carbon rich polymer-derived ceramics: processing

160
Clemson University Clemson University TigerPrints TigerPrints All Dissertations Dissertations August 2021 Carbon Rich Polymer-Derived Ceramics: Processing, Carbon Rich Polymer-Derived Ceramics: Processing, Characterization, and Properties Characterization, and Properties Michelle Greenough Clemson University, [email protected] Follow this and additional works at: https://tigerprints.clemson.edu/all_dissertations Recommended Citation Recommended Citation Greenough, Michelle, "Carbon Rich Polymer-Derived Ceramics: Processing, Characterization, and Properties" (2021). All Dissertations. 2882. https://tigerprints.clemson.edu/all_dissertations/2882 This Dissertation is brought to you for free and open access by the Dissertations at TigerPrints. It has been accepted for inclusion in All Dissertations by an authorized administrator of TigerPrints. For more information, please contact [email protected].

Upload: others

Post on 21-Feb-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Carbon Rich Polymer-Derived Ceramics: Processing

Clemson University Clemson University

TigerPrints TigerPrints

All Dissertations Dissertations

August 2021

Carbon Rich Polymer-Derived Ceramics: Processing, Carbon Rich Polymer-Derived Ceramics: Processing,

Characterization, and Properties Characterization, and Properties

Michelle Greenough Clemson University, [email protected]

Follow this and additional works at: https://tigerprints.clemson.edu/all_dissertations

Recommended Citation Recommended Citation Greenough, Michelle, "Carbon Rich Polymer-Derived Ceramics: Processing, Characterization, and Properties" (2021). All Dissertations. 2882. https://tigerprints.clemson.edu/all_dissertations/2882

This Dissertation is brought to you for free and open access by the Dissertations at TigerPrints. It has been accepted for inclusion in All Dissertations by an authorized administrator of TigerPrints. For more information, please contact [email protected].

Page 2: Carbon Rich Polymer-Derived Ceramics: Processing

CARBON RICH POLYMER-DERIVED CERAMICS: PROCESSING,

CHARACTERIZATION, AND PROPERTIES

_______________________________________________________

A Dissertation

Presented to

the Graduate School of

Clemson University

_______________________________________________________

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

Materials Science and Engineering

_______________________________________________________

by

Michelle Greenough

August 2021

_______________________________________________________

Accepted by:

Dr. Rajendra Bordia, Committee Chair

Dr. Stephen Creager

Dr. Luiz Jacobsohn

Dr. Jianhua Tong

Dr. Marek Urban

Page 3: Carbon Rich Polymer-Derived Ceramics: Processing

ii

Abstract

Polymer derived ceramics is an attractive route to make Si-based ceramics with

high temperature stability in a broad range of physical forms. Partial thermal

decomposition of hybrid Si- and C-based polymers allows for the creation of unique

materials with attractive properties, such as high surface area, high electrical

conductivity, and pores in nanometer range. However, investigations of these materials

have been limited and are the focus of this research. The unique aspect of this work is

the pyrolysis temperature which is at a low/intermediate temperature range of below

1000 °C. Several systems have been investigated to explore a broad range of

compositional space. Specifically, system composed of polysiloxane and polycarbosilane

polymers as Si-based polymers and divinylbenzene (DVB) as C-based polymer have been

studied. In some cases, boron and boron nitride were incorporated as well to influence

the microstructure and chemical stability of the material. Overall, the processing,

characterization, and specific application-relevant properties of these materials were

investigated. In addition to the composition, the focus was on varying the pyrolysis

temperature. The characterization of the material was conducted using a variety of

chemical and microstructural characterization methods, including Fourier-transform

infrared spectroscopy (FTIR), thermogravimetric analysis (TGA), Raman spectroscopy, X-

ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), solid state nuclear

magnetic resonance (NMR), thermogravimetric analysis coupled with mass

Page 4: Carbon Rich Polymer-Derived Ceramics: Processing

iii

spectrometry (TGA-MS), and electrochemical impedance spectroscopy (EIS). The

research has provided insights on the chemical, structural, and microstructural

evolution, as a function of precursor composition and thermal treatment, for this new

class of materials. Materials with a unique combination of properties – high electrical

conductivity and high microporosity, are being developed for potential applications as

gas separation membranes and/or battery electrodes.

Page 5: Carbon Rich Polymer-Derived Ceramics: Processing

iv

Acknowledgements

I would like to first acknowledge my advisor, Dr. Raj Bordia, for his guidance and

mentorship throughout my Ph.D. I would like to thank him for the opportunities to

travel to several different conferences across the United States and Europe along with

the ability to collaborate with different labs. I would like to thank the Bordia group

members both past and current, including the undergraduate students, who helped

with my research.

I would also like to thank my committee members: Dr. Creager, Dr. Tong, Dr.

Jacobsohn, and Dr. Urban for their efforts on my dissertation.

I would like to thank my collaborators in Europe, where I was fortunate enough

to work with. First, to Dr. Bernard’s group in Limoges, France for graciously hosting me

during my summer abroad, especially Drs. Maxime Balestrat and David Lopez Ferber.

Thank you for teaching me more about French and supplying me with coffee. Next, I

would like to thank my collaborators in Koblenz, Germany with Dr. Quirmbach’s group

especially, Dominik Hahn, for hosting me for two weeks.

To my “best short-term roommate”, Dr. Cathrine Christiansen, thank you for

being a great roommate and friend for the semester and later showing me Denmark on

my travels. Hopefully, you will let me come back soon.

To continuing collaborators, I would like to thank Oak Ridge National Laboratory

and Dr. Anton Ievlev for the ToF-SIMS work on future projects.

I would like to thank the Clemson facilities at AMRL, especially Dr. Kelliann

Koehler for her work with XPS and George Wetzel for assistance with BET and SEM. I

would like to thank the Brick Center at Clemson, Drs. Sanders and Huygen, for their

quick work with TGA-MS.

Page 6: Carbon Rich Polymer-Derived Ceramics: Processing

v

A special thank you to Ms. Diane Swope and Ms. Tonya Bledsoe for all their

efforts in the Materials Science and Engineering Office. I would like to thank Mr. David

White who helped me fix furnaces and equipment throughout the years. And a big

thank you to Ms. Kimberly Ivey, who helped with all the FTIR and TGA samples on

campus.

I would like to thank my financial support from the National Science Foundation

Graduate Research Fellowship Program, National Science Foundation Intern funding,

EPSCOR MADE in SC NSF funding, and CNMS Oak Ridge National Laboratory.

And last but not least, I would like to thank my family and friends who have

supported me throughout my Ph.D., especially my mom (Valerie), my dad (Craig), my

brother (Ryan), and my significant other (Brandon).

Page 7: Carbon Rich Polymer-Derived Ceramics: Processing

vi

Table of Contents

Abstract………………………………………………………………………………………………………………………… ii

Acknowledgements…………………………………………………………………………………………………….. iv

List of Figures…………………………………………………………………………………………………………....... ix

List of Tables………………………………………………………………………………………………………………. xv

Chapter 1: Introduction and Motivation……………………………………………………………………… 1

Chapter 2: Literature Review………………………………………………………………………………………. 4

2.1 Polymer Derived Ceramics…………………………………………………………………………… 4

2.1.1 Polymer Precursors………………………………………………………………………. 6

2.1.2 Polymer to Ceramic Conversion……………………………………………………. 9

2.1.3 Polymer Derived Ceramics with Added Carbon or Boron……………. 11

2.2 Polymer Derived Ceramics by Pyrolysis at Low/Intermediate Temperatures 15

2.3 Focus of this Dissertation: Low/Intermediate Pyrolysis Polymer Derived Ceramics with High Surface Area and High Carbon Content................... 19

Chapter 3: Materials Processing and Characterization Techniques…………………………… 20

3.1 Materials Processing…………………………………………………………………………………. 20 3.1.1 Preceramic Polymers...................................................................... 20 3.1.2 Crosslinking..................................................................................... 21 3.1.3 Pyrolysis.......................................................................................... 22

3.2 Characterization Techniques.......................................................................... 23 3.2.1 Thermogravimetric Analysis (TGA, dTGA) and TGA coupled with

Mass Spectrometry (TGA-MS) ..................................................... 23 3.2.2 X-ray Diffraction (XRD)................................................................. 24 3.2.3 Chemical Analysis: ATR FTIR, Raman Spectroscopy, XPS, and

Elemental Analysis.................................................................. 25 3.2.4 Solid State NMR.............................................................................. 28 3.2.5 BET Surface Area Analysis…………………………………………………………… 28 3.2.6 Electrochemical Impedance Spectroscopy...................................... 29

Page

Table of Contents………………………………………………………………………………………………………….. i

Page 8: Carbon Rich Polymer-Derived Ceramics: Processing

vii

Chapter 4: Results and Discussion on Materials Processing and Characterization for Polycarbosilane………………………………………………………………………………………………………… 31

4.1 Polycarbosilane with Carbon Addition……………………………………………………… 31

4.1.1 Processing...................................................................................... 31

4.1.2 Characterization.............................................................................. 33

4.1.3 Summary and Discussion................................................................ 50

4.2 Polycarbosilane with Boron Addition and Boron/Nitrogen Additive…………. 52

4.2.1 Processing…………………………………………………………………………………… 52

4.2.2 Characterization…………………………………………………………………………. 55

4.2.3 Summary and Discussion……………………………………………………………. 60

Chapter 5: Results and Discussion on Materials Processing and Characterization for Polysiloxane……………………………………………………………………………………………………………… 62

5.1 Processing and Characterization of Polysiloxane Derived Ceramics…………… 62 5.1.1 Processing of Polysiloxanes ………………………………………………………… 62 5.1.2 Characterization of Polysiloxane Derived Ceramics…………………….. 64 5.1.3 Summary and Discussion…………………………………………………………….. 72

5.2 Polysiloxane with Added Carbon……………………………………………………………….. 73 5.2.1 Processing....................................................................................... 74 5.2.2 Characterization…………………………………………………………………………. 75 5.2.3 Summary and Discussion……………………………………………………………. 95

5.3 Polysiloxane with Nanographene Particles……………………………………………… 100

5.3.1 Processing…………………………………………………………………………………. 100

5.3.2 Characterization……………………………………………………………………….. 101

5.3.3 Summary and Discussion…………………………………………………………… 123

Chapter 6: Summary and Concluding Remarks……………………………………………………… 125

Page 9: Carbon Rich Polymer-Derived Ceramics: Processing

viii

Chapter 7: Suggested Future Work………………………………………………………………………..… 130

7.1 Time-of-Flight Secondary Ion Mass Spectrometry (ToF-SIMS)…………………. 130

7.2 Surface Area Analysis (BET)……………………………………………………………………… 130

7.3 Anodes for Li-ion Batteries……………………………………………………………………… 131

7.3.1 Battery Tape Preparation………………………………………………………….. 131

7.3.2 Battery Cell Assembly……………………………………………………………….. 132

7.4 Other Suggested Future Work………………………………………………………………… 133

Page 10: Carbon Rich Polymer-Derived Ceramics: Processing

ix

List of Figures

2.1 Categories of the main silicon based polymers for ceramics. [1]……………………………. 6

2.2 Organosilicon based polymers and their possible constituents of X, R1 and R2. [2]…… 7

2.3 Conversion of PDCs from synthesis to annealing with temperature range comparison.[2]……………………………………………………………………………………………………………. 10

2.4 Electrical conductivity measurements at room temperature as a function of pyrolysis temperature and carbon content.[3]…………………………………………………………………………. 12

2.5 Thermogravimetric analysis of AHPCS polymer and boron modified polycarbosilane ceramics with increasing amounts of boron. [4]………………………………………………………… 14

2.6 Processing of polycarbosilane boron modified ceramics with low and high boron content. [4]………………………………………………………………………………………………………………… 15

2.7 Gas separation membrane testing apparatus for He, N2, Ar, C2H6, and CO2

gases.[5]……………………………………………………………………………………………………………………… 17

3.1 BET isotherms of I through VI [6]…………………………………………………………………………… 29

4.1 Polycarbosilane samples with varying weight percent of DVB……………………………….. 31

4.2 Polycarbosilane crosslinked sample at 180°C (left), then pyrolyzed samples at 500, 700 and 800°C……………………………………………………………………………………………………………. 32

4.3 Heat treatment schedule of polycarbosilane samples with the final pyrolysis temperature being x= 300, 400, 500, 600, 700, or 800 °C…………………………………………… 32

4.4 Polycarbosilane samples at 180, 300, 400, 500, 600, 700, and 800 °C to demonstrate color change associate with pyrolysis temperatures…………………………………………………… 33

4.5 TGA with a heating rate of 10 °C/min in flowing nitrogen of polycarbosilane samples (SMP10) with 50wt% DVB pyrolyzed at different temperatures (180 to 800 °C) …………. 34

4.6 TGA with a heating rate of 10 °C/min in flowing nitrogen of polycarbosilane samples (SMP10) with 50wt% and 0% at pyrolysis temperatures ranging from 180 to 800 °C….. 35

4.7 TGA with a heating rate of 10 °C/min in flowing nitrogen of polycarbosilane samples (SMP10) with 0 to 60wt% DVB at pyrolysis temperatures ranging from 300 to 500 °C... 36

4.8 FTIR of polycarbosilane samples (SMP10) at 50wt% DVB with pyrolysis temperature of 180 to 800 °C…………………………………………………………………………………………………………. 37

Figure Page

Page 11: Carbon Rich Polymer-Derived Ceramics: Processing

x

4.9 Raman of polycarbosilane samples (SMP10) at 50 wt% DVB with 700 and 800 °C pyrolysis temperature…………………………………………………………………………………………………. 37

4.10 Raman of polycarbosilane sample (SMP10) at 0 wt% DVB with 800 °C pyrolysis temperature……………………………………………………………………………………………………………….. 38

4.11 XRD of polycarbosilane sample pyrolyzed at 800 °C…………………………………………….. 39

4.12 29Si Solid State NMR of SiOC sample depicting SiO4, SiCO3, SiO2C2, SiC4, and SiC3O. 40

4.13 29Si Solid State NMR of polycarbosilane (SMP10) sample at 500 °C pyrolysis temperature……………………………………………………………………………………………………………….. 41

4.14 29Si Solid State NMR of polycarbosilane (SMP10) sample at 500 °C pyrolysis temperature with DVB 50……………………………………………………………………………………………. 42

4.15 29Si Solid State NMR of polycarbosilane (SMP10) at 800 °C pyrolysis temperature……………………………………………………………………………………………………………….. 43

4.16 29Si Solid State NMR of polycarbosilane (SMP10) at 800 °C pyrolysis temperature with 50 wt% DVB………………………………………………………………………………………………………… 43

4.17 29Si Solid State NMR of polycarbosilane (SMP10) samples at 500 C pyrolysis temperature for DVB 0 and DVB 50 and 800 C pyrolysis temperature at DVB 0 and DVB 50……………………………………………………………………………………………………………………………….. 44

4.18 13C Solid State NMR of polycarbosilane (SMP10) at 800 °C pyrolysis temperature with 0 wt% DVB…………………………………………………………………………………………………………… 45

4.19 13C Solid State NMR of polycarbosilane (SMP10) at 500 °C pyrolysis temperature with 0 wt% DVB…………………………………………………………………………………………………………… 46

4.20 13C Solid State NMR of polycarbosilane (SMP10) at 500 °C pyrolysis temperature with 50 wt% DVB………………………………………………………………………………………………………… 46

4.21 13C Solid State NMR of polycarbosilane (SMP10) at 800 and 500 °C pyrolysis temperature with 0 and 50 wt% DVB…………………………………………………………………………. 47

4.22 13C Solid State NMR single pulse of polycarbosilane (SMP10) at 500 °C pyrolysis temperature with 50 wt% DVB……………………………………………………………………………………. 48

Figure 4.23 13C Solid State NMR single pulse of polycarbosilane (SMP10) at 800 °C pyrolysis temperature with 50 wt% DVB……………………………………………………………………… 48

Figure 4.24 13C Solid State NMR single pulse of polycarbosilane (SMP10) at 800 and 500 °C pyrolysis temperature with 50 wt% DVB for pyrolysis comparison with higher percent of carbon…………………………………………………………………………………………………………………….. 49

Page 12: Carbon Rich Polymer-Derived Ceramics: Processing

xi

4.25 Process of boron modification to AHPCS polymer with borondimethylsulfide and toluene to create a boron modified polymer.[7]………………………………………………………….. 52

4.26 Picture of apparatus set up for (a) mixing of toluene and BDMS/BDMA in an ice bath and then the drying step (b)…………………………………………………………………………........ 53

4.27 Diagram of AHPCS polymer modification to crosslinking to pyrolysis…………………… 54

4.28 (a) FTIR AHPCS B0.1 BDMA crosslinked at 60 °C (b) FTIR AHPCS B0.1 BDMS crosslinked at 60 °C…………………………………………………………………………………………………….. 55

4.29 (a) FTIR AHPCS B0.1 BDMA pyrolyzed at 500 °C (b) FTIR AHPCS B0.1 BDMS pyrolyzed at 500 °C……………………………………………………………………………………………………………………… 56

4.30 (a) FTIR AHPCS B0.1 BDMA pyrolyzed at 800 °C (b) FTIR AHPCS B0.1 BDMS pyrolyzed at 800 °C……………………………………………………………………………………………………………………… 56

4.31 Raman of (a) AHPCS BDMA 1000C, (b) AHPCS BDMS 1000C, (c) BDMS Raman 1300C, and (d) AHPCS BDMA 1300C……………………………………………………………………………. 57

4.32 XRD Comparison of B 0.1 BDMS vs BDMA at pyrolysis temperatures of 800, 1000 and 1300 °C………………………………………………………………………………………………………………… 59

5.1. Diagram of processing for MK and H44 polymer precursors to ceramers……………… 63

5.2. Heating steps for pyrolysis of polysiloxane samples……………………………………………… 64

5.3 TGA in nitrogen environment of MK samples pyrolyzed at 200 to 700 °C………………. 65

5.4 TGA in nitrogen environment of H44 samples pyrolyzed at 200 to 500 °C……………… 66

5.5 FTIR of MK samples pyrolyzed at 200-700 °C………………………………………………………… 67

5.6 Schematic of the chemical structure of MK contributing to FTIR peaks………………… 68

5.7 FTIR of H44 samples pyrolyzed at 200-500 °C………………………………………………………… 68

5.8 Schematic of the chemical structure of major groups in H44 contributing to the FTIR peaks…………………………………………………………………………………………………………………………. 69

5.9 Comparison of the FTIR spectra of MK and H44 samples crosslinked at 200 °C……… 69

5.10 Comparison of the FTIR spectra of MK and H44 samples pyrolyzed at 500 °C……… 69

5.11 Surface area measured using BET analysis of MK 200, 400, 500 and 700 samples. 70

5.12 Pore volume as a function of relative pressure for MK 200, 400, 500 and 700 samples……………………………………………………………………………………………………………………... 71

Page 13: Carbon Rich Polymer-Derived Ceramics: Processing

xii

5.13 BET Isotherm of MK 500…………………………………………………………………………………..…. 71

5.14 Comparison of SiOC crosslinked and pyrolyzed samples with a reference…….……… 75

5.15 Weight loss as a function of temperature of crosslinked samples dvbn; n = 0,1,2,3,4,5 and 6…………………………………………………………………………………………………………. 77

5.16 Derivative of the weight loss of crosslinked samples dvbn; n = 0,1,2,3,4,5 and 6…. 77

5.17 Percent weight loss vs amount of DVB from TGA of crosslinked samples for weight loss region 1 (50 to 150 °C) and weight loss region 2 (400 to 600 °C)…………………………… 79

5.18 ATR FTIR of crosslinked dvb1 sample at room temperature…………………………………. 80

5.19 ATR FTIR spectra of crosslinked dvb 0, 1 and 6 samples at room temperature……. 81

5.20 FTIR spectrum of dvb1 sample pyrolyzed at 800 °C…………………………………………….. 82

5.21 Raman of dvb1 pyrolyzed at temperatures of 700,800 and 900 °C………………………. 85

5.22 Raman spectra with best fits of dvb1 pyrolyzed at: (a) 700 °C, (b) 800 °C, and (c) 900 °C………………………………………………………………………………………………………………………………… 85

5.23 Raman of dvbn for n from 1 to 6 all pyrolyzed at 800 °C……………………………………… 87

5.24 XRD of 800 C dvb1………………………………………………………………………………………………. 88

5.25 XPS survey 800 C dvb1…………………………………………………………………………………….…… 89

5.26 XPS spectra of C 1s (a) 800C dvb1, (b) 800C dvb6, (c) 700C dvb1 and (d) 900C dvb1………………………………………………………………………………………………………………………….... 90

5.27 XPS spectra of Si 2p (a) 800C dvb1 and (b) 900C dvb1…………………………………………. 92

5.28 Electrical conductivity as a function of the temperature of sample 800C dvb1, 800C dvb6 and 900C dvb1……………………………………………………………………………………………….….. 94

5.29 800C dvb1 three different atmospheres………………………………………………………………. 95

5.30 Comparison of crosslinked polysiloxane with and without nanographene and with a reference coin.………………………………………………………………………………………………………….. 101

5.31 ATR-FTIR of crosslinked PHMS DVB1 and PHMS DVB1 with 1wt% graphene. …………………………………………………………………………………………………………………………………. 102

5.32 TGA of crosslinked DVB1 in nitrogen from section 5.2 compared to crosslinked PHMS DVB1 in helium…………………………………………………………………………………..…………… 103

5.33 TGA of crosslinked DVB6 in nitrogen from section 5.2 compared to crosslinked PHMS DVB6 in helium……………………………………………………………………………………………….. 104

Page 14: Carbon Rich Polymer-Derived Ceramics: Processing

xiii

5.34 TGA and DTG of crosslinked PHMS DVB1 and crosslinked PHMS DVB1-1wt%graphene with 10 °C/min heating rate from 45 to 1000 °C in helium……………….. 105

5.35 TGA with 10 °C heating rate from 45 to 1000 °C in helium of crosslinked PHMS DVB 0, PHMS DVB1, PHMS DVB1- 1wt% graphene, PHMS 1wt% graphene, and PHMS 5wt% graphene………………………………………………………………………………………………………………….. 106

5.36 Full scan mass spectrum at 533.4 °C of crosslinked PHMS dvb1…………………………. 108

5.37 Extracted ion chromatogram (XIC) of PHMS DVB1 with TG plot and associated XIC of components from Figure 5.36……………………………………………………………………………….. 109

5.38 3D plot of m/z vs time vs ion current of PHMS DVB1 for TGA-MS………………………. 109

5.39 Full scan mass spectrum at 533.5 °C of PHMS dvb1 1wt%g……………………………..… 110

5.40 XIC of PHMS DVB1 1wt% graphene with TG plot and associated XIC of components from Figure 5.39……………………………………………………………………………………………………….. 111

5.41 3D plot of m/z vs time vs ion current of PHMS DVB1 with 1wt% graphene for TGA-MS………………………………………………………………………………………………………………………..…… 111

5.42 Comparison XIC of PHMS DVB1 and PHMS DVB1 1wt% graphene TG plot and associated XIC of components from Figures 5.36 and 5.37………………………………………… 112

5.43 XIC of PHMS DVB 1 1wt% graphene TG plot and associated XIC to compare methyl (m/z=15), ethyl (29), propyl (43), butyl (57), pentyl (71) and hexyl (85) groups……….... 113

5.44 Full scan MS at 738.8 °C of PHMS DVB 0………………………………………………………….. 114

5.45 XIC of PHMS DVB0 with TG plot and associated XIC of components from Figure 5.44. ………………………………………………………………………………………………………………………………… 115

5.46 Full scan MS at 504.7 °C of PHMS DVB6…………………………………………………………… 116

5.47 XIC of PHMS DVB6 with TG plot and associated XIC from Figure 5.46………………. 116

5.48 3D plot of m/z vs time vs ion current of PHMS DVB6 for TGA-MS…………………….. 117

5.49 PHMS Full scan MS at 745.3 °C of PHMS 1wt% graphene……………………………….. 117

5.50 XIC of PHMS 1wt% graphene with TG plot and associated XIC from Figure 5.49.. 118

5.51 3D plot of m/z vs time vs ion current of PHMS with 1wt% graphene for TGA-MS. 118

5.52 Comparison XIC of PHMS with 1wt% and 5wt% graphene with TG plot and associated XIC of components from Figures 5.49………………………………………………………. 119

5.53 Raman spectrum of 900 °C with 1 wt% graphene with 0.1% laser power…………. 120

Page 15: Carbon Rich Polymer-Derived Ceramics: Processing

xiv

5.54 Raman of crosslinked PHMS dvb1 with 1 wt% of graphene………………………………. 121

5.55 Raman spectrum comparing pyrolysis temperature and graphene content of a) 700 dvb1 1wt% graphene and 700 dvb1 b) 800 C dvb1 with 1 wt% graphene and 800 C dvb1………………………………………………………………………………………………………………………….. 122

5.56 Raman spectrum of PHMS DVB1 with and without 1wt% graphene pyrolyzed at 700 and 800 °C…………………………………………………………………………………………………………….….. 123

6.1 Overview for Organizing Research……………………………………………………………………… 126

7.1 Diagram of battery assembly………………………………………………………………………..…….. 132

7.2 Picture of 8 channel battery analyzer…………………………………………………………..……… 133

Page 16: Carbon Rich Polymer-Derived Ceramics: Processing

xv

List of Tables

4.1 Elemental Analysis for oxygen and carbon percent of polycarbosilane samples (SMP10) at 500 °C pyrolysis temperature for with and without DVB……………… 50

4.2 Elemental analysis of O, N, and H using combustion technique……………………… 60

5.1 Correspondence between chemical bonds and FTIR peak location. [8]…………… 67

5.2 Correspondence between chemical bonds and FTIR peak location. [8]…………… 68

5.3 Designation of crosslinked and pyrolyzed samples………………………………………….. 74

5.4 From TGA data, t percent weight loss in weight loss region 1 (50 to 150 °C) and weight loss region 2 (400-600 °C) for crosslinked samples dvbn; n = 1 to 6……… 78

5.5 Results of Raman analysis in terms of D and G bands peak position, full width at half max (FWHM) and I(D)/I(G) ratio based on area (A) and height (H) of sample dvb1 pyrolyzed at 700 °C, 800 °C and 900 °C…………………………………………………… 86

5.6 Raman values of peak position for D and G, full width at half max (FWHM) and I(D)/I(G) ratio for area (A) and heigh (H) of dvbn samples pyrolyzed at 800 °C (n from1 to 6)……………………………………………………………………………………………………… 87

5.7 XPS binding energies (BE) of C1 s for samples 800C dvb1, 800C dvb6, 700C dvb1 and 900C dvb1………………………………………………………………………………………………… 90

5.8 Summary of the analysis of XPS high resolution scans of C 1s. Each scan set of results add up to 100 %........................................................................................ 91

Table Page

Page 17: Carbon Rich Polymer-Derived Ceramics: Processing

xvi

5.9 Summary of the analysis of XPS high resolution scans of Si 2p. For each scan the % add to 100 %. ……………………………………………………………………………………………… 92

Page 18: Carbon Rich Polymer-Derived Ceramics: Processing

1

Chapter 1

Introduction and Motivation

Ceramics are generally thought as the traditional materials such as porcelain,

bricks, concrete, but they have since advanced beyond these materials to encompassing

bio ceramics, dental implants, ceramic matrix composites (CMCs), electrical materials,

3D printed materials, iPhone covers, etc. Another relatively new area of ceramics in the

last 50 years is the area of Polymer Derived Ceramics (PDCs). This area has drastically

expanded the materials capable of creating ceramics because of the polymer precursors

which can be converted to ceramics. Polymer derived ceramics are typically processed

at temperatures of above 1000 °C. The polymeric nature of the precursors allows for a

wide range of compositions and different shaping techniques. The resultant ceramics

can be either dense or porous depending on the polymer precursor, additive, and

pyrolysis temperature which in turn greatly affects the applications. Applications such as

gas separation membranes and anodes for lithium-ion batteries are emerging

applications discussed in this research.

There is a pressing need for safe, high-capacity batteries. For liquid electrolyte Li-

ion batteries, one strategy to improve safety is to replace the graphitic anode with

something that is less reactive and less flammable. Several materials are being

investigated. One material being actively investigated is high carbon and silicon-based

Page 19: Carbon Rich Polymer-Derived Ceramics: Processing

2

polymer derived ceramics. This material is the focus of this research. To be used as

battery anodes, both a high electrical conductivity and high surface area are needed.

The focus of this research was to obtain a fundamental understanding of high

carbon polymer derived ceramics produced by low/intermediate temperature pyrolysis.

The focus is on the effect of the chemistry precursor and the pyrolysis temperature on

the chemical and microstructure of the resultant ceramic. Chapter two will be a

literature review of polymer derived ceramics. Chapter three will cover the materials

processing and characterization techniques used this work. Chapters four and five will

present the results and discussions of the polymer derived ceramics synthesized.

Specifically, chapter four will focus on polycarbosilane polymer derived ceramics with

added carbon and boron/ boron nitride. Chapter five is on the polysiloxane polymer

derived ceramics with added carbon and nanographene. Chapter six present preliminary

results on the polysiloxane materials with added carbon as a Li-ion battery anode

material. Chapter seven is a summary and conclusions of this work and chapter eight is

suggestions for future work.

Objectives

1. To study the chemistry and structure of low/intermediate temperature pyrolysis

polymer derived ceramics with high carbon as a function of precursor chemistry

and pyrolysis temperature.

Page 20: Carbon Rich Polymer-Derived Ceramics: Processing

3

2. Investigate the electrical and electrochemical properties of such materials and

their performance as a Li-ion battery anode material

Overarching Goal of the Research

• Fundamental understanding of ceramics obtained from pyrolysis of polymer

precursors at low/intermediate temperature.

Page 21: Carbon Rich Polymer-Derived Ceramics: Processing

4

Chapter 2

Literature Review

2.1 Polymer Derived Ceramics

Polymer derived ceramics (PDCs) are an important class of ceramics being widely

investigated because their compositions can be molecularly tailored over a broad range

and they can be processed using conventional polymer processing routes. These

attributes lead to the ability to make a wide composition of materials and material

forms. [1,2,9] PDCs are a unique class of materials in which a polymer (typically Si-based

thermoset) is used as a precursor which is subsequently converted to a ceramic by high-

temperature pyrolysis. [1]

PDCs have many other desirable characteristics, such as, high chemical

durability, high oxidation and creep resistance, and limited crystallization and phase

separation at high temperatures.[1] These desirable characteristics allow for a wide

spectrum of applications for PDCs: high temperature-resistance materials, ultra-high

temperature materials, chemical engineering materials (e.g. electrodes for batteries and

membranes for gas separation), semiconducting materials and functional materials in

micro/nanoelectronics (MEMS).[3,9,10]

Page 22: Carbon Rich Polymer-Derived Ceramics: Processing

5

PDCs are an important group of ceramics because they undergo microstructural

changes at low temperature (~1000 to 1500 °C) compared to traditional powder, in

which densification and microstructural changes occur at temperatures in the range of

1500-2000 °C (for Si-based ceramics). In traditional ceramics, the high processing

temperatures create crystalline SiC and Si3N4; however, PDCs are processed at a much

lower processing temperature leading to an amorphous or nanocrystalline ceramic.

Because of their high temperature requirements for sintering, conventional ceramics

can be difficult to process and are limited to only a few processing methods. By using

polymers to create PDCs a large variety of processing methods can be used. Some of

these methods include: sol-gel processing, polymer infiltration pyrolysis, injection

molding, coating from solvent, extrusion, or resin transfer molding.[11] Additive

manufacturing techniques, such as 3D printing, has also been a recent processing

technique of growing interest for PDCs.[12] An added benefit for this type of ceramic is

that the microstructure can easily be tailored based on three factors: polymer precursor

used, any additives to the polymer, and pyrolysis conditions (temperature and

atmosphere).[1] These are all extremely important for controlling the chemistry,

microstructure, and properties.

Page 23: Carbon Rich Polymer-Derived Ceramics: Processing

6

2.1.1 Polymer Precursors

Silicon based polymer precursors are one of the most investigated group of PDCs

and are based off the chemical formula [ Si-R1-R2-X]n. The majority of these materials

associated with PDCs are from silicon-based polymers such as polycarbosilanes,

polysilazanes, and polysiloxanes. [1]

Figure 2.1. Categories of the main silicon based polymers for ceramics. [1]

There are two parameters to adjust the chemistry of silicon based PDCs. First

changing the (X) group in the formula results in different classes of materials, as can be

seen in Figure 2.1. Common “X” elements are carbon, oxygen, nitrogen, and boron. The

second parameter is by changing the side groups of R1 and R2 as demonstrated in Figure

2.2. These are generally organic groups. By adjusting the R groups, it changes the

Page 24: Carbon Rich Polymer-Derived Ceramics: Processing

7

chemical, mechanical, electrical, and rheological properties of the PDC. [1,2]

Decomposition of organic side groups, such as methyl and phenyl occur during the

conversion process of the polymer to a ceramic (400-1300 °C). This decomposition of

organic side groups can help create a unique microstructure (e.g. materials with

hierarchical porosity).[13] They can also aid in improving mechanical properties

depending on the starting polymer precursors.

Figure 2.2 Organosilicon based polymers and their possible constituents of X, R1 and R2. [2]

Organosilicon polymers were first developed in the 1960s. Two organosilicon

polymers precursors were investigated in this research: polycarbosilane and

polysiloxane due to their desired properties and commercial availability.

Polycarbosilane are silicon-based polymer with carbon in the backbone of the

polymer (X = C with reference to Figure 2.2). The first important use of these polymers

was in their use for making SiC fibers in the 1970’s by Yajima [14–16]. The SiC fibers

created were a great improvement to the previous SiC whiskers and had high tensile

Page 25: Carbon Rich Polymer-Derived Ceramics: Processing

8

strength. Further research led to more complex shapes such as foams, coatings, and

tapes.[17] Polymer derived ceramic matrix composites were also created.[18,19]

Commercially available polymers became more prominent for applications such as

SMP10 chemically known as allyhydridopolycarbosilane (AHPCS).[3,19,20] As the

applications of ceramics using preceramic polymers expanded so did the types of

preceramic polymers being used.

The second class of polymer precursors to be discussed are polysiloxanes, which

is a silicon-based polymer characterized by the oxygen in the backbone of the polymer

structure (X= O with reference to Figure 2.2). Polysiloxanes have a high ceramic yield

because of the high crosslinking density in the precursor. [21] Polysiloxanes are of

significant interest because the ceramics derived from them are stable up to 1500 °C

and amorphous below 1200 °C, resulting in high mechanical and chemical stability. [21]

Adding excess carbon to polysiloxane can change a traditionally insulating material into

a conducting material. And they are growing in interest for their use in anodes for Li-ion

batteries.

The polymer precursor, heating rate, pyrolysis temperature and atmosphere

have a significant effect on the resultant ceramic materials and their properties and will

be discussed next.

Page 26: Carbon Rich Polymer-Derived Ceramics: Processing

9

2.1.2 Polymer to Ceramic Conversion

A crucial phase of polymer derived ceramics is their conversion from a polymer

to a ceramic which is done by crosslinking followed by pyrolysis of the material. Figure

2.3 schematically illustrates the polymer derived ceramics conversions. Synthesis and

shaping of the material are done before crosslinking. Crosslinking can be a separate step

before pyrolysis, using the preceramic polymer to form a “green body” ceramic in the

temperature range of approximately 200 to 400 °C. [2] Thermal crosslinking relies on

reactive groups in the polymer, for example Si-H. Radical initiators can be added to help

aid the crosslinking and lower the temperature. Recent studies have demonstrated

room temperature crosslinking with metal complexes.[22,23] UV crosslinking, reactive

gases or plasma have also been used to aid in crosslinking below 200 °C. Recently, PDCs

have been 3D printed by using laser to crosslink and create complex forms. [2,12]

At temperatures above 400 °C, the polymer decomposition starts leading to

mass loss and evolution of volatile species. For most polymers, the decomposition is

complete in the temperature range of 400 to 1000 °C. [1] The resultant ceramic,

produced by pyrolysis at low temperatures (below 1200 to 1500 °C) is amorphous.

Higher temperature heat treatment leads to crystallization of the amorphous

ceramic.[2]

Page 27: Carbon Rich Polymer-Derived Ceramics: Processing

10

Figure 2.3 Conversion of PDCs from synthesis to annealing with temperature range comparison. [2]

Pyrolysis temperature and atmosphere are crucial in controlling the

microstructure.[24,25] Heating rate can be crucial when creating polymer derived

ceramics, too fast and the ceramic is likely to crack due to rapid expansion of gases

evolving. And slow heating rates can produce microporous ceramics.[26] Atmosphere is

crucial as well, using an inert atmosphere allows for more stable and denser PDCs.[27]

Especially when using polysilazane, exposure to oxygen should be limited for a thermally

stable ceramic.[28,29] Depending on the polymer precursor, reactive atmosphere, can

change the surface characteristics, stability, and chemistry of the resultant

ceramic.[2,30]

Although there are many possibilities for PDCs, there is a downside in that large

volume shrinkage (50-75%) occurs.[1] Later studies resolved this problem with the use

of inert and active fillers, which, when added to the ceramic matrix, help to reduce

Page 28: Carbon Rich Polymer-Derived Ceramics: Processing

11

shrinkage.[31] These fillers can be used to create dense or low porosity ceramics.[5,32]

Other studies used the natural porosity of PDCs to their advantage for catalysts and gas

separation membranes [9]. PDCs without fillers typically have high porosity.

The crosslinking of this polymer can be done at relatively low temperatures (100-

400 °C) [21], which can be further lowered to room temperature by using crosslinking

agents [23,33]. Pyrolysis of polysiloxanes has been done typically in the temperature

range of 800 – 1500 °C, resulting in materials with high thermomechanical and chemical

stability, especially for materials pyrolyzed above 1000 °C [21,27,34,35] PDC coatings

made from particle filled polysiloxanes have been shown to provide excellent protection

in highly corrosive and oxidizing environments [36–38]. PDCs derived from polysiloxane

precursors at temperatures above 800 °C have been shown to be semi-conductor

materials. It has been shown that increasing pyrolysis temperature increases electrical

conductivity [39].

2.1.3 Polymer Derived Ceramics with Added Carbon or Boron

The addition of other chemical elements to the polymer network allows for

improvement and tailorability of key characteristics. [22,23] For example, adding excess

carbon, in adding carbon precursors such as divinyl benzene (DVB), to polycarbosilanes

leads to significant increase in the electrical conductivity for potential use in micro

electrical mechanical systems (MEMS) as seen in Figure 2.4 [3].

Page 29: Carbon Rich Polymer-Derived Ceramics: Processing

12

Figure 2.4 Electrical conductivity measurements at room temperature as a function of pyrolysis temperature and carbon content.[3]

Incorporating DVB into the system results in the formation of a “free carbon”

phase [23,35]which can be used for tailoring the electrical properties of the resultant

PDC. [13,40–43] The addition of excess carbon can aid in the thermal stability of the

material at high temperatures and limits corrosion, grain growth, creep, in addition to

tailoring electrical properties. The mechanical properties of polysiloxane derived PDC

with excess carbon have also been evaluated and it has been shown that both the

elastic modulus and the hardness decreases as the excess carbon increases. [44] The use

of a liner polymer such as polymethylhydrosiloxane (PHMS) compared to a cyclic one

helped to improve the chemical stability, when incorporating DVB. [45]

Other types of carbon forms that have been incorporated are graphene and

graphene oxides in polysiloxanes to increase electrical conductivity. [46,47]

Page 30: Carbon Rich Polymer-Derived Ceramics: Processing

13

Polysiloxanes when pyrolyzed form SiOC at certain temperatures and are intrinsically

insulating materials, but when incorporating different carbon forms, they can semi-

conducting and conducting. This can aid in coatings especially for uses in

electromagnetic shielding.[36]

Overall, added carbon can affect morphology, thermal stability, surface

characteristics, tribology, electrical properties, mechanical stability, EM shielding

abilities, and optical properties of PDCs.[2] Pyrolysis of these materials is key depending

on their applications. High temperature above 1000 °C, have been extensively studied

and can be limited on porosity. Low to intermediate temperature “hybrid” PDCs or

“ceramers” are a more recent class of materials and provide a wide variety of

applications, different from their high temperature counterparts.

Addition of other elements have also been investigated. The inclusion of boron

can increase the temperature stability of the amorphous PDCs to 2200 °C.[20,41,48,49]

The addition of boron into polycarbosilane polymer AHPCS greatly improved thermal

stability as shown in Figure 2.5.

Page 31: Carbon Rich Polymer-Derived Ceramics: Processing

14

Figure 2.5 Thermogravimetric analysis of AHPCS polymer and boron modified polycarbosilane ceramics with increasing amounts of boron. [4]

Because of this high thermal stability, boron with nitrogen additions were also

used to create SiCBN fibers at high temperatures.[50] Nitrogen additions in either the

form of polysilazane polymers, added amine groups or ammonia atmosphere were

utilized. Nitrogen allowed for a change in microstructure by producing mesoporous

boron PDCs at high temperatures.[51] And nitrogen aided in extrusion properties to

allow for the production of fibers for boron containing PDCs.

Adjusting the amount of boron not only aids in thermal stability but can create

dense or porous Si-(B)-C ceramics. Low boron content ceramics, create porous foams

with template removal processes whereas high boron content can be warm pressed and

produce dense ceramics as demonstrated in Figure 2.6. Higher boron content allows for

Page 32: Carbon Rich Polymer-Derived Ceramics: Processing

15

denser ceramics at lower pressing temperatures without sintering aids. Traditional

ceramics are processed at above 2000 °C, with higher boron content ceramics they can

be sintered at 1700 to 1800 °C without the use of a sintering aid and improved

mechanical and electrical properties.[20] By incorporating boron into PDCs, tailorable

applications can be further investigated for different types of sensors, foams, or fibers.

Figure 2.6 Processing of polycarbosilane boron modified ceramics with low and high boron content. [4]

2.2 Polymer Derived Ceramics by Pyrolysis at Low/Intermediate Temperatures

Although PDCs obtained at pyrolysis temperature higher than 1000 °C have been

widely investigated, limited research has been reported on materials processed by

pyrolyzing polysiloxanes at low/intermediate temperature (within 300-1000 °C). [52]

Materials formed by pyrolysis at the lower end of this temperature have been called

“ceramers” [53], and it has been shown that these materials have high surface area,

Page 33: Carbon Rich Polymer-Derived Ceramics: Processing

16

tailorable porosity and the ability to incorporate different catalytic elements such as

platinum and zinc [53–56]. Although amorphous, these materials retain many desired

properties of PDCs including mechanical, chemical, and thermal stability. Ceramers are

hybrid inorganic-organic materials that start as polymers and are partially converted to

a ceramic in the range of 300-700 °C.[53] This relatively low temperature range for

conversion results in a highly microporous material. The microstructure of ceramers is

also tailorable based on polymer precursor and fillers.[57] There are two types of fillers:

active and passive. In addition, by using blowing agents, “foamed ceramers” can be

created. These foamed ceramers have been investigated as substrates for catalytic

membranes.[55] Unlike conventional monolithic catalysts, these improved membranes

eliminated the problem of pore plugging.[53] This issue was resolved because the new

material allowed the hydrocarbons to freely absorb and release through the material.

The ceramer’s catalytic activity was also improved with the incorporation of platinum

and zinc nanoparticles.[54,56,58,59] To further tailor porosity and make hierarchical

porous structures, other processing methods such as freeze casting have been

developed. [60,61] Further studies changed the surface characteristics of the materials

to enhance membrane properties.[57,62] For example, with the addition of amine

groups to the material the hydrophobicity was increased.[63] A commonly used

precursor for ceramers is polysiloxane because of its tailorability and wide range of

porosity that can be realized. [9,56,64,65] Both macro and micro pores have been

realized in the same PDC to allow for fluid or gas transport of various sizes.[26] Purely

Page 34: Carbon Rich Polymer-Derived Ceramics: Processing

17

microporous ceramics can create issues with pressure drops, but by including both

micro and macro sized pores gas transport can be optimized.

Lower temperature polymer derived ceramics with high surface area properties

have been investigated for many different applications i.e. gas separation

membranes[5,66], catalytic converters, battery applications[13,67–69].

Gas separation membranes are a large area of interest for PDCs because of the

ability to tailor the pores and the thermal stability of ceramics. PDCs allow for a wide

variety of gases to be separated as shown in Figure 2.7. [5]

Figure 2.7 Gas separation membrane testing apparatus for He, N2, Ar, C2H6, and CO2 gases. [5]

Another potential application of these materials is in Li-ion batteries. Li-ion

batteries have been growing need for higher energy and power especially for their use

Page 35: Carbon Rich Polymer-Derived Ceramics: Processing

18

in electric vehicles. For anodes, graphite is the current choice for liquid electrolyte Li-ion

batteries. However, it suffers from two setbacks: safety and low capacity (theoretical

capacity being 372 mAhg-1). This makes the introduction of silicon more desirable

because of its high capacity of 3579mAhg-1. [69] Silicon has several drawbacks though,

one being its large volume changes. Combining silicon in another form is a growing field

of interest by using PDCs with silicon-based polymers in batteries. Silicon oxycarbides

(SiOC) amorphous ceramics have been of recent interest for anodes for Li-ion batteries.

[69–71]

PDCs with the ability to incorporate free carbon into the structure is the most

recent area of interest because of the ability to tailor the amount of carbon in the SiOC

matrix. Graphite incorporated with carbon-rich SiOC materials can provide a higher

capacity of 520 mAhg-1 compared to 372mAhg-1 of just graphite.[67] Nanoscale graphite

was incorporated into the SiOC at temperatures of 1100 °C with promising results.[69]

Previously materials at and above 1000 °C were investigated but limited research

has been conducted in the area of pyrolysis temperatures lower than 1000 °C. This is

the focus of this research: polymer derived ceramics with high surface area and high

carbon content.

Page 36: Carbon Rich Polymer-Derived Ceramics: Processing

19

2.3 Focus of this Dissertation: Low/Intermediate Pyrolysis Polymer Derived Ceramics

with High Surface Area and High Carbon Content

This section is the proposed area of research that is novel to the field due to the

combination of low/intermediate temperature polymer derived ceramics with added

carbon.[72] With the growing need for improved batteries, an emphasis on materials

with high surface area and high conductivity are needed. The overall goal is to create

and investigate materials with low/intermediate pyrolysis temperature to create a

higher surface area and combine it with high carbon materials to increase the

conductivity. The effect of pyrolysis temperature and precursor composition was

investigated on the chemical structure and microstructure. Techniques such as Fourier

transform infrared spectroscopy (FTIR), thermogravimetric analysis (TGA),

thermogravimetric analysis coupled with mass spectrometry (TGA-MS), Raman

spectroscopy, x-ray diffraction (XRD), x-ray photoelectron spectroscopy (XPS), solid state

nuclear magnetic resonance (NMR), elemental analysis, BET surface area analysis, and

time of flight secondary ion mass spectrometry (ToF-SIMS) were used to investigate the

chemical- and micro-structure. Electrochemical impedance spectroscopy (EIS) was used

to investigate the electrochemical properties. Half cells were assembled, cyclic

voltammetry (CV) and charge-discharge were characterized to examine the created

materials’ performance as anodes for Li-ion batteries as a proof of concept for this

application.

Page 37: Carbon Rich Polymer-Derived Ceramics: Processing

20

Chapter 3

Materials Processing and Characterization Techniques

3.1 Materials Processing

3.1.1 Preceramic Polymers

The preceramic polymers that were used are commercially available polysiloxane

and polycarbosilane.

Two different modifications were done to a polycarbosilane, specifically,

allyhydridopolycarbosilane (AHPCS). The first modification was the addition of DVB and

the second the addition of boron and boron nitride. Polycarbosilane with boron and

boron nitride modification were produced based on previous methods of Dr. Bernard’s

group in Limoges, France and described in detail in section 4.2.[48] Chapter 4 will go

more into detail of the processing of polycarbosilane samples.

For polysiloxane samples, three types of materials were investigated: (i)

polysiloxane resins MK (polymethylsiloxane, Silres ® MK, Wacker AG) and H44

(polymethylphenylsiloxane, Silres ® H44 ,Wacker AG) were without modifications; (ii)

commercially available polymethylhydrosiloxane (PHMS, 98–100 %, MW: 2100–2400,

Gelest, Morrisville, PA) was combined with divinylbenzene (DVB, technical grade 80 %,

Sigma Aldrich, St. Louis, MO) to investigate polysiloxane with increased carbon content;

(iii) incorporation of nanographene (ACS Material, LLC, Pasadena, CA) in PHMS.

Page 38: Carbon Rich Polymer-Derived Ceramics: Processing

21

3.1.2 Crosslinking

Polycarbosilane was crosslinked using two different methods. Method one was

combining commercially available AHPCS (SMP10, Starfire Systems, Inc., USA) with 1

wt% 1,1 azobis(cyclohexanecarbonitrile) (ACHCN) as a radical initiator and from 0 to 60

wt% divinylbenzene. The mixture was stirred for several minutes (~5 mins) then added

dropwise into circular Teflon molds and crosslinked at 180 °C for six hours. The samples

were then pyrolyzed in an alumina tube furnace in ultrahigh purity argon.

Method two was conducted during a research visit to Prof. Bernard’s’ Institute in

Limoges, France. Schlenk line techniques were utilized with SMP10 and boron additives:

borane dimethylsulfide (BH3.SMe2, BDMS) or borane dimethylamine (BDMA).[4] Chapter

four goes into more detail on the processing and crosslinking techniques.

Polysiloxane was also crosslinked with two different methods. In the first

method, MK (polymethylsiloxane) was added with crosslinking agent triethanolamine

(TEA) in isopropanol. To create a foamed ceramer, an aluminum catalyst (aluminum

acetylacetate) was added to H44 (polymethylphenylsiloxane) in acetone, dried at 60 °C,

then a blowing agent powder (azodicarbonamide) was added. All samples were

crosslinked at 200 °C with an air atmosphere in an oven.

The second method for crosslinking polysiloxane samples (PHMS) was done at

room temperature with platinum catalyst (platinum divinyltetramethyldisiloxane 2% in

xylene), this was for the incorporation of DVB and nanographene. The crosslinked

Page 39: Carbon Rich Polymer-Derived Ceramics: Processing

22

samples are labeled as: PHMS DVBn (n referring to the amount of DVB: PHMS (weight

ratio). PHMS, DVB and Pt complex were stirred for two minutes then thin

(approximately 0.5 mm thickness) polymer films were cast and crosslinked at room

temperature. 10 mm diameter disks were cut out of the crosslinked film which were

subsequently pyrolyzed. Samples with nanographene were with either 1 or 5wt%

nanographene and PHMS. Samples with DVB1 and 1wt% graphene were also crosslinked

utilizing the same method.

3.1.3 Pyrolysis

All polycarbosilane samples were pyrolyzed in ultrahigh purity argon (Airgas,

Radnor, PA).

Polycarbosilane samples with DVB they were pyrolyzed at 2 °C/min heating rate

to the isothermal temperature followed by a two-hour hold at the desired temperature

(300, 400, 500, 600, 700, 800 and 900 °C) and a 2 °C/min cooling rate (in an alumina

tube Lindberg/Blue Furnace).

Polycarbosilane samples with boron and boron-nitride were pyrolyzed at 800 to

1300 °C in argon (in an alumina tube horizontal resistive furnace, Carbolite,

BGHA12/450B type) in Limoges, France at 5 °C/ min. They were first placed in an

alumina boat in an argon glovebox then transferred into the furnace, which was then

evacuated under vacuum then refilled with argon.

Page 40: Carbon Rich Polymer-Derived Ceramics: Processing

23

Polysiloxane MK or H44 were pyrolyzed at temperatures ranging from 300 to 700

°C at one hundred-degree increments in argon. MK samples were ground to

approximately 125µm with mortal, pestle, and then run through a sieve.

Polysiloxane samples with carbon additives (DVB and graphene) were pyrolyzed

in ultrahigh purity argon with 1 °C/min heating rate to the isothermal temperature

followed by a two-hour hold at the desired temperature (700, 800 and 900 °C) and a 1

°C/min cooling rate (in an alumina tube Lindberg/Blue Furnace). A slower heating rate of

0.1 °C/min was also used to pyrolyze samples at 600 and 800 °C with the same cooling

rate in the same furnace with ultrahigh purity argon.

3.2 Characterization Techniques

3.2.1 Thermogravimetric Analysis (TGA, dTGA) and TGA Coupled with Mass

Spectrometry (TGA-MS)

Thermogravimetric analysis (TGA) (TGA Q500; TA Instruments, New Castle, DE)

was used to determine mass loss, as a function of temperature, of crosslinked polymers

and pyrolyzed samples. It helped to determine thermal stability of the material as well.

TGA was carried out in flowing nitrogen at a heating rate of 10 °C/min up to 1000 °C.

dTGA was used to locate high mass loss regions by taking the differential of TGA data.

Page 41: Carbon Rich Polymer-Derived Ceramics: Processing

24

TGA-MS was used as a complimentary technique to determine the volatile

compounds specific temperature ranges. It was conducted with collaboration with the

Brick Center at Clemson University. The samples were heated treated in flowing helium

at a heating rate of 10 °C/min up to 1000 °C. The MS part utilized scanning mode to

determine all volatile species that evolved during pyrolysis. The instrument used for

TGA-MS was a a Netzsch TG couple to a 403 Aëolos Quadro quadrupole mass

spectrometer with Proteus® software.

3.2.2 X-Ray Diffraction (XRD)

X-ray diffraction (Bruker AXS D8 Discover, CuKα ) was conducted in Limoges,

France to determine crystalline phases of the polymer derived polycarbosilane ceramics

with boron and boron-nitride components. The scans were performed in the range of 2θ

〈20°; 90°〉 with a step of 0.015° and an exposure time of 0.7 s. [20] Since they were

the only samples that had been heat treated to a high enough temperature, XRD was

conducted on only these samples.

Page 42: Carbon Rich Polymer-Derived Ceramics: Processing

25

3.2.3 Chemical Analysis : ATR FTIR, Raman Spectroscopy, XPS, and Elemental Analysis

ATR FTIR

Chemical bonding states of the pyrolyzed and crosslinked samples were

determined using attenuated total reflectance Fourier transform infrared spectroscopy

(ATR FTIR) (Thermo Nicolet Magna 550 FTIR, ThermoFisher, Waltham, MA) with a

diamond crystal. Samples were manually ground using a mortar and pestle then

scanned in the 4000 to 525 cm-1 range.

In Limoges, France FTIR experiments were conducted with a FTIR Nexus

(ThermoFisher) from 4000 to 400cm-1 range in transmission mode using KBr pellets. The

pellets were prepared by grinding samples of interest and then adding 5 wt% potassium

bromide in an argon glovebox. Then the samples were pressed outside the glovebox.

Liquid polymer samples were analyzed in KBr windows. The results were indexed using

published results. [8]

Raman Spectroscopy

Raman spectroscopy was conducted in collaboration with Prof. Jacobsohn for

samples produced at Clemson University. Raman spectroscopy (Horiba LabRAM HR

Evolution Raman; Horiba, Kyoto, Japan) was used to investigate the nature of the

carbon-carbon bond. A 532 nm laser source was used at 1% power together with a 600

Page 43: Carbon Rich Polymer-Derived Ceramics: Processing

26

grooves/mm diffraction grating over a scanning range of 3000 to 500 cm-1 with a 50x

objective lens. The spectra were corrected by the pre-recorded instrument-specific

response to a calibrated white light source, namely the intensity correction system (ICS).

Peaks were normalized by the 1600 cm-1 “G peak”. The 1330cm-1 “D peak” was fit with a

Lorentzian function while for the “G peak” a Breit-Wigner-Fano function (BWF) was used

consistent with previous research on amorphous carbons. [73]

For polycarbosilane samples with BDMS and BDMA samples only, the Raman

spectra of the powders were recorded at room temperature using an InVia reflex Raman

spectrometer (Renishaw, United Kingdom) equipped with a laser diode (λ =785 nm using

a 100× objective lens and calibrated to pure Si. The spectra were obtained in the range

of 900–2000 cm−1 in Limoges, France.

X-Ray Photoelectron Spectroscopy (XPS)

X-ray photoelectron spectroscopy (XPS) measurements were performed to

determine the relative ratios of three key elements: Si, C and O and to investigate the

chemical state of these elements. Spectra were obtained using a PHI VersaProbe III

Scanning ESCA Microprobe (Physical Electronics Inc, Chanhassen, MN). The PHI

VersaProbe III has a monochromatic Al KAl Kα Xray source (hν = 1486.6 eV). The

instrument base pressure was maintained lower than 1X10-7 torr. The pyrolyzed samples

(approximately 8 mm diameter) were mounted on a stage using nonconductive tape.

Page 44: Carbon Rich Polymer-Derived Ceramics: Processing

27

The XPS beam was at a 45° angle. Carbon (1s) peaks were investigated and

deconvoluted to determine binding energies (BE). A survey and high-resolution spectra

were run for each sample. Elements responsible for less than 1% of the sample

composition were excluded. For both survey and high-resolution spectra, the Al anode

was powered at 25 W and 15 kV. The survey was conducted using three sweeps (0 to

1100 eV), pass energy of 224 eV, dwell time of 50 ms and a step size of 0.8 eV to obtain

qualitative elemental analysis. The high-resolution scan was conducted over a narrow

energy range around the C(1s) using three sweeps, pass energy of 69 eV and a step size

of 0.125 eV. The scans were run over an area of 500 x 500 µm2 with a 100 µm diameter

beam. Charge compensation of samples was accomplished with an electron beam

charge neutralization. A sputtering step was used after initial data collection with 2 Kev

and 30 s of argon sputtering for the removal of adventitious carbon and oxygen

contamination. Peak positions were calibrated to 284.8 eV for carbon for analysis. Data

analysis was done using MultiPak (Physical Electronics Inc, Chanhassen, MN) and the

elemental peaks were deconvoluted to find the binding energies using CASA XPS

software (Casaxps.com, Berlin, Germany) with a Shirley background subtraction while

adopting Gaussian-Lorentzian peak shapes (GL(60)), and (FWHM) of 1.2 cm. XPS high

resolution spectra were acquired for Si, O and C species. Surface and “bulk” analysis was

compared. The “bulk” material was obtained after sputtering.

Page 45: Carbon Rich Polymer-Derived Ceramics: Processing

28

Elemental Analysis

Elemental analysis was utilized to determine the content of oxygen, hydrogen,

nitrogen, and carbon of the pyrolyzed samples. Specifically, samples of polycarbosilane

with BDMS and BDMA additive samples were collected using combustion techniques in

collaboration with the laboratory in Limoges, France. Quantitative results were obtained

based on combustion analysis with two different gases. The carbon content of powders

was determined by combustion elemental analysis (Carbon analyzer, EMIA-321 V,

Horiba, Japan) and the oxygen, nitrogen and hydrogen contents of powders were

measured using an EMGA-830 analyzer (Horiba, Japan).

3.2.4 Solid State NMR

Solid State NMR was conducted for 29Si and 13C at University of Koblenz in

Germany. Solid State NMR was utilized to determine chemical bonding information for

ceramic samples using 29Si and 13C.

3.2.5 BET Surface Area Analysis

Brunauer-Emmett-Teller (BET) was utilized to determine specific surface area of

the samples and micropore analysis, based on gas adsorption properties. Preliminary

BET Surface area analysis was done at Clemson University at the Advanced Materials

Page 46: Carbon Rich Polymer-Derived Ceramics: Processing

29

Research Laboratory (AMRL) for polysiloxane samples. Samples were first degassed at

150 °C for 10 to 12 hours. Then run for specific surface area measurements (SSA) and

micropore analysis all under nitrogen environment at 77K using in a Quantachrome

instrument. Isotherms were used for selected samples and interpreted based on Figure

3.1.

Figure 3.1 BET isotherms of I through VI. [6]

3.2.6 Electrochemical Impedance Spectroscopy (EIS)

Electrical conductivity measurements were done in collaboration with Prof.

Tong. EIS measurements were conducted for polysiloxane samples with increased

carbon. Specifically, for conductivity measurements using EIS, silver paste was applied

on both sides of the sample (~0.6 mm thickness and ~ 8 mm diameter) by screen

Page 47: Carbon Rich Polymer-Derived Ceramics: Processing

30

printing method and then dried on a hot plate at 150 °C, which worked as current

electrodes with attachment of silver mesh and gold leading wires. The electrochemical

impedance spectroscopy (EIS) (Gamry Reference 600+) of the pellets was conducted

with 10 mV amplitude and frequency a range of 5 MHz to 0.05 Hz under a dry argon

atmosphere. The samples were then heated in a Mellen Microtherm Spilt Tube furnace

from room temperature to 550 °C with a 5 °C/min heating rate while taking resistance

measurements at 50 °C intervals after a 1.5-hour hold time at each temperature point.

The EIS results were analyzed by fitting the data with ZView software to obtain the value

the intercept of ZReal-axis in the Nyquist curve which refers to the resistance (R). Thus

was converted to conductivity using Equation 1, in which l and A refer to the thickness

and the area of the pellet, respectively.

𝜎 =𝑙

𝑅 ∗ 𝐴

(Eq. 1)

Page 48: Carbon Rich Polymer-Derived Ceramics: Processing

31

Chapter 4:

Results and Discussion on Materials Processing and Characterization for

Polycarbosilane

4.1 Polycarbosilane with Carbon Addition

4.1.1 Processing

AHPCS (SMP10) was combined with DVB.[3] 1,1’-Azobis(cyclohexanecarbonitrile)

was added as a radical initiator at 1wt% to the polymer precursor. The amounts of DVB

ranged from 0 to 60 wt%. The polymer mixture was stirred for 5 minutes or until the

radial initiator was dissolved. The samples were crosslinked at 180 °C in argon. Figure

4.1 shows the image of the crosslinked samples as the amount of DVB increases, the

samples become less flexible. In addition, the tendency to form bubbles was reduced.

Figure 4.1 Polycarbosilane samples with varying weight percent of DVB.

The crosslinked samples were pyrolyzed at different temperatures and their images are

shown in Figure 4.2.

10wt% 20wt% 30wt% 40wt% 50wt% 60wt%

Page 49: Carbon Rich Polymer-Derived Ceramics: Processing

32

Figure 4.2 Polycarbosilane crosslinked sample at 180°C (left), then pyrolyzed samples at 500, 700 and 800°C.

The various temperatures of pyrolysis were from 300 to 800°C in 100°C

increments. The pyrolysis was done in argon with a 2 °C/min heating and cooling rate

and a 2-hour hold at final pyrolysis temperature. The complete heat treatment cycle,

including crosslinking is shown in Figure 4.3.

Figure 4.3 Heat treatment schedule of polycarbosilane samples with the final pyrolysis temperature being x= 300, 400, 500, 600, 700, or 800 °C.

Figure 4.4 shows the color change as a function of pyrolysis temperature for

polycarbosilane.

180°C 500°C 700°C 800°C

X 9027

X = 300, 400,

500, 600, 700,

800

Page 50: Carbon Rich Polymer-Derived Ceramics: Processing

33

Figure 4.4 Polycarbosilane samples at 180, 300, 400, 500, 600, 700, and 800 °C to demonstrate

color change associate with pyrolysis temperatures.

4.1.2 Characterization

TGA

Thermogravimetric analysis (TGA) was utilized to determine the thermal stability

and observe weight changes depending on chemical composition and temperature. All

samples were investigated at a 10 °C/min heating rate in an inert atmosphere of

nitrogen up to 1000 °C. Figure 4.5 demonstrates the change in mass of the material as a

function of pyrolysis temperature. Each sample is the same polymer precursor and

amount of DVB but has been pyrolyzed , before TGA at a different temperature. The first

samples SMP10 DVB50 180, experience the most weight loss and is the least thermally

sample as it is the crosslinked sample at 180 °C. It’s TGA is similar to that of SMP10

polymer.[3] Samples crosslinked and pyrolyzed at 180 and 300 respectively, have similar

weight loss regions, one before 200 °C and after 400 °C along with similar weight loss

percent at 16% and 15% respectively. This weight loss percent is less than that for the

polymer (SMP10), which is approximately 20 to 25% due to the increased

crosslinking.[3,4] The samples thermal stability increases for the sample pyrolyzed at

400 °C, which is expected since it is above the weight loss region of the first two

Page 51: Carbon Rich Polymer-Derived Ceramics: Processing

34

samples. As the samples’ pyrolysis temperature increase the thermal stability increases

and the weight loss percent decreases until the samples pyrolyzed at 800 °C are

thermally stable, where less than one percent of the mass is lost up to 1000 °C.

Figure 4.5 TGA with a heating rate of 10 °C/min in flowing nitrogen of polycarbosilane samples

(SMP10) with 50wt% DVB pyrolyzed at different temperatures (180 to 800 °C).

In order to investigate the effect of DVB on thermal stability, the thermal

stability of samples with 0% DVB were compared with those with 50% DVB, both set of

samples pyrolyzed at the same temperatures. Figure 4.6 compares DVB0 with DVB50 for

samples with the same temperatures for crosslinking and pyrolysis as Figure 4.5.

Initially, DVB50 has the greater weight loss compared to DVB0 although only a 3%

difference is observed at 180 °C. The samples follow similar thermal stability above 400

°C and almost negligible difference of weight loss at 600 °C pyrolyzed temperature. This

is due to the weight loss being most likely water and other solvents.

Page 52: Carbon Rich Polymer-Derived Ceramics: Processing

35

Figure 4.6 TGA with a heating rate of 10 °C/min in flowing nitrogen of polycarbosilane samples

(SMP10) with 50wt% and 0% at pyrolysis temperatures ranging from 180 to 800 °C.

The effect of DVB content on thermal stability is shown in Figure 4.7 Comparing

the range of 10 weight percent to 60 of DVB demonstrates that with increasing weight

percent of DVB there is a larger weight loss in Figure 4.7. This is evident at lower

pyrolysis temperatures at 300 °C, as the pyrolysis temperature increases this is less

crucial, the weight loss percent difference is less than one percent between different

percentages of DVB. This trend continues with higher pyrolysis temperature especially

those above 500 °C, as the material becomes thermally stable. The next section of FTIR,

compares the pyrolysis temperature with the observant chemical bonds.

Page 53: Carbon Rich Polymer-Derived Ceramics: Processing

36

Figure 4.7 TGA with a heating rate of 10 °C/min in flowing nitrogen of polycarbosilane samples

(SMP10) with 0 to 60wt% DVB at pyrolysis temperatures ranging from 300 to 500 °C.

FTIR

FTIR was conducted to observe the nature of the chemical bonds in the

crosslinked and pyrolyzed samples. As demonstrated in Figure 4.8, as pyrolysis

temperature increased intensity for bonding decreased, especially for Si-H and C-H

bonding.[3] The crosslinking mechanism is confirmed by the Si-H bond decreases and

the existence of C=C as previously found in literature, two crosslinking mechanisms

occur, hydrosilyation and DVB polymerization as the amount of DVB is increased. As

pyrolysis temperature increases beyond 500 °C, Si-C bonds dominate. Eventually only Si-

C bonding is left at 800 °C. This can be correlated back to the TGA in the previous

section, as the first major weight loss region is around 500 °C and continues until it is

thermal stable at 800 °C.

Page 54: Carbon Rich Polymer-Derived Ceramics: Processing

37

Figure 4.8 FTIR of polycarbosilane samples (SMP10) at 50wt% DVB with pyrolysis temperature of 180 to 800 °C.

Raman Spectroscopy

Raman was utilized to help determine the carbon-carbon bonding. However,

samples pyrolyzed below 800 °C fluoresced and due to the nature of the hydrogen

bonding, peaks associated with carbon were not present in Figure 4.9.

Figure 4.9 Raman of polycarbosilane samples (SMP10) at 50 wt% DVB with 700 and 800 °C

pyrolysis temperature.

C-H

Si-H

Si-H

C-H2

C=C

Si-CH2-Si

Page 55: Carbon Rich Polymer-Derived Ceramics: Processing

38

Samples pyrolyzed at 800 °C has clear D and G peaks associated with graphitic

carbon, as demonstrated in Figure 4.9 and 4.10. As the pyrolysis temperature is

increased, these peaks will sharpen, and subsequent peaks will be present as 2D and D’.

It is clear from the two figures that graphitic carbon is present in both samples with and

without DVB.

Figure 4.10 Raman of polycarbosilane sample (SMP10) at 0 wt% DVB with 800 °C pyrolysis

temperature.

XRD

XRD was used to identify crystalline phases. However, all the samples were

amorphous. The XRD patter for SMP10 heated treated at 800 °C is shown in Figure 4.11.

D G

Page 56: Carbon Rich Polymer-Derived Ceramics: Processing

39

Figure 4.11 XRD of polycarbosilane sample pyrolyzed at 800 °C.

Solid State NMR

Solid state NMR was conducted at the University of Koblenz-Landau in Koblenz,

Germany in collaboration with Prof. Quirmbach. Both silicon (29Si) and carbon (13C) solid

state NMR were utilized to compare the bonding mechanisms based on the chemical

shifts in ceramics at pyrolyzed temperatures of 500 and 800 °C. The effect of DVB was

also investigated. A 500 JEOL instrument with 8kHz for solid state NMR was utilized for

both 29Si and 13C along with single pulse, magic angle spinning (MAS) and cross

polarization (CP) variation with a differing number of scans and time per scan depending

on the sample. Solid State NMR was conducted on crushed ceramic powders using 29Si

and 13C Solid State NMR, 1H NMR was not investigated due to the assumption that the

majority of the hydrogen bonds would have decomposed due to pyrolysis. One

downside to the 29Si Solid State NMR, is the long time period of testing due to the long

relaxation times associated with silicon. For reference of different peak positioning for

Page 57: Carbon Rich Polymer-Derived Ceramics: Processing

40

chemical shifts Figure 4.12 from literature was used. In addition to Si and C bonds, Si-O

and C-O bonds were also evaluated since there is almost always some O2 contamination

during processing.

Figure 4.12 29Si Solid State NMR of SiOC sample depicting SiO4, SiCO3, SiO2C2, SiC4, and SiC3O.[35]

Figures 4.13 and 4.14 show the 29 Si NMR spectra SMP10 and SMP10 with 50wt%

DVB pyrolyzed at 500 °C, respectively. Five major peaks were found, the chemical shift

of these peaks in ppm are reported in literature based on SiOC ceramics as SiO4, SiO3C,

SiO2C2, SiOC3, and SiC4 at approximately -100, -70, -30, 5, and -10 ppm respectively

according to Figure 4.12 and in literature.[30,35,74] The major peak is the SiO4 peak at

Page 58: Carbon Rich Polymer-Derived Ceramics: Processing

41

approximately -100 ppm, although very broad due to its amorphous nature and could

also include SiO3C peak. In Figure 4.14, the SiO3C is at approximately -65ppm as a

shoulder. The next set of peaks that can be clearly identified as they shift towards the

downfield for the polycarbosilane samples pyrolyzed at 500 °C (Figure 4.13 and Figure

4.14) are those associated with SiC4 at approximately -15 ppm. Literature reports the

SiO2C2 peaks in the -30ppm range although in Figure 4.13 and Figure 4.14 they are

shifted more to the -15 and -20 ppm range which could also be associated with SiHC3

based on literature for SMP10 solid state NMR. [4] Figure 4.13 has the final peak at

approximately 1ppm which can be associated with the SiOC3 peak normally around 5 or

7ppm but has been shifted due to possible SiHn(CH2)n influence. For Figure 4.14 the last

peak has more of this influence associate with either SiH(CH2)3and Si(CH2)4 due to the

effect of excess DVB in the structure.

Figure 4.13 29Si Solid State NMR of polycarbosilane (SMP10) sample at 500 °C pyrolysis

temperature.

(tho

usan

dths

)0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

X : parts per Million : Silicon29

100.0 0 -100.0 -200.0

1.30

8

-15.

998

-89.

852

-106

.671

-118

.127

29Si single pulse dec 8kHz 50s 4882 scansSMP10 500C

Page 59: Carbon Rich Polymer-Derived Ceramics: Processing

42

Figure 4.14 29Si Solid State NMR of polycarbosilane (SMP10) sample at 500 °C pyrolysis

temperature with DVB 50.

In comparison, Figures 4.15 and 4.16 demonstrate samples (SMP10 and SMP10

with 50wt% DVB) that have been pyrolyzed at 800 °C and also have the first peak in the

upfield associated with SiO4 and most likely overlapping with SiCO3 at approximately -

70ppm. In contrast to the previous figures, the 800 °C samples have a more

distinguishable difference between samples with and without DVB. For without DVB,

Figure 4.15, the SiC3O peak is present in contrast to the sample with DVB, Figure 4.16.

The peak is also broad enough to include possible SiC4. Figure 4.16 prominently displays

the SiO4 and SiCO3 with only a minor SiC4 peak at approximately -20 ppm.

(tho

usan

dths

)0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

X : parts per Million : Silicon29

100.0 0 -100.0 -200.0

6.79

2

-19.

288

-65.

843

-91.

437

29Si single pulse dec 8kHz 50s 4885 scansSMP10 500C DVB50

Page 60: Carbon Rich Polymer-Derived Ceramics: Processing

43

Figure 4.15 29Si Solid State NMR of polycarbosilane (SMP10) at 800 °C pyrolysis temperature.

Figure 4.16 29Si Solid State NMR of polycarbosilane (SMP10) at 800 °C pyrolysis temperature with

50 wt% DVB.

Figure 4.17 presents the four samples together. Starting with the polycarbosilane

(SMP10) pyrolyzed at 800 °C with DVB at 50wt%, one main peak is prominent (SiO4), and

(tho

usan

dths

)0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

X : parts per Million : Silicon29

100.0 0 -100.0 -200.0

-7.9

54

-69.

865

-93.

021

-102

.039

-118

.736

29Si single pulse dec 8kHz 50s 3400 scansSMP10 800C

(tho

usan

dths

)0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

10.0

X : parts per Million : Silicon29

100.0 0 -100.0 -200.0

-19.

166

-74.

496

-88.

633

-106

.427

SMP10 800C DVB5029Si single pulse dec 8kHz 50s 4800 scans

Page 61: Carbon Rich Polymer-Derived Ceramics: Processing

44

it is shifted more upfield than the other samples, although there is a shoulder for the

SiCO3 at -70ppm. The other samples have the SiCO3 and SiO4 peaks combined. This

means that the SMP10 800 DVB50 sample oxidized more than others. SMP10 without

DVB at 800 °C pyrolysis has another peak related to the remaining SiC2O2, SiC3O and

SiC4. SMP10 at 500 C does not seem to have the SiO2C2 and SiC3O but does have SiC4 and

possibly peaks relating to hydrogen as well. SMP10 at 500 C with DVB has another

shoulder at -60ppm and another peak at -20ppm relating to CH2 with possible shoulder

for SiC4 and without the SiO3C. The 500 C samples were influenced more by the methyl

groups that still remain in the ceramic before they are decomposed at 800 C.

Figure 4.17 29Si Solid State NMR of polycarbosilane (SMP10) samples at 500 C pyrolysis

temperature for DVB 0 and DVB 50 and 800 C pyrolysis temperature at DVB 0 and DVB 50.

(tho

usan

dths

)0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

10.0

X : parts per Million : Silicon29

140.0 120.0 100.0 80.0 60.0 40.0 20.0 0 -20.0 -40.0 -60.0 -80.0 -100.0 -120.0 -140.0 -160.0 -180.0 -200.0 -220.0 -240.0

11.3

02

0.21

1

-7.3

45

-20.

019

-66.

331

-74.

496

-91.

558

-108

.499

SMP10 500C

SMP10 800C SMP10 500C DVB50

SMP10 800C DVB50

Page 62: Carbon Rich Polymer-Derived Ceramics: Processing

45

13C solid state cross polarization (CP) magic angle spinning (MAS) NMR was used

to investigate the carbon bonding states in the same four samples that were used for

29Si NMR studies. For sample SMP10 800 C, in Figure 4.18, the two major peaks are

associated with sp2 and sp3 hybridized carbon. First is for sp2 hybridized carbon at

approximately 125ppm, which in literature is found at 130ppm. [68] The two asterisks

denote spinning side bands next to this peak. And the second peak is associated with sp3

hybridized carbon from CSi4 at 12ppm, which is the same as reported in literature.[74]

Figure 4.18 13C Solid State NMR of polycarbosilane (SMP10) at 800 °C pyrolysis temperature with

0 wt% DVB.

SMP10 at 500 °C only had one peak at approximately 10ppm similar to the sp3

hybridized carbon of CSi4 in Figure 4.19. SMP10 with DVB 50 pyrolyzed at 500 °C has the

second peak with the two spinning side bands and is shifted over towards 135 ppm

(Figure 4.20). The shift of the sp2 peak is due to the incorporation of ally groups (CH2) as

(tho

usan

dths

)0

0.1

0.2

0.3

0.4

0.5

0.6

X : parts per Million : Carbon13

200.0 100.0 0

125.

797

12.0

07

13C cpmas 8kHz 5s 29800 scans

SMP10 800C

*

*

Page 63: Carbon Rich Polymer-Derived Ceramics: Processing

46

reported in literature [4]. The upfield shift of the sp3 peak is also due to the

incorporation of H2 in the form of SiCH2 and SiCH3. [4]

Figure 4.19 13C Solid State NMR of polycarbosilane (SMP10) at 500 °C pyrolysis temperature with

0 wt% DVB.

Figure 4.20 13C Solid State NMR of polycarbosilane (SMP10) at 500 °C pyrolysis temperature with

50 wt% DVB.

(tho

usan

dths

)0

1.0

2.0

3.0

4.0

5.0

6.0

X : parts per Million : Carbon13

200.0 100.0 0

10.1

07

SMP10 500C

13C cpmas 8kHz 5s 12700 scans

(tho

usan

dths

)0

1.0

2.0

3.0

4.0

5.0

X : parts per Million : Carbon13

200.0 100.0 0

138.

972

127.

574

28.6

7420

.095

1.65

1

13C cmpas 8kHz 5s 17300 scansSMP10 500C DVB50

**

Page 64: Carbon Rich Polymer-Derived Ceramics: Processing

47

Figure 4.21 compares all four samples. The SMP10 with 50wt% DVB sample

pyrolyzed at 800 °C had a very weak signal as shown in Figure 4.21. Additional NMR

investigates were done on this sample as described below.

Figure 4.21 13C Solid State NMR of polycarbosilane (SMP10) at 800 and 500 °C pyrolysis

temperature with 0 and 50 wt% DVB.

In order to clarify the C bonding states in the SMP10 with 50wt% DVB sample pyrolyzed

at 800 °C, single pulse 13C NMR was conducted. For comparison, this was also done

SMP10 with 50wt% DVB sample pyrolyzed at 500 °C. The results are shown in Figure

4.22, Figure 4.23 and Figure 4.24.

(tho

usan

dths

)0

1.0

2.0

3.0

4.0

5.0

6.0

X : parts per Million : Carbon13

290.0 270.0 250.0 230.0 210.0 190.0 170.0 150.0 130.0 110.0 90.0 70.0 50.0 30.0 10.0 -10.0 -30.0 -50.0 -70.0 -90.0

138.

972

128.

003

124.

511

27.8

16

14.6

42

4.83

80.

548

*

SMP10 800C DVB50SMP10 800C

*

SMP10 500C

SMP10 500C DVB50

Page 65: Carbon Rich Polymer-Derived Ceramics: Processing

48

Figure 4.22 13C Solid State NMR single pulse of polycarbosilane (SMP10) at 500 °C pyrolysis

temperature with 50 wt% DVB.

Figure 4.23 13C Solid State NMR single pulse of polycarbosilane (SMP10) at 800 °C pyrolysis

temperature with 50 wt% DVB.

(tho

usan

dths

)0

1.0

2.0

X : parts per Million : Carbon13

200.0 100.0 0

128.

555

112.

807

-1.0

45

SMP10 500C DVB50

13C single pulse dec 8kHz 10s 2000 scans

* *

(tho

usan

dths

)0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

X : parts per Million : Carbon13

200.0 100.0 0

120.

466

-6.8

05

SMP10 800C DVB5013C single pulse dec 8kHz 10s 5740 scans

* *

Page 66: Carbon Rich Polymer-Derived Ceramics: Processing

49

Figure 4.24 13C Solid State NMR single pulse of polycarbosilane (SMP10) at 800 and 500 °C

pyrolysis temperature with 50 wt% DVB for pyrolysis comparison with higher percent of carbon.

As expected, Figure 4.22 has the same peaks as Figure 4.20. The 800 °C pyrolyzed

sample had a significantly weaker low ppm peak (sp3) and its upfield shift was lower

indicating loss of hydrogen. The sp2 peaks are similar to those for SMP10 pyrolyzed at

800 °C (Figure 4.18) indicating again low amount of hydrogen due to the increase in

pyrolysis temperature. The 29Si NMR results indicate that oxygen was incorporated in

the samples. 13C NMR helped to identify CH2 groups in the samples pyrolyzed at 500 °C.

Elemental analysis was conducted next to access the level of oxygen incorporation.

(tho

usan

dths

)-0

.3-0

.2-0

.10

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

1.2

1.3

1.4

1.5

1.6

1.7

1.8

1.9

2.0

2.1

2.2

2.3

2.4

2.5

2.6

X : parts per Million : Carbon13

290.0 270.0 250.0 230.0 210.0 190.0 170.0 150.0 130.0 110.0 90.0 70.0 50.0 30.0 10.0 -10.0 -30.0 -50.0 -70.0 -90.0

128.

555

121.

447

112.

807

-1.0

45-6

.008

*

SMP10 800C DVB50

SMP10 500C DVB50

*

Page 67: Carbon Rich Polymer-Derived Ceramics: Processing

50

Elemental Analysis

Elemental Analysis was also conducted on samples in Koblenz, Germany. Oxygen

and carbon elemental analysis were done using combustion methods with a LECO

instrument. The results are shown in Table 4.1 for SMP10 and SMP10 with 50wt% DVB

samples pyrolyzed at 500 °C. It was found that samples with DVB had a higher amount

of oxygen than the samples without DVB at the sample pyrolysis temperature at 500 °C.

As expected, the amount of C is higher for the samples with DVB. In the literature the

weight % of carbon for SMP10 has been reported to be 32%. [3] The lower amount in

the sample in this study may be due to the loss of carbon due to oxidation.

Table 4.1 Elemental Analysis for oxygen and carbon percent of polycarbosilane samples (SMP10) at 500 °C pyrolysis temperature for with and without DVB.

Sample Name Percent of Oxygen Percent Carbon

SMP10 500 C 7.15 27.09

SMP10 DVB50 500 C 17.67 33.34

4.1.3 Summary and Discussion

By using FTIR, the chemical bonding of crosslinked and pyrolyzed samples were

established. Using TGA, helped to determine that above 500 °C, the material becomes

thermally stable. Although those with added carbon, did have higher weight loss. For

samples pyrolyzed at 800 °C, there is minimal weight loss, less than 2 %. FTIR, helped to

Page 68: Carbon Rich Polymer-Derived Ceramics: Processing

51

correlate that this weight loss in the PDCs is most likely related to hydrocarbons, but

TGA-MS would be needed to help confirm this. Solid State NMR demonstrated signs of

oxygen incorporation, which was confirmed by elemental analysis.

A robust process for the low/intermediate temperature processing of

polycarbosilane with variable carbon content has been developed. The TGA results

indicates that the materials are stable up to the pyrolysis temperature. However, the

carbon present in these low temperature pyrolyzed materials is amorphous and, based

on the literature, unlikely to have high conductivity. In addition, due to the high oxygen

content and the carbothermal reduction that has been reported in the literature [1, 2],

these materials are expected to have low thermal stability.

For the next set of experiments, in which SMP10 was modified with boron

addition to improve thermal stability. In addition, Schlenk technique was used to

minimize oxygen contamination as discussed in the next section.

Page 69: Carbon Rich Polymer-Derived Ceramics: Processing

52

4.2 Polycarbosilane with Boron Addition and Boron/Nitrogen Additive

4.2.1 Processing

In collaboration with Prof. Bernard’s group in Limoges, France. Polycarbosilane

(AHPCS/ Smp10) was modified with boron and boron/nitrogen. Either borane

dimethylsulfide (BH3.SMe2, BDMS) or borane dimethylamine (BDMA) was added to the

polymer with an AHPCS-Si:B ratio of 30 based off previous experiments in the Bernard

group. [4,20,48,49] The boron additions have been shown to improve the high

temperature stability of the ceramics. The polymer modification can be seen in Figure

4.25.

Figure 4.25 Process of boron modification to AHPCS polymer with borondimethylsulfide and toluene to create a boron modified polymer [7]

Using standard Schlenk technique, the lines were vacuumed out for 30 minutes

then filled with argon. AHPCS was added to a three necked round bottom flask while

Page 70: Carbon Rich Polymer-Derived Ceramics: Processing

53

argon was flowing. 70ml of Toluene (99.85%, Extra Dry over Molecular Sieve,

AcroSeal(R) purchased from Acros Organics™) was added to the flask. A magnetic stir

bar helped to stir the solution as the temperature was lowered to 0 °C with an external

ice bath. Then 30ml of Toluene with BDMS or BDMA with a Si:B ratio of 30:1 was added

dropwise using a dropping funnel. The slow addition and lower temperature were to

ensure safety since the reaction is exothermic. The solution was slowly equilibrated to

room temperature and left stirring for 3 days. Figure 4.26. a and b demonstrate this

process. Next the solvent was removed with an ether bridge at 60 °C for several hours.

After solvent removal, BDMS produced a yellow solid like substance whereas BDMA was

still liquid of orange color.

a b

Figure 4.26 Picture of apparatus set up for (a) mixing of toluene and BDMS/BDMA in an ice bath

and then the drying step (b).

The samples were then transferred to the glovebox in the round bottom flask. All

chemical products were handled in an argon-filled glove box (Jacomex, Campus-type;

O2 and H2O concentrations kept at ≤ 0.1 ppm and ≤0.8 ppm, respectively). To transfer to

the horizontal furnace (Carbolite BGHA12/450B), samples were crushed and placed in

Page 71: Carbon Rich Polymer-Derived Ceramics: Processing

54

alumina boats into a sealed tube before placing in the furnace for pyrolysis to minimize

oxygen contamination since the boron modified polycarbosilane materials are very air

and moisture sensitive. Figure 4.27 demonstrates the overall conversion of polymer

modification through pyrolysis. For pyrolysis, the furnace tube with samples was first

placed under vacuum then refilled with argon to eliminate any possible oxygen

contamination. Samples were then pyrolyzed with a heating rate of 5 °C/min in flowing

argon at pyrolysis temperatures of 500, 800, 1000 and 1300 °C with a two hour hold and

the same cooling rate.

Figure 4.27 Diagram of AHPCS polymer modification to crosslinking to pyrolysis.

Crosslinking Pyrolysis

Final Product

BDMA

BDMS

Polymer Modification

Page 72: Carbon Rich Polymer-Derived Ceramics: Processing

55

4.2.2 Characterization

ATR FTIR

A ThermoFisher Nicolet nexus FTIR device using an Attenuated Total Reflection

(ATR) tool was utilized to obtain FTIR data. Which was conducted on AHPCS samples

with BDMA or BDMS. The samples were crosslinked at 60 °C and pyrolyzed at 500 and

800 °C. The samples were prepared using KBr pellets in an argon glovebox then

transferred to the FTIR. As shown in Figure 4.28, the crosslinked sample had peaks

corresponding to C-H, Si-H, CH2, Si-CH3, C-C, and Si-C. The crosslinked sample of BDMA

had a peak corresponding to a vinyl group at 1600 cm-1 due to uncrosslinked SMP10.

There were also no prominent nitrogen peaks, leading to the idea that the BDMA

addition with AHPCS did not fully crosslink by hydroboration reaction and confirms why

it was still liquid after solvent removal.

Figure 4.28 (a) FTIR AHPCS B0.1 BDMA crosslinked at 60°C (b) FTIR AHPCS B0.1 BDMs crosslinked

at 60°C

As the pyrolysis temperature increased, the intensity of the hydrogen bonds

decreases. For 500 °C samples, C-H, Si-H, CH2, Si-CH3 bonding remains although the

a b Si-H

C-H C=C C-H

Si-H

Si-CH3

Si-C, Si-CH3

Si-CH2-Si

Si-H

Si-CH3

Si-CH2-Si

Si-H, Si-C, Si-CH3

Page 73: Carbon Rich Polymer-Derived Ceramics: Processing

56

intensity decreases as the material is forming into a SiC ceramic (Figure 4.29). Between

BDMA and BDMS there is no major difference for samples pyrolyzed at 500 °C.

Figure 4.29 (a) FTIR AHPCS B0.1 BDMA pyrolyzed at 500 °C (b) FTIR AHPCS B0.1 BDMS pyrolyzed

at 500°C.

When the pyrolysis temperature reaches 800 °C, the only dominant bond is the

Si-C at the 800 cm-1 region in Figure 4.30 This is expected because as the ceramic is

further pyrolyzed it transforms to a B-modified SiC ceramic.

Figure 4.30 (a) FTIR AHPCS B0.1 BDMA pyrolyzed at 800 °C (b) FTIR AHPCS B0.1 BDMS pyrolyzed

at 800°C.

a b

a b

C-H C-H

Si-H

Si-H

Si-C Si-C

Si-C Si-C

Page 74: Carbon Rich Polymer-Derived Ceramics: Processing

57

Raman Spectroscopy

Raman spectroscopy was utilized to confirm the D and G peaks associated with

graphitic carbon for samples pyrolyzed at 1000 and 1300 °C, Figure 4.31. Using a 532 nm

laser with a Reinshaw InVia Reflex Raman apparatus. Samples pyrolyzed at 500 and 800

°C were not examined with Raman due to the inability to observe the D and G peaks in

these samples. As pyrolysis temperature increases, the peaks sharpen and the D peak

increases in respect to the G peak as the material becomes more crystalline. [20] When

comparing samples with BDMS and BDMA, there was no observable difference.

Figure 4.31 Raman of (a) AHPCS BDMA 1000C, (b) AHPCS BDMS 1000C, (c) BDMS Raman 1300C,

and (d) AHPCS BDMA 1300C.

(a)

(b)

(c)

(d) a

D D

D D G G

G G

Page 75: Carbon Rich Polymer-Derived Ceramics: Processing

58

XRD

X-ray diffraction analysis (Bruker AXS D8 Discover, CuKα radiation) was

conducted on samples with BDMS and BDMA at pyrolysis temperature of 800, 1000 and

1300 °C to observe the structure and phase of the material as it is converted to a

crystalline boron modified SiC ceramic. 800 °C samples demonstrated the predominately

amorphous nature of the ceramic at this pyrolysis temperature and only two weak

peaks were observed associated with β-SiC, as seen in Figure 4.32. As the pyrolysis

temperature increases, three peaks all associated with β-SiC: (111), (220) and (311)

appear. Sample BDMS at 1000 °C pyrolysis also showed a small (002) graphite peak.

Between the BDMS and BDMA samples the peak position was all the same. Peaks

associated with β-Si3N4 were not observed, but are usually not observed below 1700 °C

pyrolysis temperature along with C/BN (C) which is observed above 2000 °C [50]. Other

complimentary techniques will be needed to determine the presence of other elements

(other than Si and C).

Page 76: Carbon Rich Polymer-Derived Ceramics: Processing

59

Figure 4.32 XRD Comparison of B 0.1 BDMS vs BDMA at pyrolysis temperatures of 800, 1000 and

1300 °C.

Elemental Analysis

Elemental Analysis was conducted on samples with BDMS and BDMA at 800 °C

and 1000 °C to compare oxygen content and nitrogen content. The results are shown in

Table 4.2. All samples were found to have relatively low levels of oxygen (approximately

3%) due to the use of Schlenk lines and gloveboxes. For both set of samples, as pyrolysis

temperature increased, the amount of hydrogen decreased as expected. However,

samples that were produced from BDMA, were expected to have some percent of

nitrogen due to the amine groups, but according to Table 4. all samples had negligible

amounts of nitrogen. Future work will use an ammonia gas system for pyrolysis to help

incorporate more nitrogen into the system or use a different boron additive or polymer

precursor with nitrogen.

Page 77: Carbon Rich Polymer-Derived Ceramics: Processing

60

Table 4.2 Elemental analysis of O, N, and H using combustion technique.

Name Mass(g) O%(m/m) N%(m/m) H%(m/m)

AHPCSBN0.1 800C 0.030067 3.331252 0.086656 2.017824

AHPCSB0.1 800C 0.030767 3.452278 0.079123 1.497983

AHPCSBN0.1 1000C 0.032933 3.03978 0.032307 0.975938

AHPCSB0.1 1000C 0.031867 3.783689 0.03445 0.730605

4.2.3 Summary and Discussion

A method to modify polycarbosilane with boron and nitrogen to create ceramic at

500, 800, 1000 and 1300 °C was established. Two different boron additives were used

with AHPCS: BDMA and BDMS. BDMA contained amine groups, which would hopefully

incorporate nitrogen along with boron into the system to improve properties.

Using FTIR the crosslinked samples at 60 °C and pyrolyzed samples at 500 and 800 °C

were investigated. Comparing BDMS and BDMA boron additives confirmed the presence

of vinyl groups in BDMA samples indicating incomplete crosslink of SMP10 with BDMA.

Raman spectroscopy on 1000 and 1300 °C pyrolyzed samples, confirmed D and G peaks

associated with carbon-carbon bonding and the change of those peaks as the pyrolysis

temperature increases and the crystallinity of the material increases. XRD was utilized to

observe the changes of the material when converting from amorphous at 800 °C to

crystalline at 1300 °C with the β-SiC phases as the dominant crystalline phase. Finally

elemental analysis using a combustion technique was used to determine that there was

low oxygen levels in the 800 and 1000 °C samples for both BDMS and BDMA modified

precursors. This is due to the Schlenk lines and argon glovebox used while preparing the

Page 78: Carbon Rich Polymer-Derived Ceramics: Processing

61

boron modified polycarbosilanes. Elemental analysis did confirm that there was

negligible amount of nitrogen in the BDMA samples. Future work in this area will use a

new heat treatment with pyrolysis in ammonia atmosphere to help nitrogen

incorporation into the boron modified polycarbosilane system.

These materials need to be extensively characterized. The electrical properties, the

thermal stability and the surface area needs to be measured. Due to the low level of

oxygen and boron incorporation, the thermal stability is expected to be higher than the

materials reported in Section 4.1.

Page 79: Carbon Rich Polymer-Derived Ceramics: Processing

62

Chapter 5

Results and Discussion on Materials Processing and Characterization for Polysiloxane

5.1 Processing and Characterization of Polysiloxane Derived Ceramics

5.1.1 Processing of Polysiloxanes

In this section the processing of two different polysiloxanes: MK

(polymethylsiloxane) and H44 (polymethylphenylsiloxane) is discussed. These precursors

were selected to focus on precursors with and without phenyl group and the use of

blowing agents to create porous ceramics at lower temperatures.

Slurries were prepared by using a commercial Si-based ceramic precursor

(polymethylsiloxane powder (Silres ® MK, Wacker AG)), 2 wt% crosslinking agent

triethanolamine (TEA) and 30wt% solvent (isopropanol). The slurry was mixed in a ball

mill then heat treated at 80 °C for 6 hours and crosslinked at 200 °C for 24 hours. The

crosslinked material was ground using a high energy ball mill for 10 minutes and sieved

for particle size of <125µm for pyrolysis. Slurries using another commercial Si-based

ceramic precursor were prepared using polymethylphenylsiloxane powder (Silres ® H44,

Wacker AG), 2wt% crosslinking agent (aluminum acetylacetate) and 30wt% solvent

(acetone). The slurry was mixed using a round bottom beaker and a magnetic stir-bar

for 20 minutes until fully homogenized. The solvent was evaporated at 60 °C under

vacuum for several hours. Then 2wt% blowing agent (azodicaronamide) was mixed in to

generate foaming during crosslinking. The resulting material was crosslinked at 80 °C for

6 hours and 200 °C for 24 hours. This crosslinked material was later pyrolyzed as well.

Page 80: Carbon Rich Polymer-Derived Ceramics: Processing

63

Both MK and H44 ceramers were pyrolyzed in the temperature range of 300-700

°C with increments of 100 °C. These samples were pyrolyzed under ultra-high purity

argon with a heating rate of 1 °C/min up to the desired pyrolysis temperature and held

for two hours, then cooled with the same rate to room temperature. [53] The process is

schematically shown in Figure 5.1. The heating schedule for pyrolysis of the polysiloxane

samples is presented in Figure 5.2.

Figure 5.1 Diagram of processing for MK and H44 polymer precursors to ceramers.

MK H44

Cross-linking (200°C, Air)

Grinding/sieving (X < 125µm)

Catalyst/blowing agent

Pyrolysis (300, 400, 500, 600, 700 °C Ar)

Ceramer

Cross-linking/foaming (200°C air)

Page 81: Carbon Rich Polymer-Derived Ceramics: Processing

64

Figure 5.2 Heating steps for pyrolysis of polysiloxane samples.

5.1.2 Characterization

Thermogravimetric analysis (TGA)

The basic objective was to determine and understand the effect of pyrolysis

temperatures on the changes in the chemical composition and bonding. To investigate

thermal stability, TGA was conducted on MK samples that had been pyrolyzed at 200 to

700 °C. Each test was performed under nitrogen from room temperature to 1000 °C

with a heating rate of 10 °C/min. These results are shown in Figure 5.3. As expected, MK

200 sample had the most weight loss because of the lowest pyrolysis temperature. The

weight loss seemed to steadily increase from 200 °C for the first region of weight loss,

associated with hydrocarbons. The second region of weight loss was at temperatures

higher than 600 °C and is most likely associated with the loss of methyl groups. The

material is thermally stable for the temperature greater than 800 °C. MK 700 °C sample

Page 82: Carbon Rich Polymer-Derived Ceramics: Processing

65

reflects that the material is thermally stable with less than a few percent overall weight

loss. The stability of other samples is in between their two extremes.

Figure 5.3 TGA in nitrogen environment of MK samples pyrolyzed at 200 to 700 °C.

To investigate the effect of polymer composition on the thermal stability, the

TGA of H44 pyrolyzed at various temperatures is compared to MK in Figure 5.4. The

scale of Figure 5.3 is different from 5.4 to incorporate the increase in mass loss for H44.

For all samples, the weight loss of H44 is significantly, as much as 30%, higher than that

for MK. It should be noted in H44 that the blowing agent is only 2wt%. So, the difference

is not due to the weight loss associated with the blowing agent. One possible reason for

this higher weight loss is due to the bulkier phenyl groups. TGA coupled with mass

spectrometry is needed to understand this.

Page 83: Carbon Rich Polymer-Derived Ceramics: Processing

66

Figure 5.4 TGA in nitrogen environment of H44 samples pyrolyzed at 200 to 500 °C.

FTIR

FTIR was conducted to investigate the bonding in the pyrolyzed samples. Figure

5.5 compares the chemical bonds of the MK samples pyrolyzed at different

temperatures. In Table 5.1, the FTIR signatures seen in Figure 5.5 are identified and their

bonds with reference to the MK structure are shown schematically in Figure 5.6. As can

been seen, as pyrolysis temperature increases, the intensity of the peaks decreases and

in the 700 °C pyrolyzed samples, the predominant bonding is Si-O-Si.

Similar data for pyrolyzed H44 is shown in Figure 5.7, Table 5.2, and Figure 5.8.

The results are related to MK in terms of effect of temperature by the loss of

hydrocarbons. The primary difference is the presence of phenyl groups in these

samples. The intensity of the phenyl groups decreases and is absent in the 500 °C

pyrolyzed samples.

Page 84: Carbon Rich Polymer-Derived Ceramics: Processing

67

In Figure 5.9 and 5.10, the FTIR spectra of H44 and MK pyrolyzed at 200 °C and

500 °C, respectively, are compared. As shown in Figure 5.9, for the 200 °C pyrolyzed

samples, the significant difference is the presence of peaks corresponding to phenyl

groups in H44. For samples pyrolyzed at 500 °C, the two spectra are similar in their

decrease in bonds (Figure 5.10). But not until comparing MK 700 with H44 500 do they

appear similar in only having Si-O-Si and Si-C bonds.

Figure 5.5 FTIR of MK samples pyrolyzed at 200-700 °C.

Table 5.1 Correspondence between chemical bonds and FTIR peak location. [8]

Peaks Wavelength (cm-1)

(Si-CH3)n

(n=3,2)

2975-2950 (C-H stretching) 1440-1400, 1290-1240

(Si-CH3 stretching), 870-760

Si-O-Si 1090-1010, 625-480

O-Si-CH3 850-840

Page 85: Carbon Rich Polymer-Derived Ceramics: Processing

68

Figure 5.6 Schematic of the chemical structure of MK contributing to FTIR peaks.

Figure 5.7 FTIR of H44 samples pyrolyzed at 200-500 °C.

Table 5.2 Correspondence between chemical bonds and FTIR peak location. [8]

Peaks Wavelength (cm-1)

(Si-CH3)n 2900, 1290-1240, 870-760

Si-O-Si 1000

O-Si-CH3 850-840

R3SiPh 3000, 1600, 1480-1420, 760-700, 660-600

Page 86: Carbon Rich Polymer-Derived Ceramics: Processing

69

Figure 5.8 Schematic of the chemical structure of major groups in H44 contributing to

the FTIR peaks.

Figure 5.9 Comparison of the FTIR spectra of MK and H44 samples crosslinked at 200 °C.

Figure 5.10 Comparison of the FTIR spectra of MK and H44 samples pyrolyzed at 500 °C.

methyl phenyl

Page 87: Carbon Rich Polymer-Derived Ceramics: Processing

70

Surface Area Analysis

BET analysis was used to characterize the surface area of selected samples. MK

200, 400, 500 and 700 were examined using a multi-point BET test. Figure 5.11 shows

the surface area measurements for the samples. MK 200 and 700 were observed to

have very low surface area compared to the other two samples. MK 500 demonstrates

microporosity based on its high surface. Figure 5.12 shows BET results of volume vs

relative pressure for the four samples. More data points will be needed to determine if

the material is microporous based on if it shows a type I or IV isotherm.

Figure 5.11 Surface area measured using BET analysis of MK 200, 400, 500 and 700 samples.

11.143

165.854

430.95

5.8690

50

100

150

200

250

300

350

400

450

500

MK 200 MK 400 MK 500 MK 700

Surf

ace

Area

(m2 /g

)

Samples

B E T S u r fa c e A r e a

Page 88: Carbon Rich Polymer-Derived Ceramics: Processing

71

Figure 5.12 Pore volume as a function of relative pressure for MK 200, 400, 500 and 700 samples.

Figure 5.13 BET Isotherm of MK 500.

In order to investigate if the samples were microporous, BET adsorption and

desorption runs were obtained for MK 500. The results are shown in Figure 5.13.

Further investigations are needed to confirm and measure pore size.

MK 500 sample had a high specific surface area of 430m2/g. Samples with a

pyrolysis temperature of 700 °C had low specific surface area, due to the shrinkage of

pores.

0

20

40

60

80

100

120

0.00E+00 5.00E-02 1.00E-01 1.50E-01 2.00E-01 2.50E-01 3.00E-01

Volu

me

(cm

3 /g) a

t STP

Relative Pressure (P/Po)

MK 500

MK 200

MK 700

MK 400

Page 89: Carbon Rich Polymer-Derived Ceramics: Processing

72

5.1.3 Summary and Discussion

The aim of this research in this section is to add to the scientific understanding of

the polysiloxane precursors as they transition to ceramics. This material group also has a

wide range of potential application such as gas membranes, high area absorbents,

catalysts, and materials with low dielectric constants. This section focused on the Si-

based polysiloxane polymers which act as a baseline for the hybrid materials reported in

the rest of this chapter.

Processing techniques for MK (polymethylsiloxane) and H44

(polymethylphenylsiloxane), including low temperature pyrolysis, were established.

TGA and FTIR was used to investigate the thermal decomposition and thermal stability

of the two polymers. For the same heat treatment, as expected, the mass loss of H44 is

higher due to the loss of the bulkier phenyl group. This was confirmed by FTIR. Under

suitable conditions, high surface area thermally stable materials have been produced by

low temperature pyrolysis of polysiloxanes.

Page 90: Carbon Rich Polymer-Derived Ceramics: Processing

73

5.2 Polysiloxane with Added Carbon

Polysiloxane with added carbon, focuses on the chemistry, thermal stability, and

electrical conductivity of low/intermediate pyrolysis temperature (700-900 °C)

polysiloxane derived ceramics. These ceramics were modified with additional carbon

derived from divinylbenzene (DVB) added to the precursor. Their electrical properties

were investigated for potential uses in micro-electrical mechanical systems (MEMS) and

anodes for lithium batteries. The microstructure and chemical composition were

investigated by attenuated total reflectance Fourier transform infrared spectroscopy

(ATR-FTIR), Raman spectroscopy, and x-ray photoelectron spectroscopy (XPS);

thermogravimetric analysis (TGA) provided insight into the thermal stability; and

electrochemical impedance spectroscopy (EIS) into the electrical properties of the

material. The increase of pyrolysis temperature and carbon content lead to an

enhancement of the electrical conductivity, higher than previously reported values for

intermediate pyrolysis temperature SiOC polymer derived ceramics. A limit of the

amount of DVB that can be added to PHMS to produce a hybrid precursor has also been

obtained

Page 91: Carbon Rich Polymer-Derived Ceramics: Processing

74

5.2.1 Processing

Commercially available polymethylhydrosiloxane (PHMS) (98–100 %, MW: 2100–

2400, Gelest, Morrisville, PA) and divinylbenzene (DVB) (technical grade 80 %, Sigma

Aldrich, St. Louis, MO) were combined with platinum complex crosslinking agent

(platinum divinyltetramethyldisiloxane 2% in xylene) (Gelest, Morrisville, PA) that was

diluted to 0.1 % in xylene (Semiconductor grade, Alfa Aesar, Ward Hill, MA). 40 μL of Pt

complex was used as a crosslinking agent per 1 g of PHMS. The research was conducted

on a series of samples in which the weight ratio of DVB to PHMS was varied from 0 to 6.

The crosslinked samples are labeled as: dvbn (n referring to the ratio of DVB: PHMS

(weight ratio). For the pyrolyzed samples, the pyrolysis temperature was added to the

sample designation (Table 5.3).

Table 5.3 Designation of crosslinked and pyrolyzed samples.

Sample designation for crosslinked samples

dvbn (n = 0–6)

Sample designation for pyrolyzed samples

700C dvbn, 800C dvbn, or 900C dvbn (n = 0–6)

PHMS, DVB and Pt complex were stirred for two minutes, and thin

(approximately 0.5 mm thickness) polymer films were cast and crosslinked at room

temperature. 10 mm diameter disks were cut out of the crosslinked film and pyrolyzed

Page 92: Carbon Rich Polymer-Derived Ceramics: Processing

75

in ultrahigh purity argon (Airgas, Radnor, PA) with 1 °C/min heating rate to the

isothermal temperature followed by a two-hour hold at the desired temperature (700,

800 and 900 °C) and a 1 °C/min cooling rate (in an alumina tube in a Lindberg/Blue

Furnace). Figure 5.14 shows the conversion of crosslinked polysiloxane to the pyrolyzed

sample at 800 °C. Crosslinked samples underwent a shrinkage of approximately 17 %

after 800 °C pyrolysis. This shrinkage was caused by the loss of organics (as is

demonstrated in succeeding sections) in the material and elimination of microporosity.

Figure 5.14 Comparison of SiOC crosslinked and pyrolyzed samples with a reference.

5.2.2 Characterization

TGA and dTGA

TGA, in inert environment, was used to investigate the effect of DVB:PHMS ratio

(n) on the thermal stability and the ceramic yield of the crosslinked samples. The

samples were heat treated in nitrogen atmosphere with a heating rate of 10 °C/minute.

Thermal stability of the samples was determined by comparing the ceramic yields and

Page 93: Carbon Rich Polymer-Derived Ceramics: Processing

76

decomposition temperature. As shown in Figure 5.15, for all samples with n > 2, there

are two regions of weight loss.

This is confirmed by Figure 5.16, in which the derivative of the weight loss is

plotted as a function of temperature. The first region of interest between 50−150 °C

where initial weight loss occurs due to the loss of hydrogen and water [75]. It is clear

that as the amount of DVB increases beyond 200 % of PHMS, the weight loss in this

range is significant. In the range of 400 °C-600 °C, the weight loss corresponded to the

loss of incorporated hydrocarbons. After 600 °C, the samples stabilize and there is no

further weight loss. Table 5.4 demonstrates that the dvb1 and dvb2 samples show

minimal weight loss in region one, showing that all the DVB is being incorporated into

the precursor material. For higher levels of DVB, starting from dvb3 there is a significant

weight loss in region one showing that not all the DVB is incorporated into the siloxane

precursor possibly due to the steric hinderance of DVB and limited crosslinking sites. In

addition, for n > 3, the low temperature weight loss increases as the amount of DVB

increases indicating that the free DVB contributes to incorporation of water, from

environment, and this water is lost in the low temperature regime. This is likely due to

the fact that the crosslinked samples were exposed to lab air for a period of time

between 24–48 hours before the TGA analysis was conducted. It was shown in previous

literature that aging of the samples increases the weight loss in the low temperature

regime due to water from environment. [75]

Page 94: Carbon Rich Polymer-Derived Ceramics: Processing

77

Figure 5.15 Weight loss as a function of temperature of crosslinked samples dvbn; n =

0,1,2,3,4,5 and 6.

Figure 5.16 Derivative of the weight loss of crosslinked samples dvbn; n = 0,1,2,3,4,5 and

6.

This trend can further be illustrated by plotting the weight loss in the two regions

as a function of DVB (Figure 5.17). The weight loss in region 1 increases linearly as the

DVB content increases beyond DVB3. The weight loss in region 2 is maximum for DVB3

Page 95: Carbon Rich Polymer-Derived Ceramics: Processing

78

and then decreases as the amount of DVB increases. However, the variation is small (in

the range of 20 to 25 %). The TGA results clearly shows that up to a DVB:PHMS ratio of

2, the DVB becomes part of the hybrid precursor and beyond that DVB remains

unreacted.

Table 5.4 From TGA data, t percent weight loss in weight loss region 1 (50 to 150 °C) and

weight loss region 2 (400-600 °C) for crosslinked samples dvbn; n = 1 to 6.

Weight loss region 1 Weight loss region 2

Temperature range 50−150 (°C) 400−600 (°C)

dvb1 0.01 19.65

dvb2 0.01 20.08

dvb3 11.89 26.94

dvb4 21.75 25.83

dvb5 34.95 21.07

dvb6 45.4 16.98

Page 96: Carbon Rich Polymer-Derived Ceramics: Processing

79

Figure 5.17 Percent weight loss vs amount of DVB from TGA of crosslinked samples for

weight loss region 1 (50 to 150 °C) and weight loss region 2 (400 to 600 °C).

FTIR

ATR-FTIR was used to characterize chemical bonding and identify crosslinking

mechanisms. Based on previous research hyrdosilyation reactions were suggested for

crosslinking mechanisms based on changes in Si-H bonding. DVB polymerization

reactions were also investigated for with and without DVB (n=1 or n=0 respectively).

Different pyrolysis temperatures were then investigated to correlate to previous TGA

data.

ATR FTIR was used to characterize the effects of addition of DVB and pyrolysis

conditions on the chemical bonding, specifically the crosslinking of the samples.

Previously, FTIR has been used to gain insight into the crosslinking mechanism of

polysiloxane with DVB and the effect of pyrolysis temperature [45]. We expect that the

crosslinking mechanism will be the same as previously determined and for the pyrolyzed

Page 97: Carbon Rich Polymer-Derived Ceramics: Processing

80

samples the number and intensity of organic bonds (i.e. C-C, C-H) will decrease [45].

Crosslinked sample with divinyl benzene shows, as seen in Figure 5.18, several distinct

chemical bonds: C-Hn bonds in the 2960-2870cm-1 range, Si-H stretching bond at

2160cm-1 , (Si-H2)n wagging and bend-scissors modes at 843 and 890 cm−1, respectively,

Si-CH3 stretching bond at 1260cm-1 and rocking at 760cm-1 , Si-CH2-Si bond at 1180cm-1 ,

Si-O bonds at 1030cm-1 and 1080cm-1 , and C-C bonds in the 1400-1625cm-1 range [45].

Figure 5.18 ATR FTIR of crosslinked dvb1 sample at room temperature.

The FTIR spectra of samples crosslinked, at room temperature, with and without

DVB are compared in Figure 5.19 for representative n values. The lack of DVB is

associated with absence of 1180cm-1 peak for Si-CH2-Si and 1400 to 1625cm-1 bands for

C-C bonding. When DVB is present, a hydrosilylation reaction takes place forming Si-CH2-

Si bonds instead of unsaturated carbon creating a crosslinked network, as previously

reported in literature [45]. This role of DVB in crosslinking of PHMS leads to a hybrid

Page 98: Carbon Rich Polymer-Derived Ceramics: Processing

81

PHMS-DVB precursor. In summary, the increase of DVB leads to a relative decrease of Si-

H peak at 2160cm-1 and of the (Si-H2)n bend-scissors peak at ~840 cm-1 concomitant to

an enhancement of the C-C bonds around 1500 cm-1 , Si-CH2-Si bonds at 1180cm-1 , and

Si-O peak at 1180 cm-1 .

Figure 5.19. ATR FTIR spectra of crosslinked dvb 0, 1 and 6 samples at room

temperature.

The changes in the chemical bonding were also investigated after pyrolysis at

800 °C. Only Si-O-Si and Si-C bonds remain as can be seen in Figure 5.20 the chemical

bonding changes have been correlated with the TGA results.

Page 99: Carbon Rich Polymer-Derived Ceramics: Processing

82

Figure 5.20 FTIR spectrum of dvb1 sample pyrolyzed at 800 °C.

The microstructural transformations due to the addition of DVB were evaluated

by ATR FTIR with emphasis on the crosslinking. Samples containing various percentages

of divinyl benzene from n = 0 to 6 were compared to determine the effect DVB has on

the chemical bonding. ATR FTIR results reveal a decrease of (Si-H2)n linear polymeric

bonds at 836 cm-1 and Si-CH3 terminal bonds at 755 and 870 cm-1 while Si-CH2-Si linear

polymeric bonds at 1180 cm-1 increase. The elimination of terminal bonds is suggestive

of crosslinking. Si-H is consumed during crosslinking with divinylbenzene due to the

hydrosilylation reaction leading to a hybrid precursor. As the amount of DVB increases,

due to the steric hindrance of divinylbenzene there will be less available Si-H bonding

and a new reaction between the carbon bonds in divinylbenzene will occur as DVB

amounts are increased [3]. Further, Si-O bonds at 1083 cm-1 and C-C bonds around

Page 100: Carbon Rich Polymer-Derived Ceramics: Processing

83

1400-1500 cm-1 increase from n = 0 to n = 1 and remain unchanged after further

increase of DVB. Oxygen is not released until above 1000 °C as shown in literature [76].

Comparing the ATR FTIR with TGA helps to determine which components of the material

are lost in different temperature ranges.

Raman Spectroscopy

Raman was used to investigate the carbon-carbon bonding. Crosslinked samples

were examined as well as ceramic samples. The ideal temperature range of ceramic

pyrolysis at which D and G bonding was observed was 700 °C to 900 °C. Different laser

powers were also investigated, although 1% was deemed to be ideal and used to

compare the SiOC ceramics at different pyrolysis temperatures. Different carbon

amounts were also investigated, but no significant change was found between ID/IG

ratios. For differing pyrolysis temperatures, the ID/IG ratios were seen to change. ID/IG

ratios were determined by deconvolution of the D and G peaks using as described in

methods. Peaks were normalized by intensity of the 1600 cm−1 “G peak”. The 1330

cm−1 “D peak” was fit with a Lorentzian function while for the “G peak” a Breit-Wigner-

Fano function (BWF) was used consistent with previous research on amorphous carbons

[73]. With increasing pyrolysis temperature, comes sharpening of the D peak and

demonstrates a change from amorphous to nanographitic carbon, this change will

increase with temperature as seen in previous literature at 1000 °C and above.

Page 101: Carbon Rich Polymer-Derived Ceramics: Processing

84

Raman spectroscopy of carbon based materials presents two major bands, the

G-band, centered around 1570 cm−1, that is associated with the optically allowed E2g

zone center of crystalline graphite, and the D-band, at around 1400 cm−1, associated

with disorder-allowed zone edge modes of graphite that become Raman active due to

the lack of long-range order [77]. The effect of the pyrolysis temperatures and carbon

content on the D and G bands of amorphous carbon materials has already been

investigated [73,78]. In this work, Raman spectroscopy was used to characterize the

effects of the pyrolysis temperature on the carbon-carbon bonding through the analysis

of the D and G bands. Figure 5.21 illustrates the effects of different pyrolysis

temperatures on the Raman spectra normalized to the G peak at 1603cm-1. The peak

position, full width at half maximum (FWHM), and peak intensity ratio I(D)/I(G)

(determined in terms of area (A) and height (H)) )) were extracted by spectral fitting as

illustrated in Fig 5.22 for PHMS dvb1. Fitting results are presented in Table 5.5 as the

pyrolysis temperature increases, the I(D)/I(G) ratio increases in both area and height.

The position of the D and G peaks remain in a similar range of 1328 to 1333cm-1 and

1606 to 1608cm-1, respectively. The D band narrowing together with the increase of the

I(D)/I(G) ratio for higher pyrolysis temperatures indicate an increase in the size and/or

number of the graphitic domains from amorphous carbon [73] .

Page 102: Carbon Rich Polymer-Derived Ceramics: Processing

85

Figure 5.21 Raman of dvb1 pyrolyzed at temperatures of 700,800 and 900 °C.

Figure 5.22 Raman spectra with best fits of dvb1 pyrolyzed at: (a) 700 °C, (b) 800 °C, and (c) 900

°C.

Page 103: Carbon Rich Polymer-Derived Ceramics: Processing

86

Table 5.5 Results of Raman analysis in terms of D and G bands peak position, full width

at half max (FWHM) and I(D)/I(G) ratio based on area (A) and height (H) of sample dvb1

pyrolyzed at 700 °C, 800 °C and 900 °C.

Sample Position D (cm−1)

FWHM (cm−1)

Position G (cm−1)

FWHM (cm−1)

I(D)/I(G) A

I(D)/I(G) H

700C dvb1

1328 194 1607 78 1.6 0.84

800C dvb1

1333 187 1608 71 1.8 0.92

900C dvb1

1329 176 1606 83 1.8 1.1

The comparison of samples with different of amounts of DVB pyrolyzed at 800 °C

is presented in Figure 5.23, where spectra were normalized to the G peak maximum.

The peak position, FWHM, and I(D)/I(G) ratios are reported in Table 5.6 for samples with

different amounts of DVB pyrolyzed at 800 °C. As DVB increase there is no major change

of the I(D)/I(G) ratio as can inferred from Figure 5.23 and Table 5.6. The increase of

ID/IG ratio has been associated with the increase in size and/or number of the sp2

domains [79]. The lack of observed change of the ID/IG ratio indicates that there is no

substantial change in the structure of carbon in these materials.

Page 104: Carbon Rich Polymer-Derived Ceramics: Processing

87

Figure 5.23 Raman of dvbn for n from 1 to 6 all pyrolyzed at 800 °C.

Table 5.6 Raman values of peak position for D and G, full width at half max (FWHM) and I(D)/I(G) ratio for area (A) and heigh (H) of dvbn samples pyrolyzed at 800 °C (n from1 to

6).

Sample Position D (cm−1)

FWHM (cm−1)

Position G (cm−1)

FWHM (cm−1)

I(D)/I(G) A

I(D)/I(G) H

800C dvb1

1333 187 1608 71 1.8 0.92

800C dvb2

1334 220 1597 70 3.1 1.1

800C dvb3

1328 248 1605 72 3.4 1.1

800C dvb4

1328 219 1605 62 3.1 0.98

800C dvb5

1323 218 1602 67 3.0 1.0

800C dvb6

1325 228 1602 74 3.0 1.1

Page 105: Carbon Rich Polymer-Derived Ceramics: Processing

88

The effect of pyrolysis temperature is evaluated as a function of the temperature

and carbon content using Raman spectroscopy. The D band narrowing together with the

increase of the I(D)/I(G) ratio for higher pyrolysis temperatures indicate an increase in

the size and/or number of the graphitic domains from amorphous carbon [22].

XRD

X-ray diffraction (XRD) was performed on sample 800 C dvb1 as shown in Figure

5.24, this sample is amorphous. It is expected that the other samples will also be

amorphous as well.

Figure 5.24 XRD of 800 C dvb1.

Page 106: Carbon Rich Polymer-Derived Ceramics: Processing

89

XPS

XPS was used to determine the chemical bonding of C in selected samples (800C

dvb1, 800C dvb6, 700C dvb1 and 900C dvb1). A survey spectrum (Figure 5.25) was

obtained for XPS from there C and Si spectra were created. This was done by comparing

binding energies (BE) and percent abundance within the C (1s) spectrum. Figure 5.26 a–

d shows C (1 s) bonding that includes Si-C, Si-O-C, C=C (sp2), C-C/H (sp3) and C-O bonds.

Their binding energies are identified in Table 5.7 based on literature values [76,80,81].

The effect of increase in carbon (i.e., higher DVB content) and pyrolysis temperature

were analyzed in terms of C-Si, C-Si-O, C=C, C-C/H and C-O bonding relative intensities

shown as peaks in Figure 5.26.

Figure 5.25 XPS survey 800 C dvb1.

Su1s 1

x 103

2468

101214161820

CPS

1000 800 600 400 200 0Binding Energy (eV)

Page 107: Carbon Rich Polymer-Derived Ceramics: Processing

90

Figure 5.26 XPS spectra of C 1s (a) 800C dvb1, (b) 800C dvb6, (c) 700C dvb1 and (d) 900C dvb1.

Table 5.7. XPS binding energies (BE) of C1 s for samples 800C dvb1, 800C dvb6, 700C

dvb1 and 900C dvb1.

C-Si BE (eV) C-Si-O BE (eV) C=C BE (eV) C-C/H BE (eV) C-O BE (eV)

700C dvb1 282.6 283.3 284.4 285.6 286.8

800C dvb1 282.0 282.8 284.0 285.0 285.9

900C dvb1 282.5 283.2 284.6 285.5 286.5

800C dvb6 282.1 282.6 283.9 285.0 286.3

Using XPS results, Table 5.8 summarizes the chemical bond presence in terms of

abundance percent for the samples investigated.

Page 108: Carbon Rich Polymer-Derived Ceramics: Processing

91

Table 5.8 Summary of the analysis of XPS high resolution scans of C 1s. Each scan set of

results add up to 100 %.

C-Si (%) C-Si-O (%) C=C (%) C-C/H (%) C-O (%)

700C dvb1 64 22 9 3 2

800C dvb1 63 27 6 2 2

900C dvb1 52 35 6 4 3

800C dvb6 52 33 8 4 3

The XPS results show the changes in the abundance of various chemical bonds

both as a function of the DVB content in the precursor and the pyrolysis temperature.

Specifically, for the same pyrolysis temperature (800 °C), as the DVB content increases

(from n = 1 to n = 6), the most significant change is in the decrease of the abundance of

the C-Si bonds and an increase in the abundance of the C-Si-O bonds. The other C

containing bonds become slightly more abundant. For the same DVB content (n = 1), as

the pyrolysis temperature increases, the abundance of the C-Si bond decreases, the C-

Si-O bond increases, and the C=C decreases. These results are further discussed in the

discussion section in the context of the conductivity measurements.

XPS for silicon spectra Si 2p, two peaks were deconvoluted. The first being Si-C in

orange on Figure 5.27. The second being Si-O-C as the grey peak. As the pyrolysis

temperature increases from 800 to 900 °C as seen in Figure 5.27 from a to b, the Si-C

peak increases and the Si-O-C decreases. This trend is apparent in literature as the

Page 109: Carbon Rich Polymer-Derived Ceramics: Processing

92

pyrolysis temperature increases above 1000°C. [81] Table 5.9 compares the percent

abundance in comparing Si-C with Si-O-C. An interesting observation is that samples

with a higher amount of DVB, have increased percent of Si-O-C. This maybe due to the

higher affinity of DVB to oxygen, which was observed also in Chapter 4.

Figure 5.27 XPS spectra of Si 2p (a) 800C dvb1 and (b) 900C dvb1.

Table 5.9 Summary of the analysis of XPS high resolution scans of Si 2p. For each scan

the % add to 100 %.

Si-C (%) Si-O-C (%)

800C DVB1 81.23 18.77

800C DVB6 71.80 28.20

900C DVB1 90.07 9.93

030060090012001500

9799101103105

Inte

nsity

(a.u

.)

Binding Energy (eV)

900C

0

300

600

900

1200

1500

9799101103105

Inte

nsity

(a.u

.)

Binding Energy (eV)

800C

Page 110: Carbon Rich Polymer-Derived Ceramics: Processing

93

Electrochemical Impedance Spectroscopy (EIS)

Initial two-point testing was conducted and helped determined to test samples

at 800C and 900 °C pyrolysis temperature. EIS was used to measure the electrical

conductivity of the samples. All three samples (800 dvb1, 800 dvb6, and 900 dvb1)

exhibited a similar tendency that their conductivity increased with increasing

temperature from room temperature to 550 °C as seen in Figure 5.28. In addition, for

samples heat treated at the same temperature, 800 °C, samples with higher carbon

content had higher conductivity (Figure 5.28) over the entire temperature range. For

example, at 550 °C, the conductivity of dvb6 is around one order of magnitude three

times higher than that of dvb1. The effect of pyrolysis temperature can also be

compared in Figure 5.28 by comparing dvb1 samples pyrolyzed at 800 °C and 900 °C. For

the 900 °C pyrolyzed samples the electrical conductivity is almost two orders of

magnitude thirty times higher, at 500 °C, compared to the sample pyrolyzed at 800 °C.

This observation supports is in agreement with the statement that there is an increase

in electrical conductivity as a function of pyrolysis temperature up to 1000 °C, which was

reported in a literature [66].

EIS was used and demonstrated that the electrical conductivity increases with

increased carbon content by an order of magnitude and with increased pyrolysis

temperature by approximately two orders of magnitude. The change in pyrolysis

temperature had the effect on the conductivity. The increase in conductivity as a

Page 111: Carbon Rich Polymer-Derived Ceramics: Processing

94

function of temperature is explained by the change in the structure of carbon confirmed

by Raman.

Figure 5.28 Electrical conductivity as a function of the temperature of sample 800C dvb1, 800C

dvb6 and 900C dvb1.

Samples were also run-in different environments other than dry argon as seen in

Figure 5.29. Dry argon exhibits better electrical conductivity compared to wet argon

and wet hydrogen for sample 800 C dvb1.

Page 112: Carbon Rich Polymer-Derived Ceramics: Processing

95

Figure 5.29 800C dvb1 three different atmospheres.

5.2.3 Summary and Discussion

Previous research on polysiloxane derived ceramics with carbon additives

showed the microstructure to contain continuous amorphous/disordered graphitic and

SiOC domains [2]. The graphitic domain is formed from the incorporation of DVB. The

amorphous/disordered graphitic domains get more organized and possibly grow in size

as pyrolysis temperature increases and is expected to result in an overall conductivity

increase. To better understand this phenomena, detailed chemical investigation by

means of ATR FTIR, TGA, Raman, XPS and EIS were conducted. The chemical

transformations due to the addition of DVB were evaluated by ATR FTIR with emphasis

on the crosslinking. Crosslinked samples containing various percentages of divinyl

benzene, for DVB:PHMS weight ratio n from = 0-6 were compared to determine the

Page 113: Carbon Rich Polymer-Derived Ceramics: Processing

96

effect of DVB on the chemical bonding in the crosslinked polymer (Figure 5.19). ATR FTIR

results revealed that as the amount of DVB increases, there is a decrease of (Si-

H2)n linear polymeric bonds and Si-CH3 terminal bonds while Si-CH2-Si linear polymeric

bonds increased. The elimination of terminal bonds is suggestive of crosslinking. Si-H is

consumed during crosslinking with divinylbenzene due to a hydrosilylation reaction,

leading to a hybrid precursor. More specifically, Si-H and Si-CH3 bonds decrease due to

DVB substituting for H and CH3 attached to silicon in the polymer precursor. As the

amount of DVB increases, a point is reached such that the increase of DVB in the

precursor mixture no longer leads to a proportional increase of the amount of DVB

incorporated in the crosslinked polysiloxane. From the results of the mass loss, during

heat treatment (Figure 5.17), it is clear that the relative weight loss is quite different for

samples with DVB:PHMS ratio less than or equal to 2 than for higher amounts of DVB.

This figure clearly illustrates that for DVB: PHMS greater than 2, the weight loss in the

low temperature regime (50–150 °C) increases proportionately as the amount of DVB

increases. These results indicated that a DVB: PHMS weight ratio of 2 is maximum for

the being incorporated to create a hybrid DVB-PHMS polymer. Beyond this, a mixture of

the two polymers is obtained and the free DVB incorporates low temperature volatile

species (most likely water) from the environment.

The incorporation of DVB has been shown to be associated with a more

conductive ceramic[3]. This research also shows that as the amount of DVB increases,

the electrical conductivity increases (Figure 5.28). However, the increase in conductivity

Page 114: Carbon Rich Polymer-Derived Ceramics: Processing

97

is quite small. We found that SiOC ceramics without DVB were insulating and when the

ratio of DVB:PHMS was 1 or higher, the samples became semi-conducting. However, a

six-fold increase in DVB content only improves the conductivity by a factor of 3 or so (at

550 °C). This result corresponds well with the analysis of free carbon in the study

reported in [44] on polymer derived ceramics from DVB-PHMS. In this study, it was

shown that the free carbon increases as the amount of DVB increases. However, the

highest increase is for lower amounts of DVB. For example, the free carbon for 1:1 ratio

of PHMS:DVB is 30.6 % and is 54.2 for 1:6 ratio. XPS results (Figure 5.26 and Table 5.7)

also show that the difference between the percentage of C=C between samples with

DVB:PHMS ratio of 1 and 6 (both heat treated at 800 °C) is small. The effects of pyrolysis

temperature were evaluated as a function of the temperature (Figure 5.21 and Table

5.5) and carbon content (Figure 5.23 and Table 5.6) using Raman spectroscopy. The D

band narrowing together with the increase of the I(D)/I(G) ratio for higher pyrolysis

temperatures indicated an increase in the size and/or number of the graphitic domains

from amorphous carbon [73]. However, when comparing the amount of carbon in the

pyrolyzed samples at 800 °C as n increased from 1 to 6, no structural changes related to

the amorphous carbon/disordered graphitic domains were revealed by Raman

spectroscopy. Further insight into the structural evolution was obtained by XPS. Detailed

analysis of the carbon binding energy showed that between 800C dvb1 and 800C dvb6,

there was a decrease in the amount of C-Si bonding while an increase of the relative

abundance of C-Si-O and C=C bonds, as seen in Table 5.8. Finally, EIS revealed

Page 115: Carbon Rich Polymer-Derived Ceramics: Processing

98

measurable increase of the electrical conductivity (a factor of 3) due to increasing

carbon content. However, the pyrolysis temperature has a major effect. The electrical

conductivity, at high temperature increases by almost a factor of thirty when the

pyrolysis temperature is increased from 800 to 900 °C (for same DVB:PHMS ratio). There

is a small effect of the increased amount of DVB and this correlates to the increase in

small increase in free carbon in going from dvb1 to dvb6. The much larger increase in

conductivity as a function of the pyrolysis temperature is explained by the change in the

structure of carbon confirmed by Raman by focusing on the I(D)/I(G) ratios (Figure 5.21,

Figure 5.22, Table 5.5)

The chemistry, thermal stability and electrical properties of PHMS-DVB

blended/hybrid polymers was investigated as a function of stoichiometry and pyrolysis

temperature. Thermal stability was determined by TGA, in which, the ceramics retained

stability with minimal weight loss above 600 °C. Chemical changes also occurred due to

the change in pyrolysis temperature and stoichiometry as seen in FTIR, Raman and XPS.

Overall, the increase in pyrolysis temperature showed an increase in the I(D)/I(G) ratio

as seen in Raman. XPS was also able to confirm the presence of carbon bonding (both

sp2 and sp3) and carbon-oxygen bonding. The DVB-PHMS hybrid precursor leads to a

composite SiOC-C composite. It is shown that the amount of DVB that can be

incorporated in the formation of the hybrid polymer, during crosslinking, is limited to

around two times the weight of PHMS. Both the increased temperature and carbon

content increased the electrical conductivity as a result, an intrinsically insulating SiOC is

Page 116: Carbon Rich Polymer-Derived Ceramics: Processing

99

transformed to a semiconductor due to the addition of free carbon that undergoes a

structural change as temperature increases. These results show the promise of being

able to make semiconducting SiOC with increased carbon at low processing temperature

which are expected to result in high levels of microporosity and surface area. Further

investigations are needed to investigate the nature and extent of porosity in these

materials. In addition, the nature of the carbon derived from the hybrid polymer at low

temperatures for amorphous ceramics and that from unreacted DVB needs to be

investigated.

Page 117: Carbon Rich Polymer-Derived Ceramics: Processing

100

5.3 Polysiloxane with Nanographene Particles

Adding graphene has been popular method of incorporating carbon into a

structure for the improvement of properties such as electrical conductivity.[82]

Polysiloxane samples with nanographene platelets were investigated with in the aims to

enhance the conductivity of the resultant ceramic. The chemical composition,

microstructure and thermal decomposition were investigated by FTIR, TGA, TGA-MS,

and Raman.

5.3.1 Processing

Polysiloxane (PHMS) was combined with platinum complex as done in previous

section (5.2). Different weight percent of graphene nanoplatelets 1 and 5wt% (ACS

Material, LLC, Pasadena, CA) were combined then crosslinked at room temperature. The

graphene nanoplatelets were 2-10nm thickness. PHMS with DVB1 and 1wt% graphene

was also investigated. Samples were cast on glass and 10mm discs were punched out, as

done in previous sections (Figure 5.30). All samples were then pyrolyzed at 700, 800,

and 900 C at 1 °C/min with a two-hour hold at desired temperature in argon in the same

tube furnace as section 5.1 and 5.2.

Page 118: Carbon Rich Polymer-Derived Ceramics: Processing

101

Figure 5.30. Comparison of crosslinked polysiloxane with and without nanographene and with a reference coin.

In order to increase surface area, a slower heating rate of 0.1 °C/min was also

used for the same samples at 600 and 800 °C with a two hour hold in argon. Theses

samples were specifically tested with BET.

5.3.2 Characterization

FTIR

ATR-FTIR was conducted on crosslinked samples PHMS DVB1 and PHMS DVB1

with 1wt% graphene as seen in Figure 5.31 from 400 to 4000 cm-1. The 3000cm-1 region

is identified for C-H bonding, 2153cm-1 is for Si-H, 1600-1400 cm-1 is C-C and C=C, 1258 is

SiCH3, 1180 cm-1 is Si-CH2-Si, 1080 and 1020 cm-1 are Si-O-Si, and 900 to 500 cm-1 are SiC

related regions as previously discussed in section 5.2

There is no observable difference between the two samples since graphene is

undetectable with FTIR.[82] But it was confirmed that the graphene did not affect the

Page 119: Carbon Rich Polymer-Derived Ceramics: Processing

102

crosslinking with divinylbenzene because C=C were present meaning that the DVB was

able to crosslink both samples with and without graphene. The relative Si-H peaks were

also similar for both samples therefore the rate of hydrosilylation for the crosslinking

was the similar. Further analysis was conducted to investigate the graphene

incorporation in the samples.

Figure 5.31 ATR-FTIR of crosslinked PHMS DVB1 and PHMS DVB1 with 1wt% graphene.

Thermal Stability

TGA was run using a Netzsch instrument coupled with MS in the Brick Center in

Collaboration with Prof. Sanders and Dr. Huygen. Six crosslinked samples (PHMS DVB0,

PHMS DVB1, PHMS DVB6, PHMS DVB1 1wt% graphene, PHMS 1wt% graphene and

Page 120: Carbon Rich Polymer-Derived Ceramics: Processing

103

PHMS 5wt% graphene) were run using TGA with a 10 °C/min heating rate from 45 to

1000 °C in helium.

For DVB1 and DVB6, the TGA obtained in nitrogen, in section 5.2 (Figure 5.15) is

compared with the TGA in helium Figure 5.32 and 5.33, respectively.

Figure 5.32 TGA of crosslinked DVB1 in nitrogen from section 5.2 compared to crosslinked PHMS DVB1 in helium.

70

75

80

85

90

95

100

0 200 400 600 800 1000

PHMS DVB1

nitrogen

PHMS DVB1

helium

Page 121: Carbon Rich Polymer-Derived Ceramics: Processing

104

Figure 5.33 TGA of crosslinked DVB6 in nitrogen from section 5.2 compared to crosslinked PHMS DVB6 in helium.

Comparison of PHMS DVB1 with and without 1wt% graphene is shown in Figure

5.34 The samples with graphene were overall more thermally stable but the difference

between the samples with DVB1 and with/without graphene was less than a few

percent. Consequently, graphene does not reduce or improve the thermal stability of

the material.

0

20

40

60

80

100

0 200 400 600 800 1000

PHMS DVB6

nitrogen

PHMS DVB6

helium

Page 122: Carbon Rich Polymer-Derived Ceramics: Processing

105

Figure 5.34 TGA and DTG of crosslinked PHMS DVB1 and crosslinked PHMS DVB1-1wt%graphene with 10 °C/min heating rate from 45 to 1000 °C in helium.

For DVB1, Figure 5.32, the two TGA runs are almost identical with less than 2%

difference in weight loss at any temperature. However, the difference is significant for

DVB6 (Figure 5.33). The major source of this difference is the low temperature (< 100

°C) mass loss. Typically, this is associated with absorbed water and is not representative

of changes in the polymer composition.

Comparing all six samples (Figure 5.35), the samples without DVB had less

overall mass loss than samples with DVB. And the higher the amount of DVB the greater

the weight loss, whereas the effect of graphene was minimal.

Page 123: Carbon Rich Polymer-Derived Ceramics: Processing

106

Figure 5.35 TGA with 10 °C heating rate from 45 to 1000 °C in helium of crosslinked PHMS DVB 0, PHMS DVB1, PHMS DVB1- 1wt% graphene, PHMS 1wt% graphene, and PHMS

5wt% graphene.

MS full scans and extracted chromatograms were utilized to differentiate

between the different components of evolved gases as discussed in the next section.

TGA-MS

Thermal gravimetrical analysis coupled with mass spectrometry (TGA-MS) was

conducted at Brick Center at Clemson University, under the guidance of Dr. John

Sanders and Dr. Nathaniel Huygen. With TGA-MS, the chemical units that off gassed

during thermal decomposition of the crosslinked polymer could be observed. [75] This

Page 124: Carbon Rich Polymer-Derived Ceramics: Processing

107

gives information on the chemistry of the resultant ceramic after intermediate

temperature heat treatment. TGA-MS will also help to give insight into role of DVB in

the material. Helping to determine if DVB is decomposed early in the in pyrolysis or gets

incorporated in the Si-based polymer during crosslinking as assumed in section 5.2.

Samples were run with a 10 °C/min heating rate in helium from 45 °C up to 1000 °C.

Approximately 15mg of sample were used in a crucible that was evacuated under

vacuum then refilled with helium. The instrument that was utilized was a Netzsch TG

couple to a 403 Aëolos Quadro quadrupole mass spectrometer with Proteus® software.

This was used to determine the different volatile components that are evolved at

different temperatures decomposition during thermal treatment for six crosslinked

samples: PHMS DVB0, PHMS DVB1, PHMS DVB6, PHMS DVB1-1wt% nanographene,

PHMS 1 wt% nanographene and PHMS 5wt% nanographene.

Full scans of mass to charge (m/z) versus relative intensity were taken at the

maximum peak of the total ion chromatogram (TIC). In Figure 5.36, nine major m/z

numbers were recorded: 16, 28, 39, 51, 65,77, 91, 105 and 117. M/z of 16 is the most

prominent throughout the spectra and is associated with the loss of CH4, which also

encompasses the m/z of 12, 13,14, and 15 and is lost throughout the decomposition of

the sample. Ethane loss is associated with the m/z of 28 and smaller peaks of 26 and 27

and has a major weight loss after 500 °C. Toluene at 91m/z is another major peak

intensity and is also lost at 500 °C. 39, 51, 65, 77 and 91 are derivations of benzene. 105

and 117 can be associated with methyl and ethyl benzene as well as silanes when they

Page 125: Carbon Rich Polymer-Derived Ceramics: Processing

108

are broken down with methyl groups and oxygen as well. In Figure 5.38, total mass loss

for PHMS DVB1 is broken up into the nine components. The peaks at 105 and 117 are

relatively small compared to methane as seen in Figure 5.36 and 5.37. The peak for

water was subtracted from a background scan in cycle 1. There is no attributed mass

loss associated with water at this temperature range. The m/z values are based on

literature[75] and paired against NIST software. Figure 5.38 depicts a 3D plot of m/z vs

time vs ion current.

Figure 5.36 Full scan mass spectrum at 533.4 °C of crosslinked PHMS dvb1.

(Text File) unknown

10 20 30 40 50 60 70 80 90 100 110 120 1300

50

10016

28

39

5165 77

91

105

117

xxxx

m/z

Relative

Intensity

Page 126: Carbon Rich Polymer-Derived Ceramics: Processing

109

Figure 5.37 Extracted ion chromatogram (XIC) of PHMS DVB1 with TG plot and associated XIC of

components from Figure 5.36.

Figure 5.38 3D plot of m/z vs time vs ion current of PHMS DVB1 for TGA-MS.

Page 127: Carbon Rich Polymer-Derived Ceramics: Processing

110

Experiments on TGA-MS of PHMS DVB1 with 1% graphene was also run with the

same m/z numbers as shown in Figure 5.39. The components were then used to create

an extracted chromatogram (Figure 5.40). A 3D plot was also utilized to compare m/z,

time, and ion current (Figure 5.41).

Figure 5.39 Full scan mass spectrum at 533.5 °C of PHMS dvb1 1wt%g.

m/z

Relative

Intensity

(Text File) unknown

10 20 30 40 50 60 70 80 90 100 110 120 1300

50

10016

28

39

5165 77

91

105

117

xxxx

Page 128: Carbon Rich Polymer-Derived Ceramics: Processing

111

Figure 5. XIC of PHMS DVB1 1wt% graphene with TG plot and associated XIC of components from

Figure 5.39.

Figure 5.41 3D plot of m/z vs time vs ion current of PHMS DVB1 with 1wt% graphene for TGA-

MS.

Page 129: Carbon Rich Polymer-Derived Ceramics: Processing

112

In Figure 5.42, the thermal decomposition of PHMS DVB1 without and with 1

wt% graphene is compared. Note that the x-axis is expanded in Figure 4.42 compared to

Figure 5.37 and 5.40 to highlight the differences. The weight loss for the sample with

graphene is less than that for without graphene. However, the temperature at which

the hydrocarbons are released is the same. Although there is a higher intensity of the

lighter methyl groups (m/z =15 and 16) for the samples with graphene the intensity for

the heavier groups (m/z= 91 and 105) is higher for the samples without graphene.

Figure 5.42 Comparison XIC of PHMS DVB1 and PHMS DVB 1 1wt% graphene TG plot and

associated XIC of components from figures 5.36 and 5.39.

Page 130: Carbon Rich Polymer-Derived Ceramics: Processing

113

In Figure 5.43, for PHMS DVB1 with 1 wt% graphene, the evolution of important

hydrocarbon groups is shown. Included are methyl (m/z =15), ethyl (m/z =29), propyl

(m/z= 43), butyl (m/z = 57), pentyl (m/z= 71), and hexyl (m/z= 85). The focus is on these

groups since in the literature, investigation of the decomposition functionalized

graphene has focused on these groups. [83] As can be seen, the higher molecular weight

groups evolve early in the heat treatment (at lower temperature). Methyl both at low

and high temperature. The methyl that endures at higher temperature is most likely,

part of the crosslinking of the polymer.

Figure 5.43 XIC of PHMS DVB 1 1wt% graphene TG plot and associated XIC to compare methyl

(m/z=15), ethyl (29), propyl (43), butyl (57), pentyl (71) and hexyl (85) groups.

Page 131: Carbon Rich Polymer-Derived Ceramics: Processing

114

For samples without DVB, as mentioned above their thermal stability was higher

than those with DVB. The temperature where the most mass loss occurs for PHMS with

no DVB was shifted to approximately 730 °C compared to 500 and 530 °C for samples

with DVB. The full scan MS in Figure 5.44 reflects this shift in temperature for

decomposition. The likely explanation for this is that the hydrocarbons associated with

DVB, evolve at lower temperatures.

Figure 5.44 Full scan MS at 738.8 °C of PHMS DVB 0.

For samples without DVB, the higher temperature for thermal decomposition

resulted in the decomposition components only relating to methane derivatives as

demonstrated in Figure 5.44 and 5.45.

(Text File) unknown

0 10 20 30 40 500

50

10016

18

19 27 31 44

xxxx

m/z

Relative

Intensity

Page 132: Carbon Rich Polymer-Derived Ceramics: Processing

115

Figure 5.45 XIC of PHMS DVB0 with TG plot and associated XIC of components from Figure 5.44.

In contrast, samples with the most DVB at DVB6 had a much lower peak

decomposition point at around 500 °C and more decomposition species as

demonstrated in Figure 5.46 and 5.47. M/z 91, 105 and 117 all increased significantly

compared to DVB1 and are related to the divinyl benzene components, which degrade

at a lower temperature than hydrocarbons associated with PHMS. A 3D plot (Figure

5.48) was also investigated to show the increase of benzene derivatives across different

time and m/z.

Page 133: Carbon Rich Polymer-Derived Ceramics: Processing

116

Figure 5.46 Full scan MS at 504.7 °C of PHMS DVB6.

Figure 5.47 XIC of PHMS DVB6 with TG plot and associated XIC from Figure 5.46.

(Text File) unknown

10 20 30 40 50 60 70 80 90 100 110 120 130 1400

50

100

15 27

39

44

5165

74

77

91

105

117

xxxx

m/z

Relative

Intensity

Page 134: Carbon Rich Polymer-Derived Ceramics: Processing

117

Figure 5.48 3D plot of m/z vs time vs ion current of PHMS DVB6 for TGA-MS.

The last group of samples that were investigated were those with graphene but

no DVB. These samples again had a much higher peak thermal degradation temperature

of over 700 °C. MS scans were taken at 745.3 °C as demonstrated in Figure 5.49.

Figure 5.49 PHMS Full scan MS at 745.3 °C of PHMS 1wt% graphene.

(Text File) unknown

0 10 20 30 40 500

50

10016

18

19 26 44

xxxx

m/z

Relative

Intensity

Page 135: Carbon Rich Polymer-Derived Ceramics: Processing

118

Figure 5.50 shows only the derivatives associated with methane. Which from previous

samples without DVB is expected. A 3D plot was produced in Figure 5.51 to show that

minimal amount of different m/z across different time components.

Figure 5.50 XIC of PHMS 1wt% graphene with TG plot from associated XIC from Figure 5.49.

Figure 5.51 3D plot of m/z vs time vs ion current of PHMS with 1wt% graphene for TGA-MS.

Page 136: Carbon Rich Polymer-Derived Ceramics: Processing

119

Different weight percent of graphene from 1 to 5 wt% were investigated

(Figure 5.52). And found to have minimal differences between them.

Figure 5.52 Comparison XIC of PHMS with 1wt% and 5wt% graphene with TG plot and associated

XIC of components from Figures 5.49.

Because of the interest in carbon with the prepared samples and the TGA-MS associated

with different carbon groups, other forms of chemical analysis were conducted to

investigate the carbon-carbon bonding and nature of carbon as reported in the

following sections.

Page 137: Carbon Rich Polymer-Derived Ceramics: Processing

120

Raman Spectroscopy

Raman spectroscopy was conducted to determine the carbon-carbon bonding

and the influence of graphene on the bonding structure. The spectroscopy was done

with a 532nm laser with 1% laser power and 0.01% laser power for samples with only 1

and 5wt% graphene. Samples were tested with PHMS combined with 1 and 5wt%

graphene that were pyrolyzed at 800 and 900 °C. The samples without DVB experienced

fluorescence and needed a much lower percentage of laser power to be able to obtain

spectra as seen in Figure 5.53.

Figure 5.53 Raman spectrum of 900 °C with 1 wt% graphene with 0.1% laser power.

The crosslinked sample (PHMS DVB1 with 1wt% graphene) was also investigated

as seen in Figure 5.54. The 600-690cm-1 region is associated with SiC3. From 750 to 900

0

5000

10000

15000

20000

25000

30000

35000

40000

500 1000 1500 2000 2500 3000

Inte

nsi

ty

Raman shift (cm-1)

900C 1wt%G

Page 138: Carbon Rich Polymer-Derived Ceramics: Processing

121

cm-1 is CH3 rocking. Symmetric and asymmetric deformation of CH3 are 1250 and 1400

to 1450 cm-1 whereas symmetric and asymmetric stretches are 2900 and 2960 cm-1.[84]

Figure 5.54 Raman of crosslinked PHMS dvb1 with 1 wt% of graphene.

PHMS DVB1 samples pyrolyzed at 700 and 800° C with and without 1wt%

graphene were compared. All samples had a blank background subtraction, baseline

correction and normalized to the “G” peak. Figure 5.55 and Figure 5.56 compares the

samples with and without 1 wt% graphene.

Comparing the D peak (Figure 5.55a) in the 700 °C samples, shows that sample

with 1wt% graphene has sharper D peak. Deconvolution would be needed to determine

the LD/LG ratio. There are also D’ peaks and 2D peaks within the 2500 to 3000 cm-1

region, which can refer either to the disorder in the system or graphitic regions.

-0.1

0.1

0.3

0.5

0.7

0.9

1.1

500 1000 1500 2000 2500 3000

No

rmal

ized

Inte

nsi

ty

Raman shift (cm-1)

Page 139: Carbon Rich Polymer-Derived Ceramics: Processing

122

For the 800 °C samples the opposite effect occurs (Figure 5.55b), when

comparing with and without graphene. The 800 °C dvb1 sample without 1wt% has a

sharper D peak.

Figure 5.55 Raman spectrum comparing pyrolysis temperature and graphene content of a) 700 dvb1 1wt% graphene and 700 dvb1 b) 800 C dvb1 with 1 wt% graphene and 800 C dvb1.

Most likely the at 800 °C samples are to a more crystalline material but not

completely crystalline as at 900 °C and above as demonstrated in section 5.2. All four

samples are compared in Figure 5.56. The samples without graphene transform from

amorphous to nano graphitic carbon as demonstrated in Figure 5.56. There are no

observable changes in graphene in this temperature range.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

500 1000 1500 2000 2500 3000

700C dvb1 1wt%G

700C dvb1

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

500 1000 1500 2000 2500 3000

800C dvb1 1wt%G

800C dvb1

a b

Page 140: Carbon Rich Polymer-Derived Ceramics: Processing

123

Figure 5.56 Raman spectrum of PHMS DVB1 with and without 1wt% graphene pyrolyzed at 700 and 800 °C.

The next step is to deconvolute the Raman spectra as was done in section 5.2.

5.3.3 Summary and Discussion

Polysiloxane materials with nanographene were pyrolyzed at intermediate/low

temperatures of 900 to 600 °C. Chemical analysis was used to determine bonding

mechanisms for crosslinking and carbon-carbon bonding by using FTIR and Raman,

respectively. FTIR was unable to distinguish from polysiloxane with DVB and

polysiloxane with DVB and nanographene. But it was confirmed that they had the same

crosslinking mechanism. Raman demonstrated that graphene had a slight effect on the

nature of carbon in the low temperature pyrolyzed ceramics.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

500 1000 1500 2000 2500 3000

700C dvb1 1wt%G

700C dvb1

800C dvb1 1wt%G

800C dvb1

Page 141: Carbon Rich Polymer-Derived Ceramics: Processing

124

Thermal stability and degradation were investigated for the samples produced.

TGA coupled with MS was utilized to examine thermal stability of the samples and the

evolution of volatile components during heat treatment. From TGA-MS, it was

determined that the samples with more DVB had greater weight loss as was also found

in section 5.2. From the TGA-MS, sample PHMS DVB6 had maximum weight loss

occurring at approximately 500 °C and the m/z values corresponded to divinylbenzene

derivatives. Whereas sample PHMS DVB1, had a maximum weight loss at approximately

530 °C. Samples with DVB and nanographene had similar TGA profiles and TGA-MS

demonstrated that samples with nanographene had more weight loss associated with

methane whereas PHMS DVB1 had more weight loss of benzene derivatives compared

to PHMS DVB1 with nanographene. Previous research[83] have explained TGA-MS with

graphene by the loss of methyl, propyl, butyl, etc. groups as demonstrated in Figure

5.43. However, samples without DVB had a much higher thermal degradation

temperature of approximately 700 °C and minimal weight loss overall. The weight loss in

these samples is, primarily, due to the loss of methyl derivatives.

Page 142: Carbon Rich Polymer-Derived Ceramics: Processing

125

Chapter 6

Summary and Concluding Remarks

This study has focused on the fundamental understanding of ceramics obtained

from pyrolysis of polymer precursors at low/intermediate temperatures. Specific

objectives were: (1) to study the chemistry and structure of low/intermediate

temperature pyrolysis polymer derived ceramics with high carbon content, (2)

investigate the properties including electrical and electrochemical properties and (3)

design material for application in Li-ion batteries and gas separation membranes.

Processing protocols were developed and a series of low/intermediate

temperature pyrolyzed ceramics with variable carbon content were made. The

chemistry of these “ceramers” was characterized thoroughly. The properties and

potential applications of materials investigated in this research are shown in Figure 6.1 .

Page 143: Carbon Rich Polymer-Derived Ceramics: Processing

126

Figure 6.1 Overview for Organizing Research.

Polycarbosilane and its modification are discussed in Chapter 4. The first set of

modifications was the addition of DVB. The thermal stability of materials with varying

DVB content produced by pyrolysis at different temperatures was investigated. It was

shown that the addition of DVB decreased the thermal stability of the materials. FTIR

and Raman were used to investigate the effect of EVB additions and thermal treatment

on the chemical bonds. These investigations indicated the presence of oxygen in the

samples.

Solid state NMR and elemental analysis was conducted to confirm the presence

of oxygen and the formation of Si-O bonds.

In the next set of experiments, the focus was on reducing the oxygen content by

using Schlenk technique. In addition, polycarbosilane was modified with boron and

Page 144: Carbon Rich Polymer-Derived Ceramics: Processing

127

nitrogen incorporation since both of these are proven to improve the thermal stability

of the resultant material. FTIR, Raman and elemental analysis were used to investigate

the effect of pyrolysis temperature on the chemical structure. It was confirmed that the

oxygen content was low and boron successfully incorporated. However, the samples did

not have nitrogen. Research is continuing, in collaboration with Prof. Bernard in

Limoges, France to use ammonia as the gas during pyrolysis to incorporate nitrogen.

Additional characterization including surface area measurements and thermal stability

studies, will be conducted on samples with and without nitrogen.

Polysiloxanes and their modifications with DVB are discussed in Chapter 5. The

effect of the polymer chemistry on the thermal stability was investigated by studying

the thermal stability of ceramics obtained from the low/intermediate temperature

pyrolysis of two different precursors: polymethylsiloxane (MK) and

polymethylphenylsiloxane (H44). It was shown that the ceramics derived from MK have

higher thermal stability. FTIR was used to investigate the chemical bonding in the low

temperature pyrolyzed ceramics. High surface area, greater than 400 m2/g microporous

samples were produced from MK pyrolyzed at 500 °C.

Next the chemistry, thermal stability and electrical properties of PHMS-DVB

blended/hybrid polymers was investigated as a function of stoichiometry and pyrolysis

temperature. Thermal stability was determined by TGA, in which, the ceramics retained

stability with minimal weight loss above 600 °C. Chemical changes also occurred due to

Page 145: Carbon Rich Polymer-Derived Ceramics: Processing

128

the change in pyrolysis temperature and stoichiometry as seen in FTIR, Raman and XPS.

Overall, the increase in pyrolysis temperature showed an increase in the I(D)/I(G) ratio

as seen in Raman. XPS was able to confirm the presence of both sp2 and sp3 carbon

bonding and carbon-oxygen bonding. The DVB-PHMS hybrid precursor leads to a

composite SiOC-C composite. It is shown that the amount of DVB that can be

incorporated in the formation of the hybrid polymer, during crosslinking, is limited to

around two times the weight of PHMS. Both the increased temperature and carbon

content increased the electrical conductivity as a result, an intrinsically insulating SiOC is

transformed to a semiconductor due to the addition of free carbon that undergoes a

structural change as pyrolysis temperature increases. These results show the promise of

being able to make semiconducting SiOC with increased carbon at low processing

temperature which are expected to result in high levels of microporosity and surface

area. Further investigations are needed for to determine the extent of porosity in these

materials. In addition, the nature of the carbon derived from the hybrid polymer at low

temperatures for amorphous ceramics and that from unreacted DVB needs to be

investigated. This was later investigated in Chapter 5.3 from TGA-MS.

Finally, to further increase the electrical conductivity, especially in materials

pyrolyzed at low temperatures, nanographene was added. TGA coupled with mass

spectrometry was used to investigate the thermal stability of these materials and the

chemical structure was investigated using FTIR and Raman. These materials are being

currently investigated as summarized Chapter 7. The surface chemistry is being

Page 146: Carbon Rich Polymer-Derived Ceramics: Processing

129

investigated using ToF-SIMS and XPS. In addition, the surface area and micropore size is

being investigated. Lastly, the performance of these materials as Li-ion battery anodes is

being investigated as discussed in Chapter 7.

Page 147: Carbon Rich Polymer-Derived Ceramics: Processing

130

Chapter 7

Suggested Future Work

7.1 Time-of-Flight Secondary Ion Mass Spectrometry (ToF-SIMS)

In order to analyze the chemical structure of the polymer and ceramics surfaces,

time-of-flight secondary ion mass spectrometry (ToF-SIMS) analysis experiments are

being conducted. A proposal to do these tests at Oak Ridge National Labs (ORNL) was

approved. These tests are now being conducted by Dr. Anton Ievlev at ORNL. A ToF-

SIMS 5-NSC instrument (Ion-ToF Gmb, Germany) is being used to characterize the

surface chemistry with a Bi3+ ion beam of ~5µm, 20keV energy and current of 30nA.

[85,86] A 2D plot pf the surface chemistry will be the result. Experiments are being done

on crosslinked and samples pyrolyzed at 800 °C for PHMS DVB1 with and without 1wt%

graphene.

7.2 Surface Area Analysis (BET)

Brunauer-Emmett-Teller method (BET) is being done at the Advanced Material

Research Laboratory (AMRL) at Clemson University with Dr. Kelliann Koehler to find

specific surface area and meso/micropore analysis. A degas cycle is being run first up to

150 °C for 12 hours then surface area and micropore analysis will be conducted using a

Quantachrome instrument. Samples will then be run under absorption/ desorption

Page 148: Carbon Rich Polymer-Derived Ceramics: Processing

131

measurements at 77K in nitrogen to obtain micropore volume and size. These tests are

being done on samples pyrolyzed at 800 °C and 600 °C with a slower heating rate of 0.1

°C/min.

7.3 Anodes for Li-ion Batteries

Selected polysiloxane samples, prepared as discussed in Chapter 5, are being

evaluated as anode materials of liquid electrolyte Li-ion batteries.

7.3.1 Battery Tape Preparation

The selected polysiloxane samples were ground using a planetary ball mill for 24

hours and are being used as active material. Materials were grounded then combined

with 85wt% sample (active material), 10wt% binder polyvinylidene fluoride (PVdF,

Sigma Aldrich) and 5wt% Carbon black Super P® with equal weight percent of material

and of N-methyl-2pyrrolidone (NMP) solvent to create a slurry. The slurry was mixed for

several hours then tape cast onto aluminum and copper foil at 0.1mm thickness using a

doctor blade. The tapes were them dried in a vacuum oven at 80 °C for 24 hours to

eliminate any excess solvent. The dried tapes were then punched into 12mm diameter

discs. The punched disks were then taken into an argon glovebox for battery assembly.

Page 149: Carbon Rich Polymer-Derived Ceramics: Processing

132

7.3.2 Battery Cell Assembly

Half cells were assembled as seen in below Figure 6.1. The materials sequence is:

top case; stainless steel spring; 15.5mm stainless steel spacer; 12mm diameter cut

cathode sample (as described from above section); electrolyte (lithium

hexafluorophosphate solution in ethylene carbonate and dimethyl carbonate, 1.0 M

LiPF6 in EC/DMC=50/50 (v/v), battery grade); Celgard Li-ion battery separator cut to

18mm diameter; Li-anode; 15.5 mm stainless steel spacer; and bottom case. All

materials were purchased from MTI corporation with the exception of the electrolyte

which was purchased from Sigma Aldrich. Half cells are then crimped using a hydraulic

press and used for testing. A 8-channel battery analyzer ((BST8-MA, MTI.corp) (Figure

6.2).

Figure 7.1 Diagram of battery assembly.

Page 150: Carbon Rich Polymer-Derived Ceramics: Processing

133

Figure 7.2 Picture of 8 channel battery analyzer.

The materials being investigated are PHMS with varying amounts of DVB and

pyrolysis temperatures along with samples with graphene. The planned tests are cyclic

voltammetry and charge-discharge runs at different C-rates.

7.4 Other Suggested Future Work

This research was designed to be a fundamental study of polymer derived

ceramics with high carbon pyrolyzed at low/intermediate temperatures. Processing and

characterization of the polymer derived ceramics with added carbon with different

polymer precursors such as polycarbosilane, polycarbosilane with modified boron,

polysiloxane and polysiloxane with nanographene have been completed.

The following topics should also be investigated:

Page 151: Carbon Rich Polymer-Derived Ceramics: Processing

134

• Microstructural analysis of ceramics

• Investigation of the properties of low temperature pyrolyzed nitrogen and boron

modified polycarbosilanes.

• Possibility of using the developed materials as gas separation membranes

Page 152: Carbon Rich Polymer-Derived Ceramics: Processing

135

References

[1] P. Colombo, G. Mera, R. Riedel, G.D. Sorarù, Polymer-derived ceramics: 40 Years of research and innovation in advanced ceramics, J. Am. Ceram. Soc. (2010). https://doi.org/10.1111/j.1551-2916.2010.03876.x.

[2] Q. Wen, Z. Yu, R. Riedel, The fate and role of in situ formed carbon in polymer-derived ceramics, Prog. Mater. Sci. (2020). https://doi.org/10.1016/j.pmatsci.2019.100623.

[3] F. Dalcanale, J. Grossenbacher, G. Blugan, M.R. Gullo, A. Lauria, J. Brugger, H. Tevaearai, T. Graule, M. Niederberger, J. Kuebler, Influence of carbon enrichment on electrical conductivity and processing of polycarbosilane derived ceramic for MEMS applications, J. Eur. Ceram. Soc. (2014). https://doi.org/10.1016/j.jeurceramsoc.2014.06.002.

[4] M. Schmidt, C. Durif, E.D. Acosta, C. Salameh, H. Plaisantin, P. Miele, R. Backov, R. Machado, C. Gervais, J.G. Alauzun, G. Chollon, S. Bernard, Molecular-Level Processing of Si-(B)-C Materials with Tailored Nano/Microstructures, Chem. - A Eur. J. (2017). https://doi.org/10.1002/chem.201703674.

[5] T. Konegger, C.C. Tsai, H. Peterlik, S.E. Creager, R.K. Bordia, Asymmetric polysilazane-derived ceramic structures with multiscalar porosity for membrane applications, Microporous Mesoporous Mater. (2016). https://doi.org/10.1016/j.micromeso.2016.06.027.

[6] J.C. Vickerman, I.S. Gilmore, Surface Analysis - The Principal Techniques: Second Edition, 2009. https://doi.org/10.1002/9780470721582.

[7] S. Kaur, G. Mera, R. Riedel, E. Ionescu, Effect of boron incorporation on the phase composition and high-temperature behavior of polymer-derived silicon carbide, J. Eur. Ceram. Soc. (2016). https://doi.org/10.1016/j.jeurceramsoc.2015.11.037.

[8] G. Socrates, Infrared and Raman characteristic group frequencies. Tables and charts, 2001.

[9] B. V Manoj Kumar, Y.-W. Kim, Processing of polysiloxane-derived porous ceramics: a review, Sci. Technol. Adv. Mater. (2010). https://doi.org/10.1088/1468-6996/11/6/069801.

[10] R. Harshe, C. Balan, R. Riedel, Amorphous Si(Al)OC ceramic from polysiloxanes: Bulk ceramic processing, crystallization behavior and applications, J. Eur. Ceram. Soc. (2004). https://doi.org/10.1016/j.jeurceramsoc.2003.10.016.

Page 153: Carbon Rich Polymer-Derived Ceramics: Processing

136

[11] Y.W. Kim, J.H. Eom, C. Wang, C.B. Park, Processing of porous silicon carbide ceramics from carbon-filled polysiloxane by extrusion and carbothermal reduction, J. Am. Ceram. Soc. (2008). https://doi.org/10.1111/j.1551-2916.2008.02280.x.

[12] Z.C. Eckel, C. Zhou, J.H. Martin, A.J. Jacobsen, W.B. Carter, T.A. Schaedler, Additive manufacturing of polymer-derived ceramics, Science (80-. ). (2016). https://doi.org/10.1126/science.aad2688.

[13] G. Liu, J. Kaspar, L.M. Reinold, M. Graczyk-Zajac, R. Riedel, Electrochemical performance of DVB-modified SiOC and SiCN polymer-derived negative electrodes for lithium-ion batteries, Electrochim. Acta. (2013). https://doi.org/10.1016/j.electacta.2013.05.064.

[14] S. Yajima, J. Hayashi, J. Omori, K. Okamura, Development of a Silicon carbide fibre with high tensile strength, Nature. (1976). https://doi.org/10.1038/261683a0.

[15] S. Yajima, J. Hayashi, M. Omori, CONTINUOUS SILICON CARBIDE FIBER OF HIGH TENSILE STRENGTH, Chem. Lett. (1975). https://doi.org/10.1246/cl.1975.931.

[16] S. YAJIMA, K. OKAMURA, J. HAYASHI, M. OMORI, Synthesis of Continuous Sic Fibers with High Tensile Strength, J. Am. Ceram. Soc. (1976). https://doi.org/10.1111/j.1151-2916.1976.tb10975.x.

[17] P. Greil, Polymer derived engineering ceramics, Adv. Eng. Mater. (2000). https://doi.org/10.1002/1527-2648(200006)2:6<339::AID-ADEM339>3.0.CO;2-K.

[18] R. Jones, A. Szweda, D. Petrak, Polymer derived ceramic matrix composites, Compos. Part A Appl. Sci. Manuf. (1999). https://doi.org/10.1016/S1359-835X(98)00151-1.

[19] D. King, Z. Apostolov, T. Key, C. Carney, M. Cinibulk, Novel processing approach to polymer-derived ceramic matrix composites, Int. J. Appl. Ceram. Technol. (2018). https://doi.org/10.1111/ijac.12805.

[20] M. Balestrat, E. Diz Acosta, O. Hanzel, N. Tessier-Doyen, R. Machado, P. Šajgalík, Z. Lenčéš, S. Bernard, Additive-free low temperature sintering of amorphous Si[sbnd]B[sbnd]C powders derived from boron-modified polycarbosilanes: Toward the design of SiC with tunable mechanical, electrical and thermal properties, J. Eur. Ceram. Soc. (2020). https://doi.org/10.1016/j.jeurceramsoc.2019.12.037.

[21] C. Stabler, E. Ionescu, M. Graczyk-Zajac, I. Gonzalo-Juan, R. Riedel, Silicon oxycarbide glasses and glass-ceramics: “All-Rounder” materials for advanced structural and functional applications, J. Am. Ceram. Soc. (2018). https://doi.org/10.1111/jace.15932.

Page 154: Carbon Rich Polymer-Derived Ceramics: Processing

137

[22] H.J. Kleebe, Y.D. Blum, SiOC ceramic with high excess free carbon, J. Eur. Ceram. Soc. (2008). https://doi.org/10.1016/j.jeurceramsoc.2007.09.024.

[23] Y.D. Blum, D.B. MacQueen, H.J. Kleebe, Synthesis and characterization of carbon-enriched silicon oxycarbides, J. Eur. Ceram. Soc. (2005). https://doi.org/10.1016/j.jeurceramsoc.2004.07.019.

[24] M. Schubert, M. Wilhelm, S. Bragulla, C. Sun, S. Neumann, T.M. Gesing, P. Pfeifer, K. Rezwan, M. Bäumer, The Influence of the Pyrolysis Temperature on the Material Properties of Cobalt and Nickel Containing Precursor Derived Ceramics and their Catalytic Use for CO2 Methanation and Fischer–Tropsch Synthesis, Catal. Letters. (2017). https://doi.org/10.1007/s10562-016-1919-y.

[25] M. Černý, Z. Chlup, A. Strachota, M. Halasová, Š. Rýglová, J. Schweigstillová, J. Svítilová, M. Havelcová, Changes in structure and in mechanical properties during the pyrolysis conversion of crosslinked polymethylsiloxane and polymethylphenylsiloxane resins to silicon oxycarbide glass, Ceram. Int. (2015). https://doi.org/10.1016/j.ceramint.2015.01.034.

[26] H. Schmidt, D. Koch, G. Grathwohl, P. Colombo, Micro-/Macroporous Ceramics from Preceramic Precursors, J. Am. Ceram. Soc. (2001). https://doi.org/10.1111/j.1151-2916.2001.tb00997.x.

[27] R.M. Morcos, A. Navrotsky, T. Varga, Y. Blum, D. Ahn, F. Poli, K. Müller, R. Raj, Energetics of SixOyCz polymer-derived ceramics prepared under varying conditions, J. Am. Ceram. Soc. (2008). https://doi.org/10.1111/j.1551-2916.2008.02543.x.

[28] R. Chavez, E. Ionescu, C. Balan, C. Fasel, R. Riedel, Effect of ambient atmosphere on crosslinking of polysilazanes, J. Appl. Polym. Sci. (2011). https://doi.org/10.1002/app.32777.

[29] K.T. Strong, S.A. Arreguin, R.K. Bordia, Controlled atmosphere pyrolysis of polyureasilazane for tailored volume fraction Si3N4/SiC nanocomposites powders, J. Eur. Ceram. Soc. (2016). https://doi.org/10.1016/j.jeurceramsoc.2016.03.015.

[30] A.H. Tavakoli, M.M. Armentrout, M. Narisawa, S. Sen, A. Navrotsky, White Si-O-C ceramic: Structure and thermodynamic stability, J. Am. Ceram. Soc. (2015). https://doi.org/10.1111/jace.13233.

[31] T. Konegger, J. Torrey, O. Flores, T. Fey, B. Ceron-Nicolat, G. Motz, F. Scheffler, M. Scheffler, P. Greil, R.K. Bordia, Ceramics for Sustainable Energy Technologies with a Focus on Polymer-Derived Ceramics, in: Nov. Combust. Concepts Sustain. Energy Dev., 2014. https://doi.org/10.1007/978-81-322-2211-8_22.

Page 155: Carbon Rich Polymer-Derived Ceramics: Processing

138

[32] P. Colombo, Cellular ceramics with hierarchical porosity from preceramic polymers, in: IOP Conf. Ser. Mater. Sci. Eng., 2011. https://doi.org/10.1088/1757-899X/18/1/012002.

[33] G.D. Sorarù, F. Dalcanale, R. Campostrini, A. Gaston, Y. Blum, S. Carturan, P.R. Aravind, Novel polysiloxane and polycarbosilane aerogels via hydrosilylation of preceramic polymers, J. Mater. Chem. (2012). https://doi.org/10.1039/c2jm00020b.

[34] C. Stabler, A. Reitz, P. Stein, B. Albert, R. Riedel, E. Ionescu, Thermal properties of SiOC glasses and glass ceramics at elevated temperatures, Materials (Basel). (2018). https://doi.org/10.3390/ma11020279.

[35] G. Mera, A. Navrotsky, S. Sen, H.J. Kleebe, R. Riedel, Polymer-derived SiCN and SiOC ceramics-structure and energetics at the nanoscale, J. Mater. Chem. A. (2013). https://doi.org/10.1039/c2ta00727d.

[36] G. Barroso, Q. Li, R.K. Bordia, G. Motz, Polymeric and ceramic silicon-based coatings-a review, J. Mater. Chem. A. (2019). https://doi.org/10.1039/c8ta09054h.

[37] K. Wang, M. Günthner, G. Motz, R.K. Bordia, High performance environmental barrier coatings, Part II: Active filler loaded SiOC system for superalloys, J. Eur. Ceram. Soc. (2011). https://doi.org/10.1016/j.jeurceramsoc.2011.05.047.

[38] K. Wang, J. Unger, J.D. Torrey, B.D. Flinn, R.K. Bordia, Corrosion resistant polymer derived ceramic composite environmental barrier coatings, J. Eur. Ceram. Soc. (2014). https://doi.org/10.1016/j.jeurceramsoc.2014.05.036.

[39] J. Cordelair, P. Greil, Electrical conductivity measurements as a microprobe for structure transitions in polysiloxane derived Si-O-C ceramics, J. Eur. Ceram. Soc. (2000). https://doi.org/10.1016/S0955-2219(00)00068-6.

[40] J. Kaspar, M. Graczyk-Zajac, S. Choudhury, R. Riedel, Impact of the electrical conductivity on the lithium capacity of polymer-derived silicon oxycarbide (SiOC) ceramics, Electrochim. Acta. (2016). https://doi.org/10.1016/j.electacta.2016.08.121.

[41] M. Graczyk-Zajac, G. Mera, J. Kaspar, R. Riedel, Electrochemical studies of carbon-rich polymer-derived SiCN ceramics as anode materials for lithium-ion batteries, J. Eur. Ceram. Soc. (2010). https://doi.org/10.1016/j.jeurceramsoc.2010.07.010.

[42] V.S. Pradeep, D.G. Ayana, M. Graczyk-Zajac, G.D. Soraru, R. Riedel, High Rate Capability of SiOC Ceramic Aerogels with Tailored Porosity as Anode Materials for Li-ion Batteries, Electrochim. Acta. (2015). https://doi.org/10.1016/j.electacta.2015.01.088.

Page 156: Carbon Rich Polymer-Derived Ceramics: Processing

139

[43] H. Sun, K. Zhao, Atomistic Origins of High Capacity and High Structural Stability of Polymer-Derived SiOC Anode Materials, ACS Appl. Mater. Interfaces. (2017). https://doi.org/10.1021/acsami.7b10906.

[44] G.D. Sorarù, L. Kundanati, B. Santhosh, N. Pugno, Influence of free carbon on the Young’s modulus and hardness of polymer-derived silicon oxycarbide glasses, J. Am. Ceram. Soc. (2019). https://doi.org/10.1111/jace.16131.

[45] P. Dibandjo, S. Diré, F. Babonneau, G.D. Soraru, Influence of the polymer architecture on the high temperature behavior of SiCO glasses: A comparison between linear- and cyclic-derived precursors, J. Non. Cryst. Solids. (2010). https://doi.org/10.1016/j.jnoncrysol.2009.10.006.

[46] C. Shen, E. Barrios, L. Zhai, Bulk Polymer-Derived Ceramic Composites of Graphene Oxide, ACS Omega. (2018). https://doi.org/10.1021/acsomega.8b00492.

[47] R. Sujith, P.K. Chauhan, J. Gangadhar, A. Maheshwari, Graphene nanoplatelets as nanofillers in mesoporous silicon oxycarbide polymer derived ceramics, Sci. Rep. (2018). https://doi.org/10.1038/s41598-018-36080-1.

[48] A. Viard, D. Fonblanc, D. Lopez-Ferber, M. Schmidt, A. Lale, C. Durif, M. Balestrat, F. Rossignol, M. Weinmann, R. Riedel, S. Bernard, Polymer Derived Si–B–C–N Ceramics: 30 Years of Research, Adv. Eng. Mater. (2018). https://doi.org/10.1002/adem.201800360.

[49] M. Wynn, D. Lopez-Ferber, A. Viard, D. Fonblanc, M. Schmidt, F. Rossignol, P. Carles, G. Chollon, C. Gervais, S. Bernard, Tuning of the high temperature behaviour of Si-C-N ceramics via the chemical crosslinking of poly(vinylmethyl-co-methyl)silazanes with controlled borane contents, Open Ceram. (2021) 100055. https://doi.org/10.1016/J.OCERAM.2021.100055.

[50] S. Bernard, M. Weinmann, P. Gerstel, P. Miele, F. Aldinger, Boron-modified polysilazane as a novel single-source precursor for SiBCN ceramic fibers: Synthesis, melt-spinning, curing and ceramic conversion, J. Mater. Chem. (2005). https://doi.org/10.1039/b408295h.

[51] O. Majoulet, J.G. Alauzun, L. Gottardo, C. Gervais, M.E. Schuster, S. Bernard, P. Miele, Ordered mesoporous silicoboron carbonitride ceramics from boron-modified polysilazanes: Polymer synthesis, processing and properties, Microporous Mesoporous Mater. (2011). https://doi.org/10.1016/j.micromeso.2010.09.008.

Page 157: Carbon Rich Polymer-Derived Ceramics: Processing

140

[52] Z.D. Apostolov, E.P. Heckman, T.S. Key, M.K. Cinibulk, Effects of low-temperature treatment on the properties of commercial preceramic polymers, J. Eur. Ceram. Soc. (2020). https://doi.org/10.1016/j.jeurceramsoc.2020.02.030.

[53] M. Wilhelm, C. Soltmann, D. Koch, G. Grathwohl, Ceramers - Functional materials for adsorption techniques, J. Eur. Ceram. Soc. 25 (2005) 271–276. https://doi.org/10.1016/j.jeurceramsoc.2004.08.008.

[54] M. Adam, M. Bäumer, M. Schowalter, J. Birkenstock, M. Wilhelm, G. Grathwohl, Generation of Pt- and Pt/Zn-containing ceramers and their structuring as macro/microporous foams, Chem. Eng. J. (2014). https://doi.org/10.1016/j.cej.2014.02.063.

[55] M. Adam, S. Kocanis, T. Fey, M. Wilhelm, G. Grathwohl, Hierarchically ordered foams derived from polysiloxanes with catalytically active coatings, J. Eur. Ceram. Soc. (2014). https://doi.org/10.1016/j.jeurceramsoc.2013.12.011.

[56] M. Adam, M. Wilhelm, G. Grathwohl, Polysiloxane derived hybrid ceramics with nanodispersed Pt, Microporous Mesoporous Mater. (2012). https://doi.org/10.1016/j.micromeso.2011.10.037.

[57] T. Prenzel, T.L.M. Guedes, F. Schlüter, M. Wilhelm, K. Rezwan, Tailoring surfaces of hybrid ceramics for gas adsorption - From alkanes to CO2, Sep. Purif. Technol. (2014). https://doi.org/10.1016/j.seppur.2014.03.029.

[58] P. Sonström, M. Adam, X. Wang, M. Wilhelm, G. Grathwohl, M. Bäumer, Colloidal nanoparticles embedded in ceramers: Toward structurally designed catalysts, J. Phys. Chem. C. (2010). https://doi.org/10.1021/jp1058897.

[59] M. Wilhelm, M. Adam, M. Bäumer, G. Grathwohl, Synthesis and properties of porous hybrid materials containing metallic nanoparticles, in: Adv. Eng. Mater., 2008. https://doi.org/10.1002/adem.200800019.

[60] Y. Grebenyuk, H.X. Zhang, M. Wilhelm, K. Rezwan, M.E. Dreyer, Wicking into porous polymer-derived ceramic monoliths fabricated by freeze-casting, J. Eur. Ceram. Soc. (2017). https://doi.org/10.1016/j.jeurceramsoc.2016.11.049.

[61] H. Zhang, P. D’Angelo Nunes, M. Wilhelm, K. Rezwan, Hierarchically ordered micro/meso/macroporous polymer-derived ceramic monoliths fabricated by freeze-casting, J. Eur. Ceram. Soc. (2016). https://doi.org/10.1016/j.jeurceramsoc.2015.09.018.

[62] T. Prenzel, M. Wilhelm, K. Rezwan, Pyrolyzed polysiloxane membranes with tailorable hydrophobicity, porosity and high specific surface area, Microporous Mesoporous Mater. (2013). https://doi.org/10.1016/j.micromeso.2012.10.014.

Page 158: Carbon Rich Polymer-Derived Ceramics: Processing

141

[63] T. Prenzel, M. Wilhelm, K. Rezwan, Tailoring amine functionalized hybrid ceramics to control CO2 adsorption, Chem. Eng. J. (2014). https://doi.org/10.1016/j.cej.2013.09.039.

[64] C. Harms, M. Wilhelm, G. Grathwohl, Influence of PDMS chain length on proton conductivity in polysiloxane based membranes for HT-PEMFC application, J. Memb. Sci. (2011). https://doi.org/10.1016/j.memsci.2011.08.049.

[65] M. Jeske, C. Soltmann, C. Ellenberg, M. Wilhelm, D. Koch, G. Grathwohl, Proton conducting membranes for the high temperature-polymer electrolyte membrane-fuel cell (HT-PEMFC) based on functionalized polysiloxanes, Fuel Cells. (2007). https://doi.org/10.1002/fuce.200500226.

[66] K. Lu, Porous and high surface area silicon oxycarbide-based materials - A review, Mater. Sci. Eng. R Reports. (2015). https://doi.org/10.1016/j.mser.2015.09.001.

[67] J. Kaspar, M. Graczyk-Zajac, R. Riedel, Lithium insertion into carbon-rich SiOC ceramics: Influence of pyrolysis temperature on electrochemical properties, J. Power Sources. (2013). https://doi.org/10.1016/j.jpowsour.2012.11.086.

[68] L.M. Reinold, Y. Yamada, M. Graczyk-Zajac, H. Munakata, K. Kanamura, R. Riedel, The influence of the pyrolysis temperature on the electrochemical behavior of carbon-rich SiCN polymer-derived ceramics as anode materials in lithium-ion batteries, J. Power Sources. (2015). https://doi.org/10.1016/j.jpowsour.2015.02.074.

[69] J. Kaspar, M. Graczyk-Zajac, S. Lauterbach, H.J. Kleebe, R. Riedel, Silicon oxycarbide/nano-silicon composite anodes for Li-ion batteries: Considerable influence of nano-crystalline vs. nano-amorphous silicon embedment on the electrochemical properties, J. Power Sources. (2014). https://doi.org/10.1016/j.jpowsour.2014.06.089.

[70] M. Graczyk-Zajac, L. Toma, C. Fasel, R. Riedel, Carbon-rich SiOC anodes for lithium-ion batteries: Part I. Influence of material UV-pre-treatment on high power properties, in: Solid State Ionics, 2012. https://doi.org/10.1016/j.ssi.2011.12.007.

[71] J. Kaspar, M. Graczyk-Zajac, R. Riedel, Carbon-rich SiOC anodes for lithium-ion batteries: Part II. Role of thermal cross-linking, in: Solid State Ionics, 2012. https://doi.org/10.1016/j.ssi.2012.01.026.

[72] M. Greenough, Z. Zhao, L.G. Jacobsohn, J. Tong, R.K. Bordia, Low/intermediate temperature pyrolyzed polysiloxane derived ceramics with increased carbon for electrical applications, J. Eur. Ceram. Soc. (2021). https://doi.org/10.1016/j.jeurceramsoc.2021.04.007.

Page 159: Carbon Rich Polymer-Derived Ceramics: Processing

142

[73] A. Ferrari, J. Robertson, Interpretation of Raman spectra of disordered and amorphous carbon, Phys. Rev. B - Condens. Matter Mater. Phys. (2000). https://doi.org/10.1103/PhysRevB.61.14095.

[74] S.J. Widgeon, S. Sen, G. Mera, E. Ionescu, R. Riedel, A. Navrotsky, 29Si and 13C Solid-state NMR spectroscopic study of nanometer-scale structure and mass fractal characteristics of amorphous polymer derived silicon oxycarbide ceramics, Chem. Mater. (2010). https://doi.org/10.1021/cm1021432.

[75] D. Hourlier, S. Venkatachalam, M.R. Ammar, Y. Blum, Pyrolytic conversion of organopolysiloxanes, J. Anal. Appl. Pyrolysis. (2017). https://doi.org/10.1016/j.jaap.2016.11.016.

[76] A.S. Rudenkov, A. V. Rogachev, A.N. Kupo, P.A. Luchnikov, N. Chicherina, The phase composition and structure of silicon-carbon coatings, Mater. Sci. Forum. (2019). https://doi.org/10.4028/www.scientific.net/MSF.970.283.

[77] TUINSTRA F, KOENIG JL, RAMAN SPECTRUM OF GRAPHITE, J. Chem. Phys. (1970). https://doi.org/10.1063/1.1674108.

[78] L.G. Jacobsohn, R. Prioli, F.L. Freire, G. Mariotto, M.M. Lacerda, Y.W. Chung, Comparative study of anneal-induced modifications of amorphous carbon films deposited by dc magnetron sputtering at different argon plasma pressures, in: Diam. Relat. Mater., 2000. https://doi.org/10.1016/S0925-9635(99)00341-6.

[79] R.O. Dillon, J.A. Woollam, V. Katkanant, Use of Raman scattering to investigate disorder and crystallite formation in as-deposited and annealed carbon films, Phys. Rev. B. (1984). https://doi.org/10.1103/PhysRevB.29.3482.

[80] M.D. Irwin, C.G. Pantano, P. Gluche, E. Kohn, Bias-enhanced nucleation of diamond on silicon dioxide, Appl. Phys. Lett. (1997). https://doi.org/10.1063/1.119839.

[81] K. Wang, B. Ma, X. Li, Y. Wang, L. An, Structural evolutions in polymer-derived carbon-rich amorphous silicon carbide, J. Phys. Chem. A. (2015). https://doi.org/10.1021/jp5093916.

[82] C. Hintze, K. Morita, R. Riedel, E. Ionescu, G. Mera, Facile sol-gel synthesis of reduced graphene oxide/silica nanocomposites, J. Eur. Ceram. Soc. (2016). https://doi.org/10.1016/j.jeurceramsoc.2015.11.033.

[83] J.M. Englert, K.C. Knirsch, C. Dotzer, B. Butz, F. Hauke, E. Spiecker, A. Hirsch, Functionalization of graphene by electrophilic alkylation of reduced graphite, Chem. Commun. (2012). https://doi.org/10.1039/c2cc31181j.

Page 160: Carbon Rich Polymer-Derived Ceramics: Processing

143

[84] C. Carteret, A. Labrosse, Vibrational properties of polysiloxanes: From dimer to oligomers and polymers. 1. structural and vibrational properties of hexamethyldisiloxane (CH3)3SiOSi(CH3)3, J. Raman Spectrosc. (2010). https://doi.org/10.1002/jrs.2537.

[85] F.L. Esteban Florez, A.A. Trofimov, A. Ievlev, S. Qian, A.J. Rondinone, S.S. Khajotia, Advanced characterization of surface-modified nanoparticles and nanofilled antibacterial dental adhesive resins, Sci. Rep. (2020). https://doi.org/10.1038/s41598-020-66819-8.

[86] C.M. Chan, L.T. Weng, Y.T.R. Lau, Polymer surface structures determined using ToF-SIMS, Rev. Anal. Chem. (2014). https://doi.org/10.1515/revac-2013-0015.