chapter 24 - Storing Energy in China—An

Overview. In Storing Energy, Letcher, T. M., Ed. Elsevier: Oxford, 2016; pp 509-527.

2. Gupta, N.; Kaur, N.; Jain, S. K.; Singh Joshal, K., Chapter 3 - Smart grid power system.

In Advances in Smart Grid Power System, Tomar, A.; Kandari, R., Eds. Academic Press: 2021; pp

47-71.

3. Breeze, P., Chapter 10 - Power System Energy Storage Technologies. In Power Generation

Technologies (Third Edition), Breeze, P., Ed. Newnes: 2019; pp 219-249.

4. Okuyama, R.; Nakashima, H.; Sano, T.; Nomura, E., The Effect of Metal Sulfides in the

Cathode on Na/S Battery Performance. Journal of Power Sources 2001, 93 (1), 50-54.

5. Heusler, K. E.; Grzegorzewski, A.; Knödler, R., Corrosion of Metallic Materials in Sodium

Polysulfide. Journal of The Electrochemical Society 1993, 140 (2), 426-431.

6. Friedrich, K.; Waehlert, M., Investigation of Layer-Forming Systems. Metall und Erz 1914,

160-167.

7. Rule, A.; Thomas, J. S., XXIII.—The Polysulphides of the Alkali Metals. Part I. The

Polysulphides of Sodium. Journal of the Chemical Society, Transactions 1914, 105 (0), 177-189.

8. Pearson, T. G.; Robinson, P. L., CLXXXIX.—The Polysulphides of the Alkali Metals. Part

I. Sodium (i). Journal of the Chemical Society (Resumed) 1930, (0), 1473-1497.

9. Tegman, R., The Crystal Structure of Sodium Tetrasulphide, Na2S4. Acta

Crystallographica Section B 1973, 29 (7), 1463-1469.

10. Gallardo-Amores, J. M.; Biskup, N.; Amador, U.; Persson, K.; Ceder, G.; Morán, E.;

Arroyo y de Dompablo, M. E., Computational and Experimental Investigation of the

Transformation of V2O5 Under Pressure. Chemistry of Materials 2007, 19 (22), 5262-5271.

11. García-Moreno, O.; Alvarez-Vega, M.; García-Jaca, J.; Gallardo-Amores, J. M.; Sanjuán,

M. L.; Amador, U., Influence of the Structure on the Electrochemical Performance of Lithium

Transition Metal Phosphates as Cathodic Materials in Rechargeable Lithium Batteries:  A New

High-Pressure Form of LiMPO4 (M = Fe and Ni). Chemistry of Materials 2001, 13 (5), 1570-1576.

12. Amador, U.; Gallardo-Amores, J. M.; Heymann, G.; Huppertz, H.; Morán, E.; Arroyo y

de Dompablo, M. E., High pressure polymorphs of LiCoPO4 and LiCoAsO4. Solid State Sciences

2009, 11 (2), 343-348.

Page 86: Structural Tuning and Spectroscopic Characterizations of

73

13. Yamaura, K.; Huang, Q.; Zhang, L.; Takada, K.; Baba, Y.; Nagai, T.; Matsui, Y.;

Kosuda, K.; Takayama-Muromachi, E., Spinel-to-CaFe2O4-Type Structural Transformation in

LiMn2O4 under High Pressure. Journal of the American Chemical Society 2006, 128 (29), 9448-

9456.

14. Arroyo-deDompablo, M. E.; Dominko, R.; Gallardo-Amores, J. M.; Dupont, L.; Mali,

G.; Ehrenberg, H.; Jamnik, J.; Morán, E., On the Energetic Stability and Electrochemistry of

Li2MnSiO4 Polymorphs. Chemistry of Materials 2008, 20 (17), 5574-5584.

15. Yoncheva, M.; Stoyanova, R.; Zhecheva, E.; Alcántara, R.; Ortiz, G.; Tirado, J. L., Effect

of the High Pressure on the Structure and Intercalation Properties of Lithium–Nickel–Manganese

Oxides. Journal of Solid State Chemistry 2007, 180 (6), 1816-1825.

16. Huang, Y.; He, Y.; Sheng, H.; Lu, X.; Dong, H.; Samanta, S.; Dong, H.; Li, X.; Kim,

D. Y.; Mao, H.-k.; Liu, Y.; Li, H.; Li, H.; Wang, L., Li-Ion Battery Material under High Pressure:

Amorphization and Enhanced Conductivity of Li4Ti5O12. National Science Review 2018, 6 (2),

239-246.

17. Wang, X.; Zhang, P.; Tang, X.; Guan, J.; Lin, X.; Wang, Y.; Dong, X.; Yue, B.; Yan,

J.; Li, K.; Zheng, H.; Mao, H.-k., Structure and Electrical Performance of Na2C6O6 under High

Pressure. The Journal of Physical Chemistry C 2019, 123 (28), 17163-17169.

18. Vegas, A.; Grzechnik, A.; Syassen, K.; Loa, I.; Hanfland, M., Reversible Phase

Transitions in Na2S under Pressure: A Comparison with the Cation Array in Na2SO4. Acta

crystallographica. Section B, Structural science 2001, 57, 151-6.

19. Wan, B.; Xu, S.; Yuan, X.; Tang, H.; Huang, D.; Zhou, W.; Wu, L.; Zhang, J.; Gou,

H., Diversities of Stoichiometry and Electrical Conductivity in Sodium Sulfides. Journal of

Materials Chemistry A 2019, 7 (27), 16472-16478.

20. El Jaroudi, O.; Picquenard, E.; Gobeltz, N.; Demortier, A.; Corset, J., Raman

Spectroscopy Study of the Reaction between Sodium Sulfide or Disulfide and Sulfur: Identity of

the Species Formed in Solid and Liquid Phases. Inorganic Chemistry 1999, 38 (12), 2917-2923.

21. El Jaroudi, O.; Picquenard, E.; Demortier, A.; Lelieur, J.-P.; Corset, J., Polysulfide

Anions II:  Structure and Vibrational Spectra of the S42- and S52- Anions. Influence of the Cations

on Bond Length, Valence, and Torsion Angle. Inorganic Chemistry 2000, 39 (12), 2593-2603.

Page 87: Structural Tuning and Spectroscopic Characterizations of

74

22. Janz, G. J.; Downey, J. R.; Roduner, E.; Wasilczyk, G. J.; Coutts, J. W.; Eluard, A.,

Raman Studies of Sulfur-Containing Anions in Inorganic Polysulfides. Sodium Polysulfides.

Inorganic Chemistry 1976, 15 (8), 1759-1763.

23. Wang, L.; Zhao, Y.; Lu, R.; Meng, Y.; Fan, Y.; Luo, H.; Cui, Q.; Zou, G., High Pressure

Raman and X-Ray Studies of Sulfur and its New Phase Transition. High‐Pressure Research in

Mineral Physics: A Volume in Honor of Syun‐iti Akimoto 1987, 299-304.

24. Häfner, W.; Kritzenberger, J.; Olijnyk, H.; Wokaun, A., Phase Transitions in Crystalline

Sulfur at P>8 GPa, Observed by Raman Spectroscopy in a Diamond Anvil Cell. High Pressure

Research 1990, 6 (1), 57-75.

25. Eckert, B.; Jodl, H. J.; Albert, H. O.; Foggi, P., Sulphur at High Pressure and Low

Temperatures. In Frontiers of High-Pressure Research, Hochheimer, H. D.; Etters, R. D., Eds.

Springer US: Boston, MA, 1991; pp 143-160.

26. Luo, H.; Ruoff, A. L., High‐Pressure Raman Scattering Study of Sulfur. AIP Conference

Proceedings 1994, 309 (1), 1527-1530.

27. Yoshioka, A.; Nagata, K., Raman Spectrum of Sulfur under High Pressure. Journal of

Physics and Chemistry of Solids 1995, 56 (3), 581-584.

28. Andrikopoulos, K. S.; Gorelli, F. A.; Santoro, M.; Yannopoulos, S. N., Elemental Sulfur

under High Hydrostatic Pressure. An up-to-date Raman Study. High Pressure Research 2013, 33

(1), 134-140.

Page 88: Structural Tuning and Spectroscopic Characterizations of

75

Chapter 5 5 Summary and future work

5.1 Summary

Lithium-ion battery (LIB) has been widely used as a power source in various applications, from

portable consumer electronics to electric vehicles, due to its high specific energy density and

desired cycle life. However, due to the increasing demand for better-functioning batteries, the

lithium-ion battery starts becoming straining with its limitation in energy density and safety issues.

In the development of new battery materials and techniques, many polysulfide materials have

received increasing attention recently. Moreover, high-pressure has been considered an effective

tool for modifying and tuning materials’ chemical, structural, mechanical, electronic, magnetic,

and phonon properties. Many battery materials, including both electrode and electrolyte materials,

have been studied under high pressure. Some of these studies show that the pressure-induced

structural transformation/evolution would affect the material's performance in battery applications.

Therefore, high-pressure studies on battery-related materials could provide valuable information

on materials’ compressibility, structural evolutions, and property changes under extreme

conditions. In this thesis, the first high-pressure study of the three polysulfide materials (P4S3, P4S10,

and Na2S4) using in-situ Raman spectroscopy and IR spectroscopy are reported.

In chapter 3, two materials (P4S3 and P4S10) were investigated under high pressure by vibrational

spectroscopies. Using Raman spectroscopy, P4S3 was found to undergo 2 phase transitions through

the compression process up to 24.87 GPa. The first possible phase transition is between 2.5 – 4

GPa, and the second phase is at about 7 GPa. The material became hard to compress when the

pressure reached about 13 GPa. No further structural change was observed up to 24.87 GPa. Our

analysis suggests that both phase transitions could be induced by the volume changes and the

distortion of the molecular structure because no bond broken was observed, and the structural

changes were reversible. In regard to the stability and compressibility, pressure-dependent cell

volume was calculated using first-principle methods. The calculated cell volume has been fitted to

the Birch-Murnaghan equation of state, from which the bulk modulus K0 and its first derivative

K0' are determined to be 9.2110 GPa and 5.6683. The pressure effect is isotropic under 3 GPa

Page 89: Structural Tuning and Spectroscopic Characterizations of

76

affecting all three parameters. When the pressure passes 3 GPa, the pressure has less effect on

along a-axis direction. Above 13 GPa, the b axis becomes the most compressible lattice direction

predominantly.

P2S5 (P4S10) were investigated by using Raman and IR spectroscopies under high pressure. P2S5

could have two transformations upon compression to 17 GPa. The first phase transition occurs

around 3 GPa, perhaps caused by the double bond broken between phosphorus and sulfur. The

second phase transition was suggested at about 7.5 GPa and could be due to the formation of a

novel metastable structure. When the pressure reached above 12 GPa, the material became

amorphous. The decompression process suggested that the high-pressure effect is irreversible once

the pressure reaches the amorphous state. Preliminary first-principle calculations suggest that high-

pressure phases are formed by P=S double bond breaking to form new structures, such as a dimer

or polymetric structure.

In chapter 4, Na2S4 was investigated under high pressure up to 30 GPa by in-situ Raman

spectroscopy. The pressure effect on the structural evolution and compressibility was also studied.

Na2S4 may undergo 3 phase transitions at about 1, 6, and 11 GPa according to the discontinuity

trend of Raman spectra, and the amorphous state could be above 16 GPa. By comparison with

high-pressure studies of material containing similar structures. The possible high-pressure phase

transition could be from tetragonal to orthorhombic. Then it would become a higher symmetry

structure before hitting the amorphous state. There is also a possibility that the starting tetragonal

structure remains unchanged until the material reaches the amorphous state.

These findings provide information on the structural evolution and compressibility of the material

under high-pressure conditions. It also proves that high-pressure technology is a way to tune the

materials' structures. Finally, the phase transition may lead to electrical property changes in the

materials, which results in further battery-related research interest.

5.2 Future Work

Based on the discussion of speculation for possible structural changes in the previous sections,

further investigation includes:

Page 90: Structural Tuning and Spectroscopic Characterizations of

77

(1) For P4S3, far-IR measurements under high pressure could be performed to confirm the

concluded phase transition from Raman spectra. Further X-ray diffraction studies are required for

clarification of phase transition boundaries and the high-pressure phases.

(2) For P2S5, further X-ray diffraction studies are still needed to clarify the material

transformations and determine the high-pressure structures.

(3) For Na2S4, far-IR measurements can be conducted to confirm the phase transitions with the

result from Raman spectra. Further X-ray diffraction studies are also needed to clarify the high-

pressure phases with boundaries. High-pressure study on electrical properties, such as in-situ

conductivity measurement, could be investigated.

Page 91: Structural Tuning and Spectroscopic Characterizations of

78

Curriculum Vitae

Name: Daiqiang Liu

Post-secondary Trent University

Education and Peterborough, Ontario, Canada

Degrees: 2010-2013 BSc. Environmental Chemistry

The University of Western Ontario

London, Ontario, Canada

2019-2021 MSc. Inorganic Chemistry

Related Work Teaching Assistant

Experience The University of Western Ontario

2019-2021

Other Work Attix Pharmaceuticals Co.

Experience 2014-2016

Course Summary

2019 Fall Term Chem 9541A Crystallography I

2020 Winter Term Chem 9546T Optical Properties of Solids

2020 Fall Term Chem 9754R Powder Diffraction

Mentorship for Undergraduate Students

Robert Murray, Chem 4491 Honor’s BSc., 2019-2020, Thesis: Structural tuning of polysulfides

at high pressure

Zibo Gao, Chem 4491 Honor’s BSc., 2020-2021, Thesis: Structural tuning of MOF

Zn2(BDC)2DABCO under high external pressure by in situ vibrational spectroscopies