You are on page 1of 143

University of South Florida

Scholar Commons
Graduate Theses and Dissertations Graduate School

March 2018

Synthesis, Modification, Characterization and


Processing of Molded and Electrospun
Thermoplastic Polymer Composites and
Nanocomposites
Tamalia Julien
University of South Florida, tjulien@mail.usf.edu

Follow this and additional works at: https://scholarcommons.usf.edu/etd


Part of the Polymer Chemistry Commons

Scholar Commons Citation


Julien, Tamalia, "Synthesis, Modification, Characterization and Processing of Molded and Electrospun Thermoplastic Polymer
Composites and Nanocomposites" (2018). Graduate Theses and Dissertations.
https://scholarcommons.usf.edu/etd/7631

This Dissertation is brought to you for free and open access by the Graduate School at Scholar Commons. It has been accepted for inclusion in
Graduate Theses and Dissertations by an authorized administrator of Scholar Commons. For more information, please contact
scholarcommons@usf.edu.
Synthesis, Modification, Characterization and Processing of Molded and Electrospun

Thermoplastic Polymer Composites and Nanocomposites

by

Tamalia Julien

A dissertation submitted in the partial fulfillment


of the requirements for the degree of
Doctor of Philosophy
Department of Chemistry
College of Arts and Sciences
University of South Florida

Major Professor: Julie P. Harmon, Ph.D.


Sylvia Thomas, Ph.D.
Jianfeng Cai, Ph.D.
Theresa Evans-Nguyen, Ph.D.

Date of Approval:
March 26th, 2018

Keywords: Polycarbonate Polyurethane, Nanoparticles, Nanosilica, Nanosilver, Carbon black,


Soft Thermoplastic Urethane

Copyright © 2018, Tamalia Julien


Dedication
This dissertation is dedicated to my father, Nicholas Anthony Lawrence Julien, who has passed
before this day could be realized but who firmly believed that I was destined for greatness since I
was a child. I love you and miss you every day Dad.
Acknowledgments

Firstly, I would like to give thanks and praise to the Lord for his unwavering love and

guidance (Psalm 27). My mom, who has been my rock, and my number one supporter, I would

like to thank you for your love and guidance and encouraging words, that has lifted me up

through the low times and kept me going. My sister Tanielle, who was always just a facetime

call away who has always filled me with laughter and energy. To Lyndon for your unwavering

support in helping to raise our son while I was completing my studies, thank you for all your

help. To my son Julien who gives me strength and motivation and to the rest of my family and

friends, I am truly thankful and appreciative of your love and support.

To my advisor Dr. Julie Harmon, I genuinely want to thank you for your support, for

being my guide and my mentor, for encouraging me to step out of my comfort zone during my

doctoral study. I would also like to thank my committee members, Dr. Jianfeng Cai, Dr. Theresa

Evans-Nguyen and Dr. Xiao (Sheryl) Li, for your time, insight and willingness to be a part of my

doctoral journey. Dr. Sylvia Thomas, I want to thank you for allowing me to be collaborate with

your lab, you always had kind words and great advice and always made me feel welcomed.

Additionally, I would like to thank the members of Dr. Harmon’s lab both past and

present for welcoming me in the lab and making the lab feel like a home away from home. Mu

Seong Kim, Parul Jain, Kenneth Kull, Alejandro Rivera, Garrett Craft, Imalka Arachchilage and

Taranjot Kaur. To my original USF girls’ posse, Tarah, Josanne-Dee and Kia, thank you ladies

for being my support group, I am so thankful to have you in my life and to have each one of you
play an integral part of my USF journey. To my friends, Santana, Anne-Claire, Taylor, Shahid,

Ryan, Bryan, and to my RFM ladies, Kyoka, Khalifa, Aleya, Rohini, Eboni, and FAMU friends,

Solange and Kalisa, you are all awesome and thank you for all the memories that we have made

together and for making my college experience more enjoyable.


Table of Contents

List of Tables ................................................................................................................................. iv

List of Figures ..................................................................................................................................v

List of Equations ..............................................................................................................................x

Abstract .......................................................................................................................................... xi

Chapter 1: Thermal and Mechanical Analysis of Sustainable Biopolymer Nanocomposites .........1

1.1 Introduction ...................................................................................................................1


1.1.2 Types of Nanocomposites .............................................................................. 3
1.2 Experimental ................................................................................................................ 5
1.2.1 Materials .........................................................................................................5
1.2.2 Purification of Polycarbonate Polyurethane ..................................................5
1.2.3 Polymer- nanoparticle composite synthesis ....................................................6
1.2.4 Preparation of molded nanocomposites ..........................................................6
1.2.5 Surface Free Energy Measurement .................................................................6
1.2.6 Fourier-Transform Infrared Spectroscopy (FTIR) ..........................................8
1.2.7 Differential Scanning Calorimetry (DSC) ......................................................9
1.2.8 Parallel Plate Rheology ...................................................................................9
1.2.9 Tensile Testing ...............................................................................................9
1.3 Results and Discussion ................................................................................................10
1.3.1 Surface Free Energy Measurement ...............................................................10
1.3.2 Fourier-Transform Infrared Spectroscopy (FTIR) ........................................15
1.3.3 Differential Scanning Calorimetry (DSC) ....................................................18
1.3.4 Parallel Plate Rheology .................................................................................22
1.3.5 Tensile Testing ..............................................................................................29
1.4 Conclusions ..................................................................................................................34
1.5 References ....................................................................................................................35

Chapter 2: Fabrication of Polycarbonate Polyurethane Nanofibers and Nanomembranes ............43

2.1 Introduction ..................................................................................................................43


2.2 Experimental ................................................................................................................46
2.2.1 Materials .......................................................................................................46
2.2.2 Electrospinning Process ................................................................................47
2.2.3 Solution Viscosity Measurements.................................................................47

i
2.2.4 Scanning Electron Microscopy (SEM) .........................................................47
2.2.5 Contact Angle Measurement.........................................................................48
2.2.6 Thermogravimetric Analysis (TGA).............................................................48
2.2.7 Differential Scanning Calorimetry (DSC) ....................................................48
2.2.8 Wide-angle X-ray Scattering (WAXS) .........................................................49
2.2.9 Small- angle X-Ray (SAXS) .........................................................................50
2.2.10 Tensile Testing ............................................................................................50
2.3 Results and Discussion ................................................................................................50
2.3.1 Solution Viscosity Measurements ................................................................50
2.3.2 Scanning Electron Microscopy (SEM) .........................................................51
2.3.3 Contact Angle Measurement.........................................................................52
2.3.4 Thermogravimetric Analysis (TGA).............................................................54
2.3.5 Differential Scanning Calorimetry (DSC) ....................................................55
2.3.6 Wide-angle X-ray Scattering (WAXS) .........................................................57
2.3.7 Small- angle X-Ray (SAXS) .........................................................................59
2.3.8 Tensile Testing ..............................................................................................60
2.3.9 PCPU/ nanocomposite membranes ..............................................................62
2.3.9.1 Solution Viscosity Measurements .........................................62
2.3.9.2 Scanning Electron Microscopy (SEM) ..................................63
2.3.9.3 Thermogravimetric Analysis (TGA)......................................64
2.3.9.4 Differential Scanning Calorimetry (DSC) .............................65
2.3.9.5 Wide-angle X-ray Scattering (WAXS) ..................................67
2.3.9.6 Tensile Testing .......................................................................70
2.4 Conclusions ..................................................................................................................73
2.5 References ....................................................................................................................74

Chapter 3: Rheological and Electrical Characterization of Thermoplastic and Thermoset Polymer


Lithium electrolyte. ........................................................................................................................79

3.1 Introduction ..................................................................................................................79


3.2 Experimental ................................................................................................................82
3.2.1 Polymer Lithium Composites .......................................................................82
3.2.2 Electrospinning of Polyimide .......................................................................83
3.2.3 Polyimide /Lithium Nanomembrane Electrolyte ..........................................83
3.2.4 Dynamic Mechanical Analysis (DMA) ........................................................83
3.2.5 Four-Point Probe ...........................................................................................84
3.2.6 Solution Rheology Measurement ..................................................................85
3.2.7 Scanning Electron Microscopy (SEM) .........................................................85
3.2.8 Fourier-Transform Infrared Spectroscopy (FTIR) ........................................86
3.2.9 Thermogravimetric Analysis (TGA).............................................................86
3.3 Results and Discussion ................................................................................................86
3.3.1 Dynamic Mechanical Analysis (DMA) ........................................................86
3.3.2 Secondary Relaxation ...................................................................................93
3.3.3 Primary Relaxation .......................................................................................94
3.3.4 Conductivity..................................................................................................99
3.3.5 Polyimide nanomembrane electrolyte.........................................................103

ii
3.3.5.1 Solution Rheology Measurement .................................................103
3.3.5.2 Scanning Electron Microscopy (SEM) ........................................104
3.3.5.3 Thermogravimetric Analysis (TGA)............................................105
3.3.5.4 Fourier-Transform Infrared Spectroscopy (FTIR) .......................106
3.3.5.5 Conductivity of Polyimide fibers .................................................107
3.4 Conclusion .................................................................................................................108
3.5 References ..................................................................................................................109

Appendix A ..................................................................................................................................114
Appendix B ..................................................................................................................................124
About the Author ............................................................................................................... End Page

iii
List of Tables

Table 1.1 Contact angle data: Contact angle measurements and surface tension for PCPU and
PCPU nanocomposites ...................................................................................................................13

Table 1.2 Activation energies of viscous flow for neat PCPU and PCPU nanocomposites ........28

Table 1.3 Tensile Data: Tensile test results for PCPU and PCPU/CB, PCPU/SiO2 & PCPU/Ag
nanocomposites. .............................................................................................................................32

Table 2.1 Glass transition temperature, melt temperature and TGA temperatures of
molded and electrospun PCPU fibers. ...........................................................................................57

Table 2.2 TGA data showing onset temperature, temperature at 99% and 50% mass for PCPU/
0.5% w/w nanoparticle fiber membranes. ....................................................................................65

Table 2.3 DSC data showing glass transition temperature (Tg) and melt temperature (Tm) for
PCPU/nanoparticle fiber membranes. ..........................................................................................66

Table 2.4 Tensile strength data on electrospun PCPU and PCPU nanocomposite fiber
membranes. ..................................................................................................................................73

Table 3.1 DMA Data: DMA Tg and tan δ for the PCPU and
PCPU/LiPF6 composites ...............................................................................................................93

Table 3.2. DMA data: Activation Energy of the β-transition for neat PCPU and PCPU lithium
composites......................................................................................................................................94

Table 3.3 Conductivity data: Conductivities for uncut and cut surfaces of PCPU and
PCPU/LiPF6 nanocomposites. .....................................................................................................103

iv
List of Figures

Figure 1.1 Schematic representation of the dispersion of nanoparticles in the


polymer matrix ................................................................................................................................1

Figure 1.2 Synthetic scheme showing the process of synthesizing PCPU ......................................2

Figure 1.3 Contact Angle data for PCPU/Ag nanocomposites using water and
cyclohexane ...................................................................................................................................12

Figure 1.4 Contact Angle data for PCPU/SiO2 nanocomposites using water and
cyclohexane....................................................................................................................................13

Figure 1.5 Contact Angle data for PCPU/carbon black nanocomposites using water and
cyclohexane....................................................................................................................................13

Figure 1.6 Water contact angle of pristine and ruptures PCPU/ Ag


nanocomposite surfaces ..............................................................................................................15

Figure 1.7 Water contact angle of pristine and ruptures PCPU/ SiO2
nanocomposite surfaces ..............................................................................................................15

Figure 1.8 Water contact angle of pristine and ruptures PCPU/ CB


nanocomposite surfaces ..............................................................................................................16

Figure 1.9 FTIR spectra for cut and uncut PCPU and PCPU/ 1 % w/w
nanocomposite ...............................................................................................................................17

Figure 1.10 FTIR of –NH group showing the change in transmittance from cut and uncut PCPU
nanocomposites ..............................................................................................................................18

Figure 1.11 FTIR of –C=O group showing the change in transmittance from cut and uncut PCPU
nanocomposites ..............................................................................................................................18

Figure 1.12 FTIR of –NH group showing the change in transmittance over a
24-hour period ................................................................................................................................19

Figure 1.13 DSC Data: DSC thermogram of neat PCPU and PCPU/SiO2 nanocomposites ........20

Figure 1.14 DSC thermogram of neat PCPU and PCPU/carbon black nanocomposites...............21

v
Figure1.15. DSC thermogram of neat PCPU and PCPU/ Ag nanocomposites .............................22

Figure 1.16. DSC glass transition (Tg) trends for PCPU and PCPU
nanocomposites .............................................................................................................................22

Figure 1.17 Complex viscosity (ŋ), Storage Modulus (G’) and Loss Modulus (G’’) vs angular
frequency (ω) at 100,120,150&175°C ...........................................................................................25

Figure 1.18 Complex viscosity vs shear rate for neat PCPU and PCPU/ Ag
nanocomposites at 150⁰C .............................................................................................................26

Figure 1.19 Complex viscosity vs shear rate for neat PCPU and PCPU/ SiO2
nanocomposites at 150⁰C...............................................................................................................27

Figure 1.20 Complex viscosity vs shear rate for neat PCPU and PCPU/ CB
nanocomposites at 150⁰C...............................................................................................................28

Figure 1.21 Melt Rheology Data: Arrhenius plot of melt viscosity for neat PCPU ......................28

Figure 1.22 Tensile Data: Stress vs Strain curve neat PCPU and PCPU/Ag nanocomposites ......31

Figure 1.23 Tensile Data: Stress vs Strain curve neat PCPU and
PCPU/SiO2 nanocomposites .........................................................................................................31

Figure 1.24 Tensile Data: Stress Strain Curve neat PCPU and PCPU/ carbon black
nanocomposites ..............................................................................................................................32

Figure 2.1 Electrospinning set up ..................................................................................................44

Figure 2.2. XRD showing detection of ordered structure in a. pre-stressed neat PCPU and b.
stressed undeformed bulk PCPU ...................................................................................................49

Figure 2.3. Solution Viscosities of 8%w/v, 10%w/v, 12%w/v, 14%w/v, 16%w/v & 18%w/v neat
PCPU solution concentrations .......................................................................................................51

Figure 2.4. SEM images a.-c., 14, 16, 18 % w/v fiber membranes; d.-f. Film-like membrane
composites......................................................................................................................................52

Figure 2.5. Illustration of contact angle at θ>90⁰, θ<90⁰ and θ=90⁰ .............................................53

Figure 2.6. Contact Angle Images and graph for the bulk PCPU film and 14,16 &18%w/v fiber
membranes. ....................................................................................................................................54

Figure 2.7. TGA graph showing the degradation of PCPU film and PCPU fiber membranes ......55

vi
Figure 2.8. DSC thermogram for the second heating cycle of neat PCPU, 14% w/v PCPU fiber,
16% w/v PCPU fiber & 18% w/v PCPU fiber ...............................................................................56

Figure 2.9. XRD showing detection of ordered structure for the 14% w/v and 16% w/v
solution cast film ...........................................................................................................................58

Figure 2.10. XRD showing detection of ordered structure of the 14%, 16% and 18% w/v fiber
membranes .....................................................................................................................................59

Figure 2.11. SAXS of bulk PCPU and PCPU fiber membranes ....................................................60

Figure 2.12. a. Stress-Strain behavior, b. Elastic modulus, c. Tensile strength and d. Strain at
break of bulk PCPU and PCPU fiber mats ....................................................................................61

Figure 2.13 Solution Viscosities of 0.5%w/w Ag, 0.5% w/w SiO2 & 0.5% w/w CB PCPU
nanocomposite solutions at 16% w/v concentration ......................................................................62

Figure 2.14. SEM images a.-c., 0.5%w/w Ag, 0.5% w/w SiO2 & 0.5% w/w CB PCPU w/v fiber
membranes; d.-f. Film-like membrane composites........................................................................63

Figure 2.15 TGA graph of neat polyimide (green) and 14(blue), 16% w/w (orange) PI/LiPF6
fiber membranes.............................................................................................................................64

Figure 2.16 DSC thermos gram of neat PCPU (red) vs 0.5%w/w silver (Blue) and 0.5% w/w
carbon black (green) and 0.5% w/w silica (purple) PCPU fiber....................................................66

Figure 2.17 WAXS of 0.25%w/w (blue), 0.5%w/w(green), 0.75%w/w (purple) & 1.0% w/w
(yellow)PCPU/ Ag nanocomposite fiber membranes ....................................................................68

Figure 2.18 WAXS of 0.25%w/w (brown), 0.5%w/w (yellow), 0.75%w/w (green) & 1.0% w/w
(blue) PCPU/ SiO2 nanocomposite fiber membranes ...................................................................69

Figure 2.19 WAXS of 0.25%w/w (green), 0.5%w/w (pink), 0.75%w/w (yellow) & 1.0% w/w
(blue) PCPU/ CB nanocomposite fiber membranes ......................................................................69

Figure 2.20 Stress vs strain curves for CB nanocomposite fiber membranes ...............................70

Figure 2.21 Stress vs strain curves for Ag nanocomposite fiber membranes ................................71

Figure 2.22 Stress vs strain curves for SiO2 nanocomposite fiber membranes.............................72

Figure 3.1 Polyimide GPI 15 .........................................................................................................81

Figure 3.2 Four Point probe set-up ................................................................................................84

Figure 3.3 DMA Data: Temperature sweep at multiple frequencies of neat PCPU ......................88

vii
Figure 3.4 DMA Data: Temperature sweep at multiple frequencies of 12 wt%
PCPU/LiPF6 composite .................................................................................................................89

Figure 3.5 DMA Data: Temperature sweep at multiple frequencies of 14wt%


PCPU/LiPF6 composite .................................................................................................................90

Figure 3.6 DMA Data: Temperature sweep at multiple frequencies of 16 wt%


PCPU/LiPF6 composite .................................................................................................................91

Figure 3.7 DMA Data: Loss modulus plot of neat PCPU and PCPU/LiPF6
composites......................................................................................................................................92

Figure 3.8 DMA Data: Tan delta (δ) plot of PCPU and PCPU/ LiPF6 composites ......................92

Figure 3.9 Plots of ln frequency vs 1000/T showing WLF behavior at Tg for a. neat PCPU, b.
12% w/w PCPU/LiPF6, c. 14%w/w PCPU/LiPF6 and
d. 16% w/w PCPU/LiPF6 ..............................................................................................................95

Figure 3.10 a. Isothermal frequency sweeps for neat PCPU and 3.10b.
the master curve ............................................................................................................................96

Figure 3.11 DMA data: WLF plot of neat PCPU .........................................................................97

Figure 3.12 DMA data: WLF plot for 12%w/w LiPF6/PCPU. .....................................................97

Figure 3.13 DMA data: WLF plot for 14%w/w LiPF6/PCPU ......................................................98

Figure 3.14 DMA data: WLF plot for 16%w/w LiPF6/PCPU ......................................................99

Figure 3.15 Conductivity data: Conductivity of PCPU/LiPF6 thin film composites ..................101

Figure 3.16 Conductivity vs time for 12wt% PCPU/LiPF6 composite for cut and
uncut surface .............................................................................................................................101

Figure 3.17 Conductivity vs time for 14wt% PCPU/LiPF6 composite for cut
and uncut surface ...................................................................................................................102

Figure 3.18 Conductivity vs time for 16wt% PCPU/LiPF6 composite for cut and
uncut surface ...........................................................................................................................102

Figure 3.19 Viscosity vs shear rate for pre-spun polyimide solutions .........................................104

Figure 3.20 A-C. SEM images of neat polyimide fibers P347, R127 & XP02691. D-F. Images of
the resultant fiber membranes collected. .....................................................................................105

viii
Figure 3.21 TGA Data: Thermogravimetric curves of 15wt% and 20wt% PI fibers
membranes ...................................................................................................................................106

Figure 3.22 FTIR spectra of neat polyimide (red) and 14%w/w (green), 16% w/w (purple)
PI/LiPF6 fiber membranes ...........................................................................................................107
Figure 3.23 Conductivity vs time graph of neat polyimide (red) and 14%w/w (green), 16% w/w
(purple) PI/LiPF6 fiber membranes .............................................................................................108

ix
List of Equations

Equation 1.1: Young’s Equation ....................................................................................................7

Equation 1.2: Dupre’s Definition of Adhesion Energy ..................................................................7

Equation 1.3: Adhesion Energy as a function of polar and dispersive


Components ..................................................................................................................................7

Equation 1.4: Fowkes Primary Equation ........................................................................................7

Equation 1.5: Condensation of Fowkes Primary Equation ............................................................8

Equation 1.6: Surface Free Energy Calculation .............................................................................8

Equation 1.7: Arrhenius Expression for Melt Viscosity ..............................................................24

Equation 3.1: Resistivity Calculation ...........................................................................................85

Equation 3.2: Arrhenius Equation ................................................................................................93

Equation 3.3: William-Landel-Ferry Equation.............................................................................94

Equation 3.4: Arrhenius Relationship to WLF .............................................................................96

x
Abstract

This dissertation focuses on the versatility and integrity of a novel, ultrasoft

polycarbonate polyurethane (PCPU) by the introduction of nanoparticles and lithium salts.

Additionally, the research takes into account the use of electrospinning as a technique to create

PCPU and polyimide (PI) fibers. These polymers are of interest as they offer a wide range of

properties and uses within the medical and industrial fields.

An industrial batch of an ultrasoft thermoplastic polyurethane (TPU) was synthesized using a

two-step process. The first was to create an end capped pre-polymer from methylene bis (4-

cyclohexylisocyanate), and a polycarbonate polyol made up of 1,6- hexanediol and 3-methyl-1,5-

pentanediol. The second step was done by reacting the pre-polymer with an excess of the

polycarbonate polyol with a chain extender, 1,4-butanediol. Biocompatibility testing such as

USP Class VI, MEM Elution Cytotoxicity and Hemolysis toxicology reported that PCPU

showed no toxicity. This novel type of polyurethane material targets growing markets of

biocompatible polymers and has been used for peristaltic pump tubing, but also can be utilized as

balloon catheters, enteral feeding tubes and medical equipment gaskets and seals. This material

is ideal for replacing materials such as soft plastisols containing diethylhexyl phthalate for use in

biomedical and industrial applications. After extensive characterization of this polymer system

another dimension was added to this research.

The addition of nanoparticles and nanofillers to polyurethane can express enhanced

mechanical, thermal and adhesion properties. The incorporation of nanoparticles such as

xi
nanosilica, nanosilver and carbon black into polyurethane materials showed improved tensile

strength, thermal performance and adhesion properties of the PCPU. Samples were characterized

using contact angle measurements, Fourier transform spectroscopy (FTIR), differential scanning

calorimetry (DSC), parallel plate rheology and tensile testing.

The second chapter entails the fabrication and characterization of PCPU nanofibers and

nanomembranes through a process known as electrospinning. The resulting PCPU

nanomembranes showed a crystalline peak from the WAXS profile which is due to electrospun

and solution strain induced crystallinity. The PCPU nanocomposite nanomembranes displayed

increased thermal stability and an increase in tensile performance at higher weight percent. The

nanomembranes were investigated using contact angle measurements, thermogravimetric

analysis (TGA), DSC, WAXS, SAXS and tensile testing.

The final chapter focuses on investigating the rheological properties of PCPU/lithium

electrolytes as well as transforming an unprocessable polyimide powder into a nanomembrane.

The PCPU/ lithium composite electrolyte showed an increase in the activation energy and

conductivity, while the PI/lithium showed increased conductivity over time. Dynamic

mechanical analysis and four-point probe was used to investigate the samples.

xii
Chapter 1:
Thermal and Mechanical Analysis of Sustainable Biopolymer Nanocomposites

1.1 Introduction

Polymer nanocomposites are created when nanoparticles are introduced to polymer

systems whether post synthetically or in-situ polymerization [1–4]. The formation of

nanocomposites by introducing nanoparticles as a reinforcing filler with the goal of enhancing

and modifying the performance and characteristics of polymers have been used for a long time in

nanotechnology [1,5-7]. Nanocomposite materials cover a wide range material from inorganic

glasses to organic polymers [8] and are important materials for many modern and future

technologies, primarily due to their wide tunability in properties and light weight [9-11].

Figure 1.1 Schematic representation of the dispersion of nanoparticles in the


polymer matrix.

Nanoparticles are extremely diverse and can be placed in many sub categories depending

on their use. They can be metallic such as Iron oxide (Fe3O4), gold (Au), Silver (Ag), carbon

derived such as single walled and multiple walled carbon nanotubes, graphene and carbon black

as well as organic nanoparticles such as dendrimers, micelles and ferritin. They can also be

1
described as zero dimensional such as spheres, one dimensional such as nano wires and nanorods,

2 dimensional such as thin films and plates and three dimensional also known as bulk

nanomaterials. Because of these attributes, nanotechnology attained acceptance by industrial

sectors due to its applications in electronic storage systems [13-14], biotechnology [15], targeted

drug delivery [16,17] and vehicles for gene therapy and drug delivery [14-18].

The polymer utilized in this study can be categorized as a thermoplastic polymer and can

be further categorized as a semi crystalline polymer which is made up of not well-defined hard

crystalline segments with a low percent crystallinity and soft amorphous region. Polycarbonate

polyurethane is an ultrasoft polymer synthesized with methylene bis (4-cyclohexylisocyanate),

1,4- butanediol as a chain extender and a polycarbonate polyol containing 1, 6-hexanediol and 3-

methyl-1, 5-pentanediol [19-21]. It displays remarkable mechanical properties together with the

proficiency to re-heal after rupture without the need for additives or implanted healing agents. A

combination of properties such as high tensile strength, ultimate elongation, toughness, abrasion,

tear resistance, low compression and tensile set, low temperature performance and resistance to

oil have allowed polyurethanes to be used in many demanding applications [22-24].

Figure 1.2 Synthetic scheme showing the process of synthesizing PCPU [20].

2
PCPU is a thermoplastic polyurethane (TPU) which is a class of polymers that possess

excellent properties including toughness, abrasion resistive, excellent hydrolytic stability,

elastomeric and durability, making it versatile for widespread uses [19-21,25–30]. Because of its

biocompatibility, non-toxicity, robustness and functionality TPUs can be used in the biomedical

industry as implantable devices (vascular grafts, pacemaker leads, blood bags, bladders and

artificial heart valves) and medical applications [31–34]. Moreover, it can also be utilized in other

industries such as electronics, automotive, sporting goods and foams [35-38].

In addition, PCPU also has properties similar to segmented polyurethanes (PUs), which

consists of hard and soft segments. It exhibits microphase separation, due to a high degree of

mixing of hard and soft segments because of hydrogen bonding between the hard segment

urethane groups and the soft segment carbonate groups [41-42]. Additionally, the mixing can be

confirmed by the small percent crystallinity with crystal lamellae d spacing of 0.45Å [20].

1.1.2 Types of Nanocomposites

Three different types of zero dimensional spherical nanoparticles are being investigated

herein; fumed nanosilica, nanosilver and carbon black. Each imparting its own unique properties

into the PCPU.

Fumed silica is produced by high-temperature hydrolysis of silicon tetrachloride in a

flame [39-40]. The use of silica with polyurethanes has been widely studied but nanosilica has not

been introduced to our novel polymer. The addition of fumed silicas to polyurethanes in solution

conveys viscosity, thixotropy and pseudoplasticity, also there is an improvement in mechanical

properties [40]. The surfaces of silica nanoparticles possess an unknown number of surface

silanol (Si-OH) groups which may participate in hydrogen bonding with amines within the

urethane unit.

3
Fumed nanosilica is known to be hydrophilic and when added to the polyurethane, the

degree of phase separation escalates because of the interaction between hydrogen-bonded silanol

groups on the nanosilica surface and soft segments of the polyurethane. This leads to an increase

in the segmental incompatibility on the polyurethane with the presence of the hydrophilic

nanosilicas. Studies have shown that these specific polymer materials have the ability to form

hydrogen bonds which result in higher phase separation due to less direct interactions between

phases. Furthermore, silanol and carbonyl group interactions are weaker than those between -NH

and ester carbonyl groups, with the addition of silica the polycarbonate chain mobility increases

and becomes more ordered relative to the neat polyurethane [43-45].

Carbon black (CB) is abundant, low cost and possesses properties such as heat stability and

electrical conductivity [46-50]. It is a versatile nanoparticle which can be incorporated in

electrochemistry and mechanical enhancements and other fields. When a conductive filler such as

carbon black is introduced into a polymer matrix a conductive polymer nanocomposite is formed

with excellent thermal and mechanical properties as well as electrical conductivity which can be

applied to sensors, conductors and anti-static materials. In this study the use of low weight percent

loading of CB nanoparticles was investigated as it is known that high amounts of CB cause poor

mechanical properties and complex loading [50-51]. Also, this particular type of carbon black

used is known to be more effective at low loadings to impart maximum enhancement of the

material’s properties.

Amongst the nanoparticles, silver (Ag) nanoparticles are of high interest as silver-based

products are continuously being sought after in the biomedical field. Silver nanoparticles allow

for enhanced thermal, mechanical and antimicrobial properties that are necessary for biomaterials

such as bandages, catheters, surgical instruments and topical ointments [52-54]. However, there

4
are adverse effects from prolonged exposure of Ag nanoparticles on human health. One way to

combat this is to entrench the silver particles into the polymer matrix. PCPU is biocompatible;

therefore, it becomes an impeccable candidate for the Ag nanoparticles which can then be used

for prosthesis as well as drug delivery devices [52, 55-56].

Herein, a series of PCPU nanosilica, nanosilver and nanoscale carbon black composites

were produced via ultrasonication. Our novel PCPU was designed to be highly flexible while

keeping its mechanical properties. This introduction of the nanoparticles will allow for increased

mobility which is expected to lead to greater thermal and mechanical performance. Also, the

literature is quite rare on rheological studies and viscoelastic behavior of PCPU nanocomposites,

therefore, the rheological properties of these PCPU nanocomposites were studied to investigate its

melt processability.

1.2 Experimental

1.2.1 Materials

Nanosilver (NTX-300ET) was purchased from Nanux Inc (Yungnam,Korea) with an

average particle size of 30-50nm. Carbon black nanoparticles (VXC72) were mass produced by

Cabot Corporation (Victoria, Australia) with a particle size of 10-30 nm. Fumed silica (nanosilica

HDK N20) was mass-produced by Wacker-Chemical Corporation (Michighan, United States).

The majority particle size in all nanosilicas used was 7 nm. All particles were used as received

without further purification. The certified ACS grade reagent tetrahydrofuran (THF) 99.9% was

purchased from Sigma-Aldrich and used without further purification.

1.2.2 Purification of polycarbonate polyurethane

100 g of PCPU chips were mixed with 1000 ml of THF in a 3L reactor for 3 hours until

the polymer was fully dissolved. Deionized water was then added to the reactor in a dropwise

5
fashion to guarantee proper crashing-out of the polymer. The water was then removed, and the

process was repeated. The purified polymer was then dried under vacuum at 60⁰C.

1.2.3 Polymer-nanoparticle composite synthesis

Polycarbonate polyurethane/nanoparticle composites were prepared by dispersing the

nanoparticles using a sonicator. 10g of the purified PCPU was dissolved in 100 ml THF (ACS

reagent, 99.9% sigma Aldrich) and then the desired amount of nanoparticles (0.1-1.0 w.t.%) were

introduced to the solution and then mixed by ultrasonication (Fisher Scientific Sonic

Dismembrator 550) at a mixing rate of 10 minutes with 3 second rest intervals for two hours to

ensure homogeneity of composites. The composites were then crashed out using deionized water

and the resulting materials were dried overnight under vacuum.

1.2.4 Preparation of molded nanocomposites

Before samples were characterized, they were compression molded using a Carver laboratory

press (model C) at 250⁰C and 3 tonnes of pressure.

1.2.5 Surface Energy Measurements

Surface energy or surface tension of solid samples gives an insight into the samples’ wettability.

This behavior is of great importance as it plays a significant role in many industrial processes

such as lubrication, painting and printing [57]. The surface energy of a solid cannot be directly

measured therefore the use of liquid on a solid surface via contact angles is employed. There are

many different theories employed in surface energy measurements such as Zisman (one

component model), Owens/Wendt (two component), Fowkes (two component) and van Oss (three

component) theories.

This study focuses on the two-component model specifically Fowkes theory.

Owens/Wendt theory and Fowkes theory are similar in that they both specify that the solid surface

6
has two components; a dispersive force and a polar (non-dispersive) force. The difference

between the two is the method with which the theories are utilized.

Fowkes theory is centered on three essential equations describing the interactions between

solid surfaces and liquids [58]. As with all theories, they begin with Young's Equation which

describes the contact angle of a liquid drop on an ideal solid surface;

𝛾𝑠𝑣 = 𝛾𝑠𝑙 + 𝛾𝑙𝑣 cos 𝜃 (1.1)

where: ɣlv = liquid-vapor interfacial tensions, ɣsv = solid-vapor interfacial tensions, ɣsl = solid-

liquid interfacial tension, and θ = the contact angle between the liquid and the solid.

Fowkes theory also considers Dupre's Definition of Adhesion Energy

𝐼𝑆𝐿 = 𝛾𝑠𝑣 + 𝛾𝑙𝑣 − 𝛾𝑠𝑙 (1.2)

where: ISL = Adhesion energy per unit area between a liquid and a solid surface.

Fowkes stipulated that the adhesive energy between a solid and a liquid can be expanded into

interactions between the dispersive components of the solid and liquid as well as the interactions

between the polar components of the two components.

2 2 2
+ √𝛾𝑙𝑣𝑝 × √𝛾𝑠𝑣
𝑑 2
𝑑 𝑝
𝐼𝑆𝐿 = 2[ √𝛾𝑙𝑣 × √𝛾𝑠𝑣 ] (1.3)

d p
where: ɣlv = dispersive component of the surface tension of the test liquid, ɣlv = polar component

d
of the surface tension of the test liquid, ɣsv = dispersive component of the surface energy of the

p
solid surface, and ɣsv = the solid surface polar component.

The combination of the three equations produces the primary equation of the Fowkes'

surface energy theory:

2 2 𝛾𝑆𝐿 (cos 𝜃+1)


+ √𝛾𝑙𝑣𝑝 × 𝛾𝑠𝑣
𝑑 𝑑 𝑝
√𝛾𝑙𝑣 × 𝛾𝑠𝑣 = (1.4)
2

7
To use the Fowkes' theory on a solid sample to determine solid surface energy, the first liquid
p d
must only have a dispersive component for example a liquid must have ɣL = 0, so that ɣL = ɣSL.

Because of this, the primary equation condenses to:

𝑑 𝛾𝑙𝑣 (cos 𝜃+1)2


𝛾𝑠𝑣 = (1.5)
4

d
and ɣsv can be calculated directly from the contact angle measured from the solid surface.

The next step is to test the solid for contact angle with another liquid which has both a dispersive

component and a polar component. The contact angle produced from the liquid on the solid and
d p
the calculated ɣsv from the previous step, ɣsv can be calculated as the only unknown in the

equation 1.4. The overall surface energy of the solid, ɣsv, is then calculated as

p d
ɣsv = ɣsv + ɣsv (1.6)

Uniform drops of deionized water and cyclohexane were displaced on the polymer

nanocomposite surface and the contact angles were measure using a KSV CAM-101 video based

optical contact angle measuring device equipped with a hamilton syringe in an environmentally

controlled chamber (KSV-1TCU). All measurements were performed in air, at room temperature.

The samples were tested and then the surface was cut away and the cut surface was retested with

the deionized water.

1.2.6 Fourier Transform Infrared Spectroscopy (FTIR)

Perkin Elmer spectrum two furnished with and Attenuated Total Reflection (ATR)

accessory was utilized for this study. A scanning range of 400-4000cm-1 set at a resolution of

4cm-1 and 16 repetitions were done on the nanocomposites. The composites were scanned, then

cut along its length and then re-scanned for 24 hours to detect any shifts of functional groups.

8
1.2.7 Differential Scanning Calorimetry (DSC)

DSC analysis was performed using TA instruments 2920 Differential Analysis

Calorimeter to obtain glass transition temperature (Tg) and melt temperature (Tm) of the polymer

and polymer nanocomposites. The heat cool heat method was utilized a starting temperature of -

70⁰C and then heated to 200⁰C at a ramp rate of 10⁰C/ minute to ensure all samples has the sample

thermal history. The sample was the cooled down to -70⁰C, using a refrigerated accessory, with

the similar ramp rate and then reheated to 200⁰C. The Tg values were taken from the second

heating run.

1.2.8 Parallel Plate Rheology

Rheological properties of the polymer nanocomposites were analyzed by TA instruments

AR 2000 equipped with parallel plates with a 25mm diameter. The plate gap was kept at 3mm

during experimentation. The measurements were done at different temperatures from 100⁰C to

190⁰C

The strain sweeps were performed from 0.01% to 100% at a constant frequency of

0.33HZ. This was done to detect the linear viscoelastic region at different temperatures.

Frequency sweeps were carried at strains within the viscoelastic region of 1% and 5%

respectively. The frequency sweeps were done from 0.05 to 100 Hz. Specimen discs were of

average of 25mm and 3mm thickness.

A temperature ramp was done at different strains of 1% and 5% at a constant frequency of 1Hz to

detect the viscoelastic behavior at high temperatures.

1.2.9 Tensile Testing

9
Tensile strength characterization was done using Shimadzu ASJ tensile tester to measure

displacement of samples. All tensile testing was done at a rate of 100mm/min at room

temperature according to ASTM D638 IV. Dog bone cut samples were of various thickness 0.8

to1mm. Five measurements for each sample were done and the average tensile strength was

recorded.

1.3 Results and Discussion

1.3.1 Surface Free Energy Measurement

The surface free energy was determined from contact angle measurements using deionized

water and cyclohexane, and is presented in Figures 1.3, 1.4 & 1.5. The total surface energy of a

polymer’s surface is equal to the sum of the dispersion component and the polar component. The

polar water creates a noticeable semicircular shape while the non-polar cyclohexane spreads

across the surface of the composite. Fowkes theory which is explained in the experimental section

was then used to calculate the dispersive component of the surfaces as well as the polar

component. Table 1.1 gives the contact angles of the PCPU nanocomposites together with the

resulting surface energies. The contact angles for the nanocomposites when the water was used

had varying trends. For the silver nanocomposites there was a decrease in the contact angle from

102⁰ for the neat PCPU to 96⁰, 85⁰, 100⁰ and 88⁰ for the 0.1%, 0.25%, 0.375% as well as the 0.5%

w/w. Then the 0.75% and 1.0% w/w increased to 107⁰. The silica nanocomposites showed a slight

decrease in contact angle to 96⁰ and the carbon black composite as seen displayed a slight

decrease in the angle from the neat PCPU. However, when the cyclohexane was used the contact

angle for the neat PCPU was 12.8⁰. The contact angle increased slightly for the Ag

nanocomposites to as high as 19⁰; it doubled for the silica composites with 24⁰ and a slight

increase to 20⁰ for the carbon black nanocomposites. There was a trend demonstrated on the

10
calculation of the dispersive component for the nanocomposites. The silver nanocomposites

trended downward with a constant decrease in the dispersive energy from 24.4 mJ/m to 24.0

mJ/m as well as the silica component from 24.4 to 23.1 mJ/m. The carbon black nanocomposites

however, trended upwards showing that the increase in the % w/w of carbon black nanoparticles

added, increased the dispersive component of the polymer matrix.

When the PCPU nanocomposite surface was cut away the use of the cyclohexane as a

testing parameter was futile. Figures 1.6, 1.7 & 1.8, show the contact angle of the pristine surface

against the ruptured surface of the nanocomposites using deionized water only. When the surface

is cut away there is a decrease in the contact angle below that of 90⁰ compared to the pristine

surface >100⁰. Moreover, the general trend downward for the silver nanoparticle composite is

noted with an increase in the Ag content resulting in increased hydrophilicity. This is evidence of

available hydrogen bonding sites, which will be explained further in the FTIR section.

11
Figure 1.3 Contact Angle data for PCPU/Ag nanocomposites using water and
cyclohexane.

12
Figure 1.4 Contact angle data for PCPU/SiO2 nanocomposites using water and
cyclohexane.

Figure 1.5 Contact angle data for PCPU/CB nanocomposites using water and
cyclohexane.

13
Table 1.1 Contact angle data: Contact angle measurements and surface tension for PCPU and
PCPU nanocomposites

SAMPLE AVG CA (H20) AVG CA ƔSV ƔSVD ƔSV


(C6H6) P

NEAT PCPU 102 ± 2 12.5 ± 1.2 25.3 24.7 0.6


0.1WT% AG 96 ± 5 18.9 ± 5.5 26.2 24.5 1.7
0.25WT%AG 85 ± 3 11.6 ± 2.7 30.0 24.5 5.5
0.375WT% AG 100 ± 4 16.9 ± 0.8 25.3 24.4 0.9
0.5WT% AG 88 ± 5 17.2 ± 2.7 28.4 24.2 4.2
0.75WT% AG 107 ± 5 17.5 ± 3.4 24.3 24.2 0.1
1.0WT% AG 107 ± 2 18.5 ± 5 24.2 24.02 0.2
0.1WT% SIO2 102 ± 3 12.6 ± 2.3 25.0 24.4 0.6
0.25WT% SIO2 100 ± 2 18.2 ± 1.8 25.5 24.4 1.1
0.375WT% SIO2 97 ± 3 17.4 ± 2.5 25.9 24.2 1.7
0.5WT% SIO2 96 ± 8 17.5 ± 2.1 26.0 24.2 1.8
0.75WT% SIO2 103 ± 1 19.6 ± 3.5 24.5 23.9 0.6
1.0WT% SIO2 96 ± 2 24.3 ± 0.7 25.1 23.1 2.0
0.1 WT % CB 94 ± 4 19.9 ± 4.8 25.6 23.2 2.4
0.25WT% CB 99 ± 5 20.0 ± 3.3 25.0 23.8 1.2
0.375WT% CB 104 ± 4 17.9 ± 3.9 24.5 24.1 0.4
0.5WT% CB 102 ± 2 20.3 ± 3.9 24.6 23.9 0.7
0.75WT% CB 92 ± 8 17.2 ± 5.5 27.1 24.2 2.9
1WT% CB 102 ± 3 18.4 ± 2.4 24.6 24.1 0.5

14
Figure 1.6 Water contact angle of pristine and ruptures PCPU/ Ag
nanocomposite surfaces.

Figure 1.7 Water contact angle of pristine and ruptures PCPU/ SiO2 nanocomposite
surfaces.
15
Figure 1.8 water contact angle of pristine and ruptures PCPU/ CB nanocomposite
surfaces.

1.3.2 Fourier-Transform Infrared Spectroscopy (FTIR)

Figure 1.9 shows the IR spectra for cut and uncut surfaces of the PCPU and its

nanocomposites. The expected absorbance for associated N-H groups is 3318-3338 cm-1 and

associated C=O 1718 cm-1 and unassociated C=O is 1740 cm-1. Yokoyama et al. 1968 [59];

Seymour et al. 1970 [60]; Boerio and Wirasate 2006 [61]; Na, Lv et al. 2009] [62] all have done

studies to demonstrate associated and unassociated hydrogen bonding using FTIR. The

characteristic peaks for nanoparticles; SiO2 (1626, 1103, 809 cm-1), CB (1630, 1114, 1118 cm-1)

and silver nanoparticles (3420 and 1638 cm-1) did not show up on any of the FTIR spectra. This

can be as a result of the low concentrations together with the nanometer size of the particles used.

Figure 1.10 and 1.11 show a magnified view of the –NH region and –C=O of the spectra.

There is a shift in the wave number of the associated -NH region from 3305 cm-1 to 3307 cm-1 on

16
the addition of the nanoparticles. This suggests interactions between the particles and the NH

groups within the polymer matrix. When the surface is cut away there is a decrease in the %

transmittance for the H-bonded region at 3305cm-1 which tells us that when the surface is cut

away hydrogen bonds become available and the H-bonded regions are reduced. This correlates

with the decrease in the contact angle when the surface is cut away. Available H-bonds allow for

lower contact angle which makes the surface hydrophilic. Figure1.12 shows what happens to

these bonds if they are left over time. After 1 hour you see the transmittance decrease and after 24

hours the scan gets closer to the original scan. This is evidence of the hydrogen bonds folding

onto themselves when left alone [21].

100

90

80 Free N-H
stretch at H-bonded
-1 N-H
70 3380cm
stretch at
-1
3305cm
%Transmittance

60
Neat PCPU uncut
50 neat PCPU cut
1wt% SiO2/PCPU
1wt% SiO2/PCPU cut
40
H- 1wt% Ag/PCPU cut
1wt% Ag/PCPU uncut
bonded
30 Free C=O 1wt% CB/PCPU uncut
C=O at 1wt% cb/pcpu cut
absorptions 1718 cm
20 -1 -1
at 1740cm

10
4000 3600 3200 2800 2400 2000 1600 1200 800 400
Wavenumber cm-1

Figure 1.9 FTIR spectra for cut and uncut PCPU and PCPU/ 1%w/w nanocomposites.

17
100

98
%Transmittance

96
Neat PCPU uncut
Neat PCPU cut
94 1wt% nanocilica uncut
1wt% nanosilica cut
1wt% silver cut
92 1wt% silver uncut
1wt% cb uncut
1wt% cb cut
90
3500 3450 3400 3350 3300 3250 3200
Wavenumber cm-1

Figure 1.10 FTIR of –NH group showing the change in transmittance from cut and uncut PCPU
nanocomposites.

100

90

80
%Transmittance

70

60

50

40

30
1800 1700 Wavenumber cm-1 1600

Figure 1.11 FTIR of –C=O group showing the change in transmittance from cut and uncut PCPU
nanocomposites.

18
100

99

98
% Transmittance

97

.25 cb cut
96
%T0.25cb uncut

95 %T-0.25cb 24hr cut

%T-0.25 1hr
94
3500 3450 3400 3350 3300 3250 3200
Wavenumber cm-1

Figure 1.12 FTIR of –NH group showing the change in transmittance over a 24-hour period.

1.3.3 Differential Scanning Calorimetry (DSC)

The glass transition, Tg and melting temperatures were recorded as seen in Figure 1.13,

1.14 and 1.15. Stevens et al demonstrated that the addition of nanoparticles would induce

plasticization of the polymer composites leading to lower Tg. The PCPU/Ag composites’ Tg

ranged from -23.5⁰C to -25.4⁰C as the % w/w of silver nanoparticles were added. The PCPU silica

composites’ Tg ranged from -24.1⁰ C to -22⁰C and the PCPU/CB composites were from -24⁰C to -

22.5⁰C.

The trends of the nanoparticles are shown in figure 1.16. The silica nanocomposites and

the carbon black nanocomposites responded similarly. There was an initial decrease in the glass

transition temperature at the lower %w/w which is due to the increase in free volume space of the

polymer chains, then at 0.5 %w/w nanoparticles, the glass transition temperature increased. This

can be due to the reduction in the motion of the polymer backbone from the additional

nanoparticles which would reduce the plasticization effect and therefore allowed a rise in the Tg.
19
Because the silica and carbon black have the ability to aggregate and agglomerate this may cause

a reduction in the free volume of the polymer backbone which leads to the increase in the glass

transition temperature. Contrastingly, nanosilver particles do not aggregate so the increase in the

amount of nanosilver added increases the ability to plasticize and which leads to a continuous

decrease in the glass transition temperature.

-22.6⁰C

-23.4⁰C

-23.8⁰C

-24.1⁰C

-23.5⁰C
-23.1⁰C
-22⁰C

Figure 1.13 DSC Data: DSC thermogram of neat PCPU and PCPU/SiO2 nanocomposites.

20
-22.5⁰C

-23.2⁰C

-23.5⁰C
-24⁰C

-22.8⁰C
-22.6⁰C

-22.6⁰C

Figure 1.14. DSC thermogram of neat PCPU and PCPU/carbon black nanocomposites.

21
-22.6⁰C

-23.5⁰C

-23.6⁰C

-23.9⁰C
-24.9⁰C
-25.3⁰C

-25.4⁰C

Figure1.15. DSC thermogram of neat PCPU and PCPU/ Ag nanocomposites.

-21.5 CB SiO2 Ag
Glass Transition temperature Tg/ ⁰C

-22

-22.5

-23

-23.5

-24

-24.5

-25

-25.5

-26
0 0.25 0.5 0.75 1
Wt% nanoparticle added

Figure 1.16. DSC glass transition (Tg) trends for PCPU and PCPU nanocomposites.

22
1.3.4 Parallel Plate Rheology

Parallel plate rheology tells about the deformation of flow for viscoelastic polymer

composites at elevated temperatures above their Tg. This is essential for the processability for the

polymers. Figure 1.17 shows the melt rheology profile of the neat PCPU at different temperatures.

The PCPU nanocomposites all showed that the viscosities of the polymer composites are

temperature dependent as well as frequency dependent and that our material is in fact viscoelastic.

This can be seen when the frequencies are lower the viscosity is proportional to the change in

frequency but as the frequencies increase the viscosity is directly proportional to the frequency.

Additionally, an increase in the temperature allowed for a decrease in the viscosity. Hence, the

polymer demonstrates pseudo plastic behavior by the decrease in viscosity with increase in

frequency and shear thinning occurs. Erdmann et al [63] also experienced lower viscosities and

modulus as the temperature is increased which is typical thermoplastic behavior. The intersection

point between the G’ and G” shifts to a higher frequency with increasing temperature because of

mobility and flexibility of the PCPU chains so that disentanglement is easier [63].

Figures 1.18-1.20 show the complex viscosities vs shear rate for the neat PCPU vs

PCPU/silver, silica and CB nanocomposites. The silver nanocomposites’ complex viscosity

decreased with increasing silver nanoparticle concentration, but this is expected as stated by

Ghosh and Maiti that at high shear rates, the viscous stresses predominate over the silver particle-

particle interactions, leading to particle alignment and, therefore, lower melt viscosity [64-66].

The silica PCPU nanocomposites showed an initial decrease in the viscosity but the 1%w/w

PCPU/silica showed a higher complex viscosity at low shear rates, but it decreased as the shear

rate increased. What is also noted is that all the composites joined at the higher shear rate at a

viscosity of 630 Pa which is not evident in the other nanocomposites. The carbon black

23
nanocomposites showed very little variation in complex viscosity compared to the neat PCPU but

as the % w/w increased the viscosity decreased.

The log of complex viscosities (n), of the nanocomposites was plot against the inverse

temperature; the result was a straight line which represents Arrhenius behavior. The Arrhenius

expression for viscosity


∆𝐸𝑎
𝑛 = 𝐴 𝑒 𝑅𝑇 (1.7)

Where n is the complex melt viscosity, A is the Arrhenius constant, ΔEa is the activation

energy of viscous flow and R is the universal gas constant. The slope of the line generated from

the plot can be used to calculate the activation energy of viscous flow. The ΔEa of viscous flow

for neat PCPU came to be 70.1kJ/mol. With the addition of the Ag nanoparticles the ΔEa

decreased but as the %w/w increased the Activation energy remained unchanged giving the aspect

that the activation energy is independent of the silver content. The silica’s activation energy was

close to that of the neat polymer which suggests homogeneity of the composite. Contrastingly the

ΔEa for the carbon black PCPU composites almost doubled at higher %w/w of carbon black

nanoparticles. This suggests that the carbon black nanocomposites are less temperature sensitive

as the nanoparticles have the packing ability, so it provides less change in free volume space as

the temperature changes as experienced by Ghosh and Plochcki [66-67].

24
Figure 1.17 Complex viscosity (ŋ), Storage Modulus (G’) and Loss Modulus (G’’) vs angular
frequency (ω) at 100,120,150&175°C.

25
Overlay

Figure 1.18 Complex viscosity vs shear rate for neat PCPU and PCPU/ Ag
nanocomposites at 150⁰C.

26
Overlay 2

Figure 1.19 Complex viscosity vs shear rate for neat PCPU and PCPU/ SiO2
nanocomposites at 150⁰C.

27
Overlay

Figure 1.20 Complex viscosity vs shear rate for neat PCPU and PCPU/ CB
nanocomposites at 150⁰C.

13
y = 8435.9x - 9.9026
R² = 0.9999
12
Ea 16.7kCal/mol
11
ln (ŋ)

10

8
0.0022 0.0023 0.0024 0.0025 0.0026 0.0027 0.0028
1/T (K-1)

Figure 1.21 Melt Rheology Data: Arrhenius plot of melt viscosity for neat PCPU.
.

28
Table 1.2 Activation energies of viscous flow for neat PCPU and PCPU nanocomposites.

SAMPLE EA OF VISCOUS FLOW EA OF VISCOUS


(KJ/MOL) FLOW(KCAL/MOL)
NEAT PCPU 70.1 16.7
0.1WT% AG 62.5 14.9
0.25WT%AG 62.5 14.9
0.375WT% AG 63.7 15.2
0.5WT% AG 64.4 15.4
0.75WT% AG 62.2 14.9
1.0WT% AG 62.3 14.9
0.1WT% SIO2 69.8 16.7
0.25WT% SIO2 72.6 17.3
0.375WT% SIO2 70.2 16.8
0.5WT% SIO2 69.4 16.6
0.75WT% SIO2 55.8 13.3
1.0WT% SIO2 60.9 14.6
0.1 WT % CB 69.8 16.7
0.25WT% CB 67 16
0.375WT% CB 63.8 16.4
0.5WT% CB 124.5 29.7
0.75WT% CB 110.8 26.4
1WT% CB 116.9 27.9

29
1.3.5 Tensile Testing

PCPU can be described as a viscoelastic material, which means they are positioned

between viscous liquids and elastic solids as mentioned above. Melt rheology gives us an idea of

the mechanical behavior of the polymer at elevated temperatures and frequencies. There are two

laws associated with the elasticity and viscosity of materials. Hooke’s Law demonstrates that for

an ideal linear elastic solid, the stress is directly proportional to the strain; and Newton’s law for

an ideal viscous liquid which states that the stress should be proportional to the rate of change of

the strain [66]. Tensile testing allows us to take into consideration the mechanical behavior at

room temperature. An ideal linear elastic solid obeys Hooke’s law, i.e. stress is proportional to

strain. An ideal viscous liquid obeys Newton’s law, i.e. stress is proportional to the rate of change

of strain.

Tensile testing was conducted to evaluate mechanical properties of the resulting composites for

the reasons discussed in the introduction. Stress-strain curves are given for the averages of 5

specimens of each sample: Neat vs Silica nanocomposites (figure 1.23), neat vs carbon black

composites (figure 1.24), neat vs silver nanocomposites (figure 1.21). The results of the tensile

test including averages and statistical data are summarized in table 2.3. The tensile curve of the

neat PCPU is similar to that of other polyurethanes [67], consisting of three distinct regions of

deformation. The first region is known as the elastic region where Young’s modulus is calculated

and the resistance to deformation of the polymer is noted. Just after the curve or yield point, there

is a plateau region associated with the plastic region where the amorphous soft segments of the

polymer deform, and the last region is known as the strain hardening region associated with strain

induced crystallinity and hard segment disassociation as demonstrated by the steep slope right

before fracture [68].

30
18
Stress vs Strain of Neat PCPU vs PCPU/Ag
16

14

12
Stress (Mpa)

10 neat pcpu
0.1wt % silver
8
0.25wt%silver
6
0.375wt silver
4 0.5 wt % silver

2 0.75 wt% silver


1.0wt% silver
0
0 300 600 900 1200
Strain %
Figure 1.22 Tensile Data: Stress vs Strain curve neat PCPU and PCPU/Ag nanocomposites.

25
Stress vs Strain Neat PCPU vs PCPU/SiO2

20
neat pcpu
0.1wt% SiO2
15
Stress Mpa

0.25wt% SiO2
0.5wt% SiO2
10
0.75wt%SiO2
0.375wt% SiO2
5 1.0wt% SiO2

0
0 300 600 900 1200
Strain %
Figure 1.23 Tensile Data: Stress vs Strain curve neat PCPU and PCPU/SiO2 nanocomposites.

31
14
Stress vs Strain Neat PCPU vs PCPU/CB
12

10
Stress Mpa

neat pcpu
6
0.1wt% cb
0.25wt% cb
4 0.375wt% cb
0.5wt% cb
2 0.75wt%cb
1.0wt%cb
0
0 300 Strain % 600 900 1200
Figure 1.24 Tensile Data: Stress Strain Curve neat PCPU and PCPU/ carbon black nanocomposites.

Table 1.3 Tensile Data: Tensile test results for PCPU and PCPU/CB, PCPU/SiO2 & PCPU/Ag
nanocomposites.

Youngs
Modulus Ultimate Tensile strength
Sample Mpa Mpa Elongation%
neat pcpu 2.3 ± 0.1 10.2 ± 0.2 1071 ± 18
0.1 wt % cb 2.9 ± 0.03 12.1 ± 0.6 777 ± 11
0.25wt% cb 2.8 ± 0.02 13.1 ± 0.2 783 ± 18
0.375wt% cb 3.3 ± 0.02 10.2 ± 0.4 720 ± 14
0.5wt% cb 3.3 ± 0.07 10.2 ± 0.8 708 ± 20
0.75wt% cb 2.9 ± 0.08 11.2 ± 0.5 737 ± 12
1wt% cb 3.2 ± 0.03 12.4 ± 0.5 708 ± 15
0.1wt% SiO2 3.1 ± 0.1 13.3 ± 0.8 864 ± 13
0.25wt% SiO2 5.4 ± 0.1 19.4 ± 0.2 661 ± 23
0.375wt% SiO2 4.4 ± 0.1 13.6 ± 0.3 873 ± 15
0.5wt% SiO2 5.3 ± 0.02 11.4 ± 0.1 847 ± 12
0.75wt% SiO2 2.2 ± 0.1 12.2 ± 0.5 922 ± 25

32
Table 1.3 continued

1.0wt% SiO2 2.1 ± 0.2 12.4 ± 0.1 656 ± 20


0.1wt% Ag 2.2 ± 0.1 12.3 ± 1.2 936.6 ± 24
0.25wt%Ag 4.8 ± 0.6 13 ± 0.4 874 ± 14
0.375wt% Ag 4.8 ± 0.2 14.9 ± 1.0 542 ± 26
0.5wt% Ag 3.0 ± 0.1 10.1 ± 0.7 602 ± 36
0.75wt% Ag 4.9 ± 0.5 15.3 ± 0.2 838 ± 11
1.0wt% Ag 3.6 ± 0.1 9.2 ± 1.2 787 ± 20

The tensile curves indicate that the addition of nanoparticles displays an increased

mechanical behavior. This is confirmed by an increase in the elastic modulus and yield strength as

well as, ultimate tensile strength but a slight reduction in the ductility as seen in the elongation %.

The increase in mechanical properties is drastic in the silica composites such as the 0.25% w/w

silica, where the Young’s modulus rose to 5.4 MPa from 2.3 MPa of the neat PCPU and the

ultimate tensile strength moved from 10.1 to 19.1 MPa. The show of superior mechanical strength

illustrates successful interfacial adhesion as well as good homogeneity and compatibility between

the silica nanoparticles and the polymer matrix. What is also demonstrated in the silica

composites is that there is an effected load amount needed. As the %w/w increased to 0.75 and

1.0% the tensile properties decreased which suggests that the interfacial adhesion is no longer

effective leading to a mechanically inferior nanocomposite. The most surprising results come

from the carbon black nanocomposites which showed less free volume space in the melt rheology

results, which should have allowed for a stronger and tougher material. The nanocomposites

showed only a small increase in tensile properties as well as ultimate tensile strength at room

temperature which could mean that higher tensile strength may be seen at higher temperatures.

On the other hand, the silver nanocomposites showed promising results as the tensile strength

doubled from 2.3 MPa to 4.8MPa and 4.9 MPa for the majority of the silver nanocomposites. The

33
tensile data showed good interfacial adhesion of the nanoparticles to the polymer matrix to create

a mechanically enhanced polymer nanocomposite.

1.4 Conclusion

This study herein proves that enhancement of the thermal and mechanical properties of

PCPU can be successfully done by the dispersion of spherical silica, silver and carbon black

nanoparticles. The spherical nature of the nanoparticles showed ease of movement of the polymer

matrix by the reduction of the glass transition temperature showing that these nanoparticles move

in association with the polymer backbone. This allows for a higher thermal performance of

PCPU. Additionally, hydrogen bonding played a key role in the properties of the PCPU

nanocomposites as evidenced by the change in the contact angle and the change in the FTIR

spectra. Characterization of the polymer nanocomposites via parallel plate rheology supports

previously stated literature that the activation energy of viscous flow is independent of the

addition of silver nanoparticles. The silica nanoparticles showed homogeneity of mixing and the

carbon black nanoparticles showed that it can increase the activation energy by its packing ability

to decrease free volume space. The rheology also showed no drastic changes in complex viscosity

on the addition of the nanoparticles. The tensile properties of the ultrasoft polymer increased on

the addition of the nanoparticles especially at the low concentrations where the Young’s modulus

was double that of the neat PCPU.

The implications of the formation and characterization of the polymer nanocomposites

suggests that polymeric materials for industrial use can be enhanced by the addition of

nanoparticles to create superior polymers with high thermal and mechanical performance.

34
1.5 References

1. I.V. Khudyakov, D. R. Zopf and N.J. Turro (2009), Designed Monomers and Polymers,

12:4, 279-290

2. S. Cheng, B. Carroll, W. Lu, F. Fan, J. Y. Carrillo, H. Martin, A.P. Holt, N.Kang, V.

Bocharova, J. W. Mays, B. G. Sumpter, M. Dadmun, and A.P. Sokolov (2017) Interfacial

Properties of Polymer Nanocomposites: Role of Chain Rigidity and Dynamic

Heterogeneity Length Scale; Macromolecules, 50, 2397−2406

3. T. J. Pinnavaia and G. W. Beall (2001), Polymer-Clay Nanocomposites. Wiley, New York,

NY.

4. M. Rosoff (2000), Nano-Surface Chemistry. Marcel Dekker, New York, NY.

5. D. R. Paul and L. M. Robeson (2008), Polymer Nanotechnology Nanocomposites,

Polymer 49, 3187-3290.

6. A. Saralegi, S. C. M Fernandes, A. Alonso-Varona, T. Palomares, E. J. Foster, C. Weder,

A. Eceiza, M.A. Corcuera (2013), Shape-memory bionanocomposites based on chitin

nanocrystals and thermoplastic polyurethane with highly crystalline soft segment,

Biomacromolecules 14, 4475-4482.

7. T. Wu, M. Frydrych, K. O’Kelly, B. Chen (2014), Poly (glycerol sebacate urethane)-

cellulose nanocomposites with wateractive shape-memory effects, Biomacromolecules,

15, 2663.

8. Moon, R. J.; Martini, A.; Nairn, J.; Simonsen, J.;Youngblood, (2011), Cellulose

nanomaterials review: structure, properties and nanocomposites , J. Chem. Soc. Rev. 40,

3941.

35
9. A. Nicharat, A. Shirole, E. J. Foster, C. Weder. (2017) Thermally activated shape

memory behavior of melt-mixed polyurethane/cellulose nanocrystal composites. J. Appl.

Polym. Sci. DOI: 10.1002/APP.45033

10. B. M. Novak, (1993) Hybrid nanocomposite materials—between inorganic glasses and

organic polymers. Adv. Mater. 5, 422

11. S. C. Tjong and Y. W. Mai (2010) Physical Properties and Applications of Polymer

Nanocomposites; Woodhead Publishing Limited: Cambridge, UK.

12. A. C. Balazs, T. Emrick and T. P. Russell (2006) Nanoparticle Polymer Composites:

Where Two Small Worlds Meet. Science, 314:5802, 1107−1110.

13. S. Cheng, S. J. Xie, J. M. Y. Carrillo, B. Carroll, H. Martin, P. F. Cao, M. D. Dadmun, B.

G. Sumpter, V. N. Novikov, K. S. Schweizer and A. P. Sokolov (2017) Big Effect of

Small Nanoparticles: A Shift in Paradigm for Polymer Nanocomposites. ACS Nano, 11:1,

752−759.

14. V.V Mody, R. Siwale, A. Singh and H.R. Mody (2010), Introduction to metallic

nanoparticles, J Pharm Bioallied Sci, 2:4: 282-289.

15. Y. S. Kang, S. Risbud, J. F. Rabolt and P. Stroeve (1996) Synthesis and characterization

of nanometer-size Fe3O4 and γ-Fe2O3 Particles. Chem Mater, 8:22, 09–11.

16. Q. A. Pankhurst, J. Connolly, S. K. Jones and J. Dobson (2003) Applications of magnetic

nanoparticles in biomedicine. J Phys D Appl Phys. 36:R167.

17. J. Dobson (2006) Gene therapy progress and prospects: magnetic nanoparticle-based gene

delivery, Gene Ther. 13:283–287.

36
18. S. Rudge, C. Peterson, C. Vessely, J. Koda, S. Stevens and L. Catterall (2001) Adsorption

and desorption of chemotherapeutic drugs from a magnetically targeted carrier (MTC) J

Elsevier 74:335–340.

19. T. Appenzeller (1991) The man who dared to think small, Science; 254:1300.

20. K.L. Kull, R.W. Bass, G. Craft, T. Julien., E. Marangon, C. Marrouat and J.P. Harmon

(2015), Synthesis and Characterization of an Ultra-soft Poly (Carbonate Urethane),

European Polymer Journal 71:510-522.

21. R.W. Bass: Ph.D. Thesis, Synthesis and Characterization of Self-Healing Poly (Carbonate

Urethane) Carbon-Nanotube Composites, University of South Florida (2011)

22. J.P. Harmon and R.W. Bass (2014), Self-Healing Polycarbonate Polyurethane Nanotube

Composites. US Patent: 8,846,801.

23. R. W. Bass, T. Julien, A. Rigolo, S. Hong, X. Li, and J. P. Harmon (2016) Thermal and

mechanical properties of single-walled and multi-walled carbon nanotube polycarbonate

polyurethane composites with a focus on self-healing International Journal of Materials

Research 107, 8, 692-702

24. E. N. Doyle (1971) The Development and Use of Polyurethane Products, McGraw-Hill

Book Company New York.

25. C. Hepburn (1992) Polyurethane Elastomers, Elsevier. London: Applied Science; 1992.

26. Z. Wirpsza and T. J. Kemp (1993), Polyurethanes: chemistry, technology, and

applications. London: E. Horwood.

27. J. Ju, Z. Shi, L. Fan, Y. Liang, W. Kang, B. Cheng (2017), Preparation of elastomeric

tree-like nanofiber membranes using thermoplastic polyurethane by one-step

electrospinning, Materials Letters 205, 190–193

37
28. N.A.M. Barakat, M.A. Kanjwal, F.A. Sheikh, H. Yong Kim (2009), Spider-net within the

N6, PVA and PU electrospun nanofiber mats using salt addition: Novel strategy in the

electrospinning process, Polymer 50 4389-4396.

29. A. Cornille, R. Auvergne, O. Figovsky, B. Boutevin, S. Caillol (2017), A Perspective

Approach to Sustainable Routes for Non-Isocyanate Polyurethanes, Eur. Polym. J. 87 535-

552.

30. S.H. Choi, D.H. Kim, A.V. Raghu, K.R. Reddy, H.I. Lee, K.S. Yoon, H.M. Jeong, B.K.

Kim (2012), Properties of graphene/waterborne polyurethane nanocomposites cast from

colloidal dispersion mixtures , J. Macromol. Sci. B 51, 197-207.

31. K.R. Reddy, A.V. Raghu, H.M. Jeong (2008), Synthesis and characterization of. novel

polyurethanes based on 4,4'-{1,4-phenylenebis[methylylidenenitrilo]}. diphenol Polym.

Bull. 60 609-616

32. M. Dasdemir, M. Topalbekiroglu, A. Demir (2013). Electrospinning of thermoplastic

polyurethane microfibers and nanofibers from polymer solution and melt J. Appl. Polym.

Sci. 127,3,1901-1908, DOI: 10.1002/APP.37503

33. C. Akduman, I.Özgüney , E. Perrin, A. Kumbasar (2016), Preparation and characterization

of naproxen-loaded electrospun thermoplastic polyurethane nanofibers as a drug delivery

system, Materials Science and Engineering C 64, 383–390

34. B.S. Gupta and J.V. Edwards (2009) Textile materials and structures for wound care

products, in: S. Rajendran (Ed.), Advanced Textiles for Wound Care, Woodhead

Publishing Limited, Cambridge pp. 48–97.

38
35. T. T. N. Huynh, K. Padois, F. Sonvico, A. Rossi, F. Zani, F. Pirot, J. Doury ad F. Falson

(2010), Characterization of a polyurethane-based controlled release system for local

delivery of chlorhexidine diacetate, Eur. J. Pharm. Biopharm. 74:2, 255–264.

36. R. J. Zdrahala and I. J. Zdrahala (1999), Biomedical applications of polyurethanes: a

review of past promises, present realities, and a vibrant future, J. Biomater. Appl. 14:1,67–

90.

37. A. Huang, X. Peng, Lih-Sheng Turng. (2018), In-situ fibrillated polytetrafluoroethylene

(PTFE) in thermoplastic polyurethane (TPU) via melt blending: Effect on rheological

behavior, mechanical properties, and microcellular foamability. Polymer 134, 263-274.

38. X.D. Wang, X. Luo (2004) A polymer network based on thermoplastic polyurethane and

ethylene–propylene–diene elastomer via melt blending: morphology, mechanical

properties, and rheology, Eur. Polym. J. 40,2391-2399.

39. H.Y. Mi, X. Jing, M.R. Salick, L.S. Turng, X.F. Peng (2014), Fabrication of

thermoplastic poly-. urethane tissue engineering scaffold by combining microcellular

injection molding. and particle leaching, J. Mater Res. 29, 911-922.

40. K. A. Ames (2004) Elastomers for shoe applications, Rubber Chem. Technol. 77 413-475.

41. R. Bode, H. Ferch, H. Fratzscher (1982), Basic Charâcteristics and Applications of

Aerosils, DEGUSSA Technical Bulletin Pigments no. 11.

42. A. Torro-Paulau, J. Fernandez-Garcia, A.C. Orgiles-Barcelo, J. Martin-Martinez (2001)

Characterization of polyurethanes containing different silicas. International. Journal of

Adhesion and Adhesives, 21, pp. 1-9, 0143-7496

39
43. P.A. Gunatillake, G. F. Meijs, S. J. McCarthy, R. Adhikari, N. Sherriff (1998), Synthesis

and characterization of a series of poly(alkylene carbonate) macrodiols and the effect of

their structure on the properties of polyurethanes. J. Appl. Polym. Sci, 69: 1621-1633.

44. Y. He, D. Xie and X. Zhang (2014) The structure, microphase-separated morphology, and

property of polyurethanes and polyureas, J Mater Sci 49:7339–7352

45. M. S. Sánchez-Adsuar, E. Papón and J. Villenave (2000). Influence of the

prepolymerization on the properties of thermoplastic polyurethane elastomers. Part I.

Prepolimerization characterization. Journal of Applied Polymer Science, 76, pp. 1596-

1601, 0021-8995.

46. R. C. R. Nunes, J. L.C. M. Fonseca, and M. R. Pereira (2000) Polymer–filler interactions

and mechanical properties of a polyurethane elastomer, Polymer Testing, 19, pp. 93-103,

0142-9418

47. R. C. R. Nunes, R. A. Pereira, J. L. C. Fonseca, and M. R. Pereira (2001) X-ray studies on

compositions of polyurethane and silica, Polymer Testing, 20, pp. 707-712, 0142-9418.

48. Y. Tien and K. Wei (2001) Hydrogen bonding and mechanical properties in segmented

montmorillonite/polyuretane nanocomposites of different hard segment ratios, Polymer,

42, pp. 3213-3221, 0032-3861

49. J. Chen, X. Cui, K. Sui, Y. Zhu and W. Jiang (2017) Balance the electrical properties and

mechanical properties of carbon black filled immiscible polymer blends with a double

percolation structure, Composites Science and Technology 140, 99-105

50. L. Sun, W.M. Huang, Z. Ding, Y. Zhao, C.C. Wang, H. Purnawali and C. Tang (2012)

Stimulus-responsive shape memory materials, Mater. Des. 33, 577-640.

40
51. W. Wang, H. Pan, B. Yu, Y. Pan, L. Song, K. Meow, Liew and Y. Hu (2015) Fabrication

of carbon black coated flexible polyurethane foam for significantly improved fire safety,

RSC Adv., 5, 55870

52. C.K. Leong and D. Chung (2004), Black Dispersions and Carbon-Silver. Combinations as

Thermal Pastes That Surpass Commercial Silver and Ceramic Pastes in. Providing High

Thermal Contact Conductance. Carbon, 42, 2323–2327.

53. F. Delor-Jestin, J. Lacoste, N. Barrois-Oudin, C. Cardinet and J. Lemaire (2000), Photo-,

thermal and natural ageing of ethylene– propylene–diene monomer (EPDM) rubber used

in automotive applications. Influence of carbon black, crosslinking and stabilizing agents,

Polym. Degrad. Stab., 67, 469–477.

54. K. Cheah, M. Forsyth and G.P. Simon (2000) Processing and morphological development

of carbon black filled conducting blends using a binary host of poly(styrene co-

acrylonitrile) and poly(styrene), J. Polym. Sci. Part B Polym. Phys. 38:23, 3106-3119.

55. H. Deka, N. Karak, R. D. Kalita and A. K. Buragohain (2010) Bio-based thermostable,

biodegradable and biocompatible hyperbranched polyurethane/Ag nanocomposites with

antimicrobial activity, Polymer Degradation and Stability 95,1509-1517

56. X. Chen and H. J. Schluesener (2008) Nanosilver: A nano product in medical application.

Toxicol Lett, 176:1-12.

57. A. Travan, C. Pelillo, I. Donati, E. Marsich, M. Benincasa, T. Scarpa, et al. (2009)

Noncytotoxic silver nanoparticle-polysaccharide nanocomposites with antimicrobial

activity, Biomacromolecules, 10:1429-35

58. H. Planck, I. Syre, M. Dauner and G. Egbers (1987) Polyurethane in biomedical

engineering II, progress in biomedical engineering. Amsterdam: Elsevier.

41
59. Y. Yuan and T. R. Lee (2013) Contact Angle and Wetting Properties, Surface Science

Techniques, Springer Series in Surface Sciences 51, 1-33.

60. KRÜSS GmbH, Borsteler Chaussee 85, 22453 Hamburg, Germany www.kruss.de 4

61. T. Tanaka, T. Yokoyama, et al. (1968) Quantitative Study on Hydrogen Bonding Between

Urethane Compound and Ethers by Infrared Spectroscopy, Journal of Polymer Science

Part a-1-Polymer Chemistry 6 (8PA1): 2137-2152.

62. R. W. Seymour, G. M. Estes, et al. (1970) Infrared Studies of Segmented Polyurethane

Elastomers. 1. Hydrogen Bonding, Macromolecules 3:5, 579-583.

63. F. J. Boerio, and S. Wirasate (2006) Measurements of the Chemical Characteristics of

Polymers and Rubbers by Vibrational Spectroscopy, John Wiley & Sons, Ltd

64. B. Na, R. Lv, et al. (2009) Enhanced hydrogen bonding and its dramatic impact on

deformation behaviors in a biomedical poly(carbonate urethane) with fluorinated chain

extender, Journal of Polymer Science Part B: Polymer Physics 47:22, 2198-2205

65. R. Erdmann, S. Kabasci, J. Kurek and S. Zepnik (2014) Study of Reactive Melt

Processing Behavior of Externally Plasticized Cellulose Acetate in Presence of Isocyanate

Materials, Materials, 7, 7752-7769.

66. K. Ghosh and S. N. Maiti (1997) Melt Rheological Properties of Silver-PowderFilled

Polypropylene Composites, Polymer—Plastics Technology and Engineering, 36:5, 703-

722

67. A. F. Plochcki, in Polymer Blends, edited by D. R. Paul and S. Newman, Academic Press,

New York, 1978, Vol. 2.

68. A. K. Bhowmick (2002) Mechanical Properties of Polymers, Materials Science and

Engineering Vol. I

42
Chapter 2

Fabrication of Polycarbonate Polyurethane Nanofibers and Nanomembranes

2.1 Introduction

In 1900 John Francis Cooley was the first to file a patent for electrospinning. But, the first

presence of electrostatic attraction of a liquid, dates back to 1600 which was observed by William

Gilbert. By 1745, Bose used the electrical potentials on the surface of droplets in order to develop

aerosols. Christian Friedrich Schӧnbein was able to produce highly nitrated cellulose in 1846 and

then in 1887 Charles Vernon Boys described the process in a paper on nano-fiber manufacture.

After the filing of the first patent occurred published work on the behavior of fluid droplets began

by John Zeleny in 1914 and since then patent and publications about electrospinning have

increased each year [1].

Electrospinning is a well-established versatile technique that produces fibers with

diameter sizes in the micro and nanometer range [3-4]. This methodology employs an

electrostatic potential to create nanofibers from viscoelastic polymer solutions or polymer melts

[5-7]. Electrospinning ensues when a polymer solution or melt emits a charged fluid jet in the

presence of an electric field. When the electric field force reaches a certain threshold, the charged

polymer overcomes the surface tension and the jet undergoes a series of vigorous stretching and

splaying until it reaches a grounded target, thereby completing the circuit [5,9]. Material

properties, such as viscosity, conductivity, molecular weight and surface tension, as well as

43
processing parameters, such as applied electric field, the distance between the tip and the collector

feeding rate, air temperature and humidity, all influence spinnability and tuning properties of the

fabricated products [10-11].

Figure 2.1. Electrospinning set up [2].

The electrospinning process has the ability to produce highly porous membranes with

structural integrity. The nanofiber scaffolds possess extremely high surface to volume ratio,

tunable porosity and malleability and can be produced in a wide variety of sizes and shapes [13].

Due to these advantages, nanofiber membranes are being used in various fields like bio-medical,

pharmaceutical, nanotechnology-based industries, optical electronics, environmental engineering

and the defense industry [2, 11]. Electrospun materials have been proven to be an excellent

candidate for tissue engineering, drug delivery, vascular grafts, protective clothing systems, and

44
wound dressing as well as filtration systems for sub atomic particles [12-24].

Over the years a wide variety of polymer fibers including polyethylene (PE), polystyrene

(PS), polylactides, polyurethanes, poly(vinylidene fluoride), and polyamides have been generated

by electrospinning for various research findings [20,25]. Gazzano et. al., used semicrystalline

polymers such as, polyacrylonitrile (PAN), Nylon 6,6 (NYL) and poly(ethylene oxide) (PEO) to

investigate the structural and morphological properties of the fibers [26]. Song et al modified a

polyurethane using polyhedral oligomeric silsesquioxane (POSS) to look at improved blood

compatibility [27]. Most recently Li et. al, researched tensile properties in

silica/polydimethysiloxane (PDMS) and polymethyl methacrylate(PMMA) fibrous mats [5].

These examples show the broad range in which such a simple technique can create different

research outcomes.

The ability to electrospin our novel polycarbonate polyurethane to fabricate nanofibers

and nanomembranes, has broadened its applications in the medical and industrial fields for sensor

fabrications, drug delivery systems and scaffolds for tissue engineering where lightweight

material with larger surface area is needed. Moreover, the addition of nanoparticles to the polymer

can aid in improved catalytic activity and selectivity by silica nanoparticles [28-30], silver

nanoparticles can decrease surface inflammation and promote zinc utilization in wound healing,

[13] and carbon black nanoparticles added to a polymer matrix can create an electroactive

polymer nanomembrane [31]. These factors are in addition to its enhanced thermal and

mechanical properties already imposed on the PCPU material.

Our previous research [31-33] on the novel PCPU has shown that our polymer is ultrasoft

with a shore A hardness of 70, has high tensile strength and shows partial self-healing, which can

be enhanced with the addition of carbon nanotubes [31,33] . Now it is imperative to investigate

45
the effect of transforming the PCPU to fiber membranes to investigate whether the thermal and

mechanical properties are affected as well as how the addition of the nanoparticles affects the

overall properties of the polymer. In these studies, the utilization of thermal and mechanical

testing were completed to compare the characteristics of electrospun polymers with the bulk

PCPU and the electrospun neat PCPU with PCPU nanocomposite fibers. This research looks at

the thermomechanical and structural properties of electrospun PCPU and PCPU nanocomposite

nanofibers. Thermal properties were examined by differential scanning calorimetry (DSC).

Structural analysis was performed by Wide Angle X-ray scattering (WAXS) and Small Angle X-

ray scattering (SAXS), at room temperatures. Mechanical characterization was done by tensile

testing and morphological studies on the fiber membranes were done using Scanning Electron

Microscopy (SEM).

2.2 Experimental

2.2.1 Materials

The certified ACS grade reagent dimethylformamide (DMF) 99.9%, chloroform, tetrahydrofuran

(THF) with 250 ppm BHT inhibitor 99.9% and Ethyl Acetate (EtOAc) was purchased from

Sigma-Aldrich and used without further purification. The Polycarbonate polyurethane was

synthesized as described earlier [32-34] and the bulk sample was produced by hot pressing in a

carver press to create a film. PCPU was dissolved in the following solvent mixtures: ChCl3:

EtOAc, 90:10, THF: EtOAC, 90:10 and Dimethyformamide-ethylacetate (DMF:EtOAc) solution

in a 90:10 ratio. Polymer concentrations varied from 8-18% w/v. For the polymer

nanocomposites, the 16%w/v solution was used as the control. 0.25%, 0.5%, 0.75% 1.0% w/w of

silver, silica and carbon black nanoparticles were then added to the solution by ultrasonication for

2 hours to produce a composite polymer solution.

46
2.2.2 Electrospinning process

For the electrospinning process, each of the solutions was filled up in a 3ml syringe. A 20-gauge

needle with a flat tip was used as a spinneret. A collection distance of 20cm was done and a fixed

voltage of 20kV was applied at a feeding rate of 50μl/min.

2.2.3. Solution Viscosity Measurement

A Fungilab alpha series rotational viscometer equipped with a TR8 spindle was used to

measure the viscosity of the polymer solutions. 10ml of each solution was placed in the

cylindrical chamber and the spindle is immersed in the solution. The rotational speed was varied,

and viscosity readings were recorded. All measurements were performed at 21⁰C.

2.2.4 Scanning Electron Microscopy (SEM).

Scanning Electron Microscopy (SEM) is typically used to study surface topology such

as, the shapes and sizes of nanoparticles, the dispersion of nanoparticles within materials as well

as phase boundaries in polymer blends [35]. Research in polymer nanocomposite fiber

membranes has utilized SEM to investigate the shape and fiber diameter of electrospun polymers.

In SEM there is an electron beam and a beam in a cathode ray tube which is simultaneously

scanned across the surface of the sample. A signal is then produced by scattered electrons

resulting in an image with a three-dimensional appearance [35].

A Hitachi S-800 SEM was used in my work which has a magnification power of

300,000 times the actual sample size. SEM gives us information about fiber diameter, bead

formation, quality of the fibers or fiber membranes. The SEM is located in the Nanomaterials and

Nanomanufacturing Research Center in the Department of Engineering at the University of South

Florida.

47
2.2.5 Contact Angle Measurement

Uniform drops of the deionized water were deposited on the bulk PCPU and PCPU

electrospun membrane surface and the contact angles were measured using KSV CAM-101

video-based optical contact angle measuring device equipped with a Hamilton syringe in an

environmentally controlled chamber (KSV-1 TCU). All measurements were performed in air, at a

temperature of 25°C. Five right angles and five left angle measurements were recorded for each

sample.

2.2.6 Thermogravimetric Analysis (TGA)

A TA instruments Q500/50 TGA equipped with a standard furnace was used to detect the

degradation temperature of the polymer and polymer composite membranes. A sample weight

between 15-20mg was placed on a tared 100 ml platinum pan. Measurements were carried out

under nitrogen atmosphere as the samples were heated to 600⁰C at a heating ramp rate of 10⁰C per

minute.

2.2.7 Differential Scanning Calorimetry (DSC)

DSC analysis was performed using a TA instruments 2920 Differential Analysis

Calorimeter to obtain glass transition temperature (Tg) and melt temperature (Tm) of the polymer

and polymer membranes. The heat cool heat method was utilized a starting temperature of -70⁰C

and then heated to 200⁰C at a ramp rate of 10⁰C/ minute to ensure all samples had their thermal

history erased. The sample was then cooled down to -70⁰C, using a refrigerated accessory, with

the similar ramp rate and then reheated to 200⁰C. The Tg values were taken from second heating

run.

48
2.2.8 Wide Angle X-ray Scattering (WAXS)

WAXS powder X-ray diffraction was used to investigate the degree of crystallization of

the PCPU polymer segments. In our earlier paper [32] the degree of crystallinity was measured

for the bulk PCPU there was no visible crystal peak until the polymer was stressed for 12 hours

and a peak at 12.8⁰. X-ray diffraction patterns were collected with a Bruker D8 Focus x-ray

diffractometer. The data collection was recorded in the range of 5–80° with a step of 0.010° at

25°C. Pre-electrospun polymer film was cast from the solvent and measured. These scans were

compared to scans on electrospun mats.


19.5°

19.7°

a.
b.
12.8°

Figure 2.2. XRD showing detection of ordered structure in a. pre-stressed neat PCPU and b.

stressed undeformed bulk PCPU [32].

49
2.2.9 Small Angle X-Ray Scattering (SAXS)

SAXS data was collected on a Rigaku with a MicroMax-002+ generator, a Cu anode

tube and a 120mmDET detector and a wavelength of 1.45Å. The bulk PCPU film and the fiber

membranes were tested.

2.2.10 Tensile Testing

Tensile strength characterization was done using Shimadzu AGS-J tensile tester equipped

with a 50N load cell to measure displacement of samples. All tensile testing was done at a

crosshead speed of 100mm/min at room temperature according to ASTM D638. Dog bone cut

samples from compression molded and fiber membrane samples had thicknesses 0.15mm;

triplicate samples were tested, and average was reported.

2.3. Results and Discussion

Preliminary electrospinning tests were run on the polymer solutions with varying solvents.

The results here within are based on the polymer solutions using DMF: EtOAc in a 90:10 ratio.

The CHCl3:EtOAc as well as the THF:EtOAc combinations caused sputtering of the fibers during

the electrospinning process.

2.3.1 Solution Viscosity

Solution viscosity is important in electrospinning of polymer solutions in order to create

quality fibers. The solution viscosities at varying shear rates of the pre-spun polymers are shown

in figure 2. As the concentration of the polymer increased the viscosity overall increases. For the

8%w/v solution at the lowest shear rate of 9 s-1, the viscosity was 126mPa, and for the 18%w/v at

the same shear rate, the viscosity was found to be more than 5 times the viscosity at 729 mPa.

50
Overall it shows that the viscosity of the polymer solution decreased with increasing shear rate.

Even the 8% w/v solution had a small but consistent decrease from 126mPa to 122 mPa with

increasing shear rate. This demonstrates the pseudo plasticity of the PCPU in solution meaning a

change in the shear rate changes its viscosity.

800

700

600
8wt%
Viscosity (mPa)

500
10wt%
400
12wt%
300
14wt%
200 16wt%
100 18wt%

0
0.0 20.0 40.0 60.0 80.0 100.0 120.0 140.0 160.0 180.0
Sheat rate (1/s)

Figure 2.3. Solution Viscosities of 8% w/v, 10% w/v, 12% w/v, 14% w/v, 16% w/v & 18% w/v

neat PCPU solution concentrations.

2.3.2 Scanning Electron Microscopy (SEM)

When fibers were spun with concentrations varying from 8% w/v to 18%, lower

concentrations (8, 10 & 12 % w/v) exhibited beading and the materials were too frail to be tensile

tested. This is partially attributed to low viscosities (<400mPa). Samples spun from 14, 16, and

18% w/v for the DMF:EtOAc easily released from the collecting plate and were fully

characterized and discussed below.

Figure 3A-C shows SEM images of 14, 16 and 18% w/v PCPU fibers. Figure 3D-F shows

the PCPU fiber material as a film with different thickness. The 14% w/v material was lightweight

and had a thickness of 0.09mm, the 16% w/v had a thickness of 0.12 mm and 18% w/v material

51
felt tough has a thickness of 0.17mm. The SEM images show the material’s porosity and the

randomness of the fibers. An increase in the concentration of the solution resulted in less beading

and thinner fiber formation. For the 14% w/v, the SEM image showed some beading which may

arise from the presence of solvent during fiber formation resulting in large flat fibers with higher

fiber diameters in the 900-2.8µm range. When the concentration was increased to 16% w/v the

fiber diameter range was between 900nm-1µm. At 18% w/v, the fiber diameter ranged from 600-

950 nm, resulting in a tighter diameter range across fibers.

Figure 2.4. SEM images a.-c., 14, 16, 18 % w/v fiber membranes; d.-f. Film-like membrane
composites.

2.3.3 Contact Angle Measurement

A contact angle below 90° indicates the material is easily wetted by the test liquid. When

water is used as the test liquid, the indication is that the material is hydrophilic. Contact angles

greater than 90° indicate a resistance to wetting by the water and a hydrophobic surface. This is

depicted in Figure 4.

52
Figure 2.5. Illustration of contact angle at θ>90⁰, θ<90⁰ and θ=90⁰ [2].

Pictoral and numeric measurements for contact angles are shown in figure 5. An increase

in the contact angle occurred from the change of bulk PCPU film to PCPU fiber membrane. In

addition, an increase in contact angle also occurred as the concentration of the polymer solution

increased. The contact angles for 14% w/v, 16% w/v and 18% w/v are 103⁰, 105⁰ and 110⁰

respectively. The contact angle for the molded PCPU film was 102⁰. The slight increase in

contact angle with % w/v of polymer in the electrospun samples may indicate trace amounts of

solvent trapper in the fibers; However, TGA studies did not reveal significant weight changes.

The use of electrospinning to create fiber membranes has allowed for interesting results for future

investigations of PCPU tunable surface wettability of the polymer.

53
Figure 2.6. Contact Angle Images and graph for the bulk PCPU film and 14,16
&18%w/v fiber membranes.

2.3.4 TGA (Thermogravimetric Analysis)

The thermal stability and residual solvent dryness of the PCPU fibers were evaluated by

TGA measurements under nitrogen atmosphere. The thermograms of the molded and electrospun

PCPU at different % w/v are shown in Figure 5. The corresponding temperature at 5% weight loss

and the onset temperature are reported in Table 1. Moving from molded PCPU to the fiber

membranes we see a decrease in the onset temperature as well as the temperature at 95% mass.

This can be due to the porosity of the fibers and trapping of solvent.

54
The TGA graph shows that the thermal stability is reduced due to remnant DMF:EtoAc

solutions remaining on the fibers but as the concentration of polymer solution increases the

thermal stability also increases. The trend shows that the 18% w/v closely mimics the shape of the

neat PCPU mold and the 14% w/v and 16% w/v closely mirror each other. This suggests that the

removal of solvent leads to a more thermally stable polymer fiber mat.

100 2.5
18wt% neat PCPU fiber
90
16wt% neat PCPU fiber
2
80 14wt% neat PCPU fiber

70 neat pcpu

Deiv. Weight (%/⁰C)


1.5
60
Weight (%)

50 1

40
0.5
30

20
0
10

0 -0.5
0 100 200 300 400 500 600
Temperature (⁰C)

Figure 2.7. TGA graph showing the degradation of PCPU film and PCPU fiber membranes.

2.3.5 Differential Scanning Calorimetry (DSC)

Differential Scanning Calorimetry gives us an insight into the movement of the polymer

backbone. Results are summarized in table 1. The Glass transition temperature tells us how the

soft amorphous segments move and the melt temperature (Tm) gives us an idea of when the hard

segments melt. The glass transition temperature for the 14, 16 & 18% w/v polyurethane fibers as

well as the molded PCPU remained at -22.9⁰C. This lets us know that the soft segment

55
interactions have not changed during the electrospinning process. Wong et al mentioned that Tg

values of fiber membranes do not show a significant difference due to intermolecular coupling of

the polymer chains. DSC indicates that there is not an appreciable amount of solvent in the fibers.

The melt temperature, Tm, for the neat PCPU was at 70.5⁰C, the 14% w/v increased to 78.9⁰C,

16% to 74.4 and 18% showed two melting temperatures at 78.4 and 145⁰C. The shift to a higher

temperature means the presence of purer or more ordered crystals within the polymer matrix. The

18% w/v PCPU fiber showed a second melting which may be due to hard segment melting. There

was also an increase in the enthalpy of melt from 1.3 in the neat PCPU to 1.6 in the PCPU fibers

which can lend itself to a higher percent crystallinity not present in the bulk polymer mold.

18 % w/v

21.9 ⁰C
16 % w/v

22.8 ⁰C
14 % w/v 76.4 ⁰C
145.6
neat 22.9 ⁰C
⁰C 74.4 ⁰C
22.9
⁰C
78.9 ⁰C

70.5
⁰C

Figure 2.8. DSC thermogram for the second heating cycle of neat PCPU, 14% w/v PCPU fiber,
16% w/v PCPU fiber & 18% w/v PCPU fiber.
56
Table 2.1: Glass transition temperature, melt temperature and TGA temperatures of

molded and electrospun PCPU fibers.

Sample name First Second Tm (⁰C) ΔHm TGA Temp at 99%


Tg Tg (J/g) onset (⁰C) mass (⁰C)
(⁰C) (⁰C)

neat PCPU -24.8 -22.9 70.5 1.3 292 297

14% w/v Neat TPU -25.9 -22.9 78.9 2.7 279.0 287

fiber

16% w/v neat TPU -25.3 -22.8 74.4 1.5 280.0 288

fiber

18% w/v Neat TPU -24.8 -21.9 76.4/145.6 1.6/0.8 284.0 293

fiber

2.3.6 Wide-angle X-ray scattering (WAXS)

The DSC runs indicate that the PCPU is not highly crystalline. Indeed, this polymer is

ultrasoft and we expect and have known it to have limited crystallinity [33]. Wide-angle X-ray

scattering was used to determine the effect of electrospinning on the crystallization of PCPU.

Figure 8 Shows the WAXS intensity profile of the PCPU films cast from solution; Figure 9 is data

on the electrospun fiber. These scans exhibit more order than that found earlier in melt processed

samples [33]. In the solvent cast films, the scans showed a small peak associated with

crystallization at around 12 ° on the 2 theta scale in the 14w/v % and the 16% w/v but it

disappears in the 18% w/v PCPU film. This suggests that the DMF:EtOAc solvents influences

crystallinity in the polymer chains and as the concentration was increased (reduce the amount of

57
solvent used) the ability to create the ordered segments is reduced. Sharper reflections were noted

in the 40-50 ° regions. The electrospun polymers intensity profile shown in Figure 9, illustrates

distinct crystal peaks for all the samples. This suggests that the solvent used as well as the

electrospinning process allows for ordering of the crystal segments. Melt processed PCPU

exhibits no peak which indicates mixing of the hard crystal lamellae and the soft amorphous

regions. This is divergent to the concept that electrospinning reduces the crystallinity of the fibers

[36]. Lee et al [37] reported that the crystalline structure in fibers can be developed in ductile

polymers. Polymers that have low Tg values such as our novel PCPU (Tg~ -22.9°C) takes a

longer time to crystallize [36]. Overall, the sharper reflections in electrospun samples indicate

crystallization occurs during the electrospinning process when the polymer fibers are elongating

and possibly after the fibers have dried and solidified.

1200

14wt% neat film


1000
16wt% neat film

800 18 wt% neat film


Counts

600

400

200

0
0 20 40 60 80
2θ scale

Figure 2.9. XRD showing detection of ordered structure for the 14% w/v and 16% w/v solution

cast film.

58
7000
14wt%
6000
16wt%
5000 18wt%

4000
Counts

3000

2000

1000

0
0 20 40 60 80
2θ scale

Figure 2.10. XRD showing detection of ordered structure of the 14%, 16% and 18% w/v fiber

membranes.

2.3.7 Small-angle X-ray (SAXS)

Small angle x-ray data was recorded to understand structural features that are larger than

unit cell dimensions. Specifically, Fig [10]. There exhibits a peak at 0.14 A indicating an inter

lamellae spacing of 45Å. This was reported in our earlier paper on melt processed polymer [33].

We expected more distinct peaks in the electrospun samples, however, amplitude of the 0.14Å

peaks decreased as the concentration of the polymer solution used in spinning increased. It

appears as though alignment during the spinning process does not result thicker more well-

developed lamellae. The 45 Å structure are less plentiful in spun samples as well.

59
0.14 Å

Figure 2.11. SAXS of bulk PCPU and PCPU fiber membranes.

2.3.8 Tensile Testing

Representative stress-strain curves are depicted in Fig.12. Note that samples were run in

triplicate. Tensile strength, elastic modulus and strain to break are average values. We see that the

formation of electrospun PCPU fibers caused the modulus to initially decrease from 2.5MPa for

the bulk PCPU mold to 0.6 MPa in the 14% w/v neat fiber and 1.6 MPa in the 16% w/v neat fiber

then it increased to a modulus of 3.3MPa in the 18% w/v neat PCPU fiber. This shows that an

increase in the concentration of the pre-spun polymer cast solution drastically increases the tensile

strength of the polymer membrane. The elastic modulus data exhibits the same trend. The

increase in Tm noted in spun samples points to better developed crystals which may enhance

tensile moduli and strength. WAXS data did show sharper reflections in spun samples as well.

Baji et al tells us that the formation of crystals in the polymer matrix lends itself to increase

60
tensile strength [36]. The authors proposed that the strength and elastic modulus of the fibers are

influenced by the crystal lamellar and amorphous fractions of the chains within the polymer

fibers. Additionally, they proposed that structural formation changes taking place in the fibers

during electrospinning, specifically crystallinity and molecular orientation impart physical

uniqueness to the material and play an important role in the deformation behavior of the fibers

[36]. Strain to break values in spun samples are somewhat lower that of neat PCPU. Our future

studies will focus on the effect of spinning variables on tensile properties.

4.5 3.5
b
Elastic Modulus (MPa)

4 a 14 wt NEAT FIBER 3 3
3.5 . 16 wt NEAT FIBER 4 2.5 .
Stress (Mpa)

3 2
18wt NEAT FIBER
2.5 1.5
2 NEAT PCPU mold
1
1.5 0.5
1
0
0.5
14 wt 16 wt 18wt neat
0
NEAT NEAT NEAT PCPU
0 100 200 300 400 FIBER FIBER FIBER
Strain %

4.5
c.
Tensile Strength (Mpa)

4 400 d.
3.5
3
300
2.5
2
1.5 200
1
0.5 100
0
14 wt 16 wt 18wt neat 0
NEAT NEAT NEAT PCPU 14 wt NEAT16 wt NEAT 18wt NEAT neat PCPU
FIBER FIBER FIBER FIBER FIBER FIBER

Figure 2.12. a. Stress-Strain behavior, b. Elastic modulus, c. Tensile strength and d. Strain
at break of bulk PCPU and PCPU fiber mats.

61
2.3.9 PCPU/ nanocomposite membranes

2.3.9.1 Solution Viscosity

The solution viscosity for the PCPU and PCPU nanocomposite solutions are given in

figure 2.13. The solution viscosities generally increased as the % w/w of nanoparticles were

added to the PCPU solutions. For example, the 5% w/v PCPU/silica solution increased to twice

that of neat PCPU from 5 at shear rate to 13 mPa. The solutions also displayed pseudoplastic

behavior as seen in the neat PCPU which suggests that the addition of the nanoparticles does not

change the fluid like behavior of the polymer.

16 wt % neat

700 Viscosity-0.5silver 16wt% 1500


Viscosity-0.5silica16 wt%
Viscosity-0.5cb 16 wt%
1400

650
Viscosity (mPa)

Viscosity (mPa)
1300

1200
600

1100

550 1000
0.0 20.0 40.0 60.0 80.0 100.0 120.0 140.0 160.0 180.0
Shear rate (1/s)

Figure 2.13 Solution Viscosities of 0.5%w/w Ag, 0.5% w/w SiO2 & 0.5% w/w CB PCPU
nanocomposite solutions at 16% w/v concentration.
62
2.3.9.2 SEM

The SEM images were done to evaluate the fiber diameters for the PCPU nanocomposite

fibers. The 16wt% neat PCPU fiber diameter ranged from 900nm to 1μm. The nanocomposite

fibers ranged from 469 nm- 855 nm for the carbon black nanoparticles, 277-541nm for the silver

nanoparticles and 498nm-1.0μm for the Silica nanoparticles. Overall the nanofiber diameters were

drastically reduced when the nanoparticles were introduced to the PCPU solution

Figure 2.14. SEM images a.-c., 0.5%w/w Ag, 0.5% w/w SiO2 & 0.5% w/w CB PCPU w/v
fiber membranes; d.-f. Film-like membrane composites.

63
2.3.9.3 TGA

The TGA graph for the PCPU and 0.5% w/w nanoparticle membranes are shown in figure 2.15.

The graph illustrates that the degradation curve is shifted. The onset temperatures shown in Table

2.2 reveal that the nanoparticles are acting as reinforcements to the PCPU fibers which allow for a

higher degradation temperature from 286⁰C for the neat PCPU fiber to 288⁰C for the PCPU/SiO2

fiber. Even at 99% mass the composites membranes were at a higher temperature than the neat

PCPU increasing from 173⁰C to as high as 261⁰C.

100

90

80
16wt% neat pcpu fiber
70

60 0.5 wt% SiO2 fiber


weight (%)

50 0.5wt% cb fiber

40
0.5wt% Ag fiber
30

20

10

0
0 100 200 300 400 500 600
Temperature (⁰C)
Figure 2.15 TGA graph of neat PCPU (pink) and 0.5 wt% Ag (blue), 0.5 wt% SiO2 (green)
and 0.5wt% CB /PCPU nanocomposite fiber membranes.

64
Table 2.2 TGA data showing onset temperature, temperature at 99% and 50% mass for
PCPU/ 0.5% w/w nanoparticle fiber membranes.
Sample name Temperature- Temperature at Temperature at
onset (⁰C) 99% mass (⁰C) 50% mass (⁰C)

neat PCPU 289.0 173.6 338.9


16% w/v neat PCPU fiber 286.0 237.6 329.5

0.5% w/w Ag fiber 286.5 234.5 332


0.5 % w/w CB fiber 287.0 222.9 337.1
0.5% w/w SiO2 fiber 288.0 261.8 334.6

2.3.9.4 DSC

The DSC scans for the PCPU composite nanomembranes as shown in figures 2.16. The

glass transition temperatures for the nanocomposite fibers have decreased which means that

plasticization of the polymer matrix remains even though the polymer was electrospun. The silica

nanofiber composite showed a downward trend for the Tg which was opposite to that of the silica

molded composite detailed previously. This suggests that the electrospinning process may have

broken up any possible aggregation that can occur for the silica composites. The melt temperature

of the fibers varied 63⁰C and 78 ⁰C so there was no apparent melt temperature illustrated. Here the

expected crystallization peak is again absent from the thermogram which is similar to that of the

neat PCPU fiber composite.

65
16%w/v neat PCPU fiber
0.5 %w/w Ag
0.5 % w/w CB
0.5 %w/w SiO2

-25⁰C
64.5⁰C
-25.9⁰C
-25.4⁰C 63⁰C
-22.9⁰C 70.1⁰C
70.7⁰C

Figure 2.16 DSC thermos gram of neat PCPU (red) vs 0.5%w/w silver (Blue) and 0.5%
w/w carbon black (green) and 0.5% w/w silica (purple) PCPU fiber membranes.

Table 2.3 DSC data showing glass transition temperature (Tg) and melt temperature (Tm)
for PCPU/nanoparticle fiber membranes.

Sample name Tg (⁰C) Tm(⁰C) ΔHm (J/g)

neat PCPU -22.9 83.9 2.9


16% w/v neat PCPU fiber -22.8 70.7 3.3

0.25% w/w Ag fiber -25.0 73.1 1.1


0.5% w/w Ag fiber -25.4 70.1 3.6
0.75%w/w Ag fiber -25.3 61.6 0.6
1.0% w/w Ag fiber -24.6 74.6 0.3
0.25 % w/w CB fiber -25.1 61.4 2.8
0.5 % w/w CB fiber -25.9 63.0 9.9
0.75 % w/w CB fiber -22.9 77 4.0

66
Table 2.3 continued

1.0 % w/w CB fiber -25.7 75.0 3.0


0.25% w/w SiO2 fiber -23.4 78.1 0.25
0.5% w/w SiO2 fiber -25.0 64.5 9.4
0.75 % w/w SiO2 fiber -25.1 65.3 0.3
1.0 % w/w SiO2 fiber -25.4 77.1 0.1

2.3.9.5 WAXS

The WAXS data for the PCPU nanocomposite membrane are shown in figures 2.17, 2.18

& 2.19. For the silver nanocomposites the indicative intensity peaks at 38.4⁰, 44.6⁰, and 64.6⁰

represent the presence of silver nanoparticles. Because of limiting testing standards, the 77.5⁰

peak is not shown. For the 0.25 & 0.5 %w/w, the crystalline nature of the polymer

nanomembrane still exists but as we increase the % w/w the crystallinity disappears which means

that there can be interactions of the silver nanoparticles interrupting the crystal lamellae. Another

peculiar result came from the 1% w/w silver composite. It shows an extra peak which is known as

an amorphous peak. This can be from the presence of moisture at the surface of the

nanomembrane from the transportation of the fiber for testing. The amorphous nature of the silica

nanocomposite became more prevalent at a lower weight percent. There was a small crystal peak

present at 12.6⁰ in the 0.25 % w/w nanomembrane but it was absent in the other % w/w

nanomembrane samples. The WAXS scans for the PCPU/ carbon black composite

nanomembranes showed a crystalline peak at 12.5⁰ and 12.7⁰ for the 0.25 %, 0.5% and 0.75 %

w/w. For the 1% w/w carbon black composite fiber, the amorphous nature from the transport of

the fibers gave amorphous peaks at 30.1 ⁰.

67
Figure 2.17 WAXS of 0.25%w/w (blue), 0.5%w/w (green), 0.75%w/w (purple) & 1.0%
w/w (yellow) PCPU/ Ag nanocomposite fiber membranes.

68
Figure 2.18 WAXS of 0.25%w/w (brown), 0.5%w/w (yellow), 0.75%w/w (green) &
1.0% w/w (blue) PCPU/ SiO2 nanocomposite fiber membranes.

Figure 2.19 WAXS of 0.25%w/w (green), 0.5%w/w (pink), 0.75%w/w (yellow) & 1.0%
w/w (blue) PCPU/ CB nanocomposite fiber membranes.

69
2.3.9.6 Tensile Testing

The graphs in figures 2.20-2.22 show the stress strain curves for the 16% w/w neat PCPU

and PCPU composite fibers. For the carbon black the tensile profile initially decreased compared

to the neat PCPU tensile profile, but as the % w/w increased the tensile strength profile of the

fiber increased. The modulus of the 1%w/w carbon black composite fiber had a modulus of

4.3MPa which is three times the modulus of the neat PCPU fiber 1.3MPa (Table 2.3). The silver

nanocomposite fibers tensile profile decreased compared to neat PCPU for 0.25% to 0.75%w/w

silver whereas the 1%w/w increased only slightly. The silica fiber samples also showed reduced

mechanical strength on the addition of the nanoparticles. This suggests that adding the

nanoparticles adversely affected the mechanical strength when the PCPU nanocomposite was

electrospun.

4.5

neat PCPU fiber


3.5
0.25wt% CB fiber
3 0.5wt% CB fiber
Stress (Mpa)

0.75wt% CB fiber
2.5
1.0wt% CB fiber

1.5

0.5

0
0 50 100 150 200 250
Strain %
Figure 2.20 Stress vs strain curves for CB nanocomposite fiber membranes.
70
2.5 neat PCPu fiber 0.25wt% Ag fiber

0.5wt% Ag fiber 0.75wt% Ag fiber

1.0wt% Ag fiber
2

1.5
Stress (Mpa)

0.5

0
0 50 100 150 200 250
Strain %

Figure 2.21 Stress vs strain curves for Ag nanocomposite fiber membranes.

71
0.25wt% SiO2 fiber

1.6 0.5wt% SiO2 fiber


0.75wt% SiO2 fiber

1.4 1.0wt% SiO2 fiber


neat PCPU fiber
1.2

1
Stress (Mpa)

0.8

0.6

0.4

0.2

0
0 50 100 150 200 250
Strain %

Figure 2.22 Stress vs strain curves for SiO2 nanocomposite fiber membranes.

72
Table 2.4 Tensile strength data on electrospun PCPU and PCPU nanocomposite fiber
membranes.
Elastic Modulus Ultimate Tensile Strength Strain at Break

Sample (MPa) (MPa) (%)

16 wt NEAT

FIBER 1.3 1.5 297.9

0.25 %w/w CB 0.5 0.53 191

0.5 %w/w CB 0.77 1.03 252

0.75 %w/w CB 0.99 1.4 219

1 %w/w CB 4.3 4.5 258

0.25 %w/w Ag 0.43 0.49 190

0.5 %w/w Ag 0.5 0.62 211

0.75 %w/w Ag 0.88 0.92 139

1 %w/w Ag 1.5 1.94 199

0.25 %w/w SiO2 0.32 0.5 209

0.5 %w/w SiO2 0.6 0.6 181

0.75 %w/w SiO2 1.01 1.05 175

1 %w/w SiO2 1.31 1.35 119

2.4 Conclusion

An ultrasoft polycarbonate polyurethane was electrospun to create fiber membranes. The

resulting material showed variations in its thermal, mechanical and molecular properties when the

polymer solution concentration was varied. The 18% w/v polymer membrane showed the highest

thermal stability with an onset of 284 ⁰C compared to the 14 % w/v at 279 ⁰C. One glass transition

73
temperature was found to be -22.9 °C and it did not change as we moved from the bulk polymer

to the fiber membranes. The enthalpy of melt showed a slight increase to suggest crystal

formation. WAXS data demonstrated the presence of clearer reflections in spun samples than

those obtained in solution cast films and bulk polymer. SAXS did not show any variations in the

interdomain spacing which remained at 0.45Å. Singular fiber SAXS yield a better idea of the

molecular orientation of the electrospun PCPU. The mechanical properties showed that the

electrospun PCPU high Young’s modulus of 3.3 MPa as well as an ultimate tensile strength of 4

MPa compared to the bulk PCPU. The 16% w/v sample had the highest elongation at break of

280% which showed that its fiber membrane has the ability to stretch to almost four times its

original length.

On the addition of the nanoparticles, the fiber diameter of the PCPU nanofibers decreased

well within the nm range. However, the tensile properties were greatly affected. Only the 1% w/w

carbon black composite fiber showed a dramatic increase in strength by having a young’s

modulus of about 3 times that of neat PCPU fiber membrane. Contrastingly the thermal

performance of the nanocomposite fibers was increased on the addition of the nanoparticles by an

increase in the thermal stability of the composite fibers and the decrease in the glass transition

temperature. These results give us a clearer impression of how processing conditions can be

modified to tune the properties of the fibers.

2.5 References

1. N. Tucker, J. J. Stanger, M. P. Staiger, H. Razzaq and K. Hofman (2012) The History

of the Science and Technology of Electrospinning from 1600 to 1995, Journal of

Engineered Fibers and Fabrics, SPECIAL ISSUE, 63-73.

74
2. H. Katakam, Master’s Thesis, Fabrication and Characterization of Polycarbonate

Polyurethane (PCPU) Nanofibers Impregnated with Nano fillers.

3. S. Wang, Y. Wan, B. Sun, L. Liu and W. Xu (2014), Mechanical and electrical

properties of electrospun PVDF/MWCNT ultrafine fibers using rotating collector,

Nanoscale Research Letters, 9:522

4. H. Mi, X. Jing, E. Yu, X. Wang, Q. Lid, L. Turng (2018), Manipulating the structure

and mechanical properties of thermoplastic polyurethane/polycaprolactone hybrid

small diameter vascular scaffolds fabricated via electrospinning using an assembled

rotating collector, Journal of the Mechanical Behavior of Biomedical Materials, 78,

433–441

5. J. Li, C. Nie, B. Duan, X. Zhao, H. Lu, W. Yin, Y. Li, X. He (2017), Tensile

performance of silica-based electrospun fibrous mats, Journal of Non-Crystalline

Solids 470, 184–188

6. D. Li, Y.N. Xia (2004), Electrospinning of Nanofibers: Reinventing the Wheel? Adv.

Mater 16, 1151–1170.

7. D. H. Reneker, I. Chun (1996), Nanometre diameter fibres of polymer, produced by

electrospinning, Nanotechnology 7, 216–223.

8. A. Greiner, J. H. Wendorff (2007), Electrospinning: a fascinating method for the

preparation of ultrathin fibers, Angew Chem. Int. Ed. 46, 5670–5703.

9. M. G. McKee, C. L. Elkins, T. E. Long (2004), Influence of self-complementary

hydrogen bonding on solution rheology/electrospinning relationships, Polymer 45,

8705–8715.

75
10. H. Karakaş, A. S. Saraç, T. Polat, E. G. Budak, S. Bayram, N. Dağ, and S. Jahangiri

(2013) World Academy of Science, Engineering and Technology, International,

Journal of Health and Medical Engineering Vol: 7, No:3.

11. E.J. Chong, T.T. Phan, IJ Lim, Y.Z. Zhang, B.H. Bay, S. Ramakrishna S. Acta Mater;

3:321– 30, 2007.

12. S. W. Thomas, N. A. Alcantar, Yanay Pais (2017) . US patent US20170157571 A1

13. L. R. Lakshman, K. T. Shalumon, S. V. Nair, R. Jayakumar and S. V. Nair (2010)

Journal of Macromolecular Science, Part A: Pure and Applied Chemistry, 47, 1012–

1018

14. G. E. Martin, I. D. Cockshott and F. J. Fildes (1977) US Patent 4,044,404.

15. H. Planck and P. Ehrler (1984) US Patent 4,474,630.

16. How TV (1985). US Patent 4,552,707.

17. D. Groitzsch and E. Fahrback (1986) US Patent 4,618,524.

18. W. Simm (1976) US Patent 3,994,258.

19. A. Pedicini and R. J. Farris (2003), Mechanical behavior of electrospun polyurethane,

Polymer 44, 6857–6862

20. J. Liu, D. Y. Lin, B. Wei and D. C. Martin (2017), Single Electrospun PLLA and PCL

Polymer Nanofibers: Increased Molecular Orientation with Decreased Fiber Diameter,

Polymer 118, 143-149

21. P. Gibson, H. Schreuder-Gibson and D. Rivin (2001), Transport properties of porous

membranes based on electrospun nanofibers, Colloids Surfaces A Physicochem Eng.

Asp. 187-188, 469-481.

76
22. W. J. Li, C. T. Laurencin, E. J. Caterson, R. S. Tuan amd F.K. Ko (2002), Electrospun

nanofibrous structure: A novel scaffold for tissue engineering, J. Biomed. Mater. Res.

60, 613-621.

23. J. A. Matthews, G. E. Wnek, D. G. Simpson and G. L. Bowlin (2002), Electrospinning

of Collagen Nanofibers, Biomacromolecules 3 232-238.

24. D. H. Reneker, A. L. Yarin, E. Zussman and H. Xu (2007) Electrospinning of

nanofibers from polymer solutions and melts, Adv. Appl. Mech. 41, 43-195.

25. M. Gazzano, C. Gualandi, A. Zucchelli, T. Sui, A. M. Korsunsky, C. Reinhard and M.

L. Focarete (2015), Structure-morphology correlation in electrospun fibers of

semicrystalline polymers by simultaneous synchrotron SAXS-WAXD, Polymer 63,

154-16.

26. X. Song, T. Li, B. Cheng and J. Xing (2016), POSS–PU electrospinning nanofibers

membrane with enhanced blood compatibility, RSC Adv., 6, 65756-65762.

27. M. Hartmann (2004), Hierarchical zeolites: a proven strategy to combine shape

selectivity with efficient mass transport, Angew Chem Int Ed, 43:5880-5882.

28. H. Ogihara, S. Takenaka , I. Yamanaka, E. Tanabe, A. Genseki, K. Otsuka (2006),

Synthesis of SiO2 Nanotubes and Their Application as Nanoscale Reactors, Chem.

Mater. 2006;18:996–1000.

29. L. Chen, J. Liao, S. Lin, Y. Chuang and Y. Fu (2009) Synthesis and characterization of

PVB/silica nanofibers by electrospinning process, Polymer 50, 3516–3521

30. S. Chuangchote, A. Sirivat and P. Supaphol1 (2007) Mechanical and electro-

rheological properties of electrospun poly(vinyl alcohol) nanofibre mats filled with

carbon black nanoparticles, Nanotechnology, 18, 145705

77
31. R. W. Bass: Ph.D. Thesis, Synthesis and Characterization of Self-Healing Poly

(Carbonate Urethane) Carbon-Nanotube Composites, University of South Florida

(2011),

32. J.P. Harmon, R.W. Bass, Self-Healing Polycarbonate Polyurethane Nanotube

Composites. US Patent: 8,846,801, (2014)

33. K. L. Kull, R. W. Bass, G. Craft, T. Julien., E. Marangon, C. Marrouat and J. P.

Harmon (2015), Synthesis and Characterization of an Ultra-soft Poly (Carbonate

Urethane), European Polymer Journal 71, 510-522.

34. R. W. Bass, T. C. M. Julien, A. Rigolo, S. Hong, X. Li, and J. P. Harmon (2016),

Thermal and mechanical properties of single-walled and multi-walled carbon nanotube

polycarbonate polyurethane composites with a focus on self-healing, International

Journal of Materials Research, 107:8, 692-702

35. M. P. Stevens (1999), Polymer Chemistry: An Introduction, 3rd Ed. New York:

Oxford.

36. A. Baji, Y. Mai, S. Wong, M. Abtahi, P. Chen (2010), Electrospinning of polymer

nanofibers: Effects on oriented morphology, structures and tensile properties,

Composites Science and Technology 70, 703–718

37. K. H. Lee, H. Y. Kim, Y. M. La, D. R. Lee and N. H. Sung (2002) Influence of a

mixing solvent with tetrahydrofuran and N, N-dimethylformamide on electrospun

poly(vinyl chloride) nonwoven mats. J Polym Sci Part B: Polym Phys, 40:2259–68.

78
CHAPTER 3

Rheological and Electrical Characterization of Thermoplastic and Thermoset Polymer

Lithium electrolyte.

3.1 Introduction

When it comes to electronic devices and other electronic consumer goods, lithium ion

batteries are the main system used to power them. This is because lithium ion batteries possess

high energy density, a long lifespan and is flexible [1]. However, lithium ions have the propensity

to form dendrites during charge-discharge routes which can lead to explosive hazards. Therefore,

the idea of using lithium ions within polymers as an alternative was proposed. The demand for

lithium ion batteries in numerous devices have increased, therefore, polymer electrolytes research

has gained widespread notoriety due to their technological uses in solid state electrochemical

devices, electrochromic devices, rechargeable lithium batteries, fuel cells, super capacitors,

biosensors [2-7]. Moreover, solid polymer electrolytes (SPEs) are an advantage over liquid or gel

electrolytes as the SPEs have the added mechanical property which would allow for better

maintenance. These electrolytes are important for electronic materials for our everyday use. The

earliest work on ions within a solid came from Armand et al [8] as well as Fenton et al [9] which

prompted further study metal salts in polymers.

Polymers range from highly crystalline stiff polymers to blend polymers, however highly

crystalline polymers are not well suited to become electrolytes as the highly crystalline nature

would not allow for the lithium ions to flow within the polymer matrix. Therefore, a

79
semicrystalline polymer with low crystallinity is better suited. Polyethylene oxide (PEO) is one of

the most sought after polymer used as an electrolyte but PEO tends to crystallize at ambient

temperatures which would hinder the migration and movement of the lithium resulting in a low

conductive electrolyte [3,10-12]. Therefore, polymers need to have good thermal stability; its

polymer backbone needs to be mobile at ambient temperatures and should not crystallize. Many

studies have been executed to show modifications of PEO to develop an improved high

conductive polymer electrolyte [6, 13-16]. These modifications came from linear polymers [17-

19, 27] comb-branched copolymers [20- 21], block copolymers [22-23], cross-linked network

polymers [24] and polymer blends [25-26]. Parveen et al [3] used an amorphous polyurethane to

incorporate into PEO to reduce the crystallinity of the polymer to create an electrolyte fiber.

PCPU has a characteristic two-phase morphology that shows mixing of its hard and soft

segments. PCPU above is Tg makes for a mobile backbone that allows the ions to flow within the

matrix. Polycarbonate polyurethane (PCPU) combined with lithium salt was analyzed for its

ability to be a highly conductive electrolyte that does not easily degrade. PCPU has a Shore A

hardness of 64, whereas not many polyurethanes can reach such soft grades and maintain

processability. This is due to the PCPU’s composition of a hard segment, providing stability, and

soft segment allowing the polymer to flow at high temperatures instead of degrading.

Polyimides are also good contenders for lithium applications. They are high-performance

polymers that possess excellent thermal stability, outstanding mechanical properties, and low

dielectric constants [28]. Moreover, highly aligned polyimide nanofiber membranes prepared by

electrospinning, showed excellent mechanical and thermal properties [33-34]. Electrospun PI

nanofiber membranes which have been intensively investigated resulted in high-performance and

multifunctional composite fiber membranes [32-34]. Cho et al [29], Yang et al [30] and Wu et al

80
[31] investigated and evaluated electrospun polyacrylonitrile (PAN) and poly(vinylidene fluoride)

(PVF) nanofiber based nonwovens in LIBs and exhibited outstanding battery performances, such

as large capacity, high-rate capability and long cycle life. PI has the ability to avoid short

circuiting because it is thermally stable at temperatures above 500⁰C whereas conventional

separators can only go as high as 150⁰C. PI nanofiber nanomembranes’ extraordinary thermal,

mechanical and electrochemical properties demonstrate its capability to be used as an ideal

separator for Lithium ion batteries to achieve high battery performance, such as large capacity,

high-rate capability and long cycle life [28].

Figure 3.1 Polyimide GPI 15.

An ultrasoft thermally stable molded and fibrous polymer electrolyte with lithium salts

was developed by direct addition of industrial grade Lithium phosphate (LiPF6) to novel

polycarbonate polyurethane (PCPU) via ultrasonication. LiPF6 of varying concentration from

12w/w% - 16w/w% was investigated which covers the industrial amount (14 w/w%) used in

electrolytes. The thermal and electrical properties of varying concentrations of PCPU/LIPF6

molded and fibrous electrolytes were investigated using Dynamic Mechanical Analysis (DMA)

and four point probe. DMA showed a high interaction between lithium salts and the polymer

matrix which resulted in a decrease in glass transition temperature of more than ten degrees.

Greater interaction of the PCPU and the lithium salts resulted in an increase in ionic conductivity

81
at room temperature. The resulting polymer electrolyte is highly flexible, thermally stable and has

biocompatible properties which can be useful as biosensors. Through incorporation of lithium

salts, lithium hexafluorophosphate (LiPF6), this composite could lead to an innovative, ultra-soft,

conductive polymer electrolyte film.

Additionally, the research herein looks at variations of an unprocessable polyimide

powder GPI 15, and transforms it into polyimide membranes. Very little literature shows the

ability to electrospun a polyimide. Rather the studies show the electrospinning of the PI precursor

which is a polyamine acid (PAA) and then imidization. Here this study illustrates the ability to

electrospin polyimide powders. The membranes formed are then doped in two concentrations of

lithium salt solutions to create polyimide lithium electrolytes. TGA was used to characterize the

thermal ability of the polyimide fiber and morphological studies were done using SEM. The

characterization methods done on the lithium polyimide fibers include, Fourier transform infrared

spectroscopy (FTIR), and Four Point Probe.

3.2 Experimental

3.2.1 Polymer-Lithium composite synthesis

The solvents used (DMF and THF), to create the polymer lithium composite were dried

using molecular sieves and sealed with parafilm until ready for use. Polycarbonate

polyurethane/lithium composites were prepared by dispersing the lithium salt using a sonicator.

10g of the purified PCPU was dissolved in 100 ml dried THF ( ACS reagent, 99.9% sigma

Aldrich) and then the desired amount of Lithium hexafluorophosphate salt (12-16% w/w) wase

introduced to the solution and then mixed by ultrasonication (Fisher Scientific Sonic

Dismembrator 550) at a mixing rate of 10 minutes with 3 second rest intervals for two hours to

ensure homogeneity of composites. The composite was the solution cast and the resulting

82
material was dried overnight under vacuum. The material was then molded using a carver press

and characterized using the following methods.

3.2.2 Electrospinning of Polyimide

For the electrospinning process, each of the solutions was filled up in a 3ml syringe. A

20 gauge needle with a flat tip was used as a spinneret. A collection distance of 20cm was done

and a fixed voltage of 23kV was applied at a feeding rate of 50μl/min. The resulting

nanomembrane was left to dry under vacuum.

3.2.3 Polyimide/Lithium nanomembrane electrolyte

The polymer electrolytes were prepared by soaking the PI fiber membranes in 14 and 16

w/v LiPF6 solutions (DMF: THF) at room temperature for 12 hours in a glove box under nitrogen

atmosphere. The resulting electrolyte was dried under vacuum for two days to ensure complete

dryness.

3.2.4 Dynamic Mechanic Analysis (DMA)

In order to investigate the dynamic viscoelasticity of the PCPU/Lithium electrolyte

composite, dynamic mechanical tests were carried out with a TA Instruments AR-2000 was used

to explore changes in the glass transition, Tg , storage modulus, G', loss modulus, G” and tan δ.

Temperature sweep tests and isothermal frequency sweep tests were performed at a strain within

the materials linear viscoelastic region.

The temperature sweep tests were planned to explore the frequency dependent glass

transition temperature Tg of the material. They were carried out from -120⁰C to 100⁰C, using a

ramp rate of 5⁰C/min, and repeated at frequencies 0.2 to 10Hz. The isothermal frequency sweep

tests were conducted in the range from 1 to 20 Hz and repeated at various temperatures ranging

from -50 to 55⁰C, with an interval of 5⁰C. The stress responses to the strain excitations in these

83
isothermal frequency sweep measurements were recorded automatically, and G’, G’’ and tan δ

were calculated from these measurements.

3.2.5 Four-Point Probe conductivity

Four-point probe is a known instrument used to perform reliable measurements of

electronic transport properties in semiconductors and electrical materials. Conventional four-point

probes are millimeter sized devices with spring loaded electrodes of tungsten carbide [35-36]. The

typical set up is shown in figure 3.2. Two of the probes are used to source the current from the

AC/DC current source and the other two inner probes are used to measure voltage from the

electrometer. The idea of using four probes eliminates measurement errors due to the probe

resistance, the spreading resistance under each probe, and the contact resistance between each

metal probe and the semiconductor material. This technique involves bringing four equally spaced

probes into contact with the material of unknown resistance.

Figure 3.2 Four Point probe set-up.

84
The volume resistivity is calculated with this equation:
𝜋 𝑉
𝜌 = 𝑙𝑛2 × 𝐼 × 𝑡 × 𝑠 (3.1)

where: ρ = volume resistivity (Ω-cm), V = the measured voltage (volts), I = the current applied

(amperes), t = the sample thickness (cm), s = the probe spacing. Conductivity can then be

calculated by taking the reciprocal of the resistivity.

Electrical examination of the PCPU/lithium composites were carried out using four point

probes attached with Kiethley 6220/ 6514 electrometer. Thin film samples were run by applying

1.0mA and voltages were recorded from the electrometer over a 60 second period. To samples of

thickness 0.1 cm, the surface of the PCPU/LiPF6 composite was cut and samples were run over a

24 hour period. The polyimide electrolytes were run over a 5 minute period.

3.2.6 Solution Rheology Measurement

A Fungilab alpha series rotational viscometer equipped with a TR8 spindle was used to

measure the viscosity of the polyimide solutions. 10ml of each solution was placed in the

cylindrical chamber and the spindle is immersed in the solution. The rotational speed was varied,

and viscosity readings were recorded. All measurements were performed at 21⁰C.

3.2.7 Surface Enhanced Microscopy (SEM)

A Hitachi S-800 SEM was used which compromises of a magnification power of

300,000 times the actual sample size. SEM gives information about fiber diameter, bead

formation, quality of the polyimide fiber membranes produced from electrospinning.

85
3.2.8 Fourier Transform Infrared Spectroscopy (FTIR)

Perkin Elmer spectrum two furnished with and Attenuated Total Reflection (ATR)

accessory was utilized for this study. A scanning range of 400-4000cm-1 set at a resolution of

4cm-1 and 16 repetitions were done on the nanocomposites. The neat PI and PI/LiPF6 nanofiber

composites were scanned.

3.2.9 Thermogravimetric Analysis (TGA)

A TA instruments Q500/50 TGA equipped with a standard furnace was used to detect the

degradation temperature of the polymer membranes. A sample weight between 15-20mg was

placed on a tared 100 ml platinum pan. Measurements were carried out under nitrogen

atmosphere as the samples were heated to 800⁰C at a heating ramp rate of 10⁰C per minute.

3.3 Results and Discussion

3.3.1 Dynamic Mechanical Analysis (DMA)

Dynamic Mechanical Analysis can be used to quantify the viscoelastic properties of

polymers. The loss modulus, G", is a measure of the ability of a material to dispel mechanical

energy by converting it into heat. This absorption of mechanical energy is often related to the

movements of molecular segments within the polymer sample [37]. There were two types of

DMA techniques employed, temperature sweep and isothermal frequency sweep. Temperature

sweep DMA was used to determine the rheological behavior of the PCPU lithium nanocomposites

as well as identify the glass transition temperature Tg which is associated with chain slippage and

movement. Tg can be determined from the peak maximum of either the G’ or tan δ.

86
Figure 3.3, 3.4, 3.5 & 3.6 shows the temperature sweep test results for neat PCPU and

PCPU composites at varying frequencies. The storage modulus of the composites displays

changes in the rubbery plateau as well as the primary alpha shift. 12% w/w and 14 %w/w

PCPU/LiPF6 maintain a longer rubbery plateau compared to the 16 %w/w and the neat PCPU.

The 16 % w/w experiences an early melt at around 50⁰C which may be due to high lithium

loading which may have affected the melting of the sample. Figure 3.7 gives us the loss modulus

overlay for neat PCPU and PCPU lithium composites which display differences in the primary

( ) and secondary (β) relaxations of the composites at low temperatures.

Figure 3.8 shows the tan δ results from the temperature sweep. It is evident that the peak

tan δ, and therefore Tg, varies as the %w/w of the PCPU/LIPF6 changes, with higher % w/w

causing a shift of Tg to lower temperatures. This is evidence of plasticization of the amorphous

region of the PCPU matrix which would suggest that the lithium salts are interacting with the soft

segments of the polyurethane as seen by [27]. Table 3.1 shows the G” values of Tg of the different

composites. There is a shift of Tg of approximately a 10 degree difference between the neat PCPU

and the 16%w/w PCPU/LiPF6 composite.

87
Neat pcpu dma neg 120 to 150-0024o

Secondary transition

Figure 3.3 DMA Data: Temperature sweep at multiple frequencies of neat PCPU.

88
DMA-12wt% LiPF6/PCPU electrolyte

Secondary transition

Rubbery plateau

Figure 3.4 DMA Data: Temperature sweep at multiple frequencies of 12


wt% PCPU/LiPF6 composite.

89
DMA- 14 WT% LIPF6/PCPU

Secondary transition

Figure 3.5 DMA Data: Temperature sweep at multiple frequencies of


14wt% PCPU/LiPF6 composite.

90
DMA-16wt% LiPF6/PCPU electrolyte

Secondary transition

Figure 3.6 DMA Data: Temperature sweep at multiple frequencies of 16


wt% PCPU/LiPF6 composite.

91
Overlay

DMA Tg

Neat PCPU
12%w/w PCPU/LiPF6/PCPU
14%w/wPCPU/LiPF6/PCPU
16%w/wPCPU/LiPF6/PCPU

Figure 3.7 DMA Data: Loss modulus plot of neat PCPU and PCPU/LiPF6
composites. Overlay

Neat PCPU DMA Tg


12%w/w PCPU/LiPF6/PCPU
14%w/wPCPU/LiPF6/PCPU
16%w/wPCPU/LiPF6/PCPU

Figure 3.8 DMA Data: tan delta plot of neat PCPU and PCPU/LiPF6
composites.
92
Table 3.1 DMA Data: DMA Tg and tan δ for the PCPU and PCPU/LiPF6 composites.

Sample DMA Tg G” (⁰C) Tan δ height Tan δ width (Δ⁰C)

Neat PCPU -19.4 0.381 55.9

12%w/w LiPF6/PCPU -23.6 0.429 44.8

14%w/w LiPF6/PCPU -28.8 0.447 44.9

16%w/w LiPF6/PCPU --29.7 0.456 50.7

3.3.2 Secondary-relaxation

The secondary relaxation or β-relaxation of neat PCPU and PCPU lithium composites, as

measured by the loss modulus G” Figure 3.7, occurs between temperature range -120 to -80 ºC. It

follows Arrhenius behavior which is characteristic of secondary relaxations in polymers.

Therefore, the activation energy can be obtained using the Arrhenius type relationship between

the experimental frequency and the temperature range in Kelvin.


∆𝐸𝑎
ln 𝑓 = ln 𝑓𝑜 − 3.2
𝑅𝑇

The resulting linear relationship gives us a slope that is used to find the activation energies

(Table 3.2). This activation energy is the amount of energy needed for the pendant methyl group

to rotate. The neat PCPU was found to have an activation energy of 34.4kJ/mol which almost

doubled to 63.7 kJ/mol for 12%w/w PCPU/LiPF6, 64.0 kJ/mol for the 14%w/w PCPU/LiPF6 and

93
65.3 kJ/mol in the 16%w/w PCPU/LiPF6 composite. This demonstrates that the lithium salts

within the backbone creates a higher need for energy in order for the side groups to rotate.

Table 3.2. DMA data: Activation Energy of the β-transition for neat PCPU and
PCPU lithium composites.

Sample Ea (kCal/mol) Ea (kJ/mol)

Neat PCPU 8.2 34.4

12%w/w PCPU/LiPF6 15.2 63.7

14%w/w PCPU/LiPF6 15.3 64.0

16%w/w PCPU/LiPF6 15.6 65.3

3.3.3 Primary-relaxation

For PCPU and its lithium composites, the alpha transition involves micro Brownian

motion in the main chain and conformational changes in the phenyl groups [38-39, 41]. The α

relaxation is apparent in tan δ vs temperature plots. The maxima in tan δ for Tg occur from -14.9

to -5.05 °C for neat PCPU,-10.8 to -9.1°C for 12% w/w PCPU/LiPF6, -4.1 to -20.3 °C for 14%

w/w PCPU/LiPF6 and from -20.7 to -20.1 °C for 16% w/w PCPU/LiPF6 (Figure 3.7A-D). The

plots of log frequency vs inverse temperature unveiled a curved behavior predicted by the

William-Landel-Ferry (WLF) equation (3.3) [40-42].

−𝐶1 (𝑇−𝑇0 )
log 𝑎 𝑇 = 3.3
𝐶2 +(𝑇−𝑇0 )

Where the aT is the shift factor that corresponds to frequency, and C1 and C2 are WLF constants.

These constants can be determined using the second DMA technique which is the isothermal

frequency sweeps.

94
1.4 0
1.3 a. b.
-0.2

Log Fmax (Hz)


Log Fmax (Hz)

1.2 -0.4
1.1
-0.6
1
-0.8
0.9
-1
0.8
-1.2
0.7
3.795 3.8 3.805 3.81 3.815
3.55 3.6 3.65 3.7 3.75
1000/T K-1
1000/T K-1

1.2 0
1 c. d.
0.8
-0.2
0.6 Log Fmax (HZ)
LogFmax (Hz)

-0.4
0.4
0.2 -0.6
0
-0.8
-0.2
-0.4 -1
-0.6
-1.2
-0.8
3.95 3.955 3.96 3.965
3.7 3.8 3.9 4
1000/T K-1
1000/T K-1

Figure 3.9 Plots of ln frequency vs 1000/T showing WLF behavior at Tg for a. neat PCPU,
b. 12% w/w PCPU/LiPF6, c. 14%w/w PCPU/LiPF6 and d. 16% w/w PCPU/LiPF6.

Figure 3.10a, shows the isothermal frequency sweep results for G’ for temperatures

between -50 to 55 °C. The graph demonstrates that at higher temperatures above the Tg from 0⁰C

to 55⁰C, there is a frequency dependence occurring which means that there is less time for

relaxation to occur but at lower temperatures a change in frequency does not change the storage

modulus. Figure 3.10b shows the master curve fit after the TTS is imposed. From this the WLF

graph is plotted (Figures 3.11-3.14), and the values of C1 and C2 were found to be 33.1, 217.7K,

for neat PCPU. The WLF equation describes the effect of temperature on the shift factor for many

95
polymers near their Tg. By putting the C1 and C2 values into equation 3.3 we would get the

following Arrhenius relationship;

𝐶
∆𝐸𝑎 = (−2.303) (𝐶1 ) 𝑅𝑇 2 (3.4)
2

Where, Ea is the activation energy, predicted by the temperature sweep experiments at the glass

transition region. The Activation energy from the WLF showed a much higher Ea because there is

a greater amount of energy needed for the polymer chains to begin moving in association with

each other.

neat pcpu freq sweep-0089o

Figure 3.10a. Isothermal frequency sweeps for neat PCPU and 3.10b. the master curve.

96
neat pcpu freq sweep-0089o

c1: 33.1850
c2: 217.029 K
Tref: -19.8247 °C
R2: 0.998615

Activation Energy: 183.482 kJ/mol


Tref: -19.8247 °C
R2: 0.999582

Figure 3.11 DMA data: WLF plot of neat PCPU.

Figure 3.12 DMA data: WLF plot for 12% w/w LiPF6/PCPU.

97
14wt Lipf6 freq sweep-0082o

c1: 41.7312
c2: 234.175 K
Tref: -30.1216 °C
R2: 0.999571

Activation Energy: 200.020 kJ/mol


Tref: -30.1216 °C
R2: 0.999987

Figure 3.13 DMA data: WLF plot for 14%w/w LiPF6/PCPU.

98
16wt Lipf6 freq sweep -0087o

c1: 41.4971
c2: 211.755 K
Tref: -35.5678 °C
R2: 0.998433

Activation Energy: 206.372 kJ/mol


Tref: -35.5678 °C
R2: 0.999874

Figure 3.14 DMA data: WLF plot for 16% w/w LiPF6/PCPU.

3.3.4 Conductivity

Resistivity and conductivity are fundamental properties for semiconductors and are critical

parameters in materials research [43]. The four-point probe measurements on the PCPU/lithium

composite electrolyte were done and the conductivity was measured. Figure 3.15 shows the

conductivity for the 12%,14% and 16% w/w PCPU/LiPF6 vs time. The neat sample had a too low

conductivity to display on the chart but it was recorded in table 3.3 as 2.4 e-10 S/cm. The graph

shows that the conductivity of the lithium composites increases as the % w/w of the lithium was

increased. In addition, the conductivity levels out overtime. This indicates that as the current is

initially passed through the polymer material there is a surge of lithium activity which then levels

off after a minute to give us the final constant voltage reading.

99
The conductivity was also taken of a circular mold with a thickness of 0.1cm which

allowed us to cut away the surface to see if there is any change in the conductive reading over

time Figures 3.16-3.18 displays the conductivity measurement over a period of 24 hours from the

pristine surface to the surface being cut away and what was the conductivity after 24 hrs. What

was observed is that when the surface is cut away the conductivity of the lithium composites

increased. This is due to the hydrogen bonding that is known to be exposed when the surface of

PCPU is cut away and this allowed for the availability of the lithium to become more accessible

which lead to the surge in the conductivity from 4.0 e-3 to 1.5e-2 S/cm in the 12%w/w, 2.0e-3 to

1.7 e-2 in the 14 % w/w and 5.0e-3 to 2.8 e-2 in the 16 %w/w. After 1 hour the conductivity

reduces drastically for the 12 and 14 %w/w down to 1.5 e-3 and 2.9 e-3 S/cm but there was a

small reduction to 2.3e-2 for the 16 w.t.%. Hence, it appears that the high lithium loading in the

16% w/w allows for the slowing of the H-bonds within the polymer matrix which permitted the

conductivity of the mold to remain high. After the 24 hour period the lithium composites

conductivity was closer to its starting conductive material.

100
5.050E-03

4.550E-03

4.050E-03

3.550E-03
Conductivity (S-cm)

3.050E-03

2.550E-03 16wt% LiPF6


14wt%LiPF6
2.050E-03
12wt% LiPF6
1.550E-03

1.050E-03

5.500E-04

5.000E-05
0 10 20 30 40 50 60
Time (sec)

Figure 3.15 Conductivity data: Conductivity of PCPU/LiPF6 thin film composites.

1.800E-02
12 wt % LiPF6/ PCPU
1.600E-02

1.400E-02
Conductivity (Scm)

1.200E-02 12 wt LiPF6 cut

1.000E-02
1 hr after surface cut
8.000E-03

6.000E-03 24 hr after surface cut

4.000E-03
12 wt% LiPF6 pcpu uncut
2.000E-03 surface

0.000E+00
0 10 20 30 40 50 60 70
time (s)

Figure 3.16 Conductivity vs time for 12 wt% PCPU/LiPF6 composite for cut and uncut
surface.

101
1.600E-02
14 wt % LiPF6/PCPU
1.400E-02

1.200E-02
Conductivity (Scm)

1.000E-02 Series1
8.000E-03 1 hr after cut
24 hr after cut
6.000E-03
14wt Lipf6 uncut surface
4.000E-03

2.000E-03

0.000E+00
0 10 20 30 40 50 60 70
time (sec)

Figure 3.17 Conductivity vs time for 14% w/w PCPU/LiPF6 composite for cut and uncut
surface.

3.000E-02
16 wt % LiPF6/PCPU mold
2.500E-02 LiPF6 surface cut
1 hr after cut
Conductivty (Scm)

2.000E-02
24 hr after cut
1.500E-02 uncut lipf6/pcpu surface

1.000E-02

5.000E-03

0.000E+00
0 10 20 30 40 50 60
time (min)

Figure 3.18 Conductivity vs time for 16wt% PCPU/LiPF6 composite for cut and uncut
surface.

102
Table 3.3 Conductivity data: Conductivities for uncut and cut surfaces of PCPU and
PCPU/LiPF6 nanocomposites.
Conductivity σ (S/cm)
Sample
Uncut Cut After 1 hr After 24 hr
neat PCPU 2.4 E-10 2.3 E-10 2.30E-10 2.30E-10
12wt% PCPU/LiPF6 9.1 E-4 2.60E-03 1.5 E-3 9.90E-04
14wt% PCPU/LiPF6 1.4 E-3 4.40E-03 2.9 E-3 1.40E-03
16wt% PCPU/ LiPF6 1.70E-03 3.2 E-3 3.2 E-3 2.0 E-3

3.3.5 Polyimide nanomembrane electrolyte

3.3.5.1 Solution Rheology

Figure 3.19 shows the solution viscosity as a function of shear rate for the 20% w/v

polyimides before electrospinning. Viscosity vs shear rate graphs tells us the different the

behavior of fluids. For a dilitant fluid there would be shear thickening with increasing shear stress

meaning that the viscosity would be increasing. For a pseudo plastic fluid there would be gradual

decrese in viscosity with changing stress through shear thinning and for a newtonian fluid , the

viscosity would be independent of the changing shear rate. The polyimide solutions demonstrates

psudoplastic behavior as the shear rates goes from 9s-1 to 170s-1 the viscosity of the polymer

solutions decrease from 97mPa (XP02691) to 12 mPa (P347). Another aspect realised by the

graph identifies an increase in Mw allows for a higher viscosity of the solutions from 25 mPa

(R127), whose Mw is ~68,000, to 37 mPa ( P347) whose Mw is ~78,000 and to 80 mPa (

XP02691) whose Mw ia 79,500. The polyimide solutions are low in viscosities which was also

evident by Liu et al [45] but they also exhibited pseudo plastic fluid behavior which suggest that

the orientation of the polymer chains influences the behavior of the fluid.

103
120

100

80
Viscosity (mPa)

60 R127
XP02691
40 P347

20

0
0.0 50.0 100.0 150.0 200.0
Shear rate (1/s)

Figure 3.19 Viscosity vs shear rate for pre-spun polyimide solutions.

3.3.5.2 Scanning Electron Microscopy (SEM)

The polyimide nanofibers were successfully electrospun and the SEM images are shown

in figure 3.20. The P347 PI showed fibers with diameters ranging from 155 to 650 nm and it was

easily plied from the foil collector. The R127 had fiber sizes of 341 nm but the membrane itself

could not be taken off of the foil to be used in the doping step. The XP0291 contained fiber

diameter sizes ranging from 297 to 800 nm but the pore sizes were too large for testing for battery

separator purposes. Hence polyimide P347 was used as the material to be characterized as well as

to be doped in lithium salt solution to form an electrolyte.

104
Figure 3.20 A-C. SEM images of neat polyimide fibers P347, R127 & XP02691. D-F.
Images of the resultant fiber membranes collected.

3.3.5.3 Thermogravimetric Analysis (TGA)

Figure 3.21 shows the thermal gravimetric curves for PI P347 at different 15 and 20wt%.

The initial weight lost is likely due to packing of the fibers in the weight pan. The weight losses in

the 100–200 ⁰C range are likely due to solvent effects, as the boiling point of DMF occurs at

153⁰C which happened to Liu et al [44]. The onset temperature of decomposition for the 15wt%

and 20wt% PI fiber ensued at 436⁰C and 475 ⁰C respectively. Additionally, the PI fibers held 40%

and 50% of their weight at the highest temperature run of 800⁰C. This demonstrates the thermal

stability of the polymer at high temperatures under nitrogen atmosphere.

105
474.5⁰C

436.6⁰C

Figure 3.21 TGA Data: Thermogravimetric curves of 15wt% and 20wt% PI fibers
membranes.

3.3.5.4 Fourier-transform Infrared Spectroscopy (FTIR)

The FTIR spectra for the polyimide fiber and fiber membrane electrolytes are shown in

figure 3.22. The spectra show the typical layout for polyimides, the absence of the N-O and –OH

peak [28]. The spectra appear very similar for the three composites but there are peaks found in

the PI/Li composite membranes. They are, 1658 cm-1 which is indicative of lithium interaction

between the C-N bonds of the PI, 1455 cm-1, 1355 cm-1, 1298, 1138 cm-1, 511 cm-1 in the

fingerprint region all may be attributed to lithium salts added that is not in the neat polyimide

fiber.

106
100
1658 cm-1
%Transmitance

90 1455 1
cm-1 2
9
8
c
80 %T-Neat PI m
%T 14wt% LiPF6/PI fiber 1355 -cm-1
1
%T-16wt%LiPF6/PI fiber

70
4000 3600 3200 2800 2400 2000 1600 1200 800 400
Wavenumber cm-1

Figure 3.22 FTIR spectra of neat polyimide (red) and 14%w/w (green), 16% w/w (purple)
PI/LiPF6 fiber membranes.

3.3.5.5 Conductivity of Polyimide fibers

The conductivity, σ, at room temperature of the PI/LiPF6 nanomembranes, over a five-

minute period, is shown in figure 3.23. It took five minutes of testing for the samples to level to a

constant voltage. The conductivity of the polyimide fibers increased tremendously with increasing

lithium hexafluorophosphate. The room temperature conductivity of the PI/LiPF6 fiber


-3
membranes with 14%w/w LiPF6 was 9.3 x 10 S/cm after the five-minute period. Moreover,

when the LiPF6 salt increased to a concentration of 16% w/w the conductivity improved to 1.3 x

10-2 S/cm. This is indicative that the addition of the lithium salts creates a conductive polyimide

fiber composite.

107
1.600E-02
Conductivity of neat PI fiber vs PI/LiPF6 fiber
1.400E-02

1.200E-02
Conductivty (Scm)

1.000E-02

8.000E-03
14wt% LiPF6/PI fiber
6.000E-03
Neat PI

4.000E-03 16wt%LiPF6/PI fiber

2.000E-03

0.000E+00
0 100 200 300 400
time (s)

Figure 3.23 Conductivity vs time graph of neat polyimide (red) and 14%w/w (green),
16% w/w (purple) PI/LiPF6 fiber membranes.

3.4 Conclusion

Lithium hexafluorophosphate salts were successfully incorporated in to an ultrasoft PCPU

as well as a polyimide nanofiber membrane. Thermal and rheological analysis of the

PCPU/lithium composite electrolyte showed Arrhenius behavior in the beta transition and WLF

behavior around the Tg. The Activation energies in the beta transition from PCPU/LiPF6

composites dramatically increased compared to the neat PCPU. The possibility exists that the

structure and arrangement of the lithium salt may have attributed to the high activation energies.

The polyimide nanofiber was successfully transformed from an unusable polyimide

powder to a nanomembrane for use a lithium battery separator. The nanofiber did not possess the

ability to be thermally characterized via DSC as there were no Tg present. However, TGA shows

108
that the polyimide membrane is thermally stable in fiber form and conductivity measurements

show that the addition of the lithium salts created a highly conductive membrane.

3.5 References

1. C. Y. Li, Crystal Engineering for High Performance Solid Polymer Electrolytes , 61st

Annual Report on Research 2016.

2. J. Kim, Y. Lee, I. Lee,T. Kim, M. Ryou, J. Wook Choi (2014), Large area multi-stacked

lithium-ion batteries for flexible and rollable applications, J. Mater. Chem. A, 2, 10862

3. J. Shahitha Parveen, S.S.M. Abdul Majeed (2013), Poly (ethylene oxide)/Polyurethane

based gel polymer electrolytes for lithium batteries, International Journal of Scientific &

Engineering Research, Volume 4, Issue 12.

4. A. I. Gopalan, P. Santhosh, K. M. Manesh, J. Hee Nho, S. Ho Kim, C. Hwang, K. Lee

(2008), Development of Electrospun PVdF – PAN membrane – based Polymer

electrolytes for Lithium batteries, Journal of Membrane Science, 325, pp. 683-690.

5. J. Choi, Y. Kang, H. Shim, D. W. Kim, E. Cha, D. Kim (2010), Cycling performance of a

Lithium – ion polymer cell assembled by in- situ chemical cross – linking with

Fluorinated phosphorous-based cross – linking agent, Journal of Power Sources, 195, pp.

6177-6181.

6. S. Wang, K. Min (2010), Solid polymer electrolytes of blends of polyurethane and

polyether modified polysiloxane and their ionic conductivity, Polymer 51 2621-2628.

7. F.M. Gray (1997) ,Polymer electrolytes. UK: The Royal Society of Chemistry.

8. M. Armand (1986), Polymer Electrolytes. Annu. Rev. Mater. Res. 16, 245–261.

9. D.E. Fenton, J.M. Parker, P.V. Wright (1973), Complexes of alkali metal ions with

poly(ethylene oxide). Polymer 14, 589.

109
10. H. Cheradame, J. F. LeNest (1980), Experimental Demonstration of the importance of the

glass transition temperature of the ionic conductivity of macromolecular networks,

Makromol. Chem. Rapid Comm., 1, p595-598

11. J. R. Mac Callum, C. A. Vincent (1987), Polymer Electrolyte Reviews 1; Elsevier

London, p 103.

12. P. P. Prosini, S. Prosini (2002), A lithium battery electrolyte based on gelled polyethylene,

Solid state ionics, 146, pp 65 – 72.

13. T. Wen, M. Wu , C. Yang (1999), Spectroscopic Investigations of

Poly(oxypropylene)glycol-Based Waterborne Polyurethane Doped with Lithium

Perchlorate, Macromolecules, 32, 2712-2720.

14. J. D. van Heumen and J. R. Stevens (1995), The Role of Lithium Salts in the Conductivity

and Phase Morphology of a Thermoplastic Polyurethane Macromolecules,28, 4268-4277.

15. F. M. Gray (1991), Solid Polymer Electrolytes-Fundamentals and Technological

Applications; VCH: Weinheim, Germany, Chapter 6.

16. Z. Zhong, Q. Cao, B. Jing, X. Wang, X. Li, H. Deng (2012), Electrospun PVdF- PVC

nanofibrous polymer electrolytes for polymer lithium – ion batteries, Materials Science

and Engineering B 177, pp 86-91.

17. M. Watanabe, M. Rikukawa, K. Sanui, K. Ogata, H. Kato, T. Kobayashi, Z. Ohtaki

(1984), Ionic-Conductivity of Polymer Complexes Formed by Poly(Ethylene Succinate)

and Lithium Perchlorate,Macromolecules,17:2902-2908.

18. C. S. Harris, D. F. Shriver, M. A. Ratner, (1986), Complex formation of polyethylenimine

with sodium triflate and conductivity behavior of the complexes Macromolecules, 19,

987– 989 DOI: 10.1021/ma00158a009.

110
19. S. Clancy, D. F. Shriver, and L. A. Ochrymowycz (1986), Preparation and

characterization of polymeric solid electrolytes from poly(alkylene sulfides) and silver

salts, Macromolecules 19:606-611.

20. D. J. Bannister, G. R. Davies, I. M. Ward, and J. F. McIntyre (1984), Ionic conductivities

of poly(methoxypolyethyleneglycol monomethacrylate), Polymer 25:1600-1602.

21. J. M. G. Cowie and R. Ferguson (1985), Glass and Subglass Transitions in a Series of

Poly(itaconate ester)s with Methyl-Terminated Poly(ethylene oxide) Side Chains, J.

Polym. Sci .Polym. Phys. Ed.;23: 2181-2191.

22. F. Alloin, J.-Y. Sanchez, M. Armand (1993), Triblock copolymers and networks

incorporating oligo (oxyethylene) chains, Solid State Ionics, 60:3-9.

23. Alloin F, Sanchez JY, Armand M. Electrochem Acta 1992;37:1729.

24. H. W. Meyer (1998), Polymer Electrolytes for Lithium-Ion Batteries, Adv Mater, 10:439.

25. B. K. Choi, Y. W. Kim and H. K. Shin (2000), Ionic conduction in PEO–PAN blend

polymer electrolytes, Electrochim Acta, 45:1371.

26. T. Wen, and W. Chen (2000) ,Blending thermoplastic polyurethanes and poly(ethylene

oxide) for composite electrolytes via a mixture design approach, J Appl Polym Sci,

77:680.

27. L. Porcarelli, C. Gerbaldi, F. Bella and J. R. Nair (2016) Super Soft All-Ethylene Oxide

Polymer Electrolyte for Safe All-Solid Lithium Batteries, Scientific Reports, 6:19892.

28. Y. E. Miao, G. N. Zhu, H. Hou, Y. Y. Xia and T. Liu (2013) Electrospun polyimide

nanofiber-based nonwoven separators for lithium-ion batteries, Journal of Power Sources,

226, 82-86.

111
29. T. H. Cho, M. Tanaka, H. Onishi, Y. Kondo, T. Nakamura, H. Yamazaki, S. Tanase and

T. Sakai (2008), Battery performances and thermal stability of polyacrylonitrile nano-

fiber-based nonwoven separators for Li-ion battery, J. Power Sources, 181, 155-160.

30. C. Yang, Z. Jia, Z. Guan and L. Wang (2009) Polyvinylidene fluoride membrane by novel

electrospinning system for separator of Li-ion batteries, J. Power Sources, 189, 716-720.

31. N. Wu, Q. Cao, X. Wang, S. Li, X. Li and H. Deng (2011), In situ ceramic fillers of

electrospun thermoplastic polyurethane/poly(vinylidene fluoride) based gel polymer

electrolytes for Li-ion batteries, J.Power Sources, 196, 9751-9756.

32. L. S. Carnell, E. J. Siochi, N. M. Holloway, R. M. Stephens, C. Rhim, L. E. Niklason and

R. L. Clark (2008), Aligned Mats from Electrospun Single Fibers, Macromolecules, 41,

5345-5349.

33. D. Chen, T. X. Liu, X. P. Zhou, W. W. Tjiu and H. Q. Hou (2009), Electrospinning

fabrication of high strength and toughness polyimide nanofiber membranes containing

multiwalled carbon nanotubes, J. Phys. Chem. B 113, 9741-9748.

34. D. Chen, R. Y. Wang,W. W.Tjiu and T. X. Liu (2011), High performance polyimide

composite films prepared by homogeneity reinforcement of electrospun nanofibers,

Compos. Sci. Technol.71, 1556-1562.

35. Keithley Instruments, Inc. (2004), Low Level Measurements Handbook, 6th ed.,

Cleveland, Ohio.

36. K. Mohomed (2006), Ph.D. Dissertation, Thermal analyses of hydrophilic polymers used

in nanocomposites and biocompatible coatings, University of South Florida.

37. G. J. Pratt and M. J. A. Smith (1986), Dielectric response of commercial polycarbonate,

Br. Polym. J., 18:2, 105-11.

112
38. H. W. Spiess (1983), Molecular dynamics of solid polymers as revealed by deuteron

NMR. Colloid Polym. Sci., 261:3, 193-209.

39. M. L. Williams, R. F. Landel, and J. D. Ferry (1955) The temperature dependence of

relaxation mechanisms in amorphous polymers and other glass-forming liquids, J. Am.

Chem. Soc., 77, 3701-3707.

40. P. Jain (2013), Ph.D. Dissertation, Characterizing Interactions between Chromophores in

Synthetic and Natural Macromolecular Films via MALDI-TOF, IBF and Dielectric

Analyzer. University of South Florida.

41. U. Gedde (1999) Polymer Physics. London: Chapman and Hall.

42. M. Shahrousvand, M. S. Hoseinian, M. Ghollasi, A. Karbalaeimahdi, A. Salimi, F. A.

Tabar (2017), Flexible magnetic polyurethane/Fe2O3 nanoparticles as organic-inorganic

nanocomposites for biomedical applications: Properties and cell behavior, Materials

Science and Engineering C, 74, 556–567.

43. CAPRES Vancouver (2002) Micro Four-Point Probe Application note, CAPRES A/S.

44. J. Liu, J. Huang, E. K. Wujcik, B. Qiu, D. Rutman, X. Zhang, E. Salazard, S. Wei, Z. Guo

(2015), Hydrophobic Electrospun Polyimide Nanofibers for Self-cleaning Materials,

Macromol. Mater. Eng. 300, 358–368.

113
APPENDIX A: CHAPTER 1
Complex Viscosity of PCPU/ Nano Ag at Different Temperatures

ŋ Pa.s
ŋ Pa.s

Ẏ (1/s)
Ẏ (1/s)
ŋ Pa.s
ŋ Pa.s

Ẏ (1/s)
Ẏ (1/s)
ŋ Pa.s
ŋ Pa.s

Ẏ (1/s) 114 Ẏ (1/s)


ŋ Pa.s

ŋ Pa.s
Ẏ (1/s) Ẏ (1/s)

ŋ Pa.s
ŋ Pa.s

Ẏ (1/s)
Ẏ (1/s)
ŋ Pa.s
ŋ Pa.s

Ẏ (1/s) Ẏ (1/s)

115
ŋ Pa.s
ŋ Pa.s

Ẏ (1/s) Ẏ (1/s)
ŋ Pa.s

ŋ Pa.s

Ẏ (1/s)
Ẏ (1/s)
ŋ Pa.s
ŋ Pa.s

Ẏ (1/s) Ẏ (1/s)

116
12
y = 7527.6x - 8.045
11 R² = 0.9996
Ea=14.9 kCal/mol
ln n (Pas)

10
9
8
7
0.0021 0.0023 0.0025 0.0027
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.1wt%Ag.

13
y = 7518.3x - 8.1573
12 R² = 0.9981
11 Ea =14.9 kCal/mol
ln ŋ (Pa-s)

10
9
8
7
6
0.0021 0.0023 0.0025 0.0027
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.25wt%Ag.

14
y = 7665.6x - 8.1405
13 R² = 0.9946
12 Ea= 15.2 kCal/mol
ln (ŋ)

11
10
9
8
0.0023 0.0025 0.0027
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.375wt%Ag.

117
13
y = 7746.5x - 8.4454
12 R² = 0.9974

11 Ea= 15.4kCal/mol
ln (ŋ)

10

8
0.0023 0.0025 0.0027
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.5wt%Ag.


13 y = 7486x - 8.2354
R² = 1
12
Ea=14.9 kCal/mol
11
ln (ŋ)

10

8
0.0023 0.0025 0.0027
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.75wt%Ag.


12
y = 7489.7x - 8.409
11.5
R² = 0.9998
11
10.5 Ea 14.9 kCal/mol
ln (ŋ)

10
9.5
9
8.5
8
0.0023 0.0024 0.0025 0.0026 0.0027
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/1.0wt%Ag.

118
13
12 y = 8398.3x - 9.8813
R² = 0.9993
11
Ea= 16.7
10
ln ŋ

9
8
7
6
0.0022 0.0023 0.0024 0.0025 0.0026 0.0027 0.0028
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.1wt%SiO2.

13
y = 8735.4x - 10.789
12 R² = 0.9964
Ea= 17.3 kCal/mol
11
ln (ŋ)

10
9
8
0.0022 0.0024 0.0026 0.0028
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.25wt%SiO2.


12
y = 8442.9x - 10.468
R² = 0.9995
11
Ea = 16.8
ln (ŋ)

10

8
0.0022 0.0024 0.0026
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.375wt%SiO2.

119
12
y = 8351.7x - 10.303
R² = 1
11
Ea= 16.6
10
ln (ŋ)

8
0.0022 0.0024 0.0026
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.5wt%SiO2.


12
y = 6716.5x - 6.4602
R² = 0.9899
11
Ea= 13.3
ln (ŋ)

10

8
0.0022 0.0024 0.0026 0.0028
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.75wt%SiO2.


13
y = 7327.7x - 7.376
12 R² = 0.9987

11 Ea=14.6 kCal/mol
ln (ŋ)

10

8
0.0022 0.0023 0.0024 0.0025 0.0026 0.0027 0.0028
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/1.0wt%SiO2.

120
14
y = 8390.8x - 10.308
12 R² = 0.999

10 Ea=16.7 kCal/mol
8
ln (η)

0
0.002 0.0022 0.0024 0.0026 0.0028
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.1wt%cb.

13
y = 8056.2x - 9.0602
12 R² = 0.9976

Ea= 16.0 kCal /mol


11
ln (η)

10

8
0.002 0.0022 0.0024 0.0026 0.0028
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.25wt%cb.

121
11.5
y = 7677.3x - 8.253
11 R² = 0.996

10.5 Ea= 16.4 kCal/mol


10
ln (ŋ)

9.5
9
8.5
8
0.00220.002250.00230.002350.00240.002450.00250.002550.0026
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.375wt%cb.

13
12.5 y = 14977x - 25.725
12 R² = 0.9949

11.5 Ea= 29.7


11 kCal/mol
ln (ŋ)

10.5
10
9.5
9
8.5
8
0.00235 0.0024 0.00245 0.0025 0.00255 0.0026
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.5wt%cb.

122
12.5 y = 13325x - 21.832
12 R² = 0.9922
11.5 Ea= 26.4 kCal/mol
11
10.5
ln (ŋ)

10
9.5
9
8.5
8
0.00235 0.0024 0.00245 0.0025 0.00255 0.0026
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/0.75wt%cb.

12.5
12
y = 14060x - 23.578
11.5 R² = 1
11
27.9 kCal/mol
10.5
ln (ŋ)

10
9.5
9
8.5
8
0.00235 0.0024 0.00245 0.0025 0.00255 0.0026
1/T (K-1)

Melt Rheology Data: Arrhenius plot of melt viscosity for PCPU/1.0wt%cb.

123
APPENDIX B: Chapter 2

5
38.4
4

3
Intensity

44.6
2 64.6 77.5

0
20 30 40
2θ (⁰)50 60 70 80 90

Figure B1. WAXS on nanosilver powder.

1.5 21.9
Intensity

0.5

0
10 15 20 2θ (⁰) 25 30 35 40

Figure B2. WAXS on nanosilica powder.

124
7

6 24.5

5
Intensity

3
43.5
2

0
20 30 40 50 2θ (⁰) 60 70 80 90

Figure B3. WAXS on carbon black nanoparticle powder.

125
ABOUT THE AUTHOR

Tamalia Julien received a B.S. Degree in Chemistry from Florida A&M University in

2011. She entered the Ph.D. Chemistry program at the University of South Florida in Fall 2011

and officially joined Dr. Julie Harmon’s polymer materials research lab in Spring 2012.

While in the Ph.D. program at the University of South Florida, Tamalia was awarded the

Fred L & Helen M Tharp Endowed Scholarship Award (2015, 2017), the NOBCChe Advancing

Science Travel Grant (2015, 2017) and placed first at the 15th Annual Raymond Castle

Conference Poster Presentation.

Tamalia has co-authored three publications in various peer-reviewed scientific journals

and has participated in nine conference presentations at several regional chemical meetings.

You might also like