You are on page 1of 215

STRUCTURE-PROPERTY RELATIONSHIPS IN GRAPHENE/POLYMER

NANOCOMPOSITES

by

Muhammad Z. Iqbal
© Copyright by Muhammad Z. Iqbal, 2016

All rights reserved


A thesis submitted to the Faculty and Board of Trustees of the Colorado School of Mines
in partial fulfillment of the requirements for the degree of Doctor of Philosophy (Chemical
Engineering).

Golden, Colorado
Date: ________________

Signed: ________________________________________
Muhammad Z. Iqbal

Signed: ________________________________________
Dr. Andrew Herring
Thesis Advisor

Signed: ________________________________________
Dr. Matthew W. Liberatore
Thesis Co-advisor

Golden, Colorado
Date: _________________

Signed: ________________________________________
Dr. David Marr
Professor and Head
Department of Chemical and Biological Engineering

ii
ABSTRACT

Graphene’s unique combination of excellent electrical, thermal, and mechanical

properties can provide multi-functional reinforcement for polymer nanocomposites. However,

poor dispersion of graphene in non-polar polyolefins limits its applications as a universal filler.

Thus, the overall goal of this thesis was to improve graphene's dispersion in graphene/polyolefin

nanocomposites and develop processing-structure-property relationships.

A new polymer matrix was synthesized by blending polyethylene (PE) with oxidized

polyethylene (OPE). Inclusion of OPE in PE produced miscible blends, but the miscibility

decreased with increasing OPE loading. Meanwhile, the Young's modulus of blends increased

with increasing OPE concentration, attributed to decreased long period order in PE and increased

crystallinity. In addition, the miscibility of OPE in PE substantially reduced the viscosity of

blends.

Using thermally reduced graphene (TRG) produced by simultaneous thermal exfoliation

and reduction of graphite oxide, electrically conductive nanocomposites were manufactured by

incorporating TRG in PE/OPE blends via solution blending. The rheological and electrical

percolations decreased substantially to 0.3 and 0.13 vol% of TRG in PE/OPE/TRG

nanocomposites compared to 1.0 and 0.3 vol% in PE/TRG nanocomposites. Improved dispersion

of TRG in blends was attributed to increased TRG/polymer interactions, leading to high aspect

ratio of the dispersed TRG. A universal Brownian dispersion mechanism for graphene was

concluded similar to that of carbon nanotubes, following the Doi-Edwards theory. Furthermore,

the improved dispersion of TRG correlated with the formation of surface fractals in

iii
PE/OPE/TRG nanocomposites, whereas the poor dispersion of TRG in PE led to the formation

of only mass fractals.

Moreover, graphene and carbon black (CB) were combined as a synergic filler for

manufacturing electrically conductive PE nanocomposites. Smaller fractals were observed at

lower CB concentration whereas increasing CB concentration produced large CB agglomerates

on the graphene surface. A remarkably high conductivity of ~1 S/cm was achieved at 15 wt%

using graphene/CB ratio of 1:1, which decreased with increasing CB relative to graphene.

Therefore, using graphene/CB 1:1 hybrid filler could reduce the cost of conductive

nanocomposites by 50%.

iv
TABLE OF CONTENTS

ABSTRACT ................................................................................................................................ iii

LIST OF FIGURES ....................................................................................................................... ix

LIST OF TABLES .........................................................................................................................xv

ACKNOWLEDGEMENT ........................................................................................................... xvi

CHAPTER-1 INTRODUCTION TO POLYMER NANOCOMPOSITES ...................................1

1.1. Polymer Nanocomposites ....................................................................................... 1

1.2. Conductive Nanocomposites .................................................................................. 3

1.3. Graphene ................................................................................................................. 6

1.3.1. Graphene/Polymer Nanocomposites ......................................................... 10

1.3.2. Dispersion of Graphene in Polymers: Issues and Answers ...................... 12

1.4. Nanocomposites Characterization ........................................................................ 17

1.4.1. Melt Rheology .......................................................................................... 17

1.4.2. Small Angle X-Ray Scattering (SAXS) .................................................... 20

1.5. Thesis Overview ................................................................................................... 22

CHAPTER-2 GRAPHENE: SYNTHESIS AND CHARACTERIZATION ...............................25

2.1. Thermally Reduced Graphene (TRG) ................................................................... 25

2.1.1. Introduction ............................................................................................... 25

2.1.2. Experimental Details ................................................................................. 27

2.1.3. Results and Discussion ............................................................................. 29

2.1.4. Summary ................................................................................................... 33

2.2. Graphene Nanoplatelets (GnP) ............................................................................. 34

Acknowledgements ........................................................................................................... 35

v
CHAPTER-3 EFFECT OF SOLVENT ON THE UNCATALYZED SYNTHESIS OF
AMINOSILANE-FUNCTIONALIZED GRAPHENE .........................................36

3.1 Introduction ........................................................................................................... 36

3.2 Experimental Details ............................................................................................. 38

3.2.1 Materials ................................................................................................... 38

3.2.2 Preparation of TRG ................................................................................... 39

3.2.3 Silylation Reaction .................................................................................... 39

3.2.4 Characterization ........................................................................................ 40

3.3 Results and Discussion ......................................................................................... 41

3.3.1 Chemical Analysis .................................................................................... 41

3.3.2 Morphological Analysis ............................................................................ 50

3.3.3 Thermal Stability and Conductivity .......................................................... 54

Conclusion ........................................................................................................................ 56

Acknowledgements ........................................................................................................... 57

CHAPTER-4 SYNTHESIS AND CHARACTERIZATION OF


POLYETHYLENE/OXIDIZED POLYETHYLENE (PE/OPE) BLENDS
AND ROLE OF OPE AS A VISCOSITY CONTROL .........................................58

4.1 Introduction ........................................................................................................... 58

4.2 Experimental ......................................................................................................... 60

4.2.1 Materials ................................................................................................... 60

4.2.2 Preparation of Blends ................................................................................ 61

4.2.3 Characterization ........................................................................................ 61

4.3 Results and Discussion ......................................................................................... 63

4.3.1 Miscibility, Phase Behavior and Thermal properties ................................ 63

4.3.2 Melt Rheological Analysis ........................................................................ 74

4.3.3 Mechanical Properties ............................................................................... 79

vi
4.3.4 SEM Morphology ..................................................................................... 80

Conclusion ........................................................................................................................ 82

Acknowledgements ........................................................................................................... 83

CHAPTER-5 PROCESSABLE CONDUCTIVE GRAPHENE/POLYETHYLENE


NANOCOMPOSITES: EFFECTS OF GRAPHENE DISPERSION AND
POLYETHYLENE BLENDING WITH OXIDIZED POLYETHYLENE ON
RHEOLOGY AND MICROSTRUCTURE ..........................................................85

5.1 Introduction ........................................................................................................... 85

5.2 Experimental Details ............................................................................................. 88

5.2.1 Materials ................................................................................................... 88

5.2.2 Synthesis of TRG ...................................................................................... 89

5.2.3 Preparation of Composites ........................................................................ 89

5.2.4 Characterization ........................................................................................ 90

5.3 Result and Discussion ........................................................................................... 92

5.3.1 Characterization of TRG and CB .............................................................. 92

5.3.2 Melt Rheological Analysis ........................................................................ 93

5.3.3 SAXS-Based Microstructure .................................................................. 104

5.3.4 Thermal Properties .................................................................................. 109

5.3.5 Mechanical Properties ............................................................................. 110

5.3.6 Electrical Properties ................................................................................ 113

5.3.7 Morphology............................................................................................. 116

Conclusions ..................................................................................................................... 119

Acknowledgements ......................................................................................................... 119

CHAPTER-6 CONDUCTIVE POLYETHYLENE/GRAPHENE/CARBON BLACK


HYBRID NANOCOMPOSITES .........................................................................121

6.1 Introduction ......................................................................................................... 121

6.2 Experimental Details ........................................................................................... 124

vii
6.2.1 Materials ................................................................................................. 124

6.2.2 Nanocomposites Synthesis...................................................................... 125

6.2.3 Characterization ...................................................................................... 125

6.3 Results and Discussion ....................................................................................... 127

6.3.1 Characterization of GnP and CB............................................................. 127

6.3.2 Microstructure and Melt Rheology ......................................................... 129

6.3.3 Electrical Conductivity ........................................................................... 136

Conclusions ..................................................................................................................... 138

Acknowledgement .......................................................................................................... 139

CHAPTER-7 CONCLUSIONS AND FUTURE RECOMMENDATIONS ..............................140

7.1 Summary and Conclusions ................................................................................. 140

7.2 Future Research Directions ................................................................................. 143

REFERENCE ..............................................................................................................................149

APPENDIX-A.......................................................................... SILANE MODIFIED GRAPHENE 172

APPENDIX-B PE/OPE MISCIBLE BLENDS ............................................................................178

APPENDIX-C PE/OPE/TRG AND PE/OPE/CB COMPOSITES ...............................................186

viii
LIST OF FIGURES

Figure 1.1 Intercalated, flocculated and exfoliated polymer/layered silicates (clay)


nanocomposites ........................................................................................................4

Figure 1.2 Representation of graphene as the building block for graphite, CNTs and
fullerenes (adapted with permission from ref.28. Copyright (2010) American
Chemical Society) ....................................................................................................7

Figure 1.3 a) 0.5 g (~0.5 mL) of GO expands to 75 mL of thermally reduced graphene


(TRG) upon rapid heating at 1050°C, b) XRD patterns of pure graphite, GO
and TRG (Reprinted with permission from ref.31. Copyright (2012)
Springer) ..................................................................................................................9

Figure 1.4 Left) TEM of TRG sheets produced from very dilute solution of TRG in
DMF and drying on amorphous lacy carbon coated Cu-grid (Reproduced
from ref.36 with permission from the Centre National de la Recherche
Scientifique (CNRS) and The Royal Society of Chemistry), right) TEM of
TRG in PMMA fractured surface revealing subsurface morphology of TRG
(Reproduced with permission from ref. 38. Copy right (2008) Nature
Publishing Group). Insert in left side shows SAED of TRG showing nicely
separated and exfoliated graphene sheets on TEM support36. ...............................10

Figure 1.5 a) Proposed reactions during isocyanate treatment of graphene oxide where
organic isocyanates reacts with the hydroxyl (left oval) and carboxylic (right
oval) groups of graphene oxide sheets to form carbamate and amide
functionalities, respectively (reproduced from ref.55, copyright (2006) with
permission from Elsevier). .....................................................................................16

Figure 1.6 Rheological parameters acting as a link between molecular structure and
final properties of a polymer (reproduced from ref.61, copyright (2001) with
permission from Elsevier). .....................................................................................18

Figure 1.7 Typical viscosity curve for polyolefins defining the range of viscosities used
in various polymer processing machines (reproduced from ref.61, copyright
(2001) with permission from Elsevier). .................................................................19

Figure 1.8 Simplified working diagram of scattering instrument ...........................................21

Figure 2.1 0.5 g of GO expands to 75 ml of TRG upon rapid heating at 1050 °C (a),
XRD patterns of pure graphite, GO and TRG (b). (reproduced from ref.88,
copyright (2012) Springer) ....................................................................................30

Figure 2.2 Raman spectra of graphite and TRG. (reproduced from ref.88, copyright
(2012) Springer) .....................................................................................................31

ix
Figure 2.3 TRG with bulk density of 17 mg/cm3 and C/O = 13: SEM (a), and HRTEM
(b). (Image reproduced from ref.88) .......................................................................32

Figure 2.4 Pore size distribution of TRG sample with bulk density of 3 g/L. Inset
shows the BET adsorption-desorption isotherm curve (reproduced from
ref.88, copyright (2012) Springer). .........................................................................33

Figure 2.5 FTIR (a) and XPS high resolution C1s spectrum (b) of TRG sample. the inset
in a) provides a structural model for TRG (reproduced from ref.37 with
permission from the Centre National de la Recherche Scientifique (CNRS)
and The Royal Society of Chemistry). ...................................................................34

Figure 2.6 TEM image of graphene nanoplatelets (GnP) .......................................................35

Figure 3.1 FTIR spectra of the pure and f-TRGp and f-TRGs ................................................42

Figure 3.2 High resolution XPS spectra of pure and f-TRG samples .....................................45

Figure 3.3 EELS Carbon K-edge a), and Oxygen K-edge b) ..................................................47

Figure 3.4 Nitrogen K-edge a) and Silicon L2-3-edge b) .......................................................48

Figure 3.5 Raman spectra of TRG samples a) D and G band and b) βG (Gˈ) band ................50

Figure 3.6 TEM images of pure TRG. (a) Graphene particle composed of overlapping
graphene sheers (inset: SAED pattern from the center of the graphene
particle), (b) HRTEM image of two overlapping graphene sheets ........................51

Figure 3.7 TEM images of functionalized graphene: (a) f-TRGp, (b) f-TRGs .......................52

Figure 3.8 TEM bright field image (a) and corresponding EFTEM maps of silicon (b)
for f-TRGp; TEM bright field image (c) and corresponding EFTEM maps of
silicon (d) for f-TRGs. ...........................................................................................53

Figure 3.9 Thermograms of TRG, f-TRGp, and f-TRGs. Samples were heated from 50-
500° C at 10° /min under nitrogen .........................................................................56

Figure 4.1 FTIR spectra of OPE (a) and PE/OPE blends at various compositions .................65

Figure 4.2 Melting temperature and melt enthalpy of the blends ...........................................67

Figure 4.3 Hoffman-Weeks plot of PE/OPE blends (a) and Tm° versus OPE
concentration (b). ...................................................................................................67

Figure 4.4 Composition dependence of the Flory-Huggins interaction parameter, χ .............69

Figure 4.5 Thermal stability of PE/OPE blends ......................................................................71

x
Figure 4.6 log-log I vs q (a) and Kratky plots of PE/OPE blends (b) at 60°C and 0%
relative humidity with Y-offset, and long period order for blends (c)...................73

Figure 4.7 WAXS patterns of PE/OPE blends with Y-offset..................................................74

Figure 4.8 Start-up transient rheology of neat PE and PE/τPE blendsμ Evolution of Gˊ
with time (a), and reduced Gˊ versus time. Data reported at T = 1θ0°C, = 1
%, ω = 1rad/s..........................................................................................................76

Figure 4.9 Complex viscosity profiles of PE/OPE blends at 160°C. The black solid
lines are the Cross model fit (left panel) and zero shear viscosity as a
function of OPE wt.% at two different temperatures (right panel) ........................78

Figure 4.10 εaster curves for Gˊ (a), G˝ (b) over the reduced frequency. Insets in (a)
and (b) show slopes in the terminal regions (at low frequencies) averaged
over all the temperatures (circles 180°C, upper triangles 160°C, lower
triangles 140°C, and squares 130°C). ....................................................................79

Figure 4.11 Mechanical properties of PE/OPE blends: (a) Stress-Strain behavior, (b)
Young’s modulus, (c) yield and break stress, and (d) yield and break strains.
Arrows in stress-strain data indicate the break points. The inset in (a) shows
stress-strain data on a linear scale. .........................................................................81

Figure 4.12 SEM micro-images of PE/OPE Blends..................................................................82

Figure 5.1 TEM image of TRG (a), and CB (b). .....................................................................93

Figure 5.2 Processing control by PE/OPE blends ...................................................................94

Figure 5.3 Representative shear modulus and complex viscosity vs. frequency curves
for PE/TRG (a, d), PE/OPE 80/20/TRG (b, e) and 60/40/TRG (c, f),
respectively. Solid lines represent the polymer matrix in each box and
numbers next to each curve correspond to weight % of TRG ...............................98

Figure 5.4 Percolation analysis of TRG filled PE, PE/OPE 80/20 and 60/40 blends.
Solid lines are linear fit to experimental data. The point of intersection of
two lines indicates the percolation threshold, Φp. Inset shows power law fit
to percolation law for PE, 80/20 and 60/40 blends-filled with TRG, and the
numbers next to each line are the power law exponents. Color schemes are
same in both panels. ...............................................................................................99

Figure 5.5 Scaling behavior of critical strain (a), and G′ (b) above percolation threshold ...102

Figure 5.6 Universal behavior for TRG-filled matrices ........................................................105

Figure 5.7 I(q)~q log-log plots of PE/TRG (a), and 60/40/TRG nanocomposites (b).
Note: The lines in all panels are stacked in the same order with Y-offset for
clarity. ..................................................................................................................107

xi
Figure 5.8 I(q) ~ q log-log SAXS data for PE/OPE 60/40 Blends-CB composites. Note:
The lines in all panels are stacked in the same order with Y-offset for clarity. ...108

Figure 5.9 Kratky analysis of TRG-filled nanocomposites (a), and CB-filled


composites (b). Color schemes in (a) show percolation of each matrix. .............110

Figure 5.10 Thermograms for PE, PE+TRG 5 wt%, PE/OPE 60/40, and 60/40+TRG 5
wt%. Inset shows derivative of TGA versus the temperature ..............................111

Figure 5.11 Comparison of the Young's moduli of TRG and CB filled composites ..............112

Figure 5.12 Mechanical properties of TRG-filled nanocomposites. Solid lines are fit to
Halpin-Tsai model, and dashed lines represent Mori-Tanaka
micromechanical model. Inset shows estimated aspect ratio (Af) of TRG for
nanocomposites from the best fit. ........................................................................114

Figure 5.13 Comparison of electrical conductivities of TRG-filled and CB-filled


composites............................................................................................................115

Figure 5.14 Electrical conductivity of TRG-filled nanocomposites. Solid lines are to


guide the eye only. Inset shows the power law fit, the numbers indicate the
power law exponents............................................................................................117

Figure 5.15 SEM images of PE/TRG (a, b), 80/20/TRG (c, d), 60/40/TRG (e, f) at 5 wt%
loading. The arrows indicate TRG sheets located in these composites ...............118

Figure 6.1 TEM morphology of GnP (a) and CB (b). Inset in (b) shows high resolution
image of CB particles...........................................................................................128

Figure 6.2 XRD patterns of graphite, GnP, and CB ..............................................................129

Figure 6.3 Rheological properties of PE/GnP nanocomposites (a), and PE/CB


composites (b). .....................................................................................................130

Figure 6.4 Effects of total filler concentration, and GnP/CB ratio on rheological
properties of PE/GnP/CB nanocomposites. .........................................................132

Figure 6.5 Comparison of viscosity of composites ...............................................................133

Figure 6.6 TEM micro-images of GnP/CB suspensions with different GnP/CB ratios: a)
1:1, b) 1:5, and c) 1:10. ........................................................................................134

Figure 6.7 SAXS I~q log-log plot for PE/hybrid filler nanocomposites at a filler
loading of a) 5 wt%, and b) 15 wt% ....................................................................135

Figure 6.8 Comparison of electrical conductivities of GnP and GnP/CB hybrid


nanocomposites ....................................................................................................137

Figure 6.9 Electrical conductivity of nanocomposites. .........................................................138


xii
Figure 7.1 Comparison of electrical conductivity of nanocomposite ...................................144

Figure 7.2 Schematics for future research plan in nanocomposites research ........................146

Figure A.1 XRD spectra of graphite, graphite oxide (GO) and TRG (a), and TEM of
pure TRG (b) ........................................................................................................174

Figure A.2 Overall XPS survey for the pure and f-TRG .......................................................175

Figure A.3 N1s and S2p high resolution XPS scans for TRGp and TRGs ...............................176

Figure B.1 Arrhenius Plot of the viscosity of OPE (at shear rate = 10 1/s). Viscosity
data was obtained from performing steady state flow test on OPE after
melting on the Peltier plate ..................................................................................178

Figure B.2 Crystallization temperature and heat of crystallization of PE/OPE blends


with different OPE wt%. ......................................................................................178

Figure B.3 Thermograms of PE/OPE blends show the melting profiles of the blends.
The inset shows the melting profile of neat OPE ................................................179

Figure B.4 Melting profile of pure PE. This area has been used for calculation of melt
enthalpy and % crystallinity.................................................................................179

Figure B.5 Complete FTIR spectrum of OPE ........................................................................180

Figure B.6 Derivative of weight loss of PE/OPE Blends ......................................................180

Figure B.7 WAXD patterns of neat PE, OPE and PE/OPE blends at 25°C with Y-offset ....181

Figure B.8 Cross model fitting to different blend compositions at various temperatures .....182

Figure B.9 Han plots for the blends .......................................................................................183

Figure B.10 Shift factor for time-temperature master curves for PE/OPE blends ...................183

Figure B.11 Time-temperature master curves for all blend compositions against the
tested temperatures...............................................................................................184

Figure B.12 Time-temperature master curves for all blend compositions against the
tested temperatures...............................................................................................185

Figure C.1 XRD of graphite, graphite oxide and TRG ..........................................................186

Figure C.2 FTIR of TRG (a, b), and OPE+TRG (c). .............................................................187

Figure C.3 DLS of TRG and CB............................................................................................188

Figure C.4 TEM of Carbon Black (CB) .................................................................................188

xiii
Figure C.5 Rheological properties of PE/OPE Blends: (a) shear and loss moduli vs
frequency, and (b) complex viscosity profiles as a function of blend
composition at 160°C. (100/0 circles; 80/20 upper triangles; 60/40 lower
triangles). Pre-factors in (a) were used to shift G′~ƒ curves vertically for
clarity. ..................................................................................................................190

Figure C.6 Rheological data of CB-filled composites ...........................................................191

Figure C.7 Representative strain sweep on PE/TRG nanocomposites at 160°C. Arrows


point out the limit of linearity used for scaling argument ....................................191

Figure C.8 Power law decaying exponent for TRG-filled nanocomposites, and CB-
filled composite, obtained from SAXS I ~ q log-log plot ....................................192

Figure C.9 SAXS patterns for CB-filled composites .............................................................192

Figure C.10 Representative Kratky plot for SAXS data ..........................................................193

Figure C.11 Thermal stability of TRG-filled, and CB-filled composites ................................193

Figure C.12 Strain at break and breakage stress of PE/TRG (a), 80/20/TRG, and
60/40/TRG (c) nanocomposites ...........................................................................194

Figure C.13 Strain at break and breakage stress of PE/CB (a), 80/20/CB (b), and
60/40/CB (c) composites......................................................................................195

Figure C.14 Young's modulus of CB-filled composites, normalized by matrix moduli.


Straight lines are linear fit to experimental data with intercept 1. .......................196

Figure C.15 Electrical conductivity of CB-filled composites. Inset shows the power law
fit to the experimental data ..................................................................................196

xiv
LIST OF TABLES

Table 1.1 Electrical percolation thresholds for selected graphene/polymer composites


and their electrical conductivities (adapted from ref.39 copyright (2011) with
permission from Elsevier) ......................................................................................14

Table 3.1 Atomic composition of pure and f-TRG as obtained from XPS ............................43

Table 3.2 High resolution fitting of C1s and O1s groups appearing in the XPS spectra of
TRG, f-TRGp, and f-TRGs .....................................................................................46

Table 5.1 Electrical and rheological percolation threshold and power law exponents
for TRG-based nanocomposites ...........................................................................117

Table A.1 High resolution fitting of functional groups in N1s and Si2p XPS scans ..............177

Table A.2 Averaged EDS results ..........................................................................................177

Table B.1 Thermal stability of blends with different compositions .....................................181

Table B.2 Cross model fitting parameters for PE/OPE Blends ............................................182

xv
ACKNOWLEDGEMENT

Though it's my name appearing on the cover of this dissertation, a great many people

have contributed to its production. I owe my gratitude to all those people who have made this

journey possible, and because of whom I will cherish my graduate experience forever.

I thank the Petroleum Institute, Abu Dhabi for funding this whole work through

Cooperative Research Partnership with Colorado School of Mines. I thank U.S. Army Research

Office through DURIP Grant No. W911NF-11-1-0306 for financial support in part of this work.

I thank U.S. DOE under the contract No. DE-AC02-06CH11357 for supporting the use of

Advanced Photon Source (APS) at Argonne National Laboratory. I thank Soenke Seifert at

Argonne National Laboratory for being so generous with his time, knowledge, and good coffee

while performing x-ray scattering experiments at APS. Jim Horan was very kind and helpful in

ensuring perfect working of conductivity instruments at Fuel Cell Center, CSM. I thank John

Dorgan for letting me use his laboratory throughout this thesis work. I could not have completed

this thesis without using his facilities. I thank Mark Grass and Mike Stadick for helping me with

many instruments and facilities. I also thank Deanna Jacobs for being so proactive, and making

my life at CSM easy by always finding a feasible solution to any issue with my documents with a

big smile on her face.

I thank Ala Bazyleva for teaching me how to use a rheometer on my first day in the

laboratory. I thank Tara Pandey and Andrew Motz for helping with x-ray scattering

measurements at APS. I especially thank Benjamin Caire for helping me with mechanical

characterization of polymer thin films. I always found him helping everyone in the laboratory

without any wrinkles on his forehead. I thank Sanket Kelkar, a great friend who made my stay at

xvi
CSM very amusing through this discussions and (very) long road trips. He was a great deal of

inspiration as a very hardworking person who always aimed at learning new things. I cannot

thank Ali Chaudhary in words for his endless discussions and new ideas, which led me to do lots

of work in this thesis. I thank undergraduate student, Travis Arnold who worked with me for

three semesters. I also thank Frederick Prehen Jr. (Fritz), graduate student in chemistry for his

ideas and knowledge of macromolecular chemistry. I always called him "Professor Fritz" for his

understanding and patience in teaching me learn chemistry of macromolecules. I thank Patrizia

Smith for letting me use her work space time to time.

I am indebted to Matthew Liberatore for his excellent guidance, and the personal

understanding he displayed during my stay at CSM. I really admire his patience in guiding me to

furnish my new ideas into purposeful documents. I wish him all the best for his new tenure at

University of Toledo. I am also indebted to Ahmed Abdala for taking me as research assistant at

Petroleum Institute, Abu Dhabi in 2009, introducing me to graphene's applications, and further

leading me to CSM for Ph.D. with Matt. Ahmed has been a great deal of inspiration for his in

depth knowledge and ideas. I could always find him either in person or through Skype for

anything anytime. I also wish him the best for transitioning to Qatar.

I thank Andrew Herring for taking me as his graduate student after Matt's leaving to

Toledo, and letting me use his laboratory facilities. I thank Stephen Boyes for helping me

understanding organic and physical chemistry. During the last six months of this thesis, I have

probably spent more time in his chemistry laboratory than in my department. I also thank Vikas

Mittal at Petroleum Institute for Skype discussions, and assisting with some experiments in his

laboratory.

xvii
I thank my family for finding time to call me every day even with a 12 hours time

difference between USA and Pakistan. Though, my parents did not live long enough to see this

day, I cannot forget their contribution to what I am today and how I reached here.

Last but not least, I am thankful to have earned my Ph.D. from CSM, and I thank all other

graduate students and faculty who helped me throughout my stay. It has not been the easiest four

years away from my family and friends, and I certainly could not have done this without my

CSM family.

xviii
CHAPTER-1

INTRODUCTION TO POLYMER NANOCOMPOSITES

This dissertation continues work on applications of thermally reduced graphene (TRG),

in collaboration with the Petroleum Institute, Abu Dhabi, to understand the interplay of

chemistry, processing, and morphology of TRG on mechanical, electrical, and processing

properties of TRG/polymer nanocomposites. The overall objective of this thesis was to better

understand reasons for poor dispersion of graphene in polyolefins, specifically polyethylene, and

further develop methods to increase graphene's dispersion, leading to improved nanocomposites.

This thesis focused on polyethylene to produce electrically conductive, mechanically robust, and

processable nanocomposites at lower graphene loadings.

1.1. Polymer Nanocomposites

Over the years, overall perspective of materials has been changed by polymers, and

nowadays we deal with various types of polymers in our daily life on regular basis. Polymers

have excellent properties: light weight, toughness, good elongation, easy processing and low

cost1, 2, 3
. However, in comparison with metals and ceramics, polymers show low stiffness,

strength, flammability, and high gas permeability as weaknesses. Despite of their low properties,

high strength and electrical conductivity in polymers are needed in some applications. Mixing

polymers with inorganic fillers imparts strength and electrical properties to the polymers.

Reinforcement with nanometer size fillers can overcome many of the weaknesses in polymers (if

fillers are well dispersed within polymers). Additionally, property enhancements (mechanical

and thermal stability) can be achieved at significantly lower weight fraction of the nanofillers

(filler loadings) than conventional micron sized fillers, resulting in light-weight nanocomposites.

1
Nanocomposites are defined as having at least one dimension of the dispersed particles

in nanometer range. The nanoparticles are classified based on their dimensions. Examples

include: 1) isodimensional nanoparticles have all dimensions on the order of nanometers, e.g.,

fullerenes, 2) elongated nanoparticles such as carbon nanotubes (CNT), carbon nanofibers (CNF)

or nano-whiskers with two dimensions on the nanometer scale and the third is larger, 3) two

dimensional nanoparticles, such as layered silicate crystals or clays and graphene, in the form of

sheets of one to a few nanometers thick and hundreds to thousands of nanometers in extent4.

Polymer/layered silicates (clay) nanocomposites attracted great interest, both in industry

and in academia, because of improved properties (mechanical and thermal stability, and gas

permeability) when compared with virgin polymers or conventional (micro- or macro-)

composites. Three most commonly used clays in manufacturing clay/polymer nanocomposites

include montmorillonite (MMT), hectorite, and saponite. The layer thickness is normally 1 nm,

and the lateral dimensions may vary from 30 nm to several microns, depending on particular type

of layered silicate. Stacking of layers leads to a regular van der Waals gap between the layers

called interlayer spacing or gallery. For manufacturing nanocomposites, layer stacking is broken

down using a solvent or inter-gallery diffusion of polymer chains in a solvent at relatively high

shear rates.

Toyota5 was the first to report the nylon-6/MMT nanocomposites for practical

applications (used in automotive parts) where a very small amount of clay substantially

improved mechanical and thermal properties of nylon-6. In the past three decades, many

researchers studied use of clay in manufacturing mechanically strong, thermally stable and flame

retardant polymer nanocomposites. Polymer/clay nanocomposites are categorized in three broad

classes based on the dispersion of layered silicates in polymers: 1) intercalated, 2) flocculated,

2
and 3) exfoliated nanocomposites (Figure 1.1). Depending on the exfoliation extent of clay and

type of clay/polymer interactions, relative increase in mechanical and thermal properties of the

nanocomposites can be achieved with increased dispersion of clay. Fu and Qutubuddin6 reported

increased mechanical stability and degradation temperature of polystyrene (PS) using ~6 wt%

MMT. Cao et al.7 showed complete exfoliation of clay in polyurethane (PU)/MMT

nanocomposites resulting 6 °C increase in glass transition temperature (Tg), and ~800%

improvement in mechanical properties.

Furthermore, compatibilizers can be used to develop interactions between clay and

polymer chains, leading to enhanced dispersion of clay. Durmus et al.8 compared maleic

anhydride-graft-polyethylene (MA-g-PE) and oxidized PE (OPE) as compatibilizers in

polyethylene (PE)/clay nanocomposites. The rheological percolation threshold for OPE-based

nanocomposites was 0.016 wt% compared with 0.01 wt% for MA-g-PE-based nanocomposites.

In another report, Durmus et al.9 showed a 50% reduction in oxygen permeability of OPE-

compatibilized PE/clay nanocomposites. The use of compatibilizers in polymer nanocomposites

has been optimized to a compatibilizer/clay ratio of 3 by Hotta et al.10 using different ratios of

the alkyl-modified clay and MA-g-PE to create PE/clay interactions whereas similar ratio was

later confirmed by Durmus and co-workers8, 9.

1.2. Conductive Nanocomposites

Conductive polymer composites are another important class of nanocomposites where

conductive fillers such as conductive carbon black (CB), CNF, CNT or graphene are dispersed in

insulating polymers. Conductive nanocomposites find use in many practical applications such as

3
electromagnetic interference shielding (EMI) effectiveness, anti-static packaging materials,

printed electronic circuits, and conductive fibers etc.

Polymer Filler

Flocculated or Intercalated Exfoliated or


micro-composites composites nanocomposites

Figure 1.1 Intercalated, flocculated and exfoliated polymer/layered silicates (clay)


nanocomposites

CNTs have attracted huge amount of interest since 199111 due to the unique combination

of mechanical, electrical and thermal properties12. Since CNT/polymer nanocomposites were

first reported by Ajayan et al.13, a lot of studies evolved on the synthesis, structure, properties,

and application of these conductive materials14. CNTs and CNF are produced by catalytic

chemical vapor deposition of a hydrocarbon (such as natural gas) on a metal surface (iron, nickel

etc.). CNFs composites are used, particularly, for EMI shielding effectiveness15. However,

compared to CNTs, CNFs have received less research attention as nanofillers because CNTs

have better mechanical properties, smaller diameter, and lower density than CNFs15, 16.

4
CNT/polymer nanocomposites are generally prepared by mixing CNTs and polymers in

solutions or dispersing CNTs in polymer melts. If polymers are not either soluble or melt-able,

the related monomers might be firstly incorporated with CNTs, followed by in situ

polymerization to produce the nanocomposites16. CNTs can be single-wall (SWCNT), double-

wall (DWCNT) or multiple-wall (MWCNT) depending on the existence of one or more

concentric cylinders present. Adhikari et al.17 studied the effects of processing conditions on

thermal stability of MWCNT/high density polyethylene (HDPE) nanocomposites. Combination

of solution mixing followed by melt extrusion of the composites improved degradation

temperature of HDPE, which was attributed to the better dispersion of MWCNTs.

Since CNTs usually agglomerate due to van der Waals forces, they are extremely difficult

to disperse and align in a polymer matrix. Thus, a significant challenge in manufacturing high

performance CNT/polymer composites is to introduce individual CNTs in a polymer matrix to

achieve better dispersion, alignment, and strong interfacial interactions, which also improve load

transfer across CNT/polymer interface. The most common method to create better dispersion of

CNTs in a polymer matrix is to use oxidized CNTs (also called purified CNTs). The purified

CNTs were reported to create good interactions in polar polymer matrices such as epoxy18, poly

(methyl methacrylate) (PMMA)19, poly (ethylene terephthalate) (PET)20 and poly (vinyl alcohol)

(PVA)21. However, in nonpolar polymers such as polyolefins, CNTs do not create strong

interface due to the inert polyolefin structure. For example, Valentino et al.22 showed ~10 orders

of magnitude increase in electrical conductivity in MWCNT/low density PE (LDPE)

nanocomposites with 8 wt% MWCNT loading. Xiao et al.23 used 10 wt% MWCNTs to improve

the Young’s modulus of PE by ~λ0%.

5
The surfactants may also help increasing the dispersion of CNTs24 but the lack of strong

CNT/polymer interface leads to high CNT loadings for the load transfer applications. For this

purpose, surface modification of CNTs has been one of the most plausible route for their better

dispersion in polymer matrices. Shanmugharaj et al.25 used amino-f-MWCNTs in manufacturing

CNT/rubber composites to improve the mechanical properties. Yang et al.26 used polypropylene-

grafted-CNT (PP-g-CNT) in PP matrix, and showed ~300% improvement in toughness at 1.5

wt% CNT loading. Thus, the surface modification of CNTs can be useful for creating strong

CNT-polymer interface, and increase dispersion.

1.3. Graphene

Graphene, a two-dimensional sp2-hybridized carbon sheet, is currently a widely studied

material. This single-atom-thick sheet of C-atoms arrayed in a honey comb pattern is the world’s

thinnest, and stiffest material which is also an excellent conductor of heat and electricity27.

Graphene is considered more promising than other nanostructured C-fillers such as 1-

dimensional CNTs, and 0-dimensional fullerenes in various applications (Figure 1.2). Due to its

high electrical conductivity (θ000 S/cm), Young’s modulus (~1TPa), and thermal conductivity

(5000 W/(m.K)) along with very high surface area (theoretically 2600 m2/g)28, graphene has

been vastly used in areas such as conductive polymer nanocomposites, field effect transistors,

transparent conductive films, gas sensors, clean energy devices and diodes, etc.29. Due to high

surface area and low bulk density (~3 mg/cm3), extremely low amounts of graphene are required

to achieve the percolation thresholds for various applications.

Since the first experimental evidence of existence of stable, free-standing graphene sheets

and their electronic applications in 200430, research was concentrated on the development of new

6
synthetic routes enabling an effective production of well-defined, and separated graphene sheets.

Recently, Kim et al28 classified graphene production methods into two major classes: bottom-up

and top-down methods. The starting material for graphene is generally the graphite, which is

composed of 2-dimensional sp2-carbon sheets (graphene sheets) stacked together, and held by

the van der Waals forces of attraction. Splitting graphite into graphene sheets can be performed

by taking graphite particles or oxidizing graphite into graphite oxide (GO) followed by splitting

GO by different methods. Nevertheless, the final physical and chemical properties of graphene

sheets depend on the main source and splitting methods.

Figure 1.2 Representation of graphene as the building block for graphite, CNTs and fullerenes
(adapted with permission from ref.28. Copyright (2010) American Chemical Society)

The size of the graphene sheets and the type of attached functional groups become

important for dispersion of graphene in polymer matrices. Therefore, a brief overview of the

7
experimental methods used for analyzing the quality of the produced graphene is provided

below.

Number of layers and size: The bulk density of graphene is a direct measure of

exfoliation of graphite or GO (Figure 1.3a). Low bulk density represents better exfoliation. The

extent of graphite exfoliation is also studied by X-ray diffraction (XRD). e.g., graphite shows a

sharp reflection at β =βθ.γ° (CuKα radiation, X-ray wavelength=1.54Å) which shifts to 14.1°-

14.9° in GO31. The XRD pattern disappears as GO exfoliates and separates into graphene single

sheets (Figure 1.3b). An indirect method to study exfoliation is surface area measurement where

theoretically, the surface area is inversely proportional to the thickness of disk-like particles. The

exfoliation must lead to a sharp increase in the surface area of graphene sheets compared with

graphite and GO. Two most common methods for surface area measurements are N 2- and

methylene blue-adsorptions. However, as reported by Schniepp et al.32, the N2-adsorption based

surface area may be influenced by the compressibility of graphene sheets. On the other hand,

methylene blue method can be highly influenced by the surface topology and chemistry of the

graphene sheets33.

Atomic force microscopy (AFM) studies the surface topology, defects, and bending

properties of graphene. Step-height scans give a complete picture of number of layers and the

lateral size of graphene sheets33. However, complications such as folded or wrinkled sheets,

adsorbed solvent or moisture, largely influence AFM measurement. Size and quality of graphene

sheets can also be determined via transmission electron microscope (TEM). In addition to the

size determination by TEM imaging, selected area electron diffraction patterns (SAED) can

clearly differentiate between single and bilayer graphene (Figure 1.4).

8
Figure 1.3 a) 0.5 g (~0.5 mL) of GO expands to 75 mL of thermally reduced graphene (TRG)
upon rapid heating at 1050°C, b) XRD patterns of pure graphite, GO and TRG (Reprinted with
permission from ref.31. Copyright (2012) Springer)

Another important tool for determining the size and morphology of platelets (graphene

sheets) is solution rheology. A reasonably precise aspect ratio of graphene can be determined via

dilute solution rheology. Recently, Tesfai et al.34 used dilute solution rheology to determine size

of aqueous graphene oxide dispersions. Moreover, the static light scattering of dilute GO

solution is also reported to give fractal dimensions of GO which, however, depends strongly on

the solvent polarity35.

Chemistry of graphene: Since, methods for mass scale production of graphene start from

GO28, the produced graphene contains small amounts of residual oxygen functionalities.

Standard combustion elemental analysis can describe overall degree of oxidation in GO, and the

level of retained oxy-groups after reduction/exfoliation. X-ray photoelectron spectroscopy (XPS)

is one of the powerful tools to identify and quantify different types of functional groups present

on graphene surface. Chemical shifts in XPS C1s and O1s spectra, e.g., confirm the existence of

C-O, C=O, O-C=O on GO, graphene, and its derivatives36. Raman spectroscopy quantifies the

9
transformation of sp3(GO) to sp2 (graphene) upon reduction of GO, and stacking in graphite

samples. The transformation from sp3 to sp2 restores the electrical conductivity; thus, indicating

that electrical conductivity is also a qualitative measure of conversion of GO to graphene31, 36. In

addition, Fourier transformed infrared spectroscopy (FTIR) can be used to identify, qualitatively,

the types of functional groups on GO and graphene36, 37.

Figure 1.4 Left) TEM of TRG sheets produced from very dilute solution of TRG in DMF and
drying on amorphous lacy carbon coated Cu-grid (Reproduced from ref.36 with permission from
the Centre National de la Recherche Scientifique (CNRS) and The Royal Society of Chemistry),
right) TEM of TRG in PMMA fractured surface revealing subsurface morphology of TRG
(Reproduced with permission from ref. 38. Copy right (2008) Nature Publishing Group). Insert in
left side shows SAED of TRG showing nicely separated and exfoliated graphene sheets on TEM
support36.

1.3.1. Graphene/Polymer Nanocomposites

Graphene/polymer conductive nanocomposites have been reported with a number of

polymers28, 39. Much of the graphene/polymer nanocomposites research parallels CNT/polymer

systems. However, compared with CNTs, graphene produced from GO contains small amounts

of hydroxyl, epoxy, and carboxylic groups on its surface37. Simultaneous thermal exfoliation and

10
reduction of GO produces thermally reduced graphene (TRG) containing pendent oxy-

functionalities31 or chemically converted graphene (CCG) produced by sonicating, and reducing

GO by chemical reducing agents that leaves other functional groups on CCG surface40. The

existence of functionalities on graphene surface facilitates dispersion of graphene in some polar

polymer matrices such as PMMA, PVA, etc.28.

Similar to CNT/polymer systems, nanocomposites of graphene can be manufactured by

solvent and melt blending, and in situ polymerization. The electrical and mechanical properties

of the composites depend on the manufacturing method.

GO, very polar in character, can exfoliate in water or protic solvents via sonication

leading to graphene oxide and therefore, the nanocomposites of graphene oxide with water-

soluble polymers (such as poly (ethylene oxide) (PEO), PVA, and PU) can be fabricated easily28.

Amine or isocyanate modification helps dispersing GO in aprotic solvents and polymers such as

polystyrene (PS), PMMA and PU. Reducing GO in the presence of a polymer has also been

performed by Stankovich et al.41 where GO/sulfonated-PS were mixed in a solvent followed by

GO reduction by hydrazine hydrate. One drawback of this method is that depending on polymer

type and reducing conditions, the polymer can also degrade.

The most economically attractive and scalable method for dispersing nanoparticles into

polymers is melt blending. However, due to the thermal instability of the modified graphene,

only a limited number of polymer composites with TRG are reported via melt blending. Owing

to very low bulk density of TRG (~3 mg/cm3), feeding TRG into melt compounding machine has

been a challenge. Torkelson and co-workers42 bypassed graphene exfoliation steps, and

exfoliated graphene in PP matrix directly via high stresses generated in solid-state shear

11
pulverization process, but XRD and TEM showed presence of small stacked graphite in the

composites. Recently, Kim et al.43 compressed TRG before melt blending with PE to tackle with

TRG’s handling issue due to low bulk density. The XRD confirmed single layer structure of

TRG after compression.

In-situ intercalative polymerization of monomers on graphene has been successfully

reported for PVA, PMMA, PU, and epoxy44, 45, 46. For PMMA/graphene composites, electrical

percolation threshold has been reported between 0.31-3.5 wt% depending on the type of

graphene used. The percolation threshold for PS/graphene composites has been reported to be

between 1-8 wt%. The in situ polymerization method gives the best mechanical and thermal

properties in nanocomposites because of the covalent bonding between graphene and polymeric

chains. However, increasing and large viscosities of the reaction mixture, make bulk-phase in

situ polymerization extremely difficult47, which also increases the system’s power requirements.

In addition, the surfactants are also required to increase graphene dispersion in high viscosity in

situ polymerization systems.

A good way to understand the dispersion of graphene in polymer matrix can be via

measuring electrical conductivity of the composites. Table 1.1 shows the electrical conductivity

thresholds for various graphene/polymer nanocomposites reported. In addition, the methods such

as TEM and melt rheology also give insights into the dispersion and percolation thresholds,

respectively.

1.3.2. Dispersion of Graphene in Polymers: Issues and Answers

Although graphene/polymer nanocomposites outperform other conductive

nanocomposites28, the required graphene loadings to reach conductive percolation for nonpolar

12
polymers, e.g., polyolefins, is significantly higher than that required for polar polymers (Table

1.1). For example, the percolation in PU/graphene composite is achieved at 0.005 vol% loading48

whereas that for PE/graphene composite is attained at about 1-8% graphene loadings39, 49. The

difference in percolation level is attributed to the poor dispersion of graphene into the nonpolar

polyolefin matrix compared to that in polar matrix. Poor dispersion does not only increase the

percolation limits but also reduces the enhancement in the mechanical properties and thermal

stability. The poor dispersion of graphene is associated with the absence of sufficient

graphene/polymer interactions. In order to address this challenge, chemical functionalization of

graphene is considered a viable solution. The attached functional groups on graphene surface can

increase dispersion of graphene, and develop interactions with polyolefins which can lower the

electrical and mechanical percolation limits.

The excellent mechanical, electrical and thermal properties of graphene are far limited by

their low solubility in aqueous and organic solvents. CCG containing smooth sheets, forms

bundles, and restacks into agglomerates when kept for some time, serving as a big disadvantage

for nanocomposite applications where dispersion is the main aim. On the other hand, TRG with

oxy-functionalities on their surface, are not dispersable in water. In order to overcome these

limitations, researchers attached different functional groups on graphene surface50 and/or have

also functionalized polymers for better dispersion and property enhancement43. In general,

because of the large volumes of polymers (90-99 %) needed in nanocomposites, functionalizing

polymers can be costly compared to functionalizing the nanofillers (0.1-9%). Since, extremely

small amounts of nanofillers, such as graphene, are required to change nanocomposites

properties, functionalizing fillers can be economical.

13
Table 1.1 Electrical percolation thresholds for selected graphene/polymer composites and their
electrical conductivities (adapted from ref.39 copyright (2011) with permission from Elsevier)

Percolation Electrical
Matrix polymer Filler type
threshold conductivity
(filler wt%) (mS/cm)
Epoxy Functionalized exfoliated graphite 1 ~50
Nylon-6 Graphite oxide 0.5 8.4
Polycarbonate Thermally exfoliated graphite oxide 0.3 500
PE Reduced graphite oxide 1-8 1300
PET Thermally exfoliated graphite oxide 1 20
PMMA Graphite nanoplatelets 0.7 1000
PP Graphite nanoplatelets 10 5
PS Functionalized reduced GO 0.2-2 150
PVA Reduced GO 0.5 100
Poly (vinyl chloride) Graphite nanoplatelets 1.4 60
Poly (vinylidene fluoride) Thermally exfoliated graphite oxide 2 0.3
PU Thermally exfoliated graphite oxide 0.6 N/A

Chemical functionalization could potentially pave ways towards the commercial

applications of graphene/polymer nanocomposites. Attaching organic molecules on graphene

surface via synthetic chemistry methods allows synthesizing p- and n-doped graphene based on

the selection of electron donating- or withdrawing-complexes bonded to graphene's sp2-carbon

network. By applying appropriate functional organic derivatives on graphene surface, graphene

can be easily processed via solvent-assisted techniques such as spin-coating, layer-by-layer

assembly, and solvent casting on glass or metal surface51.

In principle, graphene functionalization methods find their origin from CNTs due to

structural similarities between the two. The functionalization methods are classified into two

broad classes: covalent and non-covalent. In covalent functionalization, organic molecules are

covalently bonded on graphene surface whereas molecules attach to graphene via van der Waals

interactions in non-covalent methods.

14
Non-covalent or supramolecular functionalization methods primarily involve

hydrophobic, van der Waal, and electrostatic interactions between graphene and organic

molecules, and require physical adsorption on graphene surface. Non-covalent functionalization

can be achieved by polymer wrapping, adsorption of surfactant or smaller molecules, or

interactions with porphyrins or biomolecules, such as DNA and peptides. Graphene is originally

a π-system, and noncovalent intermolecular interactions involving π-systems are pivotal to the

stabilization of proteins, enzyme-drug complexes, DNA-protein complexes and organic

supramolecules, and functional nanomaterials50.

In covalent functionalization, organic molecules or polymers are grafted on graphene

surface which prevent graphene sheets from re-aggregation due to π-π-stacking. Kuilla et al.52

categorized covalent functionalization into four classes: Nucleophilic, and electrophilic

substitution reactions, condensation, and addition reactions. In nucleophililc substitution

reactions, epoxide groups on graphene surface are utilized. The amine functionality (-NH2) of

organic molecules bearing a lone pair electrons attacks the epoxy oxygen. These reactions are

easy, can occur at low temperature and in aqueous media. Essentially, all aliphatic and aromatic

amines, amino acids, amine-terminated biomolecules, and silane compounds can be used to

prepare amine-functionalized graphene.

In electrophilic substitution reactions, hydrogen atoms from graphene are replaced by an

electrophile. Spontaneous grafting of aryl diazonium salts to graphene surface is an example of

this type of reactions53. One key issue is the use of surfactants such as sodium dodecyl benzenyl

sulfonate to disperse graphene before the reaction. The functionalized graphene is highly

dispersible in DMF and other organic solvents. The electrical conductivity of sulfonated-

15
graphene has been reported by Samulski and coworkers54 to be ~1250 S/m which is much higher

than that from other functionalization techniques.

Figure 1.5 a) Proposed reactions during isocyanate treatment of graphene oxide where organic
isocyanates reacts with the hydroxyl (left oval) and carboxylic (right oval) groups of graphene
oxide sheets to form carbamate and amide functionalities, respectively (reproduced from ref.55,
copyright (2006) with permission from Elsevier).

In condensation reactions, two molecules (or functional groups) combine to form one

single molecule with the loss of entropy. The condensation reaction of graphene has been

reported with isocyanate, di-isocyante, and amine compounds through the formation of amides

and carbamate ester linkages. The isocyanate functionalized graphene was pioneered by

Stankovich et al.55 who used similar reactions with graphene oxide and GO. The resulting

functionalized graphene was highly dispersible in DMF solvent, and used for producing PU

nanocomposites48. Figure 1.5 shows the proposed schematics by Stankovich et al. for these types

of reactions.

In additional reactions, two or more molecules combine together to form a larger

molecule. Graphene has been functionalized with 1,3-dipolar cycloaddition of azomethine ylid,

and maleic anhydride (MA) where the reaction with MA proceeded with free radicals and the

16
resultant functionalized graphene was stable in THF for more than two months56, 57. Also, silane

(aminopropyl triethoxy silane, APTS) modified graphene was reported for producing different

types of polymer nanocomposites36, 58, 59, 60.

1.4. Nanocomposites Characterization

The dispersion of graphene or any nanofiller in a polymer can characterized by several

techniques. However, melt rheology and small angle x-ray scattering (SAXS) are among the

most important techniques used in this context. In the subsequent subsections, brief discussion

on the importance of rheological characterization of nanocomposites, and SAXS is presented.

1.4.1. Melt Rheology

Polymeric materials are viscoelastic (VE) by nature and therefore, exhibit distinct flow

behavior. Rheological characterization is used to analyze the melt flow properties of polymer and

nanocomposites. The knowledge and design of the flow behavior is essential in all forms of

production and processing of polymers to manufacture either small or larger parts. Therefore,

rheology has attained a key position in polymer research, being an important link between chain

structures, processing behavior and final properties of polymeric materials (Figure 1.6)61, 62.

In rheology, linear VE properties (direct relationship between stress and strain or the

range up to where viscosity or shear modulus (Gʹ) is independent of applied shear rate) are

usually used to determine the processing properties of materials. εaterial’s viscosity ( ) is the

most important parameter in defining the processing machinery used for manufacturing the

polymer products (Figure 1.7).

17
Figure 1.6 Rheological parameters acting as a link between molecular structure and final
properties of a polymer (reproduced from ref.61, copyright (2001) with permission from
Elsevier).

In a rheometer, the linear VE measurements are usually conducted at small amplitude

oscillatory strains (SAOS) where a small strain is imposed on the material in dynamic frequency

sweep experiments. The linear VE region is first determined using sinusoidal strain sweep

experiment, and a low strain is selected within the linear VE range to conduct further

experimentations.

Various groups around the world have contributed towards the standardization of the

linear VE response in polymers and nanocomposites. εacosko’s group8, 9


used critical strain

( c%) (strain at which Gʹ is equal to λ0% of its plateau value) for analysis of polymer

nanocomposites analogous to gels and colloids. For a percolating aggregated network, c% is

expected to be small, typically smaller than 10%63.

18
Figure 1.7 Typical viscosity curve for polyolefins defining the range of viscosities used in
various polymer processing machines (reproduced from ref.61, copyright (2001) with permission
from Elsevier).

In addition, being more sensitive to the structural changes in polymers, significant

increase in Gʹ at low angular frequency (ω) in SAτS experiments is another indication of

development of interfacial interactions in nanocomposites. With increasing filler loading, the

appearance of Gʹ independent of ω shows solid-like (or pseudo-solid) behavior in molten

nanocomposites which indicates strong filler/polymer interactions. The magnitude increment in

Gʹ at low ω region (Gʹp) can be used to quantify the filler dispersion64, 65 as follows:

G ' p ~ (   per )v

Here, ʋ is a power law exponent, ϕ is the volume fraction of filler, and ϕper is the

percolating volume fraction of the filler. With lower value of ϕper, dispersion of filler in a

polymer matrix is considered better. However, the accuracy of the application of the above

percolation law depends on the lowest frequency (terminal range) used to calculate the terminal

19
 0
G 'p . Typically, the polymers start degrading while acquiring data at lowest possible

frequencies. Furthermore, when the filler particle aggregate structure is sufficiently built up,

measurements of the linear VE properties can further lead to assessing the microstructure of the

dispersions. Detailed discussion on determining microstructure of graphene/polymer

nanocomposites is provided in other chapters.

1.4.2. Small Angle X-Ray Scattering (SAXS)

Electromagnetic radiation used to obtain information about materials with dimensions on

the same order as wavelength of the radiation. e.g., continuous and milky white glass of milk

appears as an emulsion of particles in ultraviolet black light because black radiation wavelength

is similar to the dimensions of the butterfat that is emulsified in milk. In electromagnetic

radiations, x-ray scattering is considered one of the most powerful characterization tools

available for both homogeneous and heterogeneous materials.

Understanding the structure of new materials on the mesoscopic scale (2-50 nm), such as

clusters, aggregates, and nanosized materials, requires suitable experimental techniques.

Scattering from x-rays is caused by the differences in electron density of the materials. Since the

larger the diffraction angle the smaller the length scale probed, wide angle x-ray scattering

(WAXS) is used to determine the crystal structure on atomic length scale, whereas the small

angle x-ray scattering (SAXS) is used to explore the microstructure on the nanometer scale.

Since studying ideal lamellae in polymers66 in 1960s, SAXS has become one of the most

widely used techniques to study morphology of solid polymers. SAXS provides indispensable

information about structural changes in polymers on supra-molecular level with dimensions in

20
the range of 10Å to a few thousands Å67. A pictorial description of SAXS working process, and

example of a scattering curve, the intensity versus the scattering vector (q) is shown in Figure

1.8. The scattering vector is calculated from the diffracting angle as q = 4π/ sin( /β). From the

absolute value of scattering intensity, the magnitude of the fluctuations of electron density can be

revealed. The shape of the curve, and the position of intensity peak or shoulder yield quantities

related to the spatial arrangement of the density fluctuations. Therefore, SAXS can supply useful

information about phase transitions if these processes are accompanied by changes in magnitude

of the fluctuation or by changes in the geometry of the scattering instrument.

sample detector

Monochromatic Beam 2ϴ q
~ 0.1-0.2nm
log (Intensity)

log (q), 1/nm or 1/Å


Figure 1.8 Simplified working diagram of scattering instrument

SAXS measurements in this thesis were obtained using a synchrotron source at Advanced

Photon Source, Argonne National Laboratory, Chicago, IL. With synchrotron x-rays, high power

density and small beam divergence of the incident x-ray beam permit design of time resolved

SAXS/WAXS experiments, and the use of very small specimens68. Since, synchrotron has high

flux, the scattering from air can be neglected, and environmentally controlled experiments can be

21
performed adequately. Details on environmentally controlled chamber used in this thesis are

described in experimental sections of specific experiments provided in proceeding chapters.

Rieker et al.69 used SAXS to study morphology of CB/PE composites prepared by

different methods. Mass fractals formed by CB in PE, and interpenetration of these fractals was

investigated, and the interpenetration of mass fractals above 20 wt% CB loading was observed

by the decreasing initial slope of log I(q) ~ log q curves. This was the first report on the effects

of polymer matrix, processing conditions, and methods on the morphology of composites.

Cattani et al.70 reviewed theories of x-ray scattering from various molecules, clusters, and

nanocomposite films. Recently, Archer's group71 used the time resolved SAXS experiments to

explain the unusual phase behavior of entangled polymers reinforced by silica nanoparticles. The

particles smaller than the radius of gyration of polymers lead to the non-Einstein flow behavior

in polymer composites72. The dynamics of polymers were slowed down due to the enthalpic and

entropic affect induced by nanoparticle size and particle-polymer interactions71.

1.5. Thesis Overview

This thesis examines the strategies to reduce percolation limits in nonpolar polymer

matrices. For this purpose, PE was selected as the main nonpolar matrix. PE is one of the most

commonly used commodity polymers, having the largest production worldwide and wide range

of practical applications.

While others have published on graphene/PE nanocomposites based on either the

functionalization of graphene or that of PE matrix, better and more economical methods to

homogenously disperse graphene in PE should be sought. I believe blending PE with OPE can

impart polar character to PE with improved processing properties, which can in return serve as a

22
better matrix for graphene dispersion. Thus, the hypothesis for this thesis is that blending non-

polar polyolefins with polar polymers will improve dispersion, processability, and conductivity

in graphene-based nanocomposites at lower graphene loadings than neat polyolefin

nanocomposites. In addition, hybrid filler synergic effects have been reported for polar polymers.

No reports have been published on using hybrid fillers for nonpolar polyolefins. I believe that the

dispersion of graphene in PE can also be enhanced by incorporating CB as a second synergic

filler to produce electrically conductive nanocomposites.

The thesis is organized in seven chapters. First, since this thesis focused on

graphene/polymer nanocomposites, details on the synthesis of TRG are discussed in chapter 2.

TRG was produced via simultaneous thermal exfoliation and reduction of GO, and characterized

by spectroscopic and surface analysis techniques. Detailed characterization procedures and

results are discussed therein.

Chapter 3 is focused on the synthesis of silane-modified TRG, where the reaction route

was altered to understand the mechanism of silane attachment onto TRG surface. Silane

modification was performed with and without the presence of an organic solvent. The

functionalized graphene were characterized by several chemical, spectroscopic, and surface

techniques. A reaction mechanism for the reaction of silane on graphene is discussed.

In chapter 4, miscible blend of PE with OPE were synthesized by solvent blending. The

role of OPE in controlling the processing properties of PE was investigated. The PE/OPE blends

were characterized by thermal, rheological, and mechanical techniques. Miscibility of OPE in PE

examined by thermal methods, and influence of OPE on the morphology of PE are discussed.

23
Chapter 5 focuses on quantification of graphene dispersion in PE and PE/OPE blends.

The nanocomposites were prepared by solvent blending. The state of graphene dispersion was

analyzed by melt viscoelasticity and SAXS, whereas the effects of increased graphene's

dispersion on electrical conductivity and mechanical properties were reported. In addition,

graphene-based nanocomposites were compared with CB-based micro-composites to illuminate

the differences between nano- and micro-composites.

Chapter 6 aims at using commercial grade graphene and CB, both together to enhance

graphene dispersion in PE. The synergic effects of graphene and CB were examined by melt

viscoelasticity and electrical conductivity measurements. Effects of increasing concentration of

hybrid fillers with different graphene/CB ratios were used to determine the most effective

graphene/CB ratio. The dispersion of hybrid fillers is also analyzed by SAXS, and electron

microscopy.

Finally, chapter 7 summarizes all key results from this thesis, and provides future

research directions. Potential applications of the developed PE/OPE blends-graphene

nanocomposites, and the use of polymer grafted-graphene in manufacturing polymer

nanocomposites are proposed at the end.

24
CHAPTER-2

GRAPHENE: SYNTHESIS AND CHARACTERIZATION

This chapter focuses on synthesizing graphene by thermal exfoliation and reduction of

GO. The graphene was synthesized at the Petroleum Institute, Abu Dhabi. The purpose of this

chapter is to give a general idea to the reader about the type of graphene used in this thesis and

their synthesis.

2.1. Thermally Reduced Graphene (TRG)

In this thesis, we have used thermally reduced graphene (TRG) as one of the main

nanofiller for polymer reinforcement. Following section provides fundamental information

regarding TRG synthesis and characterization.

2.1.1. Introduction

Selection of the synthesis route for graphene depends on the choice of potential

applications. Graphene can be synthesized from natural graphite or GO. The starting material,

and method to split that material into atomically thick graphene sheets, largely influence the final

physical and chemical properties of graphene. Recently, Kim et al.28 classified the methods for

production of graphene into two broad categories: bottom-up, and top-down methods.

In bottom-up approach, graphene is produced by methods such as chemical vapor

deposition (CVD)73, arc discharge74, epitaxial growth on SiC75, 76


, chemical conversion77,

reduction of CO78, unzipping CNTs79, and self-assembly of surfactants80. Large size and defect

free graphene is often produced from CVD and epitaxial growth on small scales. Due to small

production volumes, these methods are considered suitable for producing graphene for

25
fundamental studies and electronic applications. The nature, average size, and thickness of

graphene produced from bottom-up methods is summarized elsewhere28.

Top-down methods produce graphene or modified graphene by separation or splitting of

graphite or GO derivatives. Significant economic advantages are offered over the bottom-up

approach since starting material is graphite or its derivative. That’s why, top-down methods are

considered more suitable for mass production of graphene sheets, useful for polymer

nanocomposite applications where large quantities of fillers are usually required. Starting from

graphite, the top-down approach includes micromechanical exfoliation30, direct sonication81, 82,

electrochemical sonication83, and superacid dissolution of graphite84. On the other hand,

producing modified graphene from GO includes Li-alkylation of graphite flouride85, chemical

reduction of GO86, chemical reduction of organically modified GO86, 87


, and simultaneous

thermal exfoliation and reduction of GO28, 36, 37, 38, 88.

The simultaneous thermal exfoliation and reduction of GO36, 37, 88 results in the splitting

of a fraction of the oxygen-containing groups as CO, CO2, and small quantity of steam.

Thermally reduced graphene (TRG), a reduced form of graphene oxide, is composed of a single

or few-layer oxygen-functionalized graphene sheets containing a far less amount of oxygen-

containing groups compared to GO, and corresponding to C/O/H atomic ratio of 10/1/1. The

presence of the residual oxygen-containing groups on TRG surface can anchor various molecules

by imparting it a polar character.

In following discussion, production and characterization of TRG is reported which has

been published in several collaborative publications. Also, in some parts of this thesis, a

26
commercial graphene sample was used. A brief discussion on characterization of the commercial

graphene sample called "Graphene Nanoplatelets (GnP)" is also provided at the end.

2.1.2. Experimental Details

This section contains description of materials used, and methods of preparation of

graphene used in this study.

2.1.2.1.Materials

Natural flake graphite (-10 mesh, 99.9%, Alfa Aesar), sulfuric acid (95-97%, J.T.

Bakers), hydrochloric Acid (37%, Reidel- deHaen), hydrogen peroxide (30% solution, BDH),

potassium chlorate (Sigma Aldrich), and sodium hydroxide (Reidel-deHaen) are used as

received.

2.1.2.2.TRG Synthesis

TRG was produced following the thermal exfoliation method89. In this method, graphite

is oxidized using Staudenmaier method90 as follows: graphite (5 g) was placed in ice-cooled

flask containing a mixture of H2SO4 (90 ml) and HNO3 (45 ml). Potassium chlorate (55 g) was

added slowly to the cold reaction mixture. The reaction was stopped after 96 h by pouring the

reaction mixture into water (4 L). HCl solution (5%) was used to wash the produced graphite

oxide (GO) until no sulfite ions were detected. The mixture was then washed with water till no

chloride ions were detected. GO was dried in a vacuum over night. The dried GO was exfoliated

by rapid heating at 1000 °C in a tube furnace (Model 21100, Barnstead Thermolyne) under flow

of nitrogen for 30 s to 2 minutes.

27
2.1.2.3.Characterization of TRG

XRD (X’Pert PRτ εPD diffractometer, PAσalytical) was used to test the oxidation of

graphite, and complete exfoliation of GO into TRG. A rapid scan was conducted between 5-35o

with 0.02o/sec step size at 40 KV voltage with an intensity of β0 A using CuKα radiation of

wavelength 1.54 Å. TEM images were obtained using FEI Tecnai G20 TEM with point to point

resolution of 0.11 nm, coupled with Energy Dispersive X-Ray spectroscopy. The samples were

prepared by dispersing approximately ≈0.η mg of TRG in βη mδ of dimethyl formamide by

sonication for 10 minutes in a sonicating bath at room temperature. Two drops of the suspension

were deposited on a 400 mesh copper grid covered with thin amorphous film (lacey carbon).

SEM images were obtained using Philips FEI Quanta 200 SEM operated at high-vacuum mode.

A LabRAM HR (Horiba Scientific) was used to obtain Raman spectra. Typically, a 50x objective

was used with 633 nm excitation line. Bulk density of graphene samples was measured by

AutoTap Tap density meter (Quantachrom Instruments, Model: D-AT-3). Each sample was

tapped 2000 times before taking the measurement. Density values were recorded as average of at

least three readings. Specific surface area, pore volume and pore size distribution were

determined using Quantachrom Autosorb-1 (Quantachrom Instruments). Prior to measurement,

the sample was degassed for 16 hours at 200o C. TRG Carbon/Oxygen (C/O) ratio was

determined from CHN elemental analysis carried out by Midwest MicroLab, LLC (Indiana, US)

using a combustion analyzer at 990 °C under ultrapure oxygen; oxygen content was carried out

by Unterzaucher method at 1050 °C. Fourier transformed Infrared (FTIR) spectrum of dried

TRG with KBr pellets (spectroscopic grade, Merck) with ~0.2% TRG was obtained using

Thermo Nicolet FTIR at a resolution of 4 cm-1 and 32 scans. X-ray Photoelectron Spectroscopy

(XPS) measurement was conducted using a SSX-100 system (Surface Science Laboratories, Inc.)

28
equipped with a monochromated Al Kα X-ray source, a hemispherical sector analyzer (HAS),

and a resistive anode detector.

2.1.3. Results and Discussion

Rapid heating of GO leads to simultaneous reduction and exfoliation of GO sheets. By

controlling the exfoliation condition, the bulk density, surface area, and the C/O contents of TRG

can be varied as discussed in detail elsewhere89, 91. First direct evidence of exfoliation of GO was

the substantial volume expansion of 100-300 times due to rapid heating (Figure 2.1a).

Specifically, one gram of GO with bulk volume of about 2 cm3 yielded approximately 0.7 g TRG

with a bulk volume of about 150-180 cm3 depending on the density of the produced TRG.

The X-ray diffraction (XRD) patterns provide supporting evidence for the complete

exfoliation of GO to TRG (Figure 2.1b). A strong diffraction peak (002) was observed at 2 =

26.5° for graphite that corresponds to a d-spacing of 3.37Å. The (002) peak is an intrinsic peak

showing the stacked graphene layers in pure graphite. The oxidation of graphite led to polar

oxygen functionalities on the surface of GO. The presence of these functionalities and due to the

adsorbed water intercalation, the (002) peak shifted to 2 = 11.4° indicating interlayer expansion

to a d-spacing of 7.8Å. In contrast to the diffraction patterns in graphite and GO indicating the

ordered layered structure, TRG diffraction pattern shows no noticeable diffraction peaks

confirming the complete exfoliation of GO, and production of no stacked structure.

Raman spectroscopy probes structural and electronic features of graphitic materials.

Raman spectrum provides useful information about the defects (D band), in-plane vibrations of

sp2 carbon atoms (G band), and the stacking order (βD or Gʹ band) of graphene layers92. The

complete exfoliation of graphite into TRG was studied by Raman spectroscopy (Figure 2.2). The

29
shape of the Gʹ (or βD) peak is used to distinguish between single-layer graphene, bilayer

graphene, and the bulk graphite. Bilayer sheets or sheets with less than five layers show a

broader and symmetrical Gʹ peak, while bulk graphite exhibits a distorted peak92. The Raman

spectrum of TRG showed a broader and symmetrical Gʹ peak in the βη00-2800 cm-1 range,

indicating graphene sheets with less than five layers were obtained after thermal exfoliation and

reduction.

a) b)
TRG
GO

2
5 10 15 20 25 30

Figure 2.1 0.5 g of GO expands to 75 ml of TRG upon rapid heating at 1050 °C (a), XRD
patterns of pure graphite, GO and TRG (b). (reproduced from ref.88, copyright (2012) Springer)

The second prominent feature of the Raman spectra is the size, shape, and intensity of D

and G bands. The intensity of G band is known to increase linearly with graphene layer

thickness93. The graphite exhibited a strong and intense G band showing bulk stacked sheets as

compared to that of TRG indicating a few layered graphene. The D-band that appears as a result

of induced defects, bears an inverse relation between its intensity and number of graphene layers.

Intensity of D-band decreases with increasing graphene thickness, and it is almost invisible for

defect-less bulk graphite94. The defects are more easily introduced into the thinner graphene
92
sheets, and the intensity of these defects is governed by the method of producing graphene .

30
The differences of the spectra of graphite and TRG clearly indicates that we have successfully

produced graphene instead of multilayer graphite nano-crystallites in this work.

700
300
G
600 250
G'
200
500
Raman Intensity

150

400 100

50
300 2500 2600 2700 2800

200 D
Graphite

100 TRG

0
1100 1200 1300 1400 1500 1600 1700 1800
Frequency, cm-1

Figure 2.2 Raman spectra of graphite and TRG. (reproduced from ref.88, copyright (2012)
Springer)

The surface morphology of TRG was investigated by scanning electron microscopy

(SEM), and high resolution transmission electron microscopy (HRTEM). Though, crevice like

structure was prominent in low density TRG (Figure 2.3a), and the samples with higher densities

appear to have fluffy agglomerated structures as produced (see ref.88 for details), the HRTEM of

TRG after dispersing in DMF showed graphene layers with paper-like morphology (Figure 2.3b).

The transparent graphene sheets were clearly visible confirming that GO was successfully

exfoliated. However, the elastic corrugations, and the scrolled or folded edges often result in
95
different brightness on the surface of graphene , which was also observed in TEM image as

dark lines.

31
a) b)

10 μm

Figure 2.3 TRG with bulk density of 17 mg/cm3 and C/O = 13: SEM (a), and HRTEM (b).
(Image reproduced from ref.88)

TRG was also characterized by N2 adsorption for BET (Brunauer-Emmett-Teller) surface

area, pore volume and pore size distribution. The BET measurement indicated that the

sorption/desorption behavior of TRG followed a standard type II-b BET isotherm (inset of

Figure 2.4). The cumulative pore volume was determined using BJH cumulative desorption total

pore volume method. The TRG sample with the highest C/O ratio (17/1) and the lowest bulk

density (3 mg/mL), exhibited a surface area and pore volume of 1260 m2/g and 3.7 cm3/g,

respectively. Both the surface area and the cumulative pore volume decreased with the decrease

of C/O ratio, and increasing bulk density of graphene (see ref.88 for details).

The chemical structure of TRG was studied by FTIR and XPS to determine the type of

the oxygen groups on TRG surface. The FTIR transmission spectrum between 600 and 2000 cm -
1
(Figure 2.5a) indicated the presence of carboxylic groups (~1720 cm-1), hydroxyl groups

(~1420 cm-1, 1200-1220 cm-1) and epoxy groups (~1060 cm-1). The graphitic C=C give rise to

multiple stretching vibrations in the 1550-1680 cm-1 range where the stretching at 1600 cm-1

32
represents very small amounts of unexfoliated graphite with small amount of the adsorbed

waters96.

0.5 8

Volume Adsorbed [cc/g]


Adsorption
6 Desorption
0.4
Pore Area [cc/g-Å]

4
0.3
2
0.2
0
0.0 0.2 0.4 0.6 0.8 1.0
0.1 Relative Pressure [P/Po]

0.0
0 10 20 30 40 50 60 70
Pore Size [Å]

Figure 2.4 Pore size distribution of TRG sample with bulk density of 3 g/L. Inset shows the BET
adsorption-desorption isotherm curve (reproduced from ref.88, copyright (2012) Springer).

The high resolution C1s XPS deconvoluted spectrum is shown in Figure 2.5b. The peak at

284.57 corresponded to graphitic sp2 C=C atoms. The C-O peak observed at 285.86 eV was

attributed to the presence of epoxy and hydroxyl groups. The peak at 287.26 eV corresponded to

the carboxylic acid groups (O-C=τ). The π-π interactions between graphene layers were

reflected by the peak at 290.54 eV36, 37. These results are consistent with the findings of a recent

on identifying the structure of reduced graphene using XPS and NMR97.

2.1.4. Summary

Successful synthesis of thermally reduced graphene (TRG) is observed by simultaneous

thermal exfoliation and reduction of graphite oxide (GO). TRG contains some residual oxy-

functional groups on its surface observed by XPS and FTIR, and planar defects seen in Raman

33
spectroscopy. XRD showed the evidence of complete exfoliation, whereas Raman spectroscopy

indicated that TRG particles were composed of 4-5 atomically thick graphene sheets. The

magnitude of oxy-functional groups on graphene surface can be altered by varying the

exfoliation time and temperature, which also gives control over the surface area and bulk density

of the produced graphene.

a) 5000
284.57 b)
sp2 -C
Transmission, % (a.u.)

4000

3000

Counts
285.86
2000 C-O
287.26 290.54
C=O π*-π*
1000

5 sp2-C 2 1
0
281 283 285 287 289 291 293 295
2000 1800 1600 1400 1200 1000 800 600
Binding Energy, eV
Wavenumber (cm-1)
Figure 2.5 FTIR (a) and XPS high resolution C1s spectrum (b) of TRG sample. the inset in a)
provides a structural model for TRG (reproduced from ref.37 with permission from the Centre
National de la Recherche Scientifique (CNRS) and The Royal Society of Chemistry).

2.2. Graphene Nanoplatelets (GnP)

In parts of this thesis, a commercial graphene sample called "Graphene nanoplatelets

(GnP)" was also procured from Graphene Supermarket, and used in manufacturing polymer

nanocomposites. Therefore, a brief discussion on the characterization of GnP is presented in this

section. More details on characterization and use of GnP will also be provided in the proceeding

chapters.

TEM image of GnP was obtained by sonicating a very dilute solution of GnP in acetone,

and depositing one drop of suspension on the Cu grid with holy carbon. Similar to TRG, GnP

34
showed paper-like morphology (Figure 2.6). However, TRG sheets were smaller in lateral

dimensions whereas GnP sheets are extended over a few microns. Also, importantly, no surface

wrinkles were observed in GnP sheets which indicates the absence of any oxy-functionality on

GnP.

Figure 2.6 TEM image of graphene nanoplatelets (GnP)

Acknowledgements

The date presented in this chapter has been the part of several collaborative publications,

and many people have contributed to its production. Authors would like to thank Jamie Whelan

and Sabna Khadar from Petroleum Institute for helping with BET and SEM analysis, and Issam

Ismail from University of Minnesota for XPS experiments.

35
CHAPTER-3

EFFECT OF SOLVENT ON THE UNCATALYZED SYNTHESIS OF AMINOSILANE-

FUNCTIONALIZED GRAPHENE

This chapter is modified from a paper published in

RSC Advances1

Muhammad Z. Iqbal2, Marios S. Katsiotis3, Saeed M. Alhassan4, Matthew W.

Liberatore5, Ahmed A. Abdala6

3.1 Introduction

Since its discovery in 2004, graphene has attracted the attention of researchers owing to

its extraordinary electronic, thermal and mechanical properties30, 98, 99, 100, 101
. Due to this

combination of extraordinary properties in addition to the very high surface area, graphene is

widely used as a nano-filler in manufacturing of polymer nanocomposites with improved

mechanical, thermal and electrical properties28, 38, 102, 103, 104


. Nevertheless, homogenous

dispersion of graphene and strong interfacial interactions between graphene and the polymer

matrix is required to obtain enhanced properties. The chemical functionalization of graphene

alters the Van der Waals interactions among the nanofiller aggregates, making them easier to

disperse in a polymer matrix and can also enhance the interface between graphene and the

polymer matrix. On the other hand, the extremely high surface area of graphene makes it an ideal

1
Reprinted with permission from the Centre National de le Recherche Scientifique (CNRS) and The Royal Society
of Chemistry, (2014), 4, 6830-6839
2
Primary author and researcher at Colorado School of Mines, Golden, CO, U.S.A.
3
Co-author and postdoc at The Petroleum Institute, Abu Dhabi, U.A.E.
4
Co-author and assistant professor at The Petroleum Institute, Abu Dhabi, U.A.E.
5
Author for correspondence and associate professor at The Petroleum Institute, Abu Dhabi, U.A.E.
6
Author and advisor at Colorado School of Mines, Golden, CO, U.S.A.

36
candidate as an adsorbent for H2 storage105, removal of pollutants from water37, oil31, and

gases106. In order to enhance the compatibility of graphene with polymer matrices and increasing

its adsorption affinity functionalization of graphene may be required.

The use of residual oxygen pendent groups on thermally or chemically reduced graphene

for further attachment of the organic moieties is an attractive route exploited for covalent

functionalization of graphene50. Silane coupling agents are historically applied to surface

modification of nanofillers25, 58, 59, 60, 107. Graphene produced from reduction of graphite oxide

contains sufficient amounts of hydroxyl, epoxy and carboxyl groups on their basal planes and

edges97. These groups can be dehydrated with silane coupling agents under appropriate

conditions. Also, the functional groups of silane coupling agents can be chemically attached to

the polymer. Amongst, 3-aminopropyltriethoxysilane (APTS) has been widely reported as a

reactive coupling functionalization for CNTs25, 108 and graphene58, 59, 109. Specifically, Wang et al.

reported attachment of APTS on graphene surface and used the functionalized graphene in

graphene-epoxy nanocomposites59. The grafting of APTS was proposed to follow dehydration of

carboxylic and hydroxyl groups. In another report by Yang et al., APTS reaction with chemically

converted graphene resulted in 4% Si grafting on atomic basis58. However, the proposed

mechanism of APTS grafting was through the epoxide groups on the graphene surface. In

addition, both reports employed DCC catalyst to accelerate reaction kinetics and enhance the

grafting yield. A 3.4 atomic% Si yield has also been reported for graphene by Ganguili et al. 109.

Gasper et al. reported detailed reaction mechanism for grafting various types of silanes on

CNT108. Therefore, these studies outline the importance of silane functionalized graphene in

various applications. For the mass production of APTS-f-graphene, a method that does not utilize

a catalyst is needed. Catalyst can increase the production cost by many fold. Hence, a catalyst-

37
free method in functionalizing graphene by silane based coupling agents is highly desirable. In

addition, in order to use the APTS-f-graphene in different applications, a better understanding of

APTS chemistry is needed. This study focuses on the above mentioned question using various

characterization techniques.

In the present work, we present a simple catalyst-free method for APTS attachment on

TRG with a higher grafting yield of silane attachment. Thermally reduced graphene (TRG) is

produced via simultaneous thermal exfoliation and reduction of graphite oxide (GO). The effect

of solvent on the localization of functionalization on graphene surface is also discussed. Two

simple routes for reacting APTS with TRG are presented; 1) using an organic solvent, and 2)

using pure silane without any solvent. A mechanism involving APTS attachment onto graphene

is also proposed based on various characterization tools to reflect the effect of organic solvent on

functionalization chemistry. The resulting functionalized graphene was thoroughly characterized

using physicochemical methods to understand the nature of functionalization reaction. The

success of the simple reaction routes is expected to promote mass production of APTS-

functionalized graphene.

3.2 Experimental Details

Following sections describes details on materials used in this study, and the procedures

adapted to synthesis TRG and functionalized TRG.

3.2.1 Materials

Natural flake graphite (-10 mesh, 99.9%, Alfa Aesar), Sulfuric Acid (95-97%, J.T.

Bakers), Hydrochloric Acid (37%, Reidel- deHaen), Hydrogen Peroxide (30% solution, BDH),

38
Potassium Permanganate, Sodium Nitrate (Fisher Scientific), 3-aminopropyl triethoxysilane

(98%, Merck) and toluene (99.5%, Panreac) were used as received.

3.2.2 Preparation of TRG

TRG was produced via the thermal exfoliation of graphite oxide (GO)110. In this method,

graphite is oxidized using Staudenmaier method90 as follows: graphite (5 g) was placed in ice-

cooled flask containing a mixture of H2SO4 (90 ml) and HNO3 (45 ml) then potassium chlorate

(55 g) was slowly added to the pre-cooled (0-5° C) reaction mixture under stirring. After the

reaction proceeds for 96 h, it is stopped by pouring the reaction mixture into water (4 L). GO is

filter and washed with HCl solution (5%) until no sulfite ions are detected. Finally, the resulting

deep brown colored mixture was repeatedly washed with copious amount of water until no

chloride ions was detected. The wet GO was dried under vacuum overnight. TRG is made by

rapid heating of dry GO powder at 1050° C in a tube furnace (Model 21100, Barnstead

Thermolyne) under flow of nitrogen for 30 s.

3.2.3 Silylation Reaction

Two simple routes are proposed to reflect the effect of solvent on APTS attachment

chemistry as follows:

Method 1: 50 mg of TRG was dispersed in 30 mL of pure APTS in a 50 mL reaction

flask. The mixture was refluxed at 100° C for 3 hours under stirring then cooled to room

temperature and the functionalized TRG (f-TRG) is recovered by vacuum filtration.

Method 2: same as Method 1 except the 30 mL of pure APTS is replaced by 50 mL of 30

vol% APTS in toluene. TRG functionalized using the solution is denoted (f-TRGS).

39
To remove any physically adsorbed APTS from the surface of TRG, f-TRG was

dispersed in toluene under stirring for 15-20 minutes followed by tip-sonication for additional 5

minutes and filtering under vacuum. To ensure the complete removal of any unreacted APTS,

this washing procedure was repeated twice. The samples were then dried in at 120° C for 2 h.

3.2.4 Characterization

Transmission Electron Microscopy (TEM) analyses were performed using FEI Tecnai

G20 with 0.11 nm point resolution and operated at 200 kV. Electron Energy Loss Spectroscopy

(EELS) was performed using post column energy filtered camera (Gatan GIF 963). Energy

filtered TEM (EFTEM) mapping was applied to map the location of the elements on the surface

of f-TRG samples by measuring core, post-edge and pre-edge losses of the respective elements

using the “three-window method” technique111. X-ray Photoelectron Spectroscopy (XPS)

measurements were performed using SSX-100 system (Surface Science Laboratories, Inc.)

equipped with a monochromated Al Kα X-ray source, a hemispherical sector analyzer (HAS)

and a resistive anode detector. For the high-resolution spectra, the lowest binding energy C1s

was set at 285.0 eV and used as the reference for all of the other elements. Fourier Transformed

Infrared (FTIR) spectra of TRG, f-TRGp, and f-TRGs in the range of 400-4000 cm-1 were

collected using Thermo Nicole FTIR at a resolution of 4 cm-1 and 32 scans. The spectra of the

dried samples were obtained in KBr pellets (Merck, spectroscopic grade) containing 0.2 wt% of

TRG. A LabRAM HR (Horiba Scientific) was used to obtain Raman spectra. Typically, a 50x

objective was used with 633-nm excitation line. Thermo-gravimetric analysis was carried out by

STA coupled with a mass spectrometer (STA-QMS, 409 PC Netzsch). The temperature range

was from 35 °C to 450 °C at a ramp rate of 10 °C in an inert atmosphere of nitrogen (30

mL/min). The electrical conductivity was measured by a custom-made conductivity cell at the

40
National Renewable Energy Laboratory (Golden, CO). The samples were bath-sonicated in 10

mL of water/isopropanol solution (3/1) followed by 2 minutes of tip-sonication. The dispersed

samples were dried on copper stubs for the conductivity measurements. The resistivity was

measured using a multimeter (Waveteck 28XT, Accuracy: ±1% + 0.1 Ohm). Further details on

experimental procedures are provided in Appendix-A.

3.3 Results and Discussion

The simultaneous thermal exfoliation and reduction of GO to TRG was successfully

confirmed by XRD and TEM (see Appendix-A).

The APTS-functionalized TRG was systematically characterized to investigate the effect

of solvent on chemical, thermal, electrical and morphological properties of APTS-f-graphene.

The chemical nature of the f-TRG is studied by FTIR, XPS, Raman, and electron energy loss

(EEL) spectroscopy. A mechanism of APTS grafting on TRG has been proposed. TEM micro-

images and elemental mapping were used to study the morphology of f-TRG. The electrical

conductivity and thermal stability of f-TRG are also reported.

3.3.1 Chemical Analysis

FTIR spectroscopy (Figure 3.1) confirms the successful attachment of APTS on TRG by

the appearance of additional functional group stretches associated with APTS. The intensity of

peak D (1189 cm-1), which corresponds to –COOH groups has been reduced. This suggests that

the carboxylic groups are the active moieties for APTS attachment. Peak H (1544 cm-1), which

corresponds to C=C stretching aromatic carbons in non-graphitic domains, remains the same for

f-TRGP but shifts to a higher frequency (~1558 cm-1) for f-TRGs. This shift to a higher

41
frequency for TRGs suggests the possibility of APTS reacting with C-hexagons of the TRG

sheets in addition to the reaction with TRG’s functional groups (Also confirmed by XPS

analysis). The carbonyl group >C=O at ~1740 cm-1 (peak J) also remains the same for pure and

f-TRG. The appearance of new peaks in f-TRGP and f-TRGS is attributed to –Si-O stretching

[peak A (–Si-OH at 914 cm-1), peak B (Si-O-Si at 1015 cm-1) and peak C (–Si-O-R at 1110 cm-

1)], indicating the covalent functionalization of graphene. Clearly, the solvent has influenced

broadening of Si bonding peaks. Also, the NH- amine vibration, which is expected around 3300-

3400 cm-1 in f-TRG (peak M), is not prominent, likely due to overlapping with the characteristic

bands of adsorbed water in the same region58. Peak F (1380 cm-1) corresponds to in-plane –OH

deformation and peak G (1485 cm-1) is assigned to phenyl groups. The –NH bending vibration

mode112 (peak I at 1629 cm-1) also confirms the direct attachment of amine groups to graphene

and its very low intensity indicates the presence of a very small amount of this sequence.

Moreover, the appearance of two new peaks at ~2853 and 2925 cm-1 for the f-TRG samples is

attributed to CH2-CH2 symmetric and asymmetric vibrations from APTS108.

Figure 3.1 FTIR spectra of the pure and f-TRGp and f-TRGs

42
While FTIR confirmed the attachment of APTS to graphene, a quantitative estimate of

the level of functionalization is provided by XPS analysis. The results shown in Table 3.1 reveal

that the functionalized graphene samples contain 7.4% Si and 6.9% N for f-TRGP and 8.0% Si

and 6.5% N for f-TRGS. Not only we are able to achieve such high yield without the use of any

catalyst but also the observed yield is higher than all previously reported for nanocarbons, e.g.,

graphene58 and CNT108. The Si/N ratio for f-TRGP is close to the stoichiometric APTS ratio of 1.

However, The Si/N ratio for f–TRGS is 1.2. Based on the stoichiometric ratio, Si/N ratio greater

than unity suggests the probability of side reactions in the presence of the solvent as will be

discussed later.

Table 3.1 Atomic composition of pure and f-TRG as obtained from XPS

Atomic %
C1s O1s N1s Si2p
TRG 89.1 10.9
f-TRGp 66.8 20.0 6.9 7.4
f-TRGs 67.4 18.2 6.4 8.0

In order to further explore the structure of f-TRG, high resolution C1s, O1s, N1s and Si2p

spectra were collected (Figure 3.2 and Table 3.2). In C1s spectra, all the samples exhibit the

same graphitic C=C peak whose area percentage decreases in the f-TRG due to the appearance of

new bands. The C-O peak58 in TRG decreased in f-TRGP and f-TRGS samples. Due to the

smaller electronegativity difference between C and N compared to between C and O, the peak

for C-N is observed at a lower binding energy (BE). The relative percentage of ~285 eV peak has

increased significantly in f-TRGP and f-TRGs. The peak for C=O carboxylic groups for f-TRGS

has a reduced intensity compared to the C=O peak in TRG, it vanishes in f-TRGP. Thus, C=O is

43
a potential attachment site for APTS on TRG (i.e, the carboxylic group on TRG). The

disappearance of C=O in f-TRGP suggests selective attachment of APTS onto the carboxyl

group. Meanwhile, two additional peaks are observed in C1s spectra of f-TRGP. The first peak at

286.16 eV is assigned to aliphatic C-N group58, 113


whereas the second peak at 288.85 eV is

attributed to either N-C=O or O-C=O108. Since APTS is a bi-functional silane, its grafting to

TRG can also take place through a linkage between the amine group in APTS and the carbonyl

or carboxylic functionalities on TRG. The small peak at 288.85 eV BE is attributed to such

linkage108. However, the relative area of this peak is very small indicating that fraction of APTS

that attaches to TRG through the amine linkage, is relatively small. Therefore, we conclude that

APTS reacts bi-functionally with TRG through attacking the phenolic and carbonyl groups on

the surface. εoreover, the π-π* interactions59 are observed in TRG and f-TRGS whereas no π-π*

interactions are observed in f-TRGP possibly due to the disappearance of C=O peak in f-TRGP.

Unlike C1s core level, where the assignment of the peaks to single and multiple C–O

bonds is quite straight forward, the assignment of O1s peaks is more difficult. The carbonyl

groups C=O are expected between 531 and 532 eV, C–O bonds in ethers and hydroxyls between

532 and 533.5 eV, whereas ether oxygen atoms in esters and anhydrides appears at 533.8–534.6

eV, or higher, and the carboxilates should originate a single component at BE similar to that of

C=O groups in the case of oxidized carbon surfaces. Although C=O is observed for TRG, f-

TRGP and f-TRGS, the relative intensity of C=O is lower in the f-TRG samples as a result of

APTS attachment in agreement with the analysis of the C1s spectra. The C-O-C/O-C=O band in

TRG appears at higher BE for TRGs showing the bonding with a more electronegative structure

at this point. The peak at ~ηγγ.ηeV originates from τ - contribution of TRG due to the presence

of O-2 ions.

44
The appearance of new Si-O bands in f-TRGP and f-TRGS is sign of APTS

functionalization and its intensity and relative contribution depend on the degree of APTS

functionalization114. Moreover, the carboxylic band observed for TRG at ~534 eV disappears in

the f-TRG samples in excellent agreement with the results from C1s XPS spectra and FTIR. The

prominent changes in O1s spectra of the pure and functionalized samples clearly indicate that the

silanization has changed the structure of the graphene sheets. Table 3.2 summarizes the results of

the high resolution C1s and O1s spectra. Furthermore, the details of high resolution N1s and

Si2p XPS spectra of f-TRG are provided in Appendix-A.

Figure 3.2 High resolution XPS spectra of pure and f-TRG samples

45
Considering the XPS data and the structure of f-TRG, the APTS may attach to graphene

through the phenolic as well as the carbonyl surface groups where a similar mechanism has been

reported by Gasper et al.108 for CNTs. However, this multiple functionality may differ from the

previously reported mechanisms for silylation where the aminosilanes are proposed to attack the

epoxy groups58 on the graphene surface. In addition, the second possibility of enhanced grafting

efficiency of APTS is NH2-silicon polymerization108 on the surface of TRG.

Table 3.2 High resolution fitting of C1s and O1s groups appearing in the XPS spectra of TRG, f-TRGp,
and f-TRGs
C1s O1s
C=C C-O/ N-C=O/ C-O-C/
C-N C=O π-π* C=O Si-O O-C=O
(sp2) C-N O-C=O O-C=O
TRG 284.57 285.86 - 287.26 - 290.54 531.56 - 533.55 534.84
f-TRGp 284.57 285.37 286.16 - 288.85 531.26 532.75 - -
f-TRGs 284.67 285.67 - 287.26 - 290.14 531.66 532.65 533.74 -

Previous reports demonstrated APTS grafting efficiency on MWCNT108 up to 4.5 Si

atomic%, and 3.4%109 and 4.1% for graphene58. The current method is able to generate much

higher yield of ~ 7.4% Si attachment in f-TRGP and ~ 8% Si in f-TRGS. The increase in Si

attachment can be attributed to the reaction of APTS with not only the surface functionalities of

graphene but also with the surface C-atoms. Schematics of possible mechanisms are also

discussed later.

The effect of silane attachment on the electronic structure of graphene is also confirmed

by EELS. The core-level energy losses of C, O, N and Si (Figure 3.3 and Figure 3.4) are

measured at the areas of minimum overlapping between the graphene sheets. All collected EELS

spectra are background corrected.

46
a) f-TRGs
b)

Intensity, a.u
Intensity, a.u
f-TRGp
f-TRGs
TRG
f-TRGp

280 300 320 340 360 520 535 550 565 580 595 610 625
Energy Loss, eV Energy Loss, eV
Figure 3.3 EELS Carbon K-edge a), and Oxygen K-edge b)

The carbon K-edge (Figure 3.3a) is the characteristic of graphene115. The sp2

hybridization of TRG is clearly observed as sharp peaks corresponding to the 1s  * and 1s

 * transitions at 289 and 294 eV, respectively. Interestingly, the hybridization state of

graphene is altered in the functionalized samples. In specific, the sp2 structure becomes less

dominant in the case of f-TRGP and f-TRGS in excellent agreements with FTIR, XPS and Raman

results. The ratio was calculated using the peak-ratio method116 and found to

be 98±1.5% , 94±1.1%, and 87±1.9% for TRG, TRGp, and TRGs, respectively. A plausible

explanation for this change in hybridization is attributed to the sp3 structure of carbon atoms in

silane. Based on the current information obtained from FTIR and XPS on the successful grafting

of APTS on graphene, the apparent change in hybridization most possibly represents the

combination between the sp2 structure of graphene and the sp3 structure of silane. The oxygen K-

edge region is presented in Figure 3.3b and exhibits an overall shape equivalent to that of the N

K-edge. Two * peaks can be observed at 530 and 535 eV which are attributed to the C-O bonds.

The dominating structure however, is the absorption beyond 541 eV with a particular triangular

shape. Similar findings were observed in literature117, 118, and the aforementioned structure was

47
attributed to the increased presence of Si-O bonds originating from the APTS-attachment on

graphene.

Figure 3.4a displays the nitrogen K-edge evolution, where a strong * absorption is

observed at ~399 eV and a structure-less * absorption of triangular shape is observed at ~410

eV.

a) b)
Intensity, a.u

Intensity, a.u
f-TRGs
f-TRGs
f-TRGp

f-TRGp

380 400 420 440 460 480 500 90 110 130 150 170 190
Energy Loss, eV Energy Loss, eV
Figure 3.4 Nitrogen K-edge a) and Silicon L2-3-edge b)

The *-transition is usually attributed to unsaturated C-N bonds119. Since this type of

bonds is not present in the silane molecule, the peak at ~399 eV can be attributed to a possible

loose attachment of the amine part of silane on the graphene surface in agreement with FTIR and

XPS spectra. Regarding the * absorption, this can be attributed in the Si-N bond present in

APTS. In the case of silicon L-edge (Figure 3.4b), the EELS spectra from the functionalized

samples are very similar to the spectrum of silicon oxide120. Peak at 99 eV corresponds to

elemental silicon (Si), at 105 eV to Si-N bond (Si1+) and peaks at 108 and 116 eV to Si-O (Si2+)

and Si=O (Si4+) bonds respectively121, 122. Regarding Si L1 edge at ~155eV, strong overlapping

48
with the threefold L2-3 edge prevents the acquisition of any useful information. Regardless,

silicon EELS strongly supports the presence of silane grafted on graphene.

The effect of APTS functionalization and the degree of exfoliation in TRG was also

examined by the Raman spectroscopy (Figure 3.5), which is known for its unambiguous

nondestructive identification123. For graphitic materials, the typical Raman bands are: a defect-

induced D band at 1350 cm-1, an in-plane vibration of sp2 carbon at 1580 cm-1 (G band), and a

two-phonon double-resonance process at ca. 2700 cm-1 (2D band).

The G-band position increases with decreasing number of layers in graphene124. Herein,

the G band centred at 1770 cm-1 for TRG has been upshifted to 1778 cm-1 on functionalization

(Figure 3.5a), thus revealing the decrease in the layer stacking125, mainly due to the presence of

APTS. Moreover, compared to TRG (Figure 3.5b), the characteristic 2D band for f-TRGs and f-

TRGP is observed to be blue shifted by about 28 cm-1 and accompanied with peak broadening,

confirming that the functionalization lead to better exfoliation of TRG and formation of few-

layer graphene flakes126.

Additionally, the appearance of D peak at 1339 cm-1 has been attributed to the presence

of structural disorder127. The D band arises from the activation in the first order scattering

process of sp3 carbons in graphene, and the intensity ratio of D and G bands (ID/IG) expresses the

sp2/sp3 carbon ratio, a measure of the extent of disorder128. The ID/IG for f-TRGS and f-TRGP has

increased from 0.677 for TRG to 0.903 and 1.026, respectively. The increased ID/IG ratio is

consistent with the functionalization of graphene through covalent bonding57, 129


and also in

agreement with the EELS results. In addition, the appearance of a new peak at lower frequencies

between 500-600 cm-1 is ascribed to the Si grafting on TRG 130


. Hence, overall the Raman

49
findings confirm the successful functionalization of TRG and the exfoliation state of TRG

remains unaffected by APTS attachment.

G-Band
D-Band
b)
Raman Intensity (a.u.)

a)

Raman Intensity (a.u.)


Si f-TRGs
f-TRGp
f-TRGp
Si
f-TRGs

TRG TRG

2500 2600 2700 2800


100 600 1100 1600
Frequency, cm-1 Frequency, cm -1
Figure 3.5 Raman spectra of TRG samples a) D and G band and b) βG (Gˈ) band

3.3.2 Morphological Analysis

TEM and EELS analyses are performed in order to understand the effect of functional

moieties on the morphology of graphene and to define the bonding between APTS and TRG

which can arise due the presence or absence of organic solvent. Pure TRG is composed of very

thin sheets, ranging in thickness between 3 and 6 nm and in lateral dimension from 0.5 to 1.5 m

(Figure 3.6).

These thin sheets are composed of overlapping graphene layers; the sheets are well

defined and separated. The inset in Figure 3.6a shows a selected area electron diffraction

(SAED) pattern collected from the center of the image, which verifies the crystallographic

identity of graphene.

High resolution TEM imaging (HRTEM) performed in areas where graphene was

“folded”, showed the thickness of the sheets ranges on between β.4 and θ.η nm corresponding to

50
number of overlapping layers between 4 and 10 (Figure 3.6b); based on graphene inter-planar

distance of 0.34 nm and sheet thickness of 0.314 nm). Finally, EDS results indicate the presence

of a small amount of chlorine impurity (< 0.5%), probably originating from the HCl washing

(Appendix-A).

a) b)
Figure 3.6 TEM images of pure TRG. (a) Graphene particle composed of overlapping graphene
sheers (inset: SAED pattern from the center of the graphene particle), (b) HRTEM image of two
overlapping graphene sheets

Although f-TRGp exhibits a structure similar to that of pure TRG, however, the sheets do

not appear as well separated as were in the pure TRG (Figure 3.7). The silane creates defects in

the graphene sheets (tears on the graphene surface) (see Figure 3.7a for f-TRGp). On the other

hand, no considerable change in the morphology of graphene sheets was observed for f-TRGs.

Instead, the sheets are very well defined and separated, (Figure 3.7b). The solvent appears to

soften the severe attack of APTS on graphene and hinders breaking down of sheets to smaller

ones. We infer that using organic solvent during the course of silane functionalization reaction

helps in maintaining the intrinsic morphology of graphene. The EDS analyses (Appendix-A)

from the two samples were in good agreement with the XPS elemental analysis: on average Si/N

51
ratio is found to be equal to 1.2 for f-TRGp, and 1.3 for f-TRGs (average of EDS spectra

collected at various points for each sample).

Furthermore, in order to understand the preferable attachment sites for silane on TRG,

EFTEM elemental mapping was performed on both f-TRGp and f-TRGS. To the best of our

knowledge, there is currently no publication on EFTEM mapping of the functionalized graphene

to determine the functionalization selectivity. However, the EFTEM elemental mapping,

completed here (Figure 3.8), can be a simple and versatile method for studying the local

functionalization. Bright spots in Figure 3.8(b, d) represent Si from EFTEM Si-elemental

mapping of Figure 3.8(a, c).

a) b)

`
Figure 3.7 TEM images of functionalized graphene: (a) f-TRGp, (b) f-TRGs

It is interesting to note that, in f-TRGP, Si (and N which is not shown here) is distributed

over the entire graphene surface (Figure 3.8b) representing a homogeneous distribution of the

grafted silane on f-TRGP surface. Thus, we infer that pure silane reacted homogeneously on the

graphene surface. A small amount of localized Si is also observed on the sheet edge. Thus, the

mode of silane grafting, when TRG is reacted with pure APTS, is with the surface functional

52
groups as well as the functional groups on the sheet edges. On the other hand, in f-TRGS, the Si

is observed to localize on the sheet edges. The homogeneous surface distribution of Si was not

observed as was seen in f-TRGP. Thus, it appears that in the presence of organic solvents like

toluene, the edge-functionalization of TRG can be the preferable mode of grafting. The

selectivity of edge-functionalization might further be increased by choosing appropriate solvents

and reaction conditions. In summary, solvent can play a vital role in the selective

functionalization of TRG. Further studies will evolve better control on its chemistry for selective

applications of silane functionalized TRG.

a) b)

c) d)
Figure 3.8 TEM bright field image (a) and corresponding EFTEM maps of silicon (b) for f-
TRGp; TEM bright field image (c) and corresponding EFTEM maps of silicon (d) for f-TRGs.

53
Based on the observations presented above, proposed reaction schemes for APTS with

TRG are elucidated (Scheme 1). The TRG oxy-functionalities are distributed over the surface

and sides of the sheets where the carboxylic groups are mainly located near the edges and most

of hydroxyl and epoxy groups are distributed over the surface. In f-TRGP, APTS reacts with

most of the surface groups and there is less reaction with the side carboxylic groups. In the case

of f-TRGS, the APTS molecules form covalent bonds with the carboxylic groups present on the

sides of TRG sheets. The use of various solvents assisting silane functionalization is not

uncommon. Toluene being a non-polar solvent with a dipole moment of 0.36 D has induced

significant effects on the functionalization selectivity. Owing to the non-polarity of toluene and

TRG being polar due to the presence of oxy-functional groups on its surface, the solvent helps

assisting preferable distribution of silane on TRG surface; leading silane molecules to more

reactive areas such as the edges. On the other hand, in the absence of any solvent, the silane

(being polar) is chemically adsorbed throughout the surface and edges. Increasing solvent

polarity in silylation reaction has already been proved to decrease the rate of silane reaction with

inorganic surfaces131, 132. Thus, using different non-polar solvents such as cyclohexane, benzene,

diethyl ether, chloroform,…etc can lead to discovering interesting aspects of effective solvents

for selective silane functionalization. Nevertheless, the probability and further quantification of

the attached functional groups depending on other solvents is beyond the scope of this article.

3.3.3 Thermal Stability and Conductivity

The thermal characterization of the pure and f-TRG by TGA provides a reliable

quantification of the relative amount of the introduced functionalities. As shown in Figure 3.9,

both TRGP and TRGS have lower thermal stability compared to the pure TRG.

54
NH2
OH

O
H2N
Si
OH
OH
O

Si O O
HO
OH
O
OH HO NH
OH
H2N O

No solvent OH Si
OH

HO O
APTS Si O
O
OH
OH
OH

HO O
OH NH2 HO
O Si
H5C2 HO
HO O O
NH2
O OH Si OH
O OH
O Si
O H5C2 O OH

OH C2H5 f-TRGp
HO HO
O H2N
O
OH
HO Catalyst H2N
O OH

TRG OH
NH2

Si
O
HO OH
O O
HO Si
HO
H2N O OH

Solvent O
OH
OH

O
O
Si O
O
HO
OH O
HO O O
OH
Si OH O
HO OH
Si

H2N
f-TRGs
H2N

Scheme 1 Reaction of APTS with TRG. TRG contains carboxylic functionalities on the sides
whereas hydroxyl and epoxy groups are mainly distributed over the surface

The lower thermal stability of f-TRG samples compared to pure TRG is attributed to the

full or partial detachment of the APTS groups from the graphene surface. The observed weight

losses in the temperature range of 50-500° C are 16, 22, and 25% for TRG, f-TRGS, and f-TRGP,

respectively.

As discussed earlier XPS results indicated that f-TRGP contains 7.4 atomic% Si, which

corresponds to 15 wt.% APTS and f-TRGS contains 8 atomic% Si corresponding to 16 weight%

APTS. Therefore, about 40% of APTS in f-TRGP and 55% of APTS in f-TRGS is lost between

50 and 500° C, respectively. Moreover, the higher weight loss for f-TRGS compared to f-TRGP

might be due to the most of the attached silane groups on the edges (as observed from the

elemental mapping), and in f-TRGP, the silane moieties are distributed over the surface and on

the edges of the sheets. Thus, it can be inferred that the end-functional groups are thermally less

stable as compared to the surface distributed labile groups.

55
The electrical resistance of pure and functionalized TRG samples has been investigated

using a custom made conductivity setup. The functionalization, as expected, has appeared to

increase the resistance of graphene. The measured room temperature resistance of TRG of 42

ohms slightly increases to 50.8 ohms for f-TRGP but increased significantly to 230 ohms for f-

TRGS. The increased resistance (decrease in conductivity) is expected due to the increase in sp3-

contents in the f-TRGS as indicated by ID/IG ratio in the Raman spectra (see Figure 3.5). The

edge distributed APTS in f-TRGS seems to have a more pronounced effect on the conductivity

compared to surface distributed groups in f-TRGP. This may be due to the effect of the group

distribution on the dispersion of sp2-electronic cloud of TRG.

100
Remaining Weight, %

90

80

70

60 Pure TRG
f-TRGp
f-TRGs

50
50 100 150 200 250 300 350 400 450 500
Temperature, C

Figure 3.9 Thermograms of TRG, f-TRGp, and f-TRGs. Samples were heated from 50-500° C at
10° /min under nitrogen

Conclusion

A simple catalyst-free method for synthesizing APTS-functionalized graphene was

successfully demonstrated along with the effective role of organic solvent in directing the

functionalities over the graphene surface and to the edges. The amount of APTS attached onto

graphene was determined by XPS and TGA and the covalent linkages were confirmed by FTIR,

56
XPS, EELS and Raman spectroscopy. TEM and EELS analyses show that no substantial change

in surface morphology of graphene was observed upon APTS functionalization in solution

whereas pure APTS has appeared to attack the graphene sheets adversely and changed the

morphology considerably.

The quantification of the f-TRG showed more than 7-8 atomic % attachment of Si, which

is higher than any previously reported functionalization. The method of APTS intercalation/

incorporation has a significant effect on the structure and properties of the resulting functional

material. Based on the EFTEM elemental mapping, it can be proposed that the selective end-

functionalization can be achieved using a suitable solvent during the course of the silylation

reaction. The f-TRG has shown good thermal stability over the temperature range of 50-500 °C

where only 16% weight loss is observed for pure TRG compared to 22% for TRGp and ~25%

for TRGs. The decrease in electrical conductivity (increase in electrical resistance) of TRG with

functionalization has been attributed to the increase in sp3-contents. Further studies can evolve a

better mechanism for selective functionalization of TRG with APTS.

Acknowledgements

We would like to thank Mr. Issam Ismail at the Department of Chemical Engineering and

Materials Science, University of Minnesota for the XPS measurements, Kelly Mason at National

Renewable Energy Laboratory (Golden, CO) for helping in the conductivity measurements, and

Dr. Sunil Lonkar at the Department of Chemical Engineering, The Petroleum Institute for his

valuable comments. This project has been supported by The Petroleum Institute in Abu Dhabi

through the Cooperative Research Partnership with Colorado School of Mines.

57
CHAPTER-4

SYNTHESIS AND CHARACTERIZATION OF POLYETHYLENE/OXIDIZED

POLYETHYLENE (PE/OPE) BLENDS AND ROLE OF OPE AS A VISCOSITY CONTROL

This chapter is modified from a paper published in the journal of

Applied Polymer Science1

Muhammad Z. Iqbal2, Ahmed A. Abdala3, Matthew W. Liberatore4

4.1 Introduction

Polymer blending is an economical way of designing new materials with unique

properties, such as high impact strength by leveraging the interactions between similar and

dissimilar polymer chains. The estimated consumption of polymer blends is more than 35% of
133
the total polymer consumption around the world and is continuing to increase . The properties

of pure polymers are important in predicting the resultant blend properties. Therefore, generating

new structure-property relationships between pure polymers and their blends can lead to new
133, 134
high performance materials . In general, the bulk properties of blends depend on its phase

morphology and interfacial structures.

Polymer blends are categorized as immiscible, miscible, or partially miscible. In

immiscible blends, one polymer can form a discrete phase dispersed in the other polymer

whereas in miscible blends, a single phase is observed. Also, partially miscible blend can exhibit
135
a single phase or multiple phases . Owing to the high molecular weight and small entropy of

1
Reproduced with permission Springer, (2016), 133, 43521-43532
2
Primary author and researcher
3
co-author and senior scientist at Qatar Environment and Energy Research Institute, Doha, Qatar
4
Author for correspondence and advisor

58
136
mixing, polymer blends usually exhibit immiscibility . Examples of commercial immiscible
137
blends include polystyrene with butadiene to produce high impact polystyrene or

poly(ethylene terephthalate) with poly(vinyl alcohol) for beverage bottles with better barrier

properties 138. Conversely, a small number of polymers shows miscibility when blended, such as
139 140
polystyrene with poly(phenylene oxide) , poly(vinyl chloride) with polycaprolactone , and
141, 142
poly(vinylidene fluoride) with poly(methyl methacrylate) . Usually, the miscibility is

observed when mixing polymers leads to molecular ordering. For example, when polystyrene is

blended with poly(phenylene oxide), the phenyl rings in both polymers overlap due to π-π

interactions creating miscible blends 139.

143, 144, 145, 146, 147


Polyolefins are used in packaging films , but due to their non-polar
148
nature, exhibit poor printability and dye-ability . Polyolefin blends with thermoplastics can

create new materials that impart polarity in polyolefins by mixing, allowing for dispersing fillers,

such as talc, carbon black, and graphene 148. Introducing a small amount of polar groups onto the
149 150
polymer backbone during polymerization or via post-polymerization process can

overcome the non-polar nature of polyolefins, but costs increase substantially. On the other hand,

blending polyolefin, such as linear low density polyethylene (PE), with another polar polymer to

prepare miscible blends can impart polarity in the blend matrix.

Oxidized polyethylene (OPE), commonly known as polymer wax, is often used as a


151, 152 153
processing aid or biodegradable PE . The OPE is synthesized by oxidizing PE by
154 155 156
thermal , radiation , or chemical methods . The oxidation process breaks the polymer

chains into segments of small molecular weight while introducing polar groups (e.g., acidic and
9, 157
ester groups) at the chain ends and along the chains . OPE is weaker and more brittle than

PE, but when mixed with PE, it can ease processing by decreasing viscosity. Durmus et al. 9 used

59
OPE as a compatibilizer in PE/clay nanocomposites and compared the properties of OPE

compatibilized nanocomposites with conventional PE-grafted-maleic anhydride compatibilized

nanocomposites. They reported ~120% reduction in oxygen permeability in the OPE

compatibilized nanocomposites at 15 wt% OPE loading, which was attributed to increased

exfoliation of clay. Also, OPE compatibilized nanocomposites improved barrier properties and

lowered the clay aspect ratio 8. Thus, PE/OPE blends show potential for packaging applications.

However, a more detailed understanding of the interactions of OPE with PE is needed to

engineer properties to create new products.

In this work, a comprehensive study of PE/OPE blends prepared by solution blending is

reported. The miscibility of the blend was determined by the Flory-Huggins parameter. The

rheological, mechanical, thermal and morphological properties of the blends were characterized.

Finally, the molecular morphology was correlated with the macroscopic properties.

4.2 Experimental

Details on synthesis procedures for PE/OPE blends are described below.

4.2.1 Materials

Linear low density polyethylene (LLDPE) with molecular weight of 3.6 × 106 g/mol

(determined via gel permeation chromatography) and bulk density of 0.97 g/cm3 was purchased

from Aldrich (428078, lot#07730MEV). Oxidized polyethylene (OPE) with an acid number of 7

and density of 0.94 g/cm3 (molecular weight ~ 1x103 g/mol) was supplied by Marcus Oil and

Chemicals (Texas). The Newtonian viscosity of LLDPE at 160°C is 1x104 Pa-s, and that of OPE

60
at 145°C is 0.02 Pa-s (viscosity versus temperature is provided in supporting information). The

solvent p-xylene (purity~99%, Sigma Aldrich, 134449) was used as received.

4.2.2 Preparation of Blends

PE and OPE at various compositions were dispersed in p-xylene (10 wt% solution) at

120°C for two hours. After the formation of homogenous and transparent solution, the solution

was drop cast on a heated glass plate at 120°C, and the solvent evaporated at that temperature for

two hours. The complete solvent removal was achieved by drying in a fume hood at ambient

conditions for 24 hours followed by drying for five hours at 80°C in a convection oven. The

samples were further vacuum dried overnight. The complete solvent removal was confirmed by

running thermal scans in a differential scanning calorimeter where no solvent evaporation peak

was observed. The dried samples were cut into pieces 3-5 mm wide, heated at 160°C for 10

minutes and hot pressed for 3 minutes under 5000 kPa pressure to prepare circular discs

(thickness ~ 1-1.3 mm and diameter ~ 25 mm) for rheological testing. Similarly, thin films (100-

β00 m) were prepared for mechanical and scattering testing by the same method using more

dilute solutions of PE and OPE in p-xylene and drop casting in a petri dish at 120°C and

following the same drying process.

4.2.3 Characterization

The thermal transitions of the blends were investigated using DSC (TA Instruments Q20)

. About 5-10 mg sample was heated under nitrogen flow (20 mL/min) at 10°C/min from RT to

160°C, isothermally heated for 3 minutes to remove any previous thermal history, and cooled at

10°C/min to room temperature in order to record the crystallization temperature, Tc. The melting

temperature, Tm was determined from the second heating ramp at 10°C/min from room

61
temperature to 160°C. For the determination of equilibrium melting point, after removing

thermal history, the sample was rapidly cooled to a fixed Tc at a rapid rate of 60°C/min to avoid

premature crystallization, kept at Tc for 30 minutes, cooled to room temperature at 10°C/min,

and finally heated above the melting point at 10°C/min to record the apparent melting associated

with each Tc.

Thermal stability of the blends was determined by, TGA (TA Instruments Q50).

Approximately, 10-20 mg of the sample was loaded in TGA sampling pan, heated from room

temperature to 650°C at 20°C/min under nitrogen flow (60 mL/min).

The rheological properties of the pure and blend samples were studied using TA

Instruments AR-G2 rheometer equipped with a 20 mm parallel plate. Time sweeps were

conducted at 160°C at 1% strain and 1 rad/s angular frequency. For time sweep, the samples

were carefully loaded so that less than 80 s elapsed before data collection began. Samples were

equilibrated at 180°C for 2 minutes followed by 5 seconds at 160°C under static conditions,

pressed between the plates. The blends were subjected to time sweep experiments for 60

minutes. The linear viscoelastic region was determined by running stress sweep at each

temperature at 1 Hz. The time-temperature superimposed master curves were constructed by

conducting frequency sweeps at each temperature within the linear viscoelastic range using the

horizontal shift factors.

The mechanical properties of blend films were determined following ASTM D882, using

ARES-G2 (TA) extensional rheometer with Film & Fiber tool. Measurements were carried at

room temperature under a speed of 0.0167 mm/s using thin rectangular samples ~10 mm long

62
and ~ 5 mm wide, stretched at 0.166 mm/s-1 at room temperature. The reported mechanical

properties are the average of 5 samples.

Small angle X-ray scattering (SAXS) and wide angle X-ray scattering (WAXS)

experiments were performed at the X-ray Sciences Division, beamline 12-ID-B, at the Advanced

Photon Source at Argonne National Laboratory. Details of the measurement procedure and the
158, 159
temperature and humidity control can be found elsewhere . Wide angle X-ray diffraction

(XRD) was performed using a Phillips PW γ040/θ0 spectrometer using CuKα radiation at a scan

rate of 0.02 °C/s. The samples for SAXS, WAXS, and XRD were prepared following procedure

similar to that used for mechanical testing samples.

Fourier Transformed infrared (FTIR) spectra (32 scans at 4 cm-1 resolution) were

collected using a Nicolet FTIR with ATR. The FTIR spectra were taken from 5 different spots

on thin films and averaged.

The micro-images of the blends were obtained using JEOL-JSM-7000F field emission

scanning electron microscope (FE-SEM) at an accelerating voltage of 3 KV. The blends were

cryo-fractured in liquid nitrogen and spur coated with gold for 1-1.5 minutes.

4.3 Results and Discussion

This section contains characterization of the PE/OPE blends synthesized in this thesis.

4.3.1 Miscibility, Phase Behavior and Thermal properties

Monitoring chemical bonds via FTIR spectroscopy provided the first qualitative measure

of the interactions between PE and OPE. Neat OPE showed specific vibrations originating from

63
various O-containing groups (Figure 4.1a). In the carbonyl range (1600-1800 cm-1), symmetrical

C=O stretching was observed at ~1715-1780 cm-1, including the ketonic stretching at 1715 cm-1,

C=O stretch in esters at ~1740 cm-1, and carboxylic acid C=O stretch at ~1780 cm-1. The region

in the frequency range of 1400-1500 cm-1 was attributed to the combination of symmetric

carboxylate stretch and C-H mode 160. CH2- stretching was observed at 1470 and 1460 cm-1. C-H

rocking absorption appeared at 1380 cm-1 and C-O stretch in tertiary alcohols observed at ~1150

cm-1. Thus, OPE structure contained carbonyl C=O and ketonic C=O along with alcoholic C-O

groups. The spectral regions of OPE were consistent with the structure of OPE reported

elsewhere 9.

The spectra of neat PE and PE/OPE blends were also obtained at various compositions

(Figure 4.1b). Since neat PE does not contain oxygen, no change in transmission was observed in

the carbonyl region. Increasing the OPE concentration decreased the IR-transmission through the

carbonyl peaks (at 1715 and 1740 cm-1) in the blends, which is consistent with the location of the

OPE functional groups. Averaging over 5 spots across the films, no changes in spectral patterns

were observed, so the distribution of OPE within PE matrix appeared homogeneous.

Thermal properties of polymer blends describe application behavior of blends. Increasing

OPE decreased the melting point and melt enthalpy of the blends (Figure 4.2). The melting

temperature decreased from 123°C for pure PE to 119°C for 50/50 blend whereas there was no

pattern observed for the melt enthalpy (ΔHexp). ΔHexp was in the range of 99 J/g for neat PE to

111 J/g for 50/50 blends. The melting profile of OPE showed three consecutive melting

transitions in the range of 98-105°C indicating presence of low molecular weight fractions in the

commercial τPE samples, and ΔHexp for OPE was 75 J/g (see Appendix-B). In addition, the

64
crystallization temperature, and crystallization enthalpy are also provided in supporting

information.

The melting profile for neat PE exhibited a peak melting point of 123 °C. As OPE was

added into PE, a small shoulder appeared in the vicinity of PE’s melting peak. Since this

shoulder was less prominent in neat PE, the immiscible adsorption of OPE chains containing

functional groups onto PE might be inferred. The small shoulder in PE melting profile indicated

significant short chain branches that crystallize separately from the backbone of PE. The small

separation between two melting peaks in neat PE was an indication that some chains do co-

crystallize 145. Some of the OPE chains when blended with PE acted as diluents whereas the part

that did not mix completely in PE might create the shoulder near the melting peak of PE. The

shoulder increased with increasing OPE concentration in the blends whereas the melting peak of

PE decreased as OPE concentration increased (Appendix-B). Due to the small separation

between PE and the shoulder component, it is reasonable to believe the formation of single

endotherm, which is indication of the miscibility between PE and OPE 161.

PE
Transmission, % (a.u.)

% Transmission

1780 1715 1415 1150


1380

1740
~1740

PE/OPE 0/100

a) 1470 1460
b)
~1715

1900 1800 1700 1600 1500 1400 1300 1200 1100 1760 1750 1740 1730 1720 1710 1700 1690
Wavenumbver, cm-1 Wavenumber, cm-1

Figure 4.1 FTIR spectra of OPE (a) and PE/OPE blends at various compositions

65
Furthermore, the percent crystallinity in PE varied significantly increased with increasing

the OPE loading. The percent bulk crystallinity was calculated by considering the enthalpy of

melting of a 100% crystalline PE (ΔH°PE) to be 293 J/g 162


by the formula crystallinity%

=ΔHexp/(ΔH°PE x φPE) x 100. The measured bulk crystallinity% was 34, 42, 42, 52, 56, and 76%

(±2%) for PE/OPE 100/0, 90/10, 80/20, 70/30, 60/40, and 50/50 blends, respectively. The

significant increase in crystallinity also indicates formation of high modulus materials.

The χ-parameter has been evaluated experimentally using melting point depression

following Hoffman-Weeks analysis (Figure 4.3) and Flory-Huggins theory (Figure 4.4). The
163
Hoffman-Weeks equation serves as the fundamental method of evaluating the equilibrium

melting temperature (Tmo). The theory assumes that 1) surface effects in formation of crystals are

negligible, 2) crystals at melting point exhibit equilibrium crystal perfections, and 3) the crystal

thickening is independent of the crystallization temperature. The Hoffman-Weeks equation is

expressed as:

 1 T
Tm  Tm o  1    c
  

The observed melting temperature (Tm) was measured after isothermal crystallization at

various crystallization temperatures (Tc).

A plot between Tm and Tc gives a line and the point where the line intersects the Tm=Tc

line, gives Tm° for each blend composition. In the above equation, is the thickening coefficient

equal to lc/lg* where lc is the thickness of grown crystal and lg* is the initial thickness of a chain-

folded lamellar crystal 163.

66
140
129

127 120

125

∆Hm (J/g)
100

T m ( C)
123

121 80

119
60
117

115 40
0 10 20 30 40 50 60
OPE wt%
Figure 4.2 Melting temperature and melt enthalpy of the blends

The Tmo of neat PE and in blends was determined (Figure 4.3a). A decrease in Tmo of PE

was observed with increasing OPE concentration in PE/OPE blends, an indication of the
141, 164
miscibility . The Tmo decreased from 129°C for pure PE to 124°C for the 50/50 blend

(Figure 4.3b). The decrease in Tmo is caused by the morphological and thermodynamic effects. In

terms of the thermodynamic effects, OPE acted as a miscible diluent that decreased the chemical

potential of a crystalline polymer, which in turn decreased Tmo value 164, 165.

130
129
(a) PE/OPE 129 (b)
80/20
128
126
127
Tm ( C)
T m ( C)

Tmo 126
123
125
120 124
123
117
Tm=Tc Line 122
114 121
114 117 120 123 126 129 0 10 20 30 40 50

Tc ( C) OPE wt%

Figure 4.3 Hoffman-Weeks plot of PE/OPE blends (a) and Tm° versus OPE concentration (b).

67
Thermodynamically, the chemical potential of a crystallizable polymer decreases with

addition of miscible diluents164. The decrease in chemical potential leads to reduction in Tm°,

which is caused by the morphological and thermodynamic effects. Considering only the
166
thermodynamic effects for Tm° depression, the Flory-Huggins theory modified by Nishi-

Wang 141 can be used to determine the Flory-Huggins interaction parameter, χ as followsμ

  R V2  ln 2  1 1  
 o  o        1  1 
1 1
 Tm ( pure ) Tm ( blend )  H V1  m2  m2 m1  
2
o

Here, the subscripts 1 and 2 refer to OPE and PE, respectively. Tmo(pure) and Tmo(blend) are

the equilibrium melting points of neat PE and that of PE in the blends, respectively. V 1 and V2

are the molar volumes of the repeating units of the components, R is universal gas constant, ΔH o

is the heat of fusion of a perfectly crystallizable polymer, m1 and m2 are the degrees of

polymerization, ϕ is the volume fraction of the OPE in the blend and χ is the polymer-polymer

interaction parameter (Flory-Huggins interaction parameter).

The ΔHo value for PE is 7870 J/mol 167. Molar volumes of PE and OPE repeat units were
168
calculated by the group contribution method . The calculated molar volume of PE is V1=32.2

cm3/mol. The molar volume of OPE was calculated based on the structure reported by Durmus et
9
al. where OPE structure contained ketonic and carboxylic groups. For calculating the molar

volume, two CH2- groups, one ketonic group (-CO-) and one -O- group were considered whereas

the end carboxylic functional groups (-COOH) were ignored. Thus, V2 was calculated to be 46.8

cm3/mol. Since both the components are polymers with large values of m 1 and m2, these terms

containing ‘m’ can be neglected. Thus, the simplified version of the Flory-Huggins theory is:

68
H o V1  1 
 o  o   1
R V2  Tm ( blend ) Tm ( pure ) 
1 2

A plot between the left hand side of the above equation and (1-ϕ2)2 should give a line

with an intercept of zero, whereas the χ can be obtained from the slope of the line. The χ

obtained, thus, is a single value, averaged over all the compositions. However, a critical analysis

of above equation by Rostami 169


shows that this χ is actually composition dependent (Figure

4.4). A highly negative value of χ indicated miscibility which decreased as OPE concentration

increased in the blends. Specifically, PE/OPE 90/10 exhibited χ = -2.3 that jumped to -0.6 for

80/20 composition, and was ~0 for 50/50 blend. The composition dependent χ has also been
164
found for polyacrylamide-poly(ethylene glycol) system and the blends of poly (ethylene

oxide) with different polymers 170


. Moreover, a negative value of χ indicates miscibility 168

whereas positive χ shows incompatible blends.

0.5
Interaction Parameter, χ

-0.5

-1

-1.5

-2

-2.5
0 10 20 30 40 50 60
OPE weight %

Figure 4.4 Composition dependence of the Flory-Huggins interaction parameter, χ

69
The pure PE showed a high thermal stability, losses 25% weight at 454°C, and

decomposes completely at 500°C. However, pure OPE exhibited low thermal stability, with

weight loss starting around 200°C, 25% weight loss by 389°C, and complete decomposition at

490°C. The thermal stability of the blend decreased with the inclusion of OPE (Figure 4.5).

Since miscibility produces blends with properties in between the two pure components, the

thermal stability of PE/OPE blends was in between that of neat PE and OPE. Due to the lower

thermal stability of OPE, the blends started losing weight at about 200°C and limited the

experimental window for the rheological measurements. Therefore, the TTS master curves were

constructed from 130°C (~10°C above the melting point) to 180°C.

In addition, SAXS and WAXS explored the effect of OPE inclusion on PE structure.
171
SAXS measures the shape and size of polymers between 5 and 25 nm of the repeat distances .

The scattering vector (q) probes different structural features at various length scales (length scale

~ βπ/q) in a polymer. Increase in power law decaying exponent (log(I) ~ log(q) -α) at lower q

indicates a larger size of scattering objects. A power law decaying exponent of α = β.γ for neat
172
PE (Figure 4.6a) indicated the presence of randomly branched ideal polymer chains in PE .

Increasing exponent reflect the scattering from the mass fractals in polymers. A small increase of

exponent to α = β.4 for 80/β0 blend showed a slight increase in mass fractals in blends, whereas

an exponent of α = β.8 for 70/γ0 compositions indicated a transition of morphology from ideal
173
polymer chains to larger mass fractals . A maximum exponent of 2.9 was observed for 50/50

blend, which showed stabilization of morphology between 30 to 50% OPE concentration.

70
100 100/0
90/10
Experimental 80/20

Weight Retained, %
80 Window 70/30
60/40
50/50
0/100
60

40

20

0
100 200 300 400 500 600
Temperature, °C
Figure 4.5 Thermal stability of PE/OPE blends

The Kratky-SAXS profiles (also known as Lorentzian corrected plots) of the PE/OPE

blends showed a maxima in the low q region (Figure 4.6b). This maxima is associated with the

periodicity resulting from the presence of macro-lattices formed by the centers of adjacent
174
lamellae . The arrangement of molecular chains in folded, unfolded, or partially unfolded
175
lamellae can also be obtained from a Kratky profile . Here, the long period, also called as

weight-average value of the long period, Lp, was calculated from the maxima intensity peak in

the Kratky plot. Lp decreased substantially from 180 Å for neat PE to 136 Å for 50/50 blends.

The decrease in Lp is also attributed to the increase in microcrystallinity in the PE/OPE blends as

confirmed by thermal analysis. This effect is further discussed in the mechanical properties

section.

PE showed two distinctive peaks in the high q-range in WAXS profile at 1.49 Å-1 and

1.64 Å-1 (Figure 4.7). The small domain size (called d-spacing hereafter) associated with these

peaks was calculated from d = βπ/q. A d-spacing of 4.2 Å and 3.8 Å for peaks were calculated.

71
Interestingly, the inclusion of OPE in PE did not change the d-spacing in blends, and the blends

exhibited a constant peak ratio of 1.1 across all compositions. Since, it was not possible to

perform WAXS on OPE because it does not form a film at its own, the crystal structure of OPE

was obtained through x-ray diffraction (XRD) (see Appendix-B).

PE showed two distinctive peaks in the high q-range in WAXS profile at 1.49 Å-1 and

1.64 Å-1 (Figure 4.7). The small domain size (called d-spacing hereafter) associated with these

peaks was calculated from d = βπ/q. A d-spacing of 4.2 Å and 3.8 Å for peaks were calculated.

Interestingly, the inclusion of OPE in PE did not change the d-spacing in blends, and the blends

exhibited a constant peak ratio of 1.1 across all compositions. Since, it was not possible to

perform WAXS on OPE because it does not form a film at its own, the crystal structure of OPE

was obtained through x-ray diffraction (XRD) (see Appendix-B).

The PE/OPE blends exhibited almost similar XRD patterns as were observed in neat OPE

or PE (also observed in WAXS). A nearly constant positions of crystalline peaks, and d-spacing

indicates that the PE and OPE crystals are not different from each other, and the possibility of

co-crystallization of PE and OPE. Further studies on the co-crystallization behavior of PE/OPE

blends under isothermal and non-isothermal blends could identify the nature and type of the

formed crystals, but is outside the scope of this work.

Based on thermal and scattering results discussed above, the structure of PE and PE/OPE

blends can be summarized as follows. PE consisted of lamellar structure containing the long

period order of ~ 180Å for neat PE, which reduced to ~136Å at 50 wt% OPE loading. The

substantial decrease in long period is directly associated with the increase in crystallinity of PE

with OPE addition.

72
(a) (b)

/0
-2.3

100
I(q) q , (a.u.)

/10
90
-2.4

20
80/
70/
-2.8 30

2
0
/4
-2.9 60
/20
50

0.2 0.4 0.6 0.8 -1


1.0x10
q, (1/Å)

200
Long period order, Lp, Å

(c)
180

160

140

120

100

Figure 4.6 log-log I vs q (a) and Kratky plots of PE/OPE blends (b) at 60°C and 0% relative
humidity with Y-offset, and long period order for blends (c)

The size of crystalline regions within the crystallites (d-spacing) remained unchanged,

which indicates that OPE chains could be crystallizing with PE to produce similar crystals.

Since, no quantitative argument could be made about the change in crystal size from WAXS, the

increased crystallinity indicates that both amorphous and crystalline fractions of the long period

are affected with OPE addition. This compact structure should exhibit higher modulus and

increased brittleness compared to neat PE, which will be discussed in the mechanical properties

section.

73
100/0

Intensity (a.u.)
90/10

80/20

70/30

60/40

50/50

1.3 1.4 1.5 1.6 1.7 1.8


-1
q, Å
Figure 4.7 WAXS patterns of PE/OPE blends with Y-offset

4.3.2 Melt Rheological Analysis

One important factor in designing new polymer blends is understanding their processing

properties. Melt rheology is one of the most useful techniques to assess the processing behavior

of polymers in the molten state to mimic the real time processing issues. Herein, time-dependent

rheological properties were measured to understand the blends’ thermal stability under

processing conditions and time-temperature master curves were constructed to observe the chain

dynamics over extended time scales.

The evolution of start-up transient rheological parameters is a common practice to

understand thermal stability of polymers. Here, the evolution of Gˊ was monitored with time at

160°C at constant frequency and strain. A slight increase in Gˊ was observed with time for neat

PE (Figure 4.8a). The 90/10 and 80/20 PE/OPE blends showed a similar small increase up to

~1000 s followed by a small level of thinning. At higher OPE concentration; the blends exhibited

a stronger thinning starting below 1000 s. The thinning effect was more pronounced when
74
reduced shear modulus, Grˊ (where Grˊ=Gˊ(t)/Gˊ(t=0)) was plotted against time (Figure 4.8b).

Neat PE showed about 20% increase in Grˊ and the increment decreased for 90/10 blend

followed by pure thixotropy with increasing OPE concentration in the blends. Thus, OPE acted

as a plasticizer. The critical time (time at which rheological properties such as Gˊ, G˝ and *

drop significantly) decreased with increasing OPE concentration in blends. The power law

exponent was evaluated as Gˊ~t where Gˊ was considered in the thinning regions only. The

evaluated shear thinning exponent, , values were -0.18, -0.30, -0.38, -0.55, -0.82, -0.85 for

100/0, 90/10, 80/20, 70/30, 60/40 and 50/50 blends, respectively. As increased, the blends

showed a stronger shear thinning behavior. In subsequent sections, all samples were equilibrated

at 180° C for 2 minutes followed by equilibrating for 1 minute at a specific temperature before

carrying out any measurement. The experiments were completed within 350 s at all temperatures

before the overshoot or shear thinning exceeded more than ±15%.

The plasticization effect of OPE on PE is more evident in Figure 4.9a. Significant

decrease in complex viscosity was observed with increasing OPE amount. For example, neat PE

showed a viscosity of 10,000 Pa-s (at 0.01 Hz) that was reduced by more than one order of

magnitude to 700 Pa-s for 50/50 blend.

The zero shear viscosity (ηo) of the blends at various temperatures is determined by

fitting the complex viscosity data with Cross model 176, which will be compared next.

1    
o
 ( )  1n
o

75
Where, o is the zero shear viscosity, o is the relaxation time related to the longest

relaxation time and n is the Cross-exponent. The Cross fitting parameters and fitting profiles are

given in the supporting information.

1.6
(a) (b)
10
4
PE/OPE 100/0 1.4
PE/OPE 100/0

Gr (G(t)/G(t=0)
90/10 1.2
3
10 80/20 90/10
1.0
G, Pa

70/30 0.8
2
80/20
10 60/40
0.6
50/50
70/30
10
1 0.4 50/50
60/40
0.2
0
10 1 2 3 4 10
1 2
10
3
10
4
10
10 10 10 10
Time, s Time, s

Figure 4.8 Start-up transient rheology of neat PE and PE/τPE blendsμ Evolution of Gˊ with
time (a), and reduced Gˊ versus time. Data reported at T = 1θ0°C, = 1 %, ω = 1rad/s.

The o decreased with increasing temperature and increasing OPE concentration, which

was consistent with a plasticization effect of OPE. The significant viscosity decrease might allow

little to no change in the processing condition of PE nanocomposites compared to neat PE.

Specifically, a substantial amount of nanofillers are needed to induce percolation within the non-
177
polar PE matrix . At high filler loadings, viscosity increases several orders of magnitude and

may cause processing issues. The inclusion of OPE in PE can be used to tune the viscosity of PE

nanocomposites. More importantly, the shear thinning behavior of PE remains unperturbed by

the incorporation of OPE in PE.

The zero shear viscosity, o changed with OPE loading in the blends at two experimental

temperatures tested, i.e. 180 °C and 130 °C (Figure 4.9-b). Small increase in ηo values (~70%)

with a 50°C change in temperature from 180 to 130° C indicated temperature insensitivity of the

76
blends within the accessible temperature range. A linear relationship between log- o and OPE

weight% (Figure 4.9-b) also showed the melt miscibility of OPE in PE 143, 144, 178, 179.

It is worth mentioning that the morphology of miscible blends is also considered as a

function of viscosity ratio and capillary number 180. Since, the viscosity of OPE at 145°C (~0.02

Pa-s, from Arrhenius relation) (see Appendix-B), the viscosity ratio (p = d/ m, where d is the

viscosity of dispersed phase, and m is the viscosity of matrix) becomes extremely low (p ~ 2x10-
6
). At such low viscosity ratios (p << 1), the droplet diameter essentially remains unchanged, and
180, 181, 182
will have no effect on the blend morphology and droplet break-up processes . In

addition, very low p has been shown to require a longer mixing time (~30-60 minutes) to form a
183
consolidated polymer morphology . However, above 60 minutes, miscible blends of any low

viscosity ratio (p ~ 10-6) should produce a stable morphology, which also applies in this study

(mixing time ~ 120 minutes).

Time-temperature superposition (TTS) provided master curves for Gˊ, G˝ for the linear

viscoelastic response of the nanocomposites (Figure 4.10-a,b). The validity of TTS was tested by

plotting dynamic moduli according to Han’s proposed method 184


, which showed only minor

fluctuations in the master curves of the blends. The Han’s plot and individual Gˊ~ƒ master curves

are provided in the Appendix-B. Due to the limitations of experimentally accessible temperatures

and temperature insensitivity of blends, the TTS could not be extrapolated over more than 200 s-1

frequency. Therefore, the viscoelastic data had been shifted horizontally only. The calculated

horizontal shift factor, a from the best fit was also plotted versus temperature, and is provided in
T

Appendix-B.

77
5
4
10
10

130°C
*, Pa-s

o, Pa-s
10
3
10

100/0 3 180°C
90/10
10
80/20
2
70/30
10 60/40
50/50
2
10
, s
-2 -1 0 1 2
10 10 10 10 10 0 10 20 30 40 50
-1
OPE wt%

Figure 4.9 Complex viscosity profiles of PE/OPE blends at 160°C. The black solid lines are the
Cross model fit (left panel) and zero shear viscosity as a function of OPE wt.% at two different
temperatures (right panel)

σo Gˊ, G˝ plateau was observed within the experimental range. In the low frequency

region, Gˊ and G˝ were expected to show εaxwell behavior (Gˊ~ƒ2, G˝~ƒ1). However, neat PE

exhibited an un-relaxed behavior at low frequencies (Gˊ~ƒ1.1), which was observed in similar

high molecular weight PE melts 8, 185


. The low frequency slope increased from Gˊ~ƒ1.1 for neat

PE to Gˊ~ƒ1.7 for 50/50 blend, which approached the relaxed Maxwell behavior attributed to the

plasticization effect of OPE. Similarly, at low frequencies, neat PE exhibited G˝~ƒ0.8 behavior

which increased to G˝~ƒ1 for 50/50 blends, which was an indication of formation of the relaxed

chains compare to the neat PE. Furthermore, the formation of relaxed polymer chains with

increasing τPE concentration was more evident in the master curves of over the reduced

frequencies (see Appendix-B). The delta values increased as OPE concentration increased,

approaching ~λ0° for the completely relaxed chains.

In the high frequency regions, PE showed Gˊ~ƒ0.5, G˝~ƒ0.3 behavior that changes to

Gˊ~ƒ0.8, G˝~ƒ0.7 for the 50/50 blend. In the glassy (high frequency) region, PE showed a Rouse

78
longest relaxation time with a general expression186 Gˊ~ƒ0.5. Thus, slopes near 0.5 for Gˊ in the

high frequency region indicated Rouse dynamics. Increase in high frequency slope, further

indicated that the chains did not respond as glassy chains in high frequency region, which

confirms the existence of rubbery chains in that region.

6 6
10 PE 100/0 10
(a) 90/10 (b) PE 100/0
90/10
5 80/20 5
10 10 80/20
4 4
10 10

G, Pa
70/30
G, Pa

3 3
10 10

low  slope
60/40
low  slope
70/30
2 2
10 10
60/40 100/0 1.1 50/50 100/0 0.8
1 1
10 90/10 1.0 10 90/10 0.8
50/50 80/20 1.1 80/20 0.8
0 70/30 1.2 0 70/30 0.9
10 60/40 1.2
10 60/40 0.9
50/50 1.7 50/50 1.0
-1 -1
10 10
aT  s aT  s
-1 0 1 2 -1 0 1 2
10 10 -1
10 10 10 10 10 10
-1

Figure 4.10 εaster curves for Gˊ (a), G˝ (b) over the reduced frequency. Insets in (a) and (b)
show slopes in the terminal regions (at low frequencies) averaged over all the temperatures
(circles 180°C, upper triangles 160°C, lower triangles 140°C, and squares 130°C).

4.3.3 Mechanical Properties

Another important factor that determines the practical applications of new polymeric

materials is their mechanical performance. Among mechanical properties, tensile properties are

considered more important in the initial design of new polymers. The inclusion of OPE reduced

the strain at break for the blends(Figure 4.11a). In agreement with SAXS (Figure 4.6), the

decrease in blends’ δp led to the formation of high Young’s modulus materials. The fraction of

OPE that decreased the melting point (see thermal analysis discussion earlier) also induced

plasticization. The plasticization effect of τPE was quantified by a decrease in Young’s modulus

from 1.6 MPa for neat PE to 0.94 MPa for 90/10 blends. Further inclusion of OPE chains

79
increased the Young modulus from 0.94 MPa for 90/10 blend to 2.0 for 50/50 blend (~210%

increase), which is a direct evidence of the reduced blend elasticity (Figure 4.11b) in agreement

with the decreased long period spacing (Figure 4.6). The change in Lp is also known to influence

the elongation at break and yielding strain. A direct consequence of the decreased Lp also

resulted in decreasing the elongation at break from 21.5 for PE to 0.18 for 50/50 blends (Figure

4.11d) 187, 188. The yield stress (Figure 4.11c) increased from 12 MPa for PE to 16 MPa for 60/40

blends (~74% increase) whereas the yield strain decreased ~90% for 60/40 blends (Figure

4.11d). In conclusion, OPE in PE decreased the elongation to break for the blends and increased

the Young’s modulus, consistent with the findings from scattering and thermal analysis.

4.3.4 SEM Morphology

The cryo-fractured surface morphology of neat PE shows a smooth surface upon fracture,
189
which is expected for PE (Figure 4.12a) . When OPE is added into PE and the sample is

fractured, some fibrous structures appear to be pulled out from the fractured surfaces (Figure

4.12b). Denser fibers are observed for 80/20 (see Figure 4.12c) and 60/40 (Figure 4.12e).

However, 70/30 composition shows a distinct morphology, i.e., more abundant fibrous elements

(Figure 4.12d). At this time, we do not have a specific explanation for this behavior, but the

unique morphology might account for the higher elongation in the 70/30 blend compared with

90/10 or 80/20 blends. However, the scattering results also indicate this peculiar behavior with

70/30 composition. On the other hand, the 60/40 blends show smooth surface, which is

consistent with the reduced viscosity. The 50/50 blend represents a similar surface to the 60/40

blend, but due to the even lower viscosity of that composition, some larger fibers appear after

cryo-fracturing.

80
18 3.0

Young's Modulus, MPa


(a) (b)
15 50/50 2.5

100/0 2.0
Strees,M Pa

12

9
15
1.5

6 10 1.0
5
3 0.5
0
0.0 0.2 0.4
0 0.0
0.01 0.1 1 10 100 0 10 20 30 40 50
Strain OPE wt%
Yield and Break Stress, MPa

100

Yield and Break Strain


25 (d) Strain@ Break
(c) Stress@ Break Yield Strain
Yield Stress
20
10
15

10
1

0 0.1
0 10 20 30 40 50 0 10 20 30 40 50
OPE wt% OPE wt%

Figure 4.11 Mechanical properties of PE/OPE blends: (a) Stress-Strain behavior, (b) Young’s
modulus, (c) yield and break stress, and (d) yield and break strains. Arrows in stress-strain data
indicate the break points. The inset in (a) shows stress-strain data on a linear scale.

81
(a) Neat PE (b) 90/10

2μm 2μm

(c) 80/20 (d) 70/30

2μm 2μm

(e) 60/40 (f) 50/50

2μm 2μm

Figure 4.12 SEM micro-images of PE/OPE Blends

Conclusion

A new type of PE/OPE blends has been prepared and the molten and solid state

properties of the blends are evaluated at different OPE content. All blends exhibited a negative

82
Flory-Huggins interaction parameter indicating miscibility between PE and OPE that decrease

with increasing OPE loading. Increasing OPE concentration in the PE/OPE blends from 0 to 50

weight% decreases the melt temperature from 123 °C to 119 °C and also reduces the thermal

stability. The SAXS analysis revealed that the OPE resides inside PE lamellae, and the lamellar

thickness decreases with increasing OPE inclusion, leading to increased brittleness that is also

confirmed by the increase in modulus and decrease in the strain at break.

Time-temperature superposition master curves showed the formation of relaxed polymer

chains with increasing OPE loading and the blends exhibited positive deviation from the Rouse

dynamics at higher frequencies. Moreover, incorporation of OPE reduced the zero shear

viscosity of the blends suggesting that OPE can also be incorporated as a processing aid. The

fractured surface morphology revealed interesting results exhibiting smooth surface of blends

with increasing OPE loadings, which is in agreement with the viscosity reduction in blends. In

addition, this type of blends with controlled viscosity can also be used to create PE

nanocomposites with high filler loadings and can still be processed on neat PE processing

conditions. This topic will be the subject of a future manuscript.

Acknowledgements

This project was supported by The Petroleum Institute, Abu Dhabi through the

Cooperative Research Partnership with Colorado School of Mines. The U.S. Army Research

Office (DURIP Grant No.W911NF-11-1-0306) is also acknowledged for partial financial

support. The authors would also like to thank Prof. John Dorgan for use of his lab facilities, Tara

Pandey and Soenke Seifert for assisting in SAXS experiments. Marcus Oils and Chemicals

(Texas) are acknowledged for supplying the OPE. The use of the Advanced Photon Source, an

83
office of Science User facility operated for the U.S. Department of Energy (D.O.E.) Office of

Science by Argonne National Laboratory was supported by the U.S. DOE under Contract No.

DE-AC02-06CH11357.

84
CHAPTER-5

PROCESSABLE CONDUCTIVE GRAPHENE/POLYETHYLENE NANOCOMPOSITES:

EFFECTS OF GRAPHENE DISPERSION AND POLYETHYLENE BLENDING WITH

OXIDIZED POLYETHYLENE ON RHEOLOGY AND MICROSTRUCTURE

This chapter is modified from a paper submitted in the journal

Polymer1

Muhammad Z. Iqbal2, Ahmed A. Abdala3, Sӧnke Seifert4, Andrew M. Herring5, Matthew

W. Liberatore6

5.1 Introduction

Due to the widespread interest in nanotechnology and its applications, polymer

nanocomposites, which are polymers with nano-scale fillers, have shown enormous potential.

For example, carbon nanotubes (CNTs) and graphene28, 48, 190, 191 are promising, conductive nano-

fillers used for reinforcement of polymers as well as inducing electrical and thermal conductivity

in the nanocomposites.

Graphene is a 1-atom thick layer of sp2-hybridized carbon atoms, consisting of

honeycomb-like assembly of C-atoms. Outstanding properties of graphene/polymer

nanocomposites include mechanical enhancement192, improved electrical conductivit43, 48, 191


,

thermal conductivity99, 109, 193


, gas barrier 48
, and flame retardancy194, 195
. To achieve

1
Permission will be obtained after the paper will be published
2
Primary author and researcher
3
Co-author and senior scientist at Qatar Environment and Energy Research Institute, Doha, Qatar
4
Scientist at Advanced Photon Source, Argonne National Laboratory, Chicago, IL, USA
5
Co-author and professor at Colorado School of Mines, Golden, CO, USA
6
Author for correspondence and advisor

85
combinations of practical properties in graphene/polymer nanocomposites, a homogenous

dispersion of graphene in polymer matrix is essential. In order to obtain good graphene

dispersion, graphene agglomerates must be broken down to single layers during processing (e.g.,

solvent, melt blending, or in-situ polymerization), or by surface functionalization of graphene, or

altering chemistry of the polymer matrix. Regarding the microstructure of graphene in polymer

matrices, two open questions need to be addressed: 1) Does graphene dispersion follow

traditional theories, such as the Brownian diffusion; 2) What is the structure of the graphene

network inside polymer matrix?

Single layer graphene has exceptionally high modulus of ~1,000 GPa in tension

compared to ~1-10 GPa for polymers, and 10-800 GPa for other fillers28, 196. Such high strength

of graphene enables graphene/polymer nanocomposites to carry the applied load. Because of its

high surface area and sheet structure, graphene can, in principle, significantly alter the properties

of the polymer matrix, e.g. polymer crystalline morphology197, chain conformation198, and

dynamics through confinement effects199. Thus, many factors are responsible for the property

changes offered by graphene, and better understanding of the property improvements requires

additional studies of the behavior of graphene inside the polymer matrix.

Translating graphene’s unique properties to nanocomposites is difficult, since graphene is

known to poorly disperse in polyolefins, including polyethylene43 and polypropylene200, 201, due

to the nonpolar nature of polyolefins. Several groups are attempting to develop strategies to

homogenously disperse graphene in polyethylene, since polyethylene is one of the most widely

used commodity thermoplastics. For example, Kim et al. reported high electrical percolation

threshold of graphene (12-15 wt%) nanoplatelets in polyethylene202, which is significantly higher

than many polar polymers28. High percolation is directly associated with poor dispersion and

86
larger graphene aggregates in polyolefins43, 203. Improved dispersion of graphene in polyethylene

was reported using amine functionalized graphene204. About 150% increase in elastic modulus

was achieved at 3 wt% loading using functionalized graphene, compared to graphene oxide.

Introducing functional groups onto the polymer can improve dispersion of the nano-

fillers in polyolefins. Maleic anhydride-grafted polyethylene205, and chlorinated-polyethylene206,


207
were reported as compatibilizers for graphene/polyethylene nanocomposites. However,

transmission electron microscopy showed that chlorinated polyethylene did not reduce graphene

stacking in the nanocomposites. Functionalized polyethylene containing cyano- and amino-

functional groups, showed improved mechanical properties, but the electrical conductivity of

functionalized polyethylene/graphene nanocomposites was less than neat polyethylene/graphene

nanocomposites43.

Also, poor dispersion was reported when polyethylene chains were in-situ polymerized

on graphene surface208. Despite the goals of attaining better dispersion and high conductivity,

nanocomposites with 10 vol% graphene loading exhibited electrical conductivity of 10-4 S/cm,

which is a strikingly low conductivity at such high loading. However, using such high loading of

graphene in a nonpolar polymer also increases the viscosity by several orders of magnitude200,
209
, which limits processing operations. In a report, Pang et al.102 reported a very low electrical

percolation threshold (0.07 vol%) for graphene/polyethylene nanocomposites via formation of a

segregated structure, but other thermomechanical properties were not provided. Therefore,

polyolefin nanocomposites having reduced graphene stacks, and with controlled processing

properties is highly desirable.

87
Commonly used as a processing aid151, 152
or biodegradable polyethylene153, oxidized

polyethylene (OPE) also improved the barrier properties of polyethylene/clay nanocomposites9.

Leveraging our earlier work on blending OPE with PE to produce high modulus materials with

controlled processing properties210, we combined blends of polyethylene and OPE with graphene

as nano-filler to create nanocomposites. Overall, the purpose of this study is to better understand

the origin of the reinforcement observed in PE/OPE/graphene nanocomposites with improved

dispersion. Graphene loading was varied to understand the filler microstructure. Linking the

thermo-mechanical properties of nanocomposites with their microstructure using scattering will

also be detailed. In addition, graphene nanocomposites will be compared with carbon black

composites.

5.2 Experimental Details

This section describes the synthesis of TRG and its nanocomposites with PE/OPE blends.

5.2.1 Materials

Natural flake graphite (-10 mesh, 99.9% Alfa Aesar), carbon black(CB) (99.9%, bulk

density = 170-230 g/L, surface area = 75m2/g, Alfa Aesar), linear low density polyethylene

(LLDPE) with molecular weight 3.6 x 106 g/mol (determined via gel permeation

chromatography) and bulk density 0.97 g/cm3 (428078, lot#07730MEV, Aldrich), p-xylene

(99%, Sigma Aldrich), potassium permanganate (Fisher Scientific), sulfuric acid (95-97%, J.T.

Bakers), phosphoric acid (>99%, Aldrich), hydrogen peroxide (30% solution, BDH), HCl (37%,

Reidel-deHaen), were used as received. Oxidized polyethylene with an acid number of 7 and

density of 0.94 g/cm3 was supplied by Marcus Oil and Chemicals (Texas, USA).

88
5.2.2 Synthesis of TRG

TRG has been prepared by thermal exfoliation and reduction of graphite oxide (GO)37, 88,
110
. For synthesis of Gτ, Tour’s method211 was employed. Typically, 5 g of natural flake graphite

was dispersed in a mixture of H2SO4 (272 mL) and H3PO4 (33 mL), and stirred for 30 minutes.

About 27.8 g KMNO4 was gradually added into the mixture over the course of one hour to avoid

explosion. The mixture was stirred continuously using an overhead stirrer for three days at room

temperature. After the completion of reaction, H2O2 (30%, 17.5 mL) and deionized water (137.5

mL) were added to the reaction mixture, and stirred until the color of the mixture turned from

dark brown to bright yellowish, indicating high oxidation level of graphite. The GO was washed

three times with 1M HCl aqueous solution, and further repeatedly washed with deionized water

until a pH of 4-5 was obtained. The separation of the washed GO was carried out by a centrifuge

with a force of 10000 g followed by dialysis. The washed GO was dried under vacuum for two

days to remove any traces of water. The dried GO was thermally exfoliated and reduced by

heating rapidly at 1000° C for 30 s under nitrogen flow in a tube furnace to produce thermally

reduced graphene (TRG). The produced TRG was dried overnight under vacuum before using in

the composites.

5.2.3 Preparation of Composites

Low density polyethylene (PE) and OPE were dispersed in p-xylene at 120° C for one

hour under reflux. Meanwhile, TRG (or CB) was sonicated in p-xylene for one hour at room

temperature in a sonication bath. TRG dispersion was added into PE/OPE mixture, and solution

was further stirred for another hour to form a homogenous solution at 120° C. After the

formation of homogenous solution, the solution was drop cast on a heated glass plate at 80° C

89
and the solvent was allowed to evaporate for two hours. The complete solvent removal was

achieved by drying in a fume hood at ambient conditions for 24 h followed by drying in

convection oven at 80°C for 5 h. The complete solvent removal was confirmed by differential

scanning calorimeter, where no solvent peak was observed. The dried samples were cut into

pieces 3-5 mm wide, heated at 160°C for 10 min and hot pressed for 3 min under 5000 kPa

pressure to prepare circular discs (thickness 1-1.3 mm and diameter ~ 25 mm) for rheological

testing. Similarly, thin films (100-200 µm) for mechanical and scattering testing were prepared

by the same method using more dilute solutions of PE/OPE and TRG (or CB) in p-xylene, and

drop casting in a petri dish at 80°C, and following the same drying procedures.

For pure PE/OPE blends, the same procedure was used without TRG or CB. The PE/OPE

solution was stirred for 2 hours at 120°C under reflux before drop casting on a glass plate or in a

petri dish.

5.2.4 Characterization

Wide angle x-ray diffraction (WAXD) was performed on a Phillips PW 3040/60 with

CuKα radiation at a scanning rate of 0.0β °/s to study exfoliation of graphite. Transmission

electron microscope (TEM) images were obtained using a FEI Phillips C200 at 200 kV to

confirm the production of TRG nanosheets. For TEM measurements, a dilute dispersion of TRG

(0.1 mg/20 mL acetone) was prepared by bath sonication for 10 minutes, and one drop of

solution was deposited on a 300-mesh Cu grid with holy carbon. The micro-images of blends and

composites were obtained using JEOL-JSM-7000F field emission scanning electron microscope

(FE-SEM) at an accelerating voltage of 3 KV. The blends and composites were cryo-fractured in

liquid nitrogen, and spur coated with gold for 120 seconds. Functional groups on TRG surface

90
were analyzed by Fourier transform infra red spectrometer (FTIR) (NicoletTM iSTM 10, Thermo

scientific FTIR) with 32 scan at a resolution of 4 cm-1 (TRG in KBr). Very dilute dispersions of

CB (0.1 mg/mL) and TRG (0.025 mg/mL) were dispersed in acetone via bath sonication at room

temperature, and dynamic light scattering (DLS) was performed using particle size analyzer

(NanoBrook 90Plus, Brookhaven) at 25°C.

Rheological measurements of blends and composites were performed at 160°C using TA

Instrument AR-G2 rheometer equipped with 20 mm parallel plate. Tests were conducted in air

while samples were covered with an aluminum solvent trap to avoid temperature fluctuations

during the run. Linear viscoelastic (LVE) region was determined by conducting strain sweep in

0.01-100% strain range at 1 Hz frequency. The frequency sweep experiments were performed in

the frequency range of 0.01-100 Hz at fixed strain within LVE region.

Mechanical properties of thin films were determined following ASTM D882, using

ARES-G2 (TA Instruments) rheometer with Film & Fiber tool. Measurements were carried out

at room temperature at a stretching speed of 0.0167 mm/s using thin rectangular samples ~10

mm long and ~5 mm wide.

The in-plane electrical conductivity was calculated using electrochemical impedance

spectroscopy (EIS) using the following equation:

Conductivity 
L
R.w .t

Where, R is the film resistance, L is the distance between the sense electrodes, w is the

width of film, and t is the sample thickness. The impedance spectra were obtained over a

frequency range of 0.3 Hz to 100 kHz using a four-electrode test cell connected to a multi-

91
channel potentiostat (Biologic VMP3, Knoxville, TN). All measurements were carried in an

environmental chamber to control sample temperature and humidity (TestEquity Model 1007H,

Moorpark, CA). The samples were polished using a 100-mesh grit paper before conductivity

measurements in order to remove thin polymer layer transported to the surface during the film

formation and drying process. Due to polishing, TRG and CB composites exhibited static current

build up, which was removed by conditioning the samples at 95% relative humidity and 50 °C

for one hour. The results reported are averaged over 3 samples at 25 °C temperature and 95%

relative humidity.

Small angle X-ray scattering (SAXS) experiments were performed using X-ray Sciences

Division, beamline 12-ID-C, at the Advanced Photon Source at Argonne National Laboratory.

Data have been collected at 60°C under dry conditions, and is averaged over three exposures

times. Details of the procedure and temperature/humidity control can be found elsewhere212, 213.

5.3 Result and Discussion

This section describes the characterization of nanofillers and nanocomposites.

5.3.1 Characterization of TRG and CB

Rapid heating of GO produced a substantial increase in volume, and GO was exfoliated

into TRG. Typically, GO has a bulk density of 0.5 g/cm3 (1 g with a bulk volume of ~2 cm3),

while the TRG product has a bulk density of ~10 mg/cm3 (1 g TRG with a bulk volume of 100

cm3). An exfoliated sheet-like structure was observed when TRG was deposited on a TEM grid

from a very dilute solution (Figure 5.1a). A paper-like morphology of graphene nano-sheets were

observed, and the graphene sheets were overlapping, extended to a few hundred nanometers. The

92
actual graphene sheets are extremely transparent to the electron beam, however, the folded edges

appeared dark, as observed by others88. The CB particles are clusters of almost spherical

particles, forming fractal structures extending over 2 µm (Figure 5.1b). The size of the individual

CB particles was ~60-80 nm. Although the TEM images were taken from extremely dilute

suspension of CB in acetone, the particles still formed large aggregates. The size of these

aggregates might increase with higher CB concentration in polymers. Complete characterization

of TRG, and CB particles by XRD, TEM, FTIR, and DLS is provided in the Appendix-C.

(a) (b)

Figure 5.1 TEM image of TRG (a), and CB (b).

5.3.2 Melt Rheological Analysis

The rheological properties and their analysis is provided in this section.

(i) Rheology of Blends and TRG-Nanocomposites

Melt rheology of polymers and polymer nanocomposites is essential for understanding

their processing and developing structure-property relationships. Viscoelastic (VE) response can

be connected to the intrinsic behavior of polymers, inter-particle, and polymer-particle

interaction. In general, rheological properties depend on the input strain, whether in the linear or

non-linear viscoelastic region. The dynamic strain sweep determines the strain dependent

93
viscoelastic response and the linear viscoelastic (LVE) region. The addition of OPE in PE led to

the formation of miscible blends, where OPE resided within the amorphous region of the long

period domains of PE210. Due to miscibility and plasticization, OPE decreased the complex

viscosity of PE (complex viscosity at 0.01 Hz, Figure 5.2). An approximately 7-fold decrease in

viscosity was observed for 60/40 blend compared to neat PE. This lower viscosity can be very

advantageous when processing polyolefin nanocomposites that require high filler loadings. The

complete rheological, thermal, and structural properties of PE/OPE blends can be found in our

previous work210. Also the complete rheological properties of PE/OPE blends are provided in

Appendix-C.

12000
f = 0.01 Hz
10000
η* (Pa-s)

8000

6000

4000

2000

0
PE 80/20 60/40
Figure 5.2 Processing control by PE/OPE blends

TRG-filled PE, PE/OPE 80/20, and 60/40 blends exhibited different rheological

responses at various TRG loadings (Figure 5.3). Adding TRG to PE increased Gˊ at all TRG

loadings (Figure 5.3a). With increasing TRG, an exponential increase in Gˊ was observed. Also,

the low frequency slope of Gˊ decreased, leading to asymptotic behavior that is nearly

independent of frequency. This asymptotic behavior indicates the development of a concentrated,

percolating network214. A similar trend was observed for 80/20/TRG nanocomposites (Figure

94
5.3b). Increasing TRG loading increased G' of 80/20/TRG nanocomposites, leading to

exponential growth in G' with increased loading.

A peculiar rheological behaviour was observed for 60/40/TRG nanocomposites at low

TRG loadings (Figure 5.3c). For TRG < 1 wt%, Gˊ decreased with TRG addition compared to

the unfilled blend. At higher TGR loadings, an exponential increase in G′ was observed, similar

to the PE/TRG and 80/β0/TRG nanocomposites. For θ0/40 blends, the decrease in Gˊ for dilute

TRG loadings might be attributed to the interlayer slipperiness between TRG and polymer layers

due to a low surface friction, and possible exfoliation of TRG sheets209. Exfoliation of TRG into

smaller stacks might lead to sheet alignment, which could decrease the viscoelastic response

from the nanocomposites. Further details of this slippery behaviour will be discussed in

conjunction with scattering results later.

The viscosity of nanocomposites plays a vital role in practical processing for mass

production and product formation. PE showed a typical non-Newtonian shear thinning viscosity

profile with increasing frequency (Figure 5.3d), which increased with TRG addition. Increasing

frequency (or correspondingly the shear rate according to Cox-Merz rule215) produced a similar

shear thinning profile, attributed to the alignment of graphene sheets in the shearing direction. At

higher TRG loadings (5 wt%), a rheological response typical of a highly filled percolated

network ( * ~ ƒ-1) was observed. Similar trends were observed for 80/20/TRG nanocomposites

(Figure 5.3e). The θ0/40/TRG nanocomposites exhibited similar peculiar behavior for the * as

was observed for Gˊ~ƒ data (Figure 5.3f). The viscosity decreased at low TRG loading, and a

continuous increase in viscosity was observed with increasing TRG loading above 1 wt%. The

three matrices (PE, 80/20 and 60/40) showed similar asymptotic behavior * ~ ƒ-1 at higher TRG

loading (5 wt%). However, due to the plasticization effect of OPE, the low frequency viscosity

95
for 60/40/TRG nanocomposites was even lower than that of 80/20/TRG composites, which was

lower than PE/TRG nanocomposite viscosity. The reduced viscosity of blend nanocomposites

provides ease of processing for PE nanocomposites.

Mackay72 reported the non-Einstein viscosity behaviour in nanocomposites with

nanoparticles smaller than the radius of gyration of the polymer. The TRG used in this study,

have approximately 5-6 nm thin, and 0.6-1 µm long particles. Due to highly non-polar nature of

polyolefins leading to insufficient filler/polymer interactions, TRG sheets tend to agglomerate

into a few hundred nanometer thick particles, becoming larger than the radius of gyration of

polyolefins, which do not exhibit slipperiness (Figure 5.3a). However, in 60/40/TRG

nanocomposites, OPE might be wetting the graphene surface, breaking down TRG agglomerates

to a few nanometer thin particles, leading to slipperiness. Therefore, we hypothesize that

particles with one dimension smaller than the radius of gyration of polymer might also exhibit

Mackay's non-Einstein behaviour. However, investigation of this behaviour is beyond the scope

of this study, and will be explored separately. The microstructure developed in nanocomposites

at lower TRG loading will be discussed in detail in the scattering section.

The rheological properties of CB-filled composites showed similar trends as observed for

TRG-filled nanocomposites (see Appendix-C). However, higher loading of CB (5-10 wt%) was

required to reach the percolation in the CB-composites. At dilute loadings, the complex viscosity

of 60/40/CB composites was lower than that of neat 60/40 blends, exhibiting a similar

slipperiness as was observed in 60/40/TRG nanocomposites. However, the slippery behaviour

was less prominent in CB-composites, which might be due to the highly fractal forming nature of

CB. More details on microstructure of CB-filled composites will be discussed in scattering

section.

96
(ii) Percolation and Scaling Analysis

The amount of filler needed to create an interconnected network within a polymer matrix,

called percolation threshold, is associated with polymer-particle interactions and particle

dispersion in the polymer matrix. In order to determine percolating volume fraction (Φp), we

analyzed the increase in Gˊ in the terminal region (at 0.01 Hz, Gˊ0.01), which is presented in

Figure 5.4. With increasing TRG loading, an exponential increase in G’0.01 corresponds to the

transition to solid-like behavior, which indicates Φp. Above Φp, the Gˊ0.01 exhibited a power law

dependence on TRG volume fraction as follows:

~   p 
' 
G0.01

where α is the power law exponent. The measured Φp values were 1, 0.7, and 0.3% for

PE/TRG, 80/20/TRG, and 60/40/TRG nanocomposites, respectively. The corresponding

percolation in weight% are 2.3%, 1.6%, and 0.7 wt%, respectively.

The substantial decrease in Φp for 60/40 blends indicated improved dispersion, likely due

to the increased graphene/matrix chain interactions. Fitting to percolation power law yielded the

α values of β.λ, β.η, and β.7 for PE/TRG, 80/β0/TRG, and θ0/40/TRG nanocomposites,

respectively (inset in Figure 5.4). Typically, a stress-bearing network produces α ranging from

β.1 to γ.7η. Here, α > β.1 may indicate the establishment of graphene-graphene bridging

network, which could derive from π-π interactions in graphene sheets at higher loadings.

Generally, an α < β.1 shows a network of particles bridged with polymer chains is formed,

whereas a direct particle-particle network is established for α > γ.7η216, 217, 218. However, a three

dimensional network exhibits α ≥ γ, to form a rigid percolating network219, 220, 221.

97
7
10
6
10
5 5
10 5

G  (Pa)
4 5
10
3
10 1
2 1
10 0.16
0.16 1
PE 80/20/TRG 60/40/TRG
1
10 PE/TRG 80/20
0
(a) (b) 60/40 0.16 (c)
10
7
PE/TRG 80/20/TRG 60/40/TRG
10 (d) (e) (f)
6
5
10
* (Pa-s)

5 5
10
1
0.16 1
4
10 0.16
PE 1
3 80/20 60/40
10
0.16
2
10

 (s )
-2 -1 0 1 2 -2 -1 0 1 2 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10 10 10 10 10 10
-1

Figure 5.3 Representative shear modulus and complex viscosity vs. frequency curves for
PE/TRG (a, d), PE/OPE 80/20/TRG (b, e) and 60/40/TRG (c, f), respectively. Solid lines
represent the polymer matrix in each box and numbers next to each curve correspond to weight
% of TRG

Furthermore, the aspect ratio of graphene in these matrices (Af) can be determined from

Φp. Af is the ratio of width to the thickness of the filler (l/d), and is considered as an indicative

parameter for exfoliation and dispersion. Af is equal to 1 for perfectly spherical filler. In case of

TRG/polymer nanocomposites, Af is maximum for completely exfoliated structures depending

on the lateral size of graphene layers. Ren et al.222 proposed a formula to determine the

relationship between the number of clay layers in each tactoid above percolation and clay

concentration. This expression was further simplified to calculate the average aspect ratio of

TRG191 as follows:

3pR
Af 
2p

98
where ΦpR is the percolation threshold volume fraction for randomly packed spheres, and

Φp is the experimentally determined percolating volume fraction. Combining ΦpR ~ 0.3, a

constant for the randomly packed spheres191, and measured Φp values (from Figure 5.4), the

aspect ratio was calculated for each nanocomposite. A more exfoliated structure exhibits higher

Af values. In TRG-filled systems, Af increased from 41 to 60 as the polymer matrix changed

from neat PE to PE/OPE 80/20 blend. However, an Af of 155 was obtained for PE/OPE 60/40

matrix, indicating better exfoliation and improved dispersion of TRG in the 60/40 blend (Table

5.1).

7
10

6
10
4


5 10
G

10 2
G0.01 (Pa)


10

p
3 -4 -3 -2 -1
10 10 10 10 10 PE/TRG

p=0.3%
80/20/TRG p=0.7%
1
10
p=1.0%
-1
60/40/TRG
10 -5

TRG volume fraction, 


-4 -3 -2 -1
10 10 10 10 10

Figure 5.4 Percolation analysis of TRG filled PE, PE/OPE 80/20 and 60/40 blends. Solid lines
are linear fit to experimental data. The point of intersection of two lines indicates the percolation
threshold, Φp. Inset shows power law fit to percolation law for PE, 80/20 and 60/40 blends-filled
with TRG, and the numbers next to each line are the power law exponents. Color schemes are
same in both panels.

By determining the power law dependences of Gˊ0.01 and critical strain ( c %) on TRG

volume fraction (Φ), δVE properties of nanocomposites can quantify the microstructure of fillers

99
in polymer nanocomposites. Critical strain ( c %) is defined as the strain ( %) at which G′

decreases to 90 % of its initial value in a strain sweep experiment. Shih et al.223 proposed a

scaling model for the elastic properties of fractal networks built up by filler aggregates above Φ p.

In nanocomposites, the fractal network is considered as elastically linked filler aggregates. A

decrease in the linear-elastic behavior with increasing filler volume fraction above Φp

(Appendix-C) can be explained by the physical links between the aggregates, called the strong-

link regime. In the percolation region, the power law dependence of c and Gˊ0.01 can be

described by the following equations:

 c ~  (1 x )/(3d )
f

'
G0.01 ~  (3 x )/(3df )

where, df is the fractal dimension of the filler aggregates, and x is an exponent related to

filler Φ, and aggregate size by giving dimension of aggregate backbone, which is responsible for

elasticity. The c and Gˊ0.01 (above percolation threshold) were plotted as a function of TRG Φ

(Figure 5.5), and exponents of Φ were evaluated. Both PE/TRG and 80/β0/TRG nanocomposites

exhibited similar exponent of c%, whereas the exponent decreased for 60/40/TRG

nanocomposites. The scaling exponent of G′ was lower for PE/TRG system (γ.8), than for

80/β0/TRG and θ0/40/TRG nanocomposites (4.γ). The higher value of G′ exponent further

indicates the formation of elastic network219, which is more efficient in storing elastic energy in

blend/TRG nanocomposites compared with PE/TRG nanocomposites. In addition, G′ scaling

exponents (3.8 and 4.3) in this study are higher than that reported for CNTs/polypropylene

nanocomposites (2.8)224. This difference in scaling exponents implies that the scaling can depend

on the fillers dimensionality, and the two-dimensional platelets, such as TRG nanosheets might

100
be more effective in enhancing the elastic contribution in polymers compared to the one-

dimensional nanotubes.

The exponents scaling for c and Gˊ0.01 above percolation can be expressed as:

c ~ 
m

'
G0.01 ~  mG

Combining the scaling equations, the power law exponents become:

df  3
2
m  mG

x 3
2mG
m  mG

The calculated values for df and x were 0.6 and 6.3 for PE/TRG, 1.4 and 3.7 for 80/20-

TRG, and 1.8 and 2.2 for 60/40-TRG nanocomposites, respectively. Surprisingly, df increased

with increasing OPE loading, which is different from other reports8, 219
. Usually, df decreases

with increasing Af, which is exhibited by various clay65 and graphene filled polymers219. A

relatively high value of x for PE/TRG system indicates that most sheets in TRG flocs contribute

to network elasticity, and the structural defects are elastically inactive such as dangling ends

which rarely exist216, 219


. The value of x decreased with increasing OPE loading in

nanocomposites. However, all three values of x are still large enough to confirm that the flocs are

elastically active in the three nanocomposites219.

101
c   (b)
-2.99
(a)
Gp  
3.81

Gp  
2 PE/TRG 5

Critical Strain, c (%)


10 10 4.25

c  
-2.98 PE/TRG
80/20/TRG

G0.01 (Pa)
1 80/20/TRG 10
4
10

0 3
10 10
c  
Gp  
-2.62
4.26

10
-1 60/40/TRG 10
2
60/40/TRG

TRG  TRG 
-3 -2 -1 -3 -2 -1
10 10 10 10 10 10

Figure 5.5 Scaling behavior of critical strain (a), and G′ (b) above percolation threshold

(iii) Mechanism of TRG Dispersion

The dispersion mechanism of nanoscale, hard, and rigid disks in a polymer matrix is

governed by either Brownian motion of particles or non-Brownian interactions. According to the

Stokes-Einstein equation225 for disk-shaped particles (of length L), particle rotary diffusion

coefficient (Dr, s-1) in dilute concentration regime in a suspending medium of viscosity ( o) is

given as:

Dr 
3kBT
4o L3

Where, kB is the Boltzmann constant, and T is the temperature in absolute scale.

Considering L~600 nm (longest dimension in Figure 5.1), the Dr calculated for PE/TRG, 80/20-

TRG, and 60/40-TRG at 160° C were 5x10-9, 1x10-8, and 1x10-7 s-1, respectively. Conversely, the

diffusion time, ~1/Dr was 108, 108, and 107 s, respectively, which represents the time needed for

nanosheets to travel a distance equal to their radii (300 nm) at the processing temperature (160°

C). The diffusion time, however, seems unrealistic, since a polymeric material would degrade if

kept at that high temperatures (160° C) for such extremely long time. Thus, the diffusion

102
coefficient or diffusion time alone cannot answer whether the diffusion of graphene inside a

polymer matrix follows Brownian or non-Brownian dynamics. Therefore, the concept of Dr is

further extended to describe the dynamics of disk-shaped particles via the universal Doi-Edwards

theory, originally developed for CNTs.

The Doi-Edwards theory226 describes the dispersion of nanorods of length L, and

diameter d, in liquid suspension in dilute and semi-dilute regimes. For nanorods, v is the number

of nanorods per unit volume, which is related to volume fraction of nanorods as Φ=d 2Lvπ/4.

Considering d as the thickness of graphene nanosheets (d=5 nm for TRG, since Raman analysis

indicated that our TRG was comprised of about four 1-atomic thick sheets88). According to Doi

and Edwards, the concentration dependence of the zero shear viscosity, o for dilute regime

obeys the following scaling law:

~  L3 
c
M
1

Here, C, M are viscosity of composites, and that of matrix, respectively. The rotary

diffusion of particles is hindered by neighboring particles in semi-dilute regime, and Dr is

predicted to follow the power law Dr ~ (vL3)-2. Consequently, the viscosity scales in the semi-

dilute regime as follows:

~  L3 
c
M
3

However, in the concentrated regime, the factors such as sheet flexibility, Brownian

motion, and van der Waals interactions are strongly affected by the viscosity of the matrix227.

The Doi-Edwards theory can be applied to any nanorods (or nano-sheets, in this case), provided

103
these particles behave as solid Brownian entities in the suspending medium on the nanometer

scale. Thus, from the equations derived by Doi and Edwards, the variation of the reduced

viscosity ( C/ M) with vL3 should follow a master curve. The Brownian motion of nanosheets

(similar to nanorods) should be universal and independent of their nature. Recently,

Cassagnaue228 reported the universal master curve of CNTs, cellulose whiskers, and polymer

nanofibers in different polymer matrices, which indicates that the rotary motion of these fillers

was dominated by the Brownian motion. It is worth mentioning that even extremely dilute

suspensions of nanorods were not able to produce c/ m ~ Φ1 behaviour 227, 228.

Following the Doi-Edward theory, a universal curve of log ( c/ M) versus the particle

concentration (vL3) was created for TRG-based nanocomposites (Figure 5.6). The successful

building of the universal curve indicates that TRG dispersion in these matrices is dominated by

Brownian forces227, c/ M
228
. The TRG nanocomposites showed ~ (vL3)3 behavior in the

concentrated regime. A downshift in the master curve was observed for 60/40/TRG

nanocomposites, attributed to the extremely low viscosity of 60/40 blends. Thus, the diffusion of

graphene sheets in polymers follows Brownian diffusion. However, further investigations in the

low viscosity polymers filled with graphene nanosheets should be carried out to claim the

universality of this behavior.

5.3.3 SAXS-Based Microstructure

The microstructural parameters, such as df and x, extracted from melt rheology scaling

are averaged for each matrix over all the compositions. However, each composition has its own

microstructure, which can be analyzed using scattering methods, including small angle x-ray

scattering (SAXS). SAXS measures the shape and size of polymers between 5 and 25 nm of the

104
repeat distances171. The scattering vector (q) probes different structural features at various length

scales (length scale ~ βπ/q) in polymers and polymer nanocomposites.

3.0
PE/TRG
2.5 80/20/TRG
60/40/TRG
log (ηC/ηM ) 2.0 ~3

1.5

1.0
~0
0.5

0.0 ~0

-0.5
0.0 0.5 1.0 1.5 2.0 2.5 3.0
log (vL3)
Figure 5.6 Universal behavior for TRG-filled matrices

Increase in power law decaying exponent (log(I) ~ log(q)-α) at lower q values indicates a

larger size of scattering objects. For clay minerals, Malekani et al.229 proposed that the decaying

exponent α was directly related to the fractal dimensions, df. For α ≤ γ, the particles form mass

fractals (a substance whose surface and mass are both characterized by the fractal properties). In

this case, α directly corresponds to mass fractal dimensions, dm. If γ ≤ α ≤ 4, surface fractals are

formed where the scattering is observed from a smooth surface. The surface fractal dimensions,

ds, is calculated from α as followsμ α ≡ θ - ds where 3 < 6 - ds ≤ 4229, 230. Thus, ds varies from 3 to

2, where ds ~ β indicates a smooth surface approaching Porod's scattering (α = 4)172, and ds ~ 3

indicates uneven concentration of scattering objects at the scattering surface231.

Pure PE exhibited α = β.γ, typical for randomly oriented polymer chains 172 (Figure 5.7a).

Inclusion of τPE in PE increased α to β.λ for PE/τPE θ0/40 blends (Figure 5.7b). In the absence

105
of any filler, higher α indicates increased mass fractal concentration in blends, which is attributed

to immiscibility of OPE at higher loadings210. Addition of TRG in PE showed smaller α values

(1.5 to 2.1) until a loading of 1 wt% (Figure 5.7a). The small α values at low TRG concentration

indicate the exfoliation of TRG sheets inside PE matrix. Increasing TRG loading to 5 wt%,

although α increased but it was still below 3, showing that TRG was forming the mass fractals.

The 80/β0/TRG nanocomposites exhibited a similar pattern of α increasing and only mass

fractals being observed (Appendix-C). However, 60/40/TRG nanocomposites showed a different

behaviour when TRG ≥ γ wt%; α > γ was observed. The increased α showed the formation of

surface fractals229, which was attributed to the better exfoliation, and high Af in 60/40/TRG

nanocomposites, which verifies the finding from rheology discussed earlier.

Scattering from the nanocomposites changes above the percolation threshold. The

PE/TRG nanocomposites exhibited the rheological percolation at ~1.0 vol% (2.2 wt%), whereas

the percolation for 60/40/TRG nanocomposites was observed at 0.3 vol% (0.7 wt%). Below

percolation, low α values were measured at 1 wt% for θ0/40/TRG system (1.λ) and PE/TRG

nanocomposites (2.1), which indicated better dispersion of TRG. Also, the slippery behaviour

observed in the rheological data for 60/40/TRG nanocomposites below percolation can be

associated with this small α values. Since, we did not observe the slipperiness in PE/TRG

nanocomposites, it might also be associated with relatively low viscosity of 60/40 blends, and

shear alignment of TRG sheets upon exfoliation.

Above the percolation threshold (TRG > γ wt%), the TRG formed surface fractals (α =

γ.η) in θ0/40 blend, whereas only mass fractals were observed for PE/TRG nanocomposites (α =

β.7). The large difference in α values further indicates markedly improved exfoliation and

dispersion of TRG in 60/40 blends.

106
(b)
60/40/TRG
3 wt% SPE40_29_00036_001
'6060TGR_016_00645_00001'
=3.5 '6040TGR_1_00647_00001'
'6060TGR_3_00648_00001'
1 wt%
=1.9

I(q) [a.u.]
60/40
=2.9

0.16 wt%
=1.4

5 6 7 8 2 3 4 5 6 7 8 2 3
0.01 0.1
q [1/Å]
Figure 5.7 I(q)~q log-log plots of PE/TRG (a), and 60/40/TRG nanocomposites (b). Note: The
lines in all panels are stacked in the same order with Y-offset for clarity.

Due to the fractal forming nature of CB particles, a relatively large α = β.1 was observed

at 0.5 wt% CB loading in 60/40/CB composites (Figure 5.8). A consistent increase in α was

observed with increasing CB loading. The arrangement of CB particles to form surface fractals

was observed at 10 wt%, which indicates the build up of strong network and a good dispersion of

CB in θ0/40 blend. However, the surface fractals (α > γ) were observed only in θ0/40/CB

composites, showing better dispersion of CB in 60/40 blends. The PE/CB, and 80/20/CB

composites showed similar increase in α with increasing CB concentrations, but no surface

fractals were observed (Appendix-C). In addition, secondary structure was observed in all CB-

filled composites, which could be seen as a shoulder appearing at low q values. With increasing

CB concentration, the shoulder shifted to higher q (smaller length scales). The calculated length

scales associated with this shoulder are shown later in this section.

The SAXS-knee in the range of 0.02 to 0.08 Å-1 corresponds to the lamellar thickness in

polymers (Figure 5.7). The lamellar thickness (Lp) or interparticle spacing can be extracted from

the maxima in the Kratky plots174, 175


, also called Lorentzian corrected plots, where I(q)q2 is

107
plotted versus the scattering vector (q). The maxima is converted into length scales by Lp ~ βπ /q.

The Lp is the sum of amorphous and crystalline fractions in polymers210.

60/40/CB
60/40/CB 0.55
60/40/CB 1.16
60/40/CB 4.3
60/40/CB 6.58
60/40/CB 10

I (q) [a.u.]

=2.1

9 2 3 4 5 6 7 8 9 2 3
0.01 0.1
q [1/Å]
Figure 5.8 I(q) ~ q log-log SAXS data for PE/OPE 60/40 Blends-CB composites. Note: The lines
in all panels are stacked in the same order with Y-offset for clarity.

The Lp was calculated for the TRG-filled nanocomposites, and CB-filled composites (see

supporting information for a representative Kratky plot). Inclusion of OPE in PE has already

been shown to decrease Lp from 180 Å for neat PE to 161 Å for 80/20 blends, which further

reduced to 147 Å for 60/40 blends210. Substantial decrease in Lp was observed with increasing

TRG in PE/TRG, and 60/40/TRG nanocomposites (Figure 5.9a). Similarly, the nanofillers (TRG

or CB) do not affect the WAXS patterns of the matrices (Appendix-C). Therefore, it is

reasonable to believe that the fillers reside inside the amorphous portion of the matrix, and due to

the packing of fillers inside lamellae, overall Lp decreased. The decrease in Lp has been reported

to have a direct effect on the macroscopic mechanical properties of polymers. Specifically, the

Young's modulus of polymers increases with decreasing Lp (see mechanical properties

108
section)187, 188, 210. A constant Lp with TRG ≥ γ wt% clearly indicated the percolation behaviour

in TRG-filled nanocomposites.

For CB-filled composites, the Lp also decreased with increasing CB loading in PE and

PE/OPE blends (Figure 5.9b). However, the decrease in Lp was not as significant as it was in

TRG-filled nanocomposites. The scattering also corroborates that high CB loading leads to lower

mechanical and rheological properties than TRG-filled nanocomposites. A constant Lp was not

observed for CB-filled composites at higher CB loading, which also indicates that these

composites do not achieve percolation.

5.3.4 Thermal Properties

Neat PE is thermally stable for a polymer, completely decomposing around 500°C. The

PE/OPE 60/40 blend showed lower thermal stability compared with neat PE, which is attributed

to miscibility and plasticization effects of PE and OPE, and oxygen contents of OPE 210. On the

one hand, addition of 5 wt% TRG in PE (Figure 5.10) did not affect the stability of the

nanocomposite (decomposition temperature ~ 474°C) compared to that of neat PE (473°C). The

unaffected thermal stability of PE/TRG nanocomposites indicates the absence of any favorable

interactions between graphene and PE. On the other hand, thermal stability of 60/40 blends at 5

wt% TRG was enhanced as the decomposition temperature increased from 472° C for neat 60/40

blend to 477° C for the nanocomposite, attributed to improved graphene-matrix interactions in

these nanocomposites. Similar patterns of thermal stability were observed for CB-filled

composites (see Appendix-C). The lack of interactions between CB and PE reduced the thermal

stability of PE/CB composites at 10 wt% loading. However, due the presence of OPE, and

interactions in blends, the stability of blend-CB composites was also improved.

109
200 180
Percolated regimes PE/CB
80/20/CB
PE/TRG 60/40/CB
180 80/20/TRG
60/40/TRG 160
Lp [Å]

160

Lp [Å]
140
140

120
120

(a) (b)
100 100
0 1 2 3 4 5 0 1 2 3 4 5 6 7 8
TRG wt % CB wt %

Figure 5.9 Kratky analysis of TRG-filled nanocomposites (a), and CB-filled composites (b).
Color schemes in (a) show percolation of each matrix.

5.3.5 Mechanical Properties

One of the advantages of adding nanofiller to a polymer matrix is improving the

mechanical properties. Mechanically robust nanocomposites can be used for a large number of

practical applications. Young's modulus of PE decreased from 1.6 MPa to ~1 MPa for 80/20

blends, and increased to 1.8 MPa for 60/40 blends. The increase in Young's moduli of blends

was directly associated with the decreased Lp187, 188, 210.

The superiority of TRG over CB as filler for polymers is evident by the comparatively

improved Young's modulus of filled matrices (Figure 5.11). The PE/TRG and PE/CB exhibited

similar Young's moduli, showing the poor dispersion of these fillers in PE in the absence of any

filler/polymer interactions. On the other hand, approximately ~3-fold increase in modulus was

observed for 60/40/TRG nanocomposites compared to PE/TRG nanocomposites due to the

increased filler/matrix interactions. The melt viscosity of 60/40 blends is ~7-fold lower than that

of neat PE, whereas the increase in the Young's modulus was 3-fold higher for 60/40/TRG

nanocomposites compared to that of PE/TRG nanocomposites.

110
100
2.5

Mass Retained, %
2.0
80

dW/dT, % / C
1.5

1.0

60 0.5

0.0

-0.5
40 350 400 450 500 550

PE
20 PE/TRG 5 wt %
60/40
60/40/TRG 5 wt%
0
0 100 200 300 400 500 600
Temperature, C
Figure 5.10 Thermograms for PE, PE+TRG 5 wt%, PE/OPE 60/40, and 60/40+TRG 5 wt%.
Inset shows derivative of TGA versus the temperature

This significantly lower processing viscosity and increased mechanical property indicates

the formation of processable and mechanically robust nanocomposites, which is very remarkable

to achieve as most nanocomposite that yield enhanced mechanical properties suffer from

significant increase in viscosity.

The elastic properties of fiber/flake polymer nanocomposites can also be predicted by

numerous micromechanical models232. Generally, the elastic properties in micromechanical

models depend on particle/matrix stiffness ratio (Ep/Em), particle volume fraction (Φp), particle

aspect ratio (Af), and orientation of particles in the matrix233.

In this study, both TRG and the matrices are assumed to be linearly elastic, either of

which can be taken as isotropic or transversely isotropic. Two widely-used models are the
233
Halpin-Tsai and Mori-Tanaka models . The Halpin-Tsai (H-T)234 equation gives reasonable

estimate for effective stiffness in nanocomposites, but the Mori-Tanaka model (M-T)235 is

111
considered more accurate for high aspect ratio fillers. Considering longitudinal stiffness (E11) of

nanocomposites (obtained from tensile testing), the H-T equation is given as:

E11( H T ) 1  2 Af  p

Em 1   p
( E p / Em )  1
E / Em   2 Af

p

and Mori-Tanaka (M-T) model is represented as:

E11( M T )

 f  2 m A3  1   m  A4  1   m  A5 A
1

1
Em
2A

10
CB 2.5 wt%
Young's Modulus (MPa)

TRG 3 wt%

PE 80/20 60/40
Figure 5.11 Comparison of the Young's moduli of TRG and CB filled composites

Here, vm, and vp are the matrix and particle Poisson ratios, and constants A, A3, A4, A5 are

calculated from matrix/particle properties and components of the Eshelby tensor236, 237
, which

depend on the particle Af, and dimensionless elastic constants of the matrix. For modeling, we

used Ep as 250 GPa for TRG sheets238, 239


, vp as 0.0006240, vm as 0.48 for PE241, whereas the

112
experimentally determined Em values are used throughout. The only variable fitting parameter in

the two models is the particle aspect ratio (Af).

The tensile moduli of nanocomposites normalized by the moduli are compared to the

loading to examine the model fits (Figure 5.12). PE filled with TRG at 2.14 vol% was

approximately 50% stiffer than neat PE. On the other hand, 80/20/TRG nanocomposites showed

~150% more stiffness compared to the neat 80/20 blend, and the 60/40/TRG nanocomposites

exhibited ~500% increased stiffness compared with the neat blend. Fitting based on the effective

medium models (H-T and M-T) were applied to moduli increase versus TRG concentration

(wt%), and the effective aspect ratio (Af) of the dispersed TRG sheets were estimated. Both the

H-T and M-T models produced similar Af of TRG for PE and 80/20 blends (Figure 5.12).

However, M-T model exhibited insensitivity for the 60/40/TRG nanocomposites due to large

increase in modulus. Only, H-T model was able to capture the Af for 60/40/TRG system, which

was ~460. A very large Ep/Em (~105) might be the reason for M-T model's insensitivity above Af

> 200. However, for small Ep/Em (~102), H-T model overestimates E11/Em compared with that

obtained from M-T model233.

5.3.6 Electrical Properties

The electrical properties of graphene/polymer nanocomposites are strongly dependent on

effective dispersion of graphene and its aspect ratio, whereas the dispersion is directly connected

to graphene-polymer and inter-graphene interactions. The electrical conductivity ( ) of PE/TRG,

80/20/TRG and 60/40/TRG nanocomposites (measured at 0.01 Hz) showed significant

improvement as the concentration of graphene was increased from 0 to 5 wt%. PE/OPE 60/40

113
blends exhibited a low rheological percolation threshold, and a similar transition was observed

for the electrical conductivity (Figure 5.13).

8.0
H-T M-T
PE/TRG 21 15
6.0 80/20/TRG 155 160
60/40/TRG 456
E11/Em

4.0

2.0

0.0
0.0 0.5 1.0 1.5 2.0 2.5
TRG vol%
Figure 5.12 Mechanical properties of TRG-filled nanocomposites. Solid lines are fit to Halpin-
Tsai model, and dashed lines represent Mori-Tanaka micromechanical model. Inset shows
estimated aspect ratio (Af) of TRG for nanocomposites from the best fit.

At almost same TRG loading, 60/40/TRG nanocomposites exhibited more than 1 order of

magnitude improvement in the electrical conductivity compared with PE/TRG, and 80/20/TRG

nanocomposites. The increase in is attributed to enhanced TRG-matrix interactions, and large

Af in blend-nanocomposites due to OPE. Also, at similar loadings, the TRG nanocomposites

exhibited better electrical conductivities compared to CB composites. The electrical conductivity

of TRG nanocomposites was ~2 orders of magnitude higher than that for CB composites (Figure

5.13).

A low electrical percolation threshold also signifies the improved interactions in

nanocomposites. The increased from 7.4x10-14 S/cm to 3.6x10-4 for PE/TRG as TRG

concentration was increased from 0 to 5 wt% (Figure 5.14). Similarly, the for 80/β0 blends

114
increased from 7.6x10-14 to 1.1x10-4 S/cm, and from 7.4x10-14 to 9.6x10-4 S/cm for 60/40 blends

for the same TRG concentration range.

-3
10 CB 2.5 wt%
-4
TRG 3 wt%
10
 [S/cm] 10
-5

-6
10
-7
10
-8
10
PE 80/20 60/40
Figure 5.13 Comparison of electrical conductivities of TRG-filled and CB-filled composites

The electrical percolation threshold values (Wp ) were 0.67, 0.45, and 0.3 wt% for

PE/TRG, 80/20/TRG, and 60/40/TRG nanocomposites, respectively. The rheological percolation

thresholds were higher than the electrical percolation for all composites. Similar to the

rheological percolation, the 60/40/TRG nanocomposites exhibited the lowest percolation, which

further signifies the increased graphene-matrix interactions, and high Af of TRG in the

nanocomposites. Moreover, the CB filled nanocomposites exhibited similar increase in electrical

conductivity with increasing CB loading. Expectedly, the electrical percolation thresholds for

CB-filled composites were higher than those obtained for TRG-filled nanocomposites. Higher

percolation threshold in CB-filled composites is attributed to highly fractal, and agglomerate-

forming nature of CB. The complete electrical conductivity data of CB-composites is provided in

the supporting information.

115
Again, a power law scaling relationship can be applied to the electrical conductivity data

as follows:

  W  Wp 
t

A log-log plot between and W-Wc should give a straight line with exponent t. The inset

in Figure 5.14 shows the power law fitting, and the power law exponents (t) are shown next to

each set of data (also see Table 5.1). The exponent, t, reflects the dimensionality of the system.

Typically, 1.6 < t < 2.0 is for three dimensional network, and 1 < t < 1.3 is exhibited by the two

dimensional network in nanocomposites242, 243, 244. In this study, t > 1.3, which showed a three

dimensional network was formed by TRG in all composites. Interestingly, the t-values for CB

composites were even higher than that for TRG nanocomposites, which can be attributed to the

fractal network forming nature of CB to construct a three dimensional network in composites

(Appendix-C). However, the electrical conductivity of CB composites are lower than that TRG,

which might be due to structural imperfections in the networks formed by the CB particles.

5.3.7 Morphology

The cryogenically fractured nanocomposites at 5 wt% loadings were used to study the

morphology via FE-SEM (Figure 5.15). The SEM images were taken at two different

magnifications. The PE/TRG nanocomposites (Figure 5.15 a, b) showed graphene being pulled

out during cryo-fracturing.

116
-2
10
-3
60/40/TRG
10
-4
10

 [S/cm]
-5
10 80/20/TRG
-6
10 10
-3

-7 10
-4
10


10 t~1.7
-5
-8
10 PE/TRG 10
-6

-7 t~2.6
t~1.5
-9 10 -1 0 1
10 10 10
W-Wp
10
-10
10
0 1 2 3 4 5 6
TRG wt%
Figure 5.14 Electrical conductivity of TRG-filled nanocomposites. Solid lines are to guide the
eye only. Inset shows the power law fit, the numbers indicate the power law exponents.

Table 5.1 Electrical and rheological percolation threshold and power law exponents for TRG-
based nanocomposites

Rheological Electrical
WpR% ΦpR% α Af Wpσ% Φpσ% t
PE/TRG 2.3 1 2.9 41 0.7 0.3 2.6
80/20/TRG 1.7 0.7 2.5 60 0.5 0.2 1.5
60/40/TRG 0.7 0.3 2.7 155 0.3 0.1 1.7

The graphene stacks were more visible in the high resolution image (Figure 5.15b).

Similar morphology was exhibited by 80/20/TRG nanocomposites. High Af of TRG in these

nanocomposites compared to PE/TRG system was evident in the high-resolution image (Figure

5.15d) as smaller graphene stacks are seen in these nanocomposites. The 60/40/TRG

nanocomposites exhibited morphology similar to PE/TRG in low-resolution images. However,

under high-resolution, 60/40/TRG nanocomposites showed very different, and exfoliated

morphology. The TRG stacks in these nanocomposites were either very small (seen at a few

117
spots), or not seen at all. These results are in agreement with the rheological, mechanical, and

scattering analysis, that better exfoliation was observed in 60/40/TRG nanocomposites compared

with the other two composite systems.

(a) (b)

(c) (d)

(e) (f)

Figure 5.15 SEM images of PE/TRG (a, b), 80/20/TRG (c, d), 60/40/TRG (e, f) at 5 wt%
loading. The arrows indicate TRG sheets located in these composites

118
Conclusions

Dispersion of thermally reduced graphene (TRG) in PE, and PE/OPE blends was

investigated by thermomechanical, electrical, and scattering techniques. Reduced rheological and

electrical percolation thresholds in PE/OPE blends showed better dispersion, and exfoliation of

graphene sheets in the blends compared to neat PE. The universal behavior of TRG diffusion in

polymers followed the Doi-Edwards theory for CNTs dispersion, showing that TRG follows the

Brownian diffusion. The microstructure developed in TRG-filled nanocomposites by SAXS

showed the formation of mass and surface fractals in nanocomposites. At lower TRG loadings,

TRG showed better dispersion as indicated by small slope in I ~ q log-log plots. The dispersion

of TRG was fairly inconsistently similar at low TRG concentrations. However, a more detailed

study of SAXS with carefully synthesized nanocomposites is highly desirable to understand this

discrepancy in dispersion. The decrease in long period order (Lp) was linked with the mechanical

properties, where substantial decrease in Lp in blend nanocomposites showed remarkable

improvement in elastic moduli. Modeling and simulation of mechanical properties using

micromechanical models showed even higher exfoliation of graphene sheets in nanocomposites.

A ~3-fold increase in elastic properties of TRG-filled blends with a ~7-fold lower melt viscosity

indicates the formation of processable blend nanocomposites with improved filler/matrix

interactions. Higher electrical conductivity of blend-TRG nanocomposites than that of PE/TRG

nanocomposites is another indication of better dispersion and exfoliation of TRG in the blends.

Acknowledgements

This project has been supported by The Petroleum Institute, Abu Dhabi through the

Cooperative Research Partnership with Colorado School of Mines. Authors thank U.S. Army

119
Research Office (DURIP Grant No.W911NF-11-1-0306) for financial support in part of this

work. Prof. John Dorgan is acknowledged for his laboratory facilities. Authors also thank Tara

Pandey and Andrew Motz at Colorado School of Mines for helping with SAXS measurements,

and Marcus Oils and Chemicals (Texas) for supplying the commercial OPE samples used in this

study. Authors thank Travis Arnold for assisting with sample preparation and rheological

experiments. Use of the Advanced Photon Source, an office of Science User facility operated for

the U.S. Department of Energy (D.O.E.) Office of Science by Argonne National Laboratory, was

supported by the U.S. DOE under Contract No. DE-AC02-06CH11357.

120
CHAPTER-6

CONDUCTIVE POLYETHYLENE/GRAPHENE/CARBON BLACK HYBRID

NANOCOMPOSITES

Understanding the synergic effects of hybrid fillers in polymer reinforcement is an

exciting area in polymer nanocomposites. This chapter focuses on using commercial graphene

(called graphene nanoplatelets, GnP) and carbon black (CB) together to understand the hybrid

filler synergy in polyethylene (PE), and produce electrically conductive nanocomposites.

6.1 Introduction

Structural composite materials have attracted lots of attention due to high strength, high

or low thermal and electrical resistances, and low bulk density compared to metals1, 2, 3. One path

to composite materials is by incorporating a variety of fillers into polymer matrices, which

achieves the desired properties. In the past few decades, nanomaterial-reinforced polymers

(polymer nanocomposites) with high performance properties have attracted both academia and

industry. These advancements are likely to not only bring new technology into traditional

composite industry, but also offer cheaper and more environmentally benign production routes.

Electrically and thermally conductive polymer composites have introduced important

applications in areas such as adhesives, sensors, antistatic coatings and films, actuators, and

electromagnetic interference shielding materials for electronic devices.

Conventional micro-scale fillers such as metal powders245, 246, carbon black (CB)247, 248,

and graphite249 are required in large quantities (usually 10-50 wt%) to induce electrical and

thermal conductivities in insulating polymers. The conventional composite industry suffers from

processing and property issues when such large quantities of fillers are used. Currently,

121
nanomaterials, such as carbon nanotubes (CNTs) and graphene, are replacing conventional

micro-fillers due to their superior electrical properties, thermal conductivities, and reinforcing

effects in polymers. CNTs are 1D materials composed of sp2-carbons in the form of single or

concentric tubes, and graphene is a 2D, atomically-thick sheet of sp2-carbon atoms. A small

quantity of these nanofillers substantially alters the properties of polymers38. However, property

improvement in polymers largely depends on the characteristic nature of polymers, and that of

nanofiller28. For example, electrical percolation threshold of graphene in polar polymer/graphene

nanocomposites lies in the range of 0.05-1 wt% whereas it is higher (1-15 wt%) for the nonpolar

polymer/graphene nanocomposites. Specifically, polycarbonate191 showed rapid increase in

electrical conductivity at 0.6 wt% graphene, and thermoplastic polyurethane48 achieved the

percolation limit at 0.5 wt%. However, nonpolar polyolefins, such as polyethylene and

polypropylene, show electrical percolation at 2-12 wt% graphene loading202, 209


, which is

markedly higher than polar polymer nanocomposites.

In general, absence of sufficient filler/polymer molecular interactions in nanocomposites

results in the formation of large agglomerates of filler particles inside the polymers, which

eventually leads to high percolation limits. The polyolefin/graphene nanocomposites show

higher percolation because of graphene agglomeration which is a matter of great interest and

concern for researchers and industry around the world. Recently, we addressed the issue of poor

dispersion of graphene in polyethylene by synthesizing miscible blends of PE with OPE

(Chapters 4 and Chapter 5). The electrical percolation limit for graphene was substantially

reduced in PE/OPE blends (0.3 wt%) compared to neat PE/graphene nanocomposites (0.7 wt%).

In addition, remarkably improved mechanical properties were achieved in blend/graphene

nanocomposites, resulting from the high aspect ratio of the dispersed graphene in PE/OPE

122
blends. Herein, we address a new method of increasing graphene dispersion in polyethylene by

using CB as a synergic filler.

In the realm of polymer nanocomposites, besides manufacturing mechanically robust

materials, nanocomposites with unique electrical, thermal, magnetic, and optical properties has

become a new research highlight. Hybrid fillers250, 251, 252, consisting of two heterogeneous non-

elementary units with different properties can be used to prepare a new class of efficient,

functional, and multi-dimensional composite nanofillers. Incorporated in polymers through

different approaches, such as melt or solvent blending, these hybrid nanofillers can impart

important properties in polymers to manufacture high performance and multi-functional polymer

nanocomposites. Consequently, the next technological frontiers can evolve by designing and

optimizing the nanomaterial combinations and their synergic functions. In addition, expectedly,

the nanoparticles agglomeration and interfacial interaction issues in polymer nanocomposites

could also be reduced by using such functional hybrid nanofillers253.

Recently, in the last few years, a few reports have emerged on using hybrid nanofillers,

introducing various functional properties in polymer nanocomposites. Substantial improvement

in thermal conductivity of epoxy matrix was reported by Yang and Gu107 with modified CNTs

and silicon carbide (SiC) hybrid filler synergy. The thermal conductivity of epoxy

nanocomposites was almost doubled with CNT/SiC hybrids compared to CNTs alone. Ma et

al.254 reported CNT/CB hybrid filler in epoxy nanocomposites to improve the electrical

conductivity. A 50/50 ratio of CNT (0.2 wt%) and CB (0.2 wt%) showed synergic effects,

leading to electrical percolation in nanocomposites compared to percolation by CNTs at 0.3

wt%, and CB at 0.6 wt%, individually. In another report, marked improvement in mechanical

properties of nylon 6 was reported by Zhang et al.255 with CNT/silica hybrids at only 1 wt% of

123
the hybrid filler. However, to the best of our knowledge, no report has surfaced on using hybrid

filler to reinforce nonpolar polyolefins.

The current study is part of a larger research project focused on improving graphene's

dispersion in nonpolar polyolefins to produce electrically conductive, and mechanically robust

nanocomposites. Here, linear low density polyethylene (PE), one of the most widely used

polymers in the polyolefin family, was used as a matrix. The synergic effects of graphene and

CB were investigated by varying graphene/CB ratios and concentration in PE, mainly aiming at

improving the electrical conductivity of the nanocomposites. The study includes characterization

of the filler dispersion in nanocomposites by melt rheology, and microstructure of

nanocomposite investigated by small angle x-ray scattering.

6.2 Experimental Details

Following section provides the information on materials and synthesis procedures

adapted in this study.

6.2.1 Materials

Graphene nanoplatelets (A12, Graphene Supermarket), carbon black (99.9%, bulk

density = 170-230 g/L, surface area = 75 m2/g, Alfa Aesar), linear low density polyethylene

(LLDPE) with molecular weight 3.6 x 106 g/mol (determined via gel permeation

chromatography), and bulk density 0.97 g/cm3 (428078, lot#07730MEV, Aldrich), and p-xylene

(99%, Sigma Aldrich) were used as received.

124
6.2.2 Nanocomposites Synthesis

Graphene nanoplatelets (GnP) and/or carbon black (CB) were dispersed in p-xylene via 2

h sonication at room temperature. The dispersion was also stirred for 30 minutes after the first

hour of sonication for better mixing, followed by sonication for another hour. Low density

polyethylene (PE) was added into the dispersion at room temperature, and stirred at 120 °C for 2

h under reflux. After complete homogenization, the composite dispersion was drop cast on a

heated glass plate at 80 °C, and dried for 2 h at the same temperature. The complete solvent

removal was assured by drying the cast films for 24 h inside a fumehood at room temperature.

The dried thin films were chopped into small pieces (3-5 mm wide), heated at 160 °C for 10

minutes, and hot pressed for 3 minutes under 5000 kPa pressure to prepare circular discs

(thickness 1-1.3 mm and diameter ~ 25 mm) for rheological testing. Thin films for scattering and

electrical testing were prepared following the similar procedure using more dilute solutions in p-

xylene, and drop casting on a glass plate at 80 °C, and following the same drying procedures.

6.2.3 Characterization

X-ray diffraction patterns were obtained from a wide angle x-ray diffractometer (WAXD)

(Phillips PW γ040/θ0) with CuKα radiation at a scanning rate of 0.02 °/s. Transmission electron

microscope (TEM) images were obtained using a FEI Phillips C200 at 200 kV. For TEM

measurements, a dilute dispersion of GnP or CB or hybrid filler (0.1 mg/20mL acetone) was

prepared by bath sonication for 10 minutes, and one drop of the suspension was dried on a 300-

mesh Cu grid with holy carbon. Very dilute dispersion of CB (2.5 mg/mL), and that of TRG (1

mg/mL) were dispersed in acetone via bath sonication at room temperature, and dynamic light

125
scattering (DLS) was performed on a particle size analyzer (NanoBrook 90Plus, Brookhaven) at

25 °C.

Rheological measurements were performed at 160 °C using a TA instrument AR-G2

rheometer equipped with 20 mm parallel plate. Tests were conducted in air while samples were

covered with an aluminum shield to avoid temperature fluctuations during the run. Linear

viscoelastic (LVE) region was determined by running strain sweep in 0.01-100% strain range at

1 Hz frequency, and storage (Gˊ) and loss (G˝) moduli were recorded as a function of strain%.

The frequency sweep experiments were performed in the frequency range of 0.01-100 Hz under

a small amplitude oscillatory shear, = o sin(ωt) within the δVE region.

The in-plane electrical conductivity was calculated using electrochemical impedance

spectroscopy (EIS) using the following equation:

Conductivity 
L
R.w .t

Where, R is the film resistance, L is the distance between the sense electrodes, w is the

width of film, and t is the sample thickness. The impedance spectra were obtained over a

frequency range of 0.3 Hz to 100 kHz using a four electrode test cell connected to a multi-

channel potentiostat (Biologic VMP3, Knoxville, TN). All measurements were made in an

environmental chamber to control sample temperature and humidity (TestEquity Model 1007H,

Moorpark, CA). The samples were polished using a 100 mesh grit paper before conductivity

measurements in order to remove thin polymer layer transported on the surface during the film

formation and drying process. Due to polishing, composite films exhibited static current build

126
up, which was removed by conditioning the samples at 95% relative humidity and 50 °C for one

hour. The results reported here are averaged over 3 samples at 25 °C and 95% relative humidity.

Small angle X-ray scattering (SAXS) experiments were performed at the X-ray Sciences

Division, beamline 12-ID-C, at the Advanced Photon Source at Argonne National Laboratory.

Data has been collected at 60°C under dry conditions, and is averaged over three exposures

times. Details of the procedure, and temperature and humidity control can be found elsewhere 212,
213
.

6.3 Results and Discussion

This section provides details on the characterization of fillers and nanocomposites

synthesized in this study.

6.3.1 Characterization of GnP and CB

A typical paper-like morphology was observed for GnP (Figure 6.1a), where finely

separated graphene nanosheets were extended over a few microns in the lateral dimensions. In

addition, small overlapping areas (marked with arrows in Figure 6.1a) also showed the presence

of some smaller GnP sheets 200-300 nm wide and ~500 nm long. Thus, GnP is composed of two

types of particles. It is also worth noting that GnP sheets are not wrinkled indicating the absence

of any foreign element present on its surface88. On the other hand, CB particles are known for

their strong agglomerating nature256. Strong agglomeration/association of CB particles was also

observed in TEM (Figure 6.1b). The CB particles were spherical in nature, with a 50-60 nm

diameter (inset in Figure 6.1b). However, CB agglomerates were extended over a few microns

127
even when a very dilute suspension of CB particles was prepared in acetone, and deposited on

the Cu grid for TEM analysis.

(a) (b)

100 nm

Figure 6.1 TEM morphology of GnP (a) and CB (b). Inset in (b) shows high resolution image of
CB particles.

The quality of the commercial grade GnP and CB was also examined by XRD analysis

(Figure 6.2). Graphite showed the intrinsic (00β) peak at β ~ βθ.4°, corresponding to an

interlayer spacing of ~3.37Å. A single, sharp XRD peak in graphite further indicated the

presence of a perfectly stacked structure in graphite. The particle size of graphite was calculated

to be 145 nm from Scherrer analysis257. On the other hand, no diffraction patterns were observed

in GnP and CB compared to that in graphite. Inset in Figure 6.2 shows a small (002) peak in

GnP, corresponding to a particle size of ~35 nm. Assuming a perfect interlayer spacing of 3.37 Å

in graphene layers of ~1 nm thickness, a particle size of 35 nm shows approximately 24

graphene sheets stacked in one GnP particle. Generally, the graphene particle containing less

than 10 graphene sheets are considered high quality graphene258 for polymer reinforcement

applications. However, the large particle size indicates that this GnP might not be as effective as

high quality graphene produced by other methods; e.g., graphene synthesized by thermal

exfoliation and reduction of GO contains ~4-5 graphene layers28, 38.

128
CB

Intensity (a.u.)
GnP

26.4°
200
26.4°
100

0
20 22 24 26 28 30

Graphite

2 (°)
5 10 15 20 25 30 35 40

Figure 6.2 XRD patterns of graphite, GnP, and CB

6.3.2 Microstructure and Melt Rheology

Rheological properties of polymers are considered pertinent in designing new composite

materials. Understanding the flow behavior of nanocomposites gives direct insight into the

processing properties, and the dispersion of fillers. The rheological properties of PE/GnP

nanocomposites, and PE/CB composites were obtained within the LVE range (Figure 6.3). With

increasing GnP concentration in PE, the initial G' also increased. At 10 wt% GnP loading, a

quasi-asymptomatic behavior was observed where G' was nearly independent of the frequency.

Similarly, the initial G' increased with increasing CB loading. However, the increment was not as

significant as that was in PE/GnP nanocomposites. Strong network formation in PE/GnP

nanocomposites was observed at 10 wt% of GnP loading, and that in PE/CB composites

observed at 20 wt% of CB loading.

129
6 6
10 (a) 10 (b)
5 Increasing GnP wt% 5 Increasing CB wt%
10 10
G (Pa)

G (Pa)
4 4
10 10
Neat PE Neat PE
3 PE/GnP 0.5 wt% 3 PE/CB 0.5 wt%
10 PE/GnP 1.0 wt%
10 PE/CB 1.0 wt%
PE/GnP 1.5 wt% PE/CB 2.5 wt%
2 PE/GnP 2.5 wt% 2 PE/CB 4.3 wt%
10 PE/GnP 5.0 wt% 10 PE/CB 6.5 wt%
PE/GnP 10.0 wt% PE/CB 10.0 wt%
PE/GnP 15.0 wt% PE/CB 20.0 wt%
1 1
10 -3 -2 -1 0 1 2 10 -3

 (s )
-2 -1 0 1 2

 (s )
10 10 10 10 10 10 10 10 10 10 10 10
-1 -1

Figure 6.3 Rheological properties of PE/GnP nanocomposites (a), and PE/CB composites (b).

In order to understand the synergic effects of CB in PE/GnP nanocomposites, two

variables were selected: 1) GnP/CB ratio, and 2) total concentration of hybrid filler (Figure 6.4).

For the first variable, three representative GnP/CB ratios were used: a) 1:1, b) 1:5, and c) 1:10.

And, for the second variable, the concentration was varied from 0 to 15 wt%. The main objective

in varying the GnP/CB ratio was to fabricate conductive nanocomposites with least amount of

GnP.

Increasing concentration of nanofiller increases the rheological properties of

nanocomposites. An exponential increase in the initial G' defines the establishment of filler

network inside the polymers: an essential requirement in nanocomposites showing the effective

dispersion of filler and filler/polymer interactions. In GnP/CB 1:1 hybrid nanocomposites

(Figure 6.4a), incorporation of a small amount of the hybrid filler (1 wt%) increased the initial G'

at 0.01 Hz frequency from 42 Pa for neat PE to 191 Pa for the nanocomposites. The exponential

growth in G' was observed at 10 wt%, attributed to the network formation. In GnP/CB 1:5

(Figure 6.4b), and 1:10 (Figure 6.4c) hybrid nanocomposites, the initial increase in G' was not as

significant as that was in GnP/CB 1:1 nanocomposites at 1 wt% hybrid filler. However, the

130
rheological network formation was observed at 10 wt% hybrid filler in all the hybrid

nanocomposites. Thus, the network formation was not affected by changing GnP/CB ratio, and is

a function of filler concentration.

Since, incorporating nanofillers in polymers increases the viscosity of nanocomposites to

several orders of magnitude. The increased viscosity further indicates increased energy

requirements while processing these nanocomposites. For this purpose, the complex viscosity of

nanocomposites was compared to analyze the effect of hybrid filler on the processing properties.

Neat PE showed a non-Newtonian shear thinning viscosity profile where the viscosity

decreased with increasing frequency. Incorporation of 5 wt% GnP increased the complex

viscosity approximately 0.5 order of magnitude compared to that of PE at 0.01 Hz frequency.

Sine CB is a fractal filler which does not provide sufficient reinforcement effects, the complex

viscosity of PE/CB composites at 5 wt% CB loading was approximately double to that of neat

PE.

In hybrid fillers, the GnP:CB 1:1 hybrid/PE nanocomposites showed improved properties

where CB particles anchored to GnP sheets (later in Figure 6.6). The complex viscosity of

hybrid/PE nanocomposites was in between that of PE/GnP and PE/CB composites. Interestingly,

the viscosity of hybrid composites was closer to that of PE/CB composites which shows that

these composites can be processed at conditions similar to those for CB composites. The low

viscosity of hybrid nanocomposites was lower than that of PE/GnP nanocomposites which also

indicates ease of processing of these nanocomposites.

131
6 6
10 (a) 10 (b)
5 5
10 10

G (Pa)
G (Pa)

4 4
10 10
3 PE/Filler wt/wt 3 PE/Filler wt/wt
10 100/0 10 100/0
99/1 99/1
2 97/3 2 97/3
10 95/5 10 95/5
90/10 90/10
GnP/CB 1:1 85/15 1
GnP/CB 1:5 85/15
1
10 -3 -2 -1 0 1 2
10 -3 -2 -1 0 1 2

 (s )  (s )
10 10 10 10 10 10 10 10 10 10 10 10
-1 -1

6
10 (c)
5
10
G (Pa)

4
10
3 PE/Filler wt/wt
10 100/0
99/1
2 97/3
10 95/5
90/10
GnP/CB 1:10 85/15
1
10 -3 -2 -1 0 1 2

 (s )
10 10 10 10 10 10
-1

Figure 6.4 Effects of total filler concentration, and GnP/CB ratio on rheological properties of
PE/GnP/CB nanocomposites.

The arrangement of GnP sheets and CB particles in hybrid filler was further analyzed

using TEM imaging of the dried hybrid dispersions with different GnP/CB ratios (Figure 6.6). At

lower CB concentration in GnP/CB hybrid suspension, CB particles were observed being

anchored by the graphene sheets (Figure 6.6a). Increasing concentration of CB produced large

micron-scale agglomerates (Figure 6.6b). Interestingly, at GnP/CB ratio of 1:10, CB lost its

particle identity, and spread over graphene sheets like a viscous liquid.

132
5
10
GnP Total Filler = 5 wt%
1:1 Hybrid
CB

 , Pa-s
4
10
PE
*

3
10

, s
-2 -1 0 1 2
10 10 10 10 10
-1

Figure 6.5 Comparison of viscosity of composites

The morphology of GnP/CB hybrid suspensions is in agreement with the rheological

analysis (Figure 6.4). Due to the anchoring nature of CB particles at lower concentrations, the

initial increment in G' for GnP/CB 1:1 hybrid filler was higher compared to that in 1:5, and 1:10

hybrid filler/PE nanocomposites. However, increasing concentration of hybrid filler of any

GnP/CB ratio might produced agglomerates of CB. That’s why, the network forming

concentration of the hybrid filler was not affected by the GnP/CB ratio.

At this time, we do not have an explanation for the liquid-like merging behavior of CB

particles in GnP/CB 1:10 hybrid filler. More studies into the concentration stabilization of GnP

with varying concentration of CB might shed some light on this behavior.

The rheological analyses provide overall averaged microstructure in nanocomposites

below and above percolation65, 219. On the other hand, structural techniques such as small angle

x-ray scattering (SAXS) can show microstructure at individual compositions in nanocomposites.

SAXS measures the shape and size of polymers between 5 and 25 nm of the repeat distances 171.

133
The scattering vector (q) probes different structural features at various length scales (length scale

~ βπ/q) in polymers and polymer nanocomposites. In graphene-based nanocomposites, SAXS-

based morphology of nanocomposites can be correlated with the dispersion of graphene in

nanocomposites (Chapter 5).

(a) (b) (c)

1µm 1µm 1µm

Figure 6.6 TEM micro-images of GnP/CB suspensions with different GnP/CB ratios: a) 1:1, b)
1:5, and c) 1:10.

In the log-log intensity versus scattering vector plots, initial decaying exponent (α) at

lower q indicates the size of the scattering objects. The decaying exponent (α) is directly related

to the fractal dimensions, df of the scattering objects229. In general, for α ≤ γ, particles form mass

fractals (a substance whose surface and mass are both characterized by the fractal properties). In

mass fractals, α directly corresponds to fractal dimensions, dm. For γ ≤ α ≤ 4, surface fractals are

formed where the scattering is observed from a smooth surface. Fractal dimension, d s, is

calculated from α as followsμ α ≡ θ - ds where 3 < 6 - ds ≤ 4229, 230. Thus, ds varies from 3 to 2,

where ds ~ β is for a smooth surface approaching Porod's scattering (α = 4)172, and ds ~ 3

indicates the uneven concentration of scattering objects at the scattering surface231.

Pure PE exhibited α = β.γ, typical for randomly oriented polymer chains 172 (chapter 5).

We recently showed the increased dispersion of graphene in PE elucidated by decreasing

decaying exponent (α), where I ~ q-α log-log analyses hold for low q range. Two representative

nanocomposites were selected for SAXS analysis: below and above the rheological percolation

134
limit (Figure 6.7). For all nanocomposites, α > γ was observed, signifying the formation of

surface fractals. Using Malekani et al. analysis of fractal size, the surface fractal dimensions are

calculated as ds = 6 - α. GnP/CB 1μ1 hybrid showed ds similar to that for PE/GnP

nanocomposites. The ds increased for 1:5, and 1:10 hybrids (Figure 6.7a). Increased ds in

GnP/CB 1:5, and 1:10 nanocomposites might indicate the uneven agglomeration of CB particles

in these hybrid nanocomposites.

Increasing hybrid filler loading to 15 wt% (Figure 6.7b) showed slight decrease in α for

PE/GnP, and PE/CB composites compared to that at 5 wt% loadings (Figure 6.7a). Similarly, the

GnP/CB 1:1 showed the same exponent at 15 wt% filler loading as was observed at 5 wt% filler

loading. The exponent α decreased for GnP/CB 1μη nanocomposites from γ.4 at η wt% to γ.β at

1η wt%, whereas α increased from γ.η at η wt% to γ.θ at 1η wt% filler loading for GnP/CB 1:10

hybrid nanocomposites. However, the change in α is not significant compared to increased

concentration of fillers, and they might be indicating no change in dispersion state at two filler

loadings.

3
10 4
GnP:CB 1:5
10
(a) = 3.2 (b)
'8-PECB15_01596_00001'
'7-PEGnP15_01587_00001'
2 GnP/CB 1:10 3 '19-H15-15_01634_00001'
10 GnP:CB 1:10
= 3.5 10 = 3.6
'24-H110-15_01647_00001'
'14-H11-15_01619_00001'

1 GnP/CB 1:5 2
10 = 3.4 10
I (q) [a.u.]

PE/GnP
I(q) [a.u.]

1
0 10 = 3.0
10
0
-1 10
10 PE/CB
-1
10 = 3.0
-2 PE/GnP
10 = 3.1 -2
GnP/CB 1:1 10
-3 '5-PEGnP5_01585_00001'
= 3.2
10 '12-H11-5_01617_00001'
'17-H15-5_01627_00001' 10
-3
'22-H110-5_01649_00001'

-4 -4
10 5 6 7 2 3 4 5 6 7 2 3 10 5 6 7 2 3 4 5 6 7 2 3
0.01 0.1 0.01 0.1
q [1/Å] q [1/Å]

Figure 6.7 SAXS I~q log-log plot for PE/hybrid filler nanocomposites at a filler loading of a) 5
wt%, and b) 15 wt%

135
6.3.3 Electrical Conductivity

Electrically conductive polymer nanocomposites are used in a variety of applications.

The electrical conductivity is a function of dispersion of filler and its aspect ratio were the

effective dispersion depends on filler/polymer and inter-filler interactions. Significant

improvement in electrical conductivity of nanocomposites (measured at 0.01 Hz) was observed

as the filler concentration increased from 0 to 15 wt%. Neat PE is an insulating polymer with

conductivity ~ 7.4 x 10-14 S/cm. Incorporation of GnP increased the conductivity to ~10-3 S/cm at

15 wt% graphene loading. On the other hand, a gradual increase in electrical conductivity was

observed in PE/CB composites, where the conductivity reached its peak value at 10 wt% CB

loading, followed by a slight decrease at 15 wt%. The decreased conductivity might be

associated with the highly fractal forming and agglomerating nature of the CB particles.

In order to comment on the efficacy of GnP, a comparison between GnP and hybrid

nanocomposites is provided below. PE/GnP nanocomposites showed a conductivity of ~9x10-4

S/cm at 5 wt% GnP loading. The GnP/CB 1:1 hybrid/PE nanocomposites at 10 wt% total filler

loading exhibited conductivity of ~2x10-1 S/cm. It is worth noting that 10 wt% of 1:1 hybrid

filler contains 5 wt% GnP and 5 wt% CB. Thus, incorporating equal amounts of GnP and CB in

PE increased the electrical conductivity by about 3 orders of magnitude compared with that of

GnP and CB (Figure 6.9) at 5 wt% individual loading.

On the other hand, strong agglomeration of CB was observed in GnP/CB 1:10 hybrid/PE

nanocomposites. The conductivity of hybrid nanocomposites at GnP/CB ratios of 1:5 and 1:1

were very close, showing less agglomeration of GnP/CB in these nanocomposites.

136
1.E+00
Total Hybrid= 10 wt%
1.E-01

Conductivity, S/cm
1.E-02

1.E-03

1.E-04

1.E-05

1.E-06

1.E-07

1.E-08
PE/GnP 5 1:10 Hybrid 1:5 Hybrid 1:1 Hybrid
wt%

Figure 6.8 Comparison of electrical conductivities of GnP and GnP/CB hybrid nanocomposites

Increasing concentration of hybrid filler in PE increased the electrical conductivity of the

nanocomposites. However, the conductivity is observed to be a function of GnP/CB ratio (Figure

6.9). In GnP/CB 1:1 hybrid nanocomposites, a sharp increase in conductivity was observed

between 5 and 10 wt% total filler loading, showing the electrical percolation limit as ~7.5 wt%.

Similar observations were observed for GnP/CB 1:5 hybrid filler ratio. On the other hand, the

GnP/CB 1:10 hybrid filler showed no electrical conductivity at lower concentrations, whereas

sharp increase in conductivity was observed between 10-15 wt% showing the electrical

percolating limit at ~12.5 wt%. The high percolation concentration in GnP/CB 1:10

nanocomposites might be associated with high agglomeration of CB particles, and conversion of

CB into viscous liquid-like appearance on the GnP nanosheets (Figure 6.6). Further insights into

the liquid-like merging of CB on GnP might shed light on this behavior, and help find a better

explanation.

137
2
10
PE/GnP
0 PE/CB
10 1:1 Hybrid
1:5 Hybrid

 [S/cm]
-2 1:10 Hybrid
10
-4
10
-6
10
-8
10
-10
10
0 5 10 15 20
Total Filler wt%
Figure 6.9 Electrical conductivity of nanocomposites.

Conclusions

The concept of manufacturing conductive nanocomposites using graphene/CB hybrid

nanofiller in nonpolar polyolefin matrix is a new and technologically challenging area. This

study is the first report on producing hybrid filler/PE nanocomposites. Dispersion of hybrid filler

was studied using rheological properties, electrical conductivity, and small angle scattering. The

microstructure formed inside the nanocomposites showed that hybrid fillers formed surface

fractals inside the polymers. Rheological properties were largely unaffected by GnP/CB ratio at

higher hybrid filler concentrations, attributed to the agglomerating nature of CB particles. On the

other hand, at lower concentrations, significant improvement in rheological properties was

observed in GnP/CB 1:1 nanocomposites. The electrical percolation threshold was also observed

to be a function of GnP/CB ratio, where the percolation limit decreased with increasing GnP/CB

ratio.

138
Acknowledgement

This study is supported by The Petroleum Institute, Abu Dhabi through the Cooperative

Research Partnership with Colorado School of Mines. Authors thank U.S. Army Research Office

(DURIP Grant No.W911NF-11-1-0306) for partial financial support. Prof. John Dorgan is

acknowledged for his laboratory facilities. Authors also thank Tara Pandey and Andrew Motz at

Colorado School of Mines, and Sonke Seifert at Argonne National Laboratory for assisting with

SAXS measurements. Use of the Advanced Photon Source, an office of Science User facility

operated for the U.S. Department of Energy (D.O.E.) Office of Science by Argonne National

Laboratory, was supported by the U.S. DOE under Contract No. DE-AC02-06CH11357.

139
CHAPTER-7

CONCLUSIONS AND FUTURE RECOMMENDATIONS

This thesis is a fundamental study on synthesis and characterization of composites made

from nanoscale carbon structures, especially TRG. A number of open questions in

nanocomposites research were explored and raised some new questions. The overall summary of

the thesis is presented in this chapter. In addition, new insights for the future research in the field

of nanocomposites are also provided.

7.1 Summary and Conclusions

TRG was successfully synthesized and characterized for its physical and chemical

properties. The residual oxy-functional groups on TRG were utilized for attaching the amino-

silane on TRG surface. A simple method to attach silane onto TRG surface was demonstrated

which allowed the grafting of more than 7-8 atomic% silane on graphene surface, more than

what was reported earlier. A fundamental understanding of the attachment mechanism of amino-

silane on graphene surface was demonstrated with and without the use of an organic solvent. The

described method can be further used for selective edge-functionalization of TRG by selecting a

suitable organic solvent.

Since, polyolefins are comprised of highly nonpolar structure, the dispersion of

nanofillers such as graphene and carbon nanotubes in polyolefins is considered difficult. One

serious disadvantage of the poor dispersion is filler agglomeration, due to which the potential

reinforcements in nanocomposites cannot be achieved. In this thesis, rather than using the

functionalization approach to solubilise graphene in PE, blends of PE with OPE were

140
synthesized via solvent blending approach. OPE was found miscible in PE whereas the

miscibility decreased with increasing OPE concentration in the blends. In molten state, OPE

acted as a processing aid for optimizing the processing properties of PE; whereas in the solid

state, inclusion of OPE increased the mechanical properties of PE. Being amorphous in nature,

OPE resided within the lamellae of PE, and increased PE's crystallinity which eventually

attributed in increasing the mechanical properties of the blends.

TRG was also incorporated in PE/OPE blends to manufacture polymer nanocomposites

via solvent blending, and the dispersion of TRG was quantified. Remarkable processing and

mechanical properties were achieved when nanocomposites were manufactured with PE/OPE

blends compared to neat PE/TRG nanocomposites. The dispersion of TRG in PE/OPE blends

was substantially increased, marked by the significant reduction in rheological and electrical

percolation limits. Along with the reduction in percolation thresholds, viscosity of PE/OPE/TRG

nanocomposites was considerably lower than that of PE/TRG nanocomposites which was

remarkable since nanofillers such as graphene increase the viscosity of nanocomposites many

orders of magnitude. Higher viscosity, however, is a big disadvantage in the nanocomposite

industry since higher viscosity means more energy requirements in processing these materials.

The microstructure formed by TRG inside the nanocomposites investigated by SAXS

confirmed the formation of surface fractals when high dispersion was achieved (blends/TRG)

whereas only the mass fractals were observed when graphene was not uniformly dispersed inside

the polymer matrix (PE/TRG). The SAXS results were in excellent agreement with rheology,

mechanical and conductive properties. The contextualization of microstructure with the

performance properties of nanocomposites further helped in understanding the "slippery

behaviour" in graphene-based nanocomposites reported in many publications. In addition to

141
manufacturing mechanically robust and electrically conductive nanocomposites, fundamental

questions on the dispersion mechanism of graphene in polymers were addressed. This thesis (and

related publication) was the first report showing that TRG followed the Doi-Edwards theory of

dispersion in polymers, which was reported only for carbon nanotubes-based nanocomposites

previously.

Procuring small amounts of high quality graphene (less than 10 number of graphene

layers stacked in one graphene particle) being very costly and due to the large volumes of

graphene required in nanocomposites' mass productions, new methods to lower the overall cost

of nanocomposites were explored. Therefore, a commercial grade graphene called "graphene

nanoplatelets (GnP)" was also used in manufacturing conductive PE nanocomposites. Due to

large particles of size ~ 35 nm (~25 graphene layers stacked in one graphene particle), GnP was

not as effective as TRG for reinforcing PE (TRG is a highly quality graphene containing 4-5

graphene layers in one particle). CB was investigated as a synergic filler for GnP in PE

nanocomposites to understand the effects of GnP/CB ratio, and the total concentration of hybrid

filler on the dispersion. Higher electrical conductivities were achieved with GnP/CB 1:1 ratio at

15 wt% filler loading, compared to that of individual fillers or GnP/CB ratios of 1:5 and 1:10.

Overall, my hypothesis that TRG could be better dispersed in PE/OPE blends has proven

true. The new polyolefin matrix, PE/OPE miscible blend, proved better in manufacturing

mechanically robust, and electrically conductive nanocomposites at lower TRG loading

compared to that required for manufacturing PE/TRG nanocomposites. A fundamental

understanding was developed for the processing properties of PE/OPE blends and PE/OPE/TRG

nanocomposites. Moreover, a new concept of synergic effects of graphene and CB was tested.

The results indicate that CB could be used to increase graphene's dispersion in PE

142
nanocomposites. This idea can be very helpful in reducing the overall cost associated with

graphene-based polymer nanocomposites because CB is inexpensive and abundantly available. A

GnP/CB 1:1 ratio exhibited the best results in terms of electrical conductivity which shows that

approximately 50% cost of the filler can be reduced if CB is used along with graphene.

In addition to the work presented in this thesis, applications of TRG in environmental

science were also explored in collaborative research. The synthesis conditions for TRG

production were altered to change the residual C/O ratio, and other physical properties of TRG

such as the bulk density and surface area, and further used to adsorb spilled crude oil from the

water surface88. TRG showed an adsorption capacity of (~300 g-oil/g-TRG) for crude oil which

is the highest among the reported adsorbents so far. The adsorption behaviour was further

correlated with the physical and chemical properties of TRG. TRG was also used to remove the

harmful dyes from the colored wastewaters37. In another report, graphene oxide (graphite oxide

being exfoliated in water) was used as a thermal fluid for heat transfer applications34.

A comparison of the best results obtained in this thesis provided in Figure 7.1. The

electrical conductivity of PE was increased by approximately 1 order of magnitude when

PE/OPE 60/40 blend was used. The conductivity was substantially increased to ~1 S/cm when

GnP/CB 1:1 hybrid filler used in manufacturing nanocomposites.

7.2 Future Research Directions

Polymer nanocomposites are used in a variety of applications ranging from high-tech and

high performance materials to household daily appliances. Nanocomposites are a very rapidly

growing field, and there are many opportunities to pursue this field in future.

143
During the course of this thesis, a number of new alterations/ideas were identified to

improve the main goals delineated in the beginning. Due to the time constraints and instrument

limitations at the time, not all the ideas could be pursued properly. In the following sections, a

few strategies to improve the current research are discussed, and platforms the applications of

this research are suggested.

1
10
Conductivity, S/cm

0
10
-1
10
-2
10
-3
10
-4
10
15 wt%
5 wt%
5 wt%

-5
10
PE/TRG 60/40/TRG 1:1 Hybrid

Figure 7.1 Comparison of electrical conductivity of nanocomposite

In chapter 3, aminopropyl triethoxy silane (APTS) was attached onto TRG surface, and a

mechanism of APTS reaction with TRG was demonstrated with the help of various

characterization techniques. APTS has three ethoxy groups which were observed self

crosslinking on TRG surface; making the mechanism of attachment very difficult to understand

by inducing more structural complexities. APTS was selected in this work because APTS silane-

modified fillers have been reported in many published reports.

144
However, I believe a better structural understanding of the attachment mechanism of

silane can be developed if simpler silanes such as mono- and di-ethoxy silanes are used. The

monoethoxy silane represents the simplest member in the silane family. Also, TRG contains

residual hydroxyl, epoxy, and carboxylic acid functional groups on its surface. For clean

attachment, hydroxyl and epoxy groups may be caped so that only carboxylic acid groups would

be accessible for silane attachment. The availability of only one type of anchoring groups on

TRG surface will give single site for attaching monoethoxy silane. In addition, chapter 3

concluded that several organic solvents can be used for selective edge functionalization of TRG

with silane groups. Thus, a good starting point for doing research in this area will be to

synthesize carboxylated graphene, and use monoethoxy silane in the presence of various organic

solvents to study the silane attachment, quantitatively.

Later, PE/OPE miscible blends were synthesized via solvent blending where OPE acted

as a processing control agent. The OPE used in this study was a commercial product obtained

from the vendors directly. A first step in moving forward in this area will be to use melt blending

for fabricating PE/OPE blends, and ascertain that the results from solvent and melt blending are

similar or different. Another area to explore can be the synthesis of blends of PP with oxidized

PP (OPP) using the same methods to understand if PP/OPP blends show similar behaviour.

Since, PE and PP have structural similarities, creating miscible PP/OPP blends can also be

advantageous for nanocomposites applications.

The PE/OPE/TRG nanocomposites showed excellent rheological, electrical, and

mechanical properties. However, the nanocomposites were prepared by solvent blending which

is not as famous as melt blending for mass production of composite materials. Therefore,

corroborating solvent blending results with melt blended composites will be highly desired. A

145
facile application of this study can be in manufacturing nanocomposite packaging materials for

food and electronic packaging applications. Plate-like fillers are already established for

increasing the gas barrier of polymers259, and therefore, determining the gas barrier of

PE/OPE/TRG nanocomposites will be a very versatile area to look into. A pictorial

representation of the above suggestions is also provided in Figure 7.2260, 261, 262.

Micro-compounder

Melt
Graphene pressing

Solvent
Polymer casting

Mechanical
SAXS sample detector
Barrier Characterization
Properties
Monochromatic Beam 2ϴ q
~ 0.1-0.2nm
log (Intensity)

log (q), 1/nm or 1/Å

Figure 7.2 Schematics for future research plan in nanocomposites research

While, during the course of this thesis, several questions about the dispersion mechanism

of graphene in polymers were addressed; some new questions were also raised. For example, we

still do not understand the reasons for highly fibrous behaviour of PE/OPE 70/30 composition.

The peculiar behavior was also corroborated with other characterization techniques, but we could

not find any suitable explanation at the moment. Thus, focusing on the peculiarity of 70/30

blends might be very interesting.

146
In PE/OPE/TRG nanocomposites, 60/40 blend-TRG nanocomposites exhibited viscosity

lower than that of neat 60/40 blends, which was very unusual than what is expected in

nanocomposites. Generally, the nanocomposites' viscosity increases with gradual addition of

nanofiller. This kind of unusual behaviour is known as the non-Einstein behaviour in

nanocomposites71, 72. The non-Einstein behaviour is exhibited by filler particles having size less

than the radius of gyration of the polymer. In this thesis, TRG particles were comprised of 4-5

graphene layers (~3 nm thick) and extended over ~1µm in lateral dimensions. Incorporation of

TRG in polyolefins usually leads to the agglomeration of TRG, increasing the particle size, and

eventually no dimension of the agglomerate would be smaller than the radius of gyration of

polyolefins. Since, 60/40 blends showed highly improved TRG dispersion, individual graphene

particle might have one dimension smaller than the radius of gyration. In addition, the viscosity

of 60/40 blends was ~1 order of magnitude less than that of neat PE. That's why, 60/40 blends-

TRG nanocomposites showed the non-Einstein behaviour at very low loadings of TRG. Further

studies into the low viscosity model polymers such as PDMS, and incorporation of single sheet

graphene will be extremely helpful in understanding the non-Einstein behaviour of graphene in

polymer nanocomposites.

The synergic effects of graphene and CB in PE nanocomposites, and how CB affects

graphene's dispersion are some open questions not addressed in the published literature. The

main element in these questions is the packing of GnP and CB to produce highly conductive

nanocomposites. Higher loadings of GnP/CB hybrid fillers not only induce higher electrical

conductivity in nanocomposites, but can also produce thermally conductive nanocomposites.

Introducing thermal conductivity in nanocomposites will open up a new area of applications of

these materials. The next step in this research might be to understand thermal conductivity of

147
PE/hybrid filler nanocomposites, and correlate thermal conductivity with packing of GnP and

CB together inside PE. In addition, hybrid nanocomposites were fabricated via solvent blending.

Thus, using melt blending to manufacture hybrid nanocomposites, and quantifying filler

dispersion can be another way of moving forward in hybrid nanocomposites.

The percolation laws are used in rheological scaling analysis of nanocomposites (Chapter

5). The absence of the percolation laws dealing with packing of hybrid fillers offers a big gap in

hybrid filler nanocomposites research. The percolation laws stem from the solution rheology of

dispersions, and then applied to nanocomposites by assuming if nanoparticles are following

similar packing trends in nanocomposites as they were in colloidal assembly. Therefore, a certain

area of research can be investigating the colloidal assembly of GnP/CB hybrid assembly, and

applying these laws to hybrid nanocomposites.

Another important area of research that has not been explored in this thesis is the

extensional viscosity of nanocomposites. The uniaxial extension of thin nanocomposites films

under melt conditions is considered important for packaging applications. These experiments can

be easily performed on an ARES G2 extensional rheometer. Important parameters such as melt

strength having major importance in packaging material design can also be deduced from

extensional rheology, which could further be used to understand blowing mechanism of these

nanocomposites in manufacturing packaging films.

148
REFERENCE

1. Cowie, J. M. Polymers: chemistry and physics of modern materials; CRC Press1991.

2. Billmeyer, F. W. Textbook of polymer science. 1971.

3. Brandrup, J.; Immergut, E. H.; Grulke, E. A.; Abe, A.; Bloch, D. R. Polymer handbook;
Wiley New York1999; Vol. 1999.

4. Sinha Ray, S.; Okamoto, M. Polymer/layered silicate nanocomposites: a review from


preparation to processing. Progress in Polymer Science 2003, 28 (11), 1539-1641.

5. Fukushima, Y.; Inagaki, S.; Kamigaito, O.; Kawasumi, M.; Kurauchi, T.; Okada, A.;
Sugiyama, S.; Usuki, A. Composite material and process for manufacturing same. Google
Patents, 1988.

6. Fu, X.; Qutubuddin, S. Polymer–clay nanocomposites: exfoliation of organophilic


montmorillonite nanolayers in polystyrene. Polymer 2001, 42 (2), 807-813.

7. Cao, X.; James Lee, L.; Widya, T.; Macosko, C. Polyurethane/clay nanocomposites
foams: processing, structure and properties. Polymer 2005, 46 (3), 775-783.

8. Durmus, A.; Kasgoz, A.; Macosko, C. W. Linear low density polyethylene (LLDPE)/clay
nanocomposites. Part I: Structural characterization and quantifying clay dispersion by
melt rheology. Polymer 2007, 48 (15), 4492-4502.

9. Durmuş, A.; Woo, ε.; Kaşgöz, A.; εacosko, C. W.; Tsapatsis, ε. Intercalated linear low
density polyethylene (LLDPE)/clay nanocomposites prepared with oxidized polyethylene
as a new type compatibilizer: structural, mechanical and barrier properties. European
Polymer Journal 2007, 43 (9), 3737-3749.

10. Hotta, S.; Paul, D. Nanocomposites formed from linear low density polyethylene and
organoclays. Polymer 2004, 45 (22), 7639-7654.

11. Iijima, S. Helical microtubules of graphitic carbon. Nature 1991, 354 (6348), 56-58.

12. Wang, C.; Guo, Z.-X.; Fu, S.; Wu, W.; Zhu, D. Polymers containing fullerene or carbon
nanotube structures. Progress in Polymer Science 2004, 29 (11), 1079-1141.

13. Ebbesen, T.; Ajayan, P. Large-scale synthesis of carbon nanotubes. Nature 1992, 358
(6383), 220-222.

149
14. Bauhofer, W.; Kovacs, J. Z. A review and analysis of electrical percolation in carbon
nanotube polymer composites. Composites Science and Technology 2009, 69 (10), 1486-
1498.

15. Al-Saleh, M. H.; Sundararaj, U. A review of vapor grown carbon nanofiber/polymer


conductive composites. Carbon 2009, 47 (1), 2-22.

16. Sun, X.; Sun, H.; Li, H.; Peng, H. Developing Polymer Composite Materials: Carbon
Nanotubes or Graphene? Advanced Materials 2013, 25 (37), 5153-5176.

17. Adhikari, A. R.; Chipara, M.; Lozano, K. Processing effects on the thermo-physical
properties of carbon nanotube polyethylene composite. Materials Science and
Engineering: A 2009, 526 (1), 123-127.

18. Song, Y. S.; Youn, J. R. Influence of dispersion states of carbon nanotubes on physical
properties of epoxy nanocomposites. Carbon 2005, 43 (7), 1378-1385.

19. Stéphan, C.; Nguyen, T. P.; de la Chapelle, M. L.; Lefrant, S.; Journet, C.; Bernier, P.
Characterization of singlewalled carbon nanotubes-PMMA composites. Synthetic Metals
2000, 108 (2), 139-149.

20. Jin, S. H.; Park, Y.-B.; Yoon, K. H. Rheological and mechanical properties of surface
modified multi-walled carbon nanotube-filled PET composite. Composites Science and
Technology 2007, 67 (15-16), 3434-3441.

21. Bartholome, C.; Miaudet, P.; Derré, A.; Maugey, M.; Roubeau, O.; Zakri, C.; Poulin, P.
Influence of surface functionalization on the thermal and electrical properties of
nanotube-PVA composites. Composites Science and Technology 2008, 68 (12), 2568-
2573.

22. Valentino, O.; Sarno, M.; Rainone, N. G.; Nobile, M. R.; Ciambelli, P.; Neitzert, H. C.;
Simon, G. P. Influence of the polymer structure and nanotube concentration on the
conductivity and rheological properties of polyethylene/CNT composites. Physica E:
Low-dimensional Systems and Nanostructures 2008, 40 (7), 2440-2445.

23. Xiao, K.; Zhang, L.; Zarudi, I. Mechanical and rheological properties of carbon
nanotube-reinforced polyethylene composites. Composites Science and Technology 2007,
67 (2), 177-182.

24. Hsiao, A.-E.; Tsai, S.-Y.; Hsu, M.-W.; Chang, S.-J. Decoration of multi-walled carbon
nanotubes by polymer wrapping and its application in MWCNT/polyethylene
composites. Nanoscale research letters 2012, 7 (1), 1-5.

150
25. Shanmugharaj, A.; Bae, J.; Lee, K.; Noh, W.; Lee, S.; Ryu, S. Physical and chemical
characteristics of multiwalled carbon nanotubes functionalized with aminosilane and its
influence on the properties of natural rubber composites. Composites Science and
Technology 2007, 67 (9), 1813-1822.

26. Yang, B.-X.; Shi, J.-H.; Pramoda, K.; Goh, S. H. Enhancement of the mechanical
properties of polypropylene using polypropylene-grafted multiwalled carbon nanotubes.
Composites Science and Technology 2008, 68 (12), 2490-2497.

27. Gong, L.; Kinloch, I. A.; Young, R. J.; Riaz, I.; Jalil, R.; Novoselov, K. S. Interfacial
stress transfer in a graphene monolayer nanocomposite. Advanced Materials 2010, 22
(24), 2694-2697.

28. Kim, H.; Abdala, A.; Macosko, C. Graphene/Polymer Nanocomposites. Macromolecules


2010, 43 (16), 6515-6530.

29. Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J.; Potts, J.; Ruoff, R. Graphene and graphene
oxide: Synthesis, properties, and applications. Advanced Materials 2010.

30. Novoselov, K.; Geim, A.; Morozov, S.; Jiang, D.; Zhang, Y.; Dubonos, S.; Grigorieva, I.;
Firsov, A. Electric field effect in atomically thin carbon films. Science 2004, 306 (5696),
666-669.

31. Iqbal, M. Z.; Abdala, A. A. Oil spill cleanup using graphene. Environmental Science and
Pollution Research 2012, 20 (5), 3271-3279.

32. Schniepp, H. C.; Li, J.-L.; McAllister, M. J.; Sai, H.; Herrera-Alonso, M.; Adamson, D.
H.; Prud'homme, R. K.; Car, R.; Saville, D. A.; Aksay, I. A. Functionalized single
graphene sheets derived from splitting graphite oxide. The Journal of Physical Chemistry
B 2006, 110 (17), 8535-8539.

33. McAllister, M.; Li, J.; Adamson, D.; Schniepp, H.; Abdala, A.; Liu, J.; Herrera-Alonso,
M.; Milius, D.; Car, R.; Prud'homme, R. Single sheet functionalized graphene by
oxidation and thermal expansion of graphite. Chem. Mater 2007, 19 (18), 4396-4404.

34. Tesfai, W.; Singh, P.; Shatilla, Y.; Iqbal, M. Z.; Abdala, A. A. Rheology and
microstructure of dilute graphene oxide suspension. Journal of nanoparticle research
2013, 15 (10), 1-7.

35. Wen, X.; Garland, C. W.; Hwa, T.; Kardar, M.; Kokufuta, E.; Li, Y.; Orkisz, M.; Tanaka,
T. Crumpled and collapsed conformation in graphite oxide membranes. 1992.

151
36. Iqbal, M. Z.; Katsiotis, M. S.; Alhassan, S. M.; Liberatore, M. W.; Abdala, A. A. Effect
of solvent on the uncatalyzed synthesis of aminosilane-functionalized graphene. RSC
Advances 2014, 4 (13), 6830-6839.

37. Iqbal, M.; Abdala, A. A. Thermally reduced graphene: synthesis, characterization and
dye removal applications. RSC Adv. 2013, 3 (46), 24455-24464.

38. Ramanathan, T.; Abdala, A.; Stankovich, S.; Dikin, D.; Herrera-Alonso, M.; Piner, R.;
Adamson, D.; Schniepp, H.; Chen, X.; Ruoff, R. Functionalized graphene sheets for
polymer nanocomposites. Nature nanotechnology 2008, 3 (6), 327-331.

39. Potts, J. R.; Dreyer, D. R.; Bielawski, C. W.; Ruoff, R. S. Graphene-based polymer
nanocomposites. Polymer 2011, 52 (1), 5-25.

40. Villar-Rodil, S.; Paredes, J. I.; Martinez-Alonso, A.; Tascon, J. M. D. Preparation of


graphene dispersions and graphene-polymer composites in organic media. J. Mater.
Chem. FIELD Full Journal Title:Journal of Materials Chemistry 2009, 19 (22), 3591-
3593.

41. Stankovich, S.; Piner, R.; Chen, X.; Wu, N.; Nguyen, S.; Ruoff, R. Stable aqueous
dispersions of graphitic nanoplatelets via the reduction of exfoliated graphite oxide in the
presence of poly (sodium 4-styrenesulfonate). Journal of Materials Chemistry 2006, 16
(2), 155-158.

42. Wakabayashi, K.; Pierre, C.; Dikin, D. A.; Ruoff, R. S.; Ramanathan, T.; Brinson, L. C.;
Torkelson, J. M. Polymer-graphite nanocomposites: effective dispersion and major
property enhancement via solid-state shear pulverization. Macromolecules 2008, 41 (6),
1905-1908.

43. Kim, H.; Kobayashi, S.; AbdurRahim, M. A.; Zhang, M. J.; Khusainova, A.; Hillmyer,
M. A.; Abdala, A. A.; Macosko, C. W. Graphene/polyethylene nanocomposites: Effect of
polyethylene functionalization and blending methods. Polymer 2011.

44. Hu, H.; Wang, X.; Wang, J.; Wan, L.; Liu, F.; Zheng, H.; Chen, R.; Xu, C. Preparation
and properties of graphene nanosheets-polystyrene nanocomposites via in situ emulsion
polymerization. Chemical Physics Letters 2010, 484 (4-6), 247-253.

45. Goh, H. W.; Goh, S. H.; Xu, G. Q.; Pramoda, K. P.; Zhang, W. D. Dynamic mechanical
behavior of in situ functionalized multi-walled carbon nanotube/phenoxy resin
composite. Chemical Physics Letters 2003, 373 (3-4), 277-283.

46. Chen, G.; Weng, W.; Wu, D.; Wu, C. PMMA/graphite nanosheets composite and its
conducting properties. European Polymer Journal 2003, 39 (12), 2329-2335.

152
47. Sengupta, R.; Bhattacharya, M.; Bandyopadhyay, S.; Bhowmick, A. K. A review on the
mechanical and electrical properties of graphite and modified graphite reinforced
polymer composites. Progress in Polymer Science 2011, 36 (5), 638-670.

48. Kim, H.; Miura, Y.; Macosko, C. W. Graphene/Polyurethane Nanocomposites for


Improved Gas Barrier and Electrical Conductivity. Chemistry of Materials 2010, 22 (11).

49. Kim, H.; Kobayashi, S.; Abdur Rahim, M. A.; Zhang, M. J.; Khusainova, A.; Hillmyer,
M. A.; Abdala, A. A.; Macosko, C. W. Graphene/polyethylene nanocomposites: Effect of
polyethylene functionalization and blending methods. Polymer 2011, 52 (8), 1837-1846.

50. Georgakilas, V.; Otyepka, M.; Bourlinos, A. B.; Chandra, V.; Kim, N.; Kemp, K. C.;
Hobza, P.; Zboril, R.; Kim, K. S. Functionalization of graphene: covalent and non-
covalent approaches, derivatives and applications. Chemical reviews 2012, 112 (11),
6156-6214.

51. Chua, C. K.; Pumera, M. Covalent chemistry on graphene. Chemical Society Reviews
2013.

52. Kuila, T.; Bose, S.; Mishra, A. K.; Khanra, P.; Kim, N. H.; Lee, J. H. Chemical
functionalization of graphene and its applications. Progress in Materials Science 2012,
57 (7), 1061-1105.

53. Bahr, J.; Yang, J.; Kosynkin, D.; Bronikowski, M.; Smalley, R.; Tour, J.
Functionalization of carbon nanotubes by electrochemical reduction of aryl diazonium
salts: a bucky paper electrode. J. Am. Chem. Soc 2001, 123 (27), 6536-6542.

54. Si, Y.; Samulski, E. Synthesis of water soluble graphene. Nano letters 2008, 8 (6), 1679-
1682.

55. Stankovich, S.; Piner, R.; Nguyen, S.; Ruoff, R. Synthesis and exfoliation of isocyanate-
treated graphene oxide nanoplatelets. Carbon 2006, 44 (15), 3342-3347.

56. Georgakilas, V.; Bourlinos, A. B.; Zboril, R.; Steriotis, T. A.; Dallas, P.; Stubos, A. K.;
Trapalis, C. Organic functionalisation of graphenes. Chemical Communications 2010, 46
(10), 1766-1768.

57. Quintana, M.; Spyrou, K.; Grzelczak, M.; Browne, W. R.; Rudolf, P.; Prato, M.
Functionalization of graphene via 1, 3-dipolar cycloaddition. ACS nano 2010, 4 (6),
3527-3533.

153
58. Yang, H.; Li, F.; Shan, C.; Han, D.; Zhang, Q.; Niu, L.; Ivaska, A. Covalent
functionalization of chemically converted graphene sheets via silane and its
reinforcement. J. Mater. Chem. 2009, 19 (26), 4632-4638.

59. Wang, X.; Xing, W.; Zhang, P.; Song, L.; Yang, H.; Hu, Y. Covalent functionalization of
graphene with organosilane and its use as a reinforcement in epoxy composites.
Composites Science and Technology 2012, 72 (6), 737-743.

60. Ma, P.; Kim, J.; Tang, B. Functionalization of carbon nanotubes using a silane coupling
agent. Carbon 2006, 44 (15), 3232-3238.

61. Gahleitner, M. Melt rheology of polyolefins. Progress in Polymer Science 2001, 26 (6),
895-944.

62. Dealy, J. M.; Wissbrun, K. F. Melt rheology and its role in plastics processing: theory
and applications; Springer1999.

63. Larson, R. G. The structure and rheology of complex fluids; Oxford university press New
York1999; Vol. 702.

64. Jeon, H.; Rameshwaram, J.; Kim, G.; Weinkauf, D. Characterization of polyisoprene–
clay nanocomposites prepared by solution blending. Polymer 2003, 44 (19), 5749-5758.

65. Vermant, J.; Ceccia, S.; Dolgovskij, M.; Maffettone, P.; Macosko, C. Quantifying
dispersion of layered nanocomposites via melt rheology. Journal of Rheology (1978-
present) 2007, 51 (3), 429-450.

66. Vonk, C.; Kortleve, G. X-ray small-angle scattering of bulk polyethylene. Kolloid-
Zeitschrift und Zeitschrift für Polymere 1967, 220 (1), 19-24.

67. Fischer, E. Small angle x-ray scattering studies of phase transitions in polymeric and
oligomeric systems. Pure and Applied Chemistry 1971, 26 (3-4), 385-422.

68. Chu, B.; Hsiao, B. S. Small-angle X-ray scattering of polymers. Chemical reviews 2001,
101 (6), 1727-1762.

69. Rieker, T. P.; Hindermann-Bischoff, M.; Ehrburger-Dolle, F. Small-angle X-ray


scattering study of the morphology of carbon black mass fractal aggregates in polymeric
composites. Langmuir 2000, 16 (13), 5588-5592.

70. Cattani, M.; Salvadori, M.; Teixeira, F. SAXS Structural Characterization of


Nanoheterogeneous Conducting Thin Films. A Brief Review of SAXS Theories. arXiv
preprint arXiv:0907.3131 2009.

154
71. Mangal, R.; Srivastava, S.; Archer, L. A. Phase stability and dynamics of entangled
polymer-nanoparticle composites. Nature communications 2015, 6.

72. Mackay, M. E.; Dao, T. T.; Tuteja, A.; Ho, D. L.; Van Horn, B.; Kim, H.-C.; Hawker, C.
J. Nanoscale effects leading to non-Einstein-like decrease in viscosity. Nature materials
2003, 2 (11), 762-766.

73. Wang, X.; You, H.; Liu, F.; Li, M.; Wan, L.; Li, S.; Li, Q.; Xu, Y.; Tian, R.; Yu, Z.;
Xiang, D.; Cheng, J. Large-Scale Synthesis of Few-Layered Graphene using CVD.
Chemical Vapor Deposition 2009, 15 (1-3), 53-56.

74. Li, N.; Wang, Z.; Zhao, K.; Shi, Z.; Gu, Z.; Xu, S. Large scale synthesis of N-doped
multi-layered graphene sheets by simple arc-discharge method. Carbon 2010, 48 (1),
255-259.

75. Rollings, E.; Gweon, G.-H.; Zhou, S.; Mun, B.; McChesney, J.; Hussain, B.; Fedorov, A.;
First, P.; De Heer, W.; Lanzara, A. Synthesis and characterization of atomically thin
graphite films on a silicon carbide substrate. Journal of Physics and Chemistry of Solids
2006, 67 (9), 2172-2177.

76. De Heer, W. A.; Berger, C.; Wu, X.; First, P. N.; Conrad, E. H.; Li, X.; Li, T.; Sprinkle,
M.; Hass, J.; Sadowski, M. L. Epitaxial graphene. Solid State Communications 2007, 143
(1), 92-100.

77. Yang, X.; Dou, X.; Rouhanipour, A.; Zhi, L.; Räder, H. J.; Müllen, K. Two-dimensional
graphene nanoribbons. Journal of the American Chemical Society 2008, 130 (13), 4216-
4217.

78. Kim, C.-D.; Min, B.-K.; Jung, W.-S. Preparation of graphene sheets by the reduction of
carbon monoxide. Carbon 2009, 47 (6), 1610-1612.

79. Kosynkin, D. V.; Higginbotham, A. L.; Sinitskii, A.; Lomeda, J. R.; Dimiev, A.; Price, B.
K.; Tour, J. M. Longitudinal unzipping of carbon nanotubes to form graphene
nanoribbons. Nature 2009, 458 (7240), 872-876.

80. Zhang, W.; Cui, J.; Tao, C. a.; Wu, Y.; Li, Z.; Ma, L.; Wen, Y.; Li, G. A Strategy for
Producing Pure Single‐Layer Graphene Sheets Based on a Confined Self‐Assembly
Approach. Angewandte Chemie 2009, 121 (32), 5978-5982.

81. Bourlinos, A. B.; Georgakilas, V.; Zboril, R.; Steriotis, T. A.; Stubos, A. K. Liquid‐phase
exfoliation of graphite towards solubilized graphenes. Small 2009, 5 (16), 1841-1845.

155
82. Hernandez, Y.; Nicolosi, V.; Lotya, M.; Blighe, F. M.; Sun, Z.; De, S.; McGovern, I. T.;
Holland, B.; Byrne, M.; Gun'Ko, Y. K.; Boland, J. J.; Niraj, P.; Duesberg, G.;
Krishnamurthy, S.; Goodhue, R.; Hutchison, J.; Scardaci, V.; Ferrari, A. C.; Coleman, J.
N. High-yield production of graphene by liquid-phase exfoliation of graphite. Nat.
Nanotechnol. FIELD Full Journal Title:Nature Nanotechnology 2008, 3 (9), 563-568.

83. Liu, N.; Luo, F.; Wu, H.; Liu, Y.; Zhang, C.; Chen, J. One‐step ionic‐liquid‐assisted
electrochemical synthesis of ionic‐liquid‐functionalized graphene sheets directly from
graphite. Advanced Functional Materials 2008, 18 (10), 1518-1525.

84. Behabtu, N.; Lomeda, J. R.; Green, M. J.; Higginbotham, A. L.; Sinitskii, A.; Kosynkin,
D. V.; Tsentalovich, D.; Parra-Vasquez, A. N. G.; Schmidt, J.; Kesselman, E.
Spontaneous high-concentration dispersions and liquid crystals of graphene. Nature
nanotechnology 2010, 5 (6), 406-411.

85. Worsley, K.; Ramesh, P.; Mandal, S.; Niyogi, S.; Itkis, M.; Haddon, R. Soluble graphene
derived from graphite fluoride. Chemical Physics Letters 2007, 445 (1-3), 51-56.

86. Stankovich, S.; Dikin, D.; Piner, R.; Kohlhaas, K.; Kleinhammes, A.; Jia, Y.; Wu, Y.;
Nguyen, S.; Ruoff, R. Synthesis of graphene-based nanosheets via chemical reduction of
exfoliated graphite oxide. Carbon 2007, 45 (7), 1558-1565.

87. Lomeda, J.; Doyle, C.; Kosynkin, D.; Hwang, W.; Tour, J. Diazonium functionalization
of surfactant-wrapped chemically converted graphene sheets. Journal of the American
Chemical Society 2008, 130 (48), 16201-16206.

88. Iqbal, M. Z.; Abdala, A. A. Oil spill cleanup using graphene. Environmental Science and
Pollution Research 2013, 20 (5), 3271-3279.

89. Prud'homme, R. K.; Aksay, I. A.; Adamson, D.; Abdala, A. Thermally exfoliated
graphite oxide. US Patent 7,658,901, 2010.

90. Staudenmaier, L. Method for the preparation of graphitic acid. Ber. Dtsch. chem. Ges.
1898, 31, 1481-87.

91. Aksay, I. A.; Milius, D. L.; Korkut, S.; Prud'Homme, R. K. Functionalized graphene
sheets having high carbon to oxygen ratios. US Patent 2011/0114897, 2011.

92. Ni, Z.; Wang, Y.; Yu, T.; Shen, Z. Raman spectroscopy and imaging of graphene. Nano
Research 2008, 1 (4), 273-291.

93. Wang, Y.; Ni, Z.; Shen, Z.; Wang, H.; Wu, Y. Interference enhancement of Raman signal
of graphene. Appl. Phys. Lett. 2008, 92, 043121.

156
94. Gupta, A.; Chen, G.; Joshi, P.; Tadigadapa, S.; Eklund, P. Raman scattering from high-
frequency phonons in supported n-graphene layer films. Nano Lett. 2006, 6 (12), 2667-
2673.

95. Meyer, J. C.; Geim, A.; Katsnelson, M.; Novoselov, K.; Booth, T.; Roth, S. The structure
of suspended graphene sheets. Nature 2007, 446 (7131), 60-63.

96. Xu, Y.; Bai, H.; Lu, G.; Li, C.; Shi, G. Flexible graphene films via the filtration of water-
soluble noncovalent functionalized graphene sheets. Journal of the American Chemical
Society 2008, 130 (18), 5856-5857.

97. Gao, W.; Alemany, L.; Ci, L.; Ajayan, P. New insights into the structure and reduction of
graphite oxide. Nature Chemistry 2009, 1 (5), 403-408.

98. Cheekati, S. L. Graphene based anode materials for Lithium-ion batteries. MS Thesis,
Wright State University Dayton, Ohio2011.

99. Balandin, A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C.
Superior thermal conductivity of single-layer graphene. Nano letters 2008, 8 (3), 902-
907.

100. Dikin, D. A.; Stankovich, S.; Zimney, E. J.; Piner, R. D.; Dommett, G. H. B.;
Evmenenko, G.; Nguyen, S. B. T.; Ruoff, R. S. Preparation and characterization of
graphene oxide paper. Nature 2007, 448 (7152), 457-460.

101. Geim, A.; Novoselov, K. The rise of graphene. Nature Materials 2007, 6 (3), 183-191.

102. Pang, H.; Chen, T.; Zhang, G.; Zeng, B.; Li, Z. An electrically conducting
polymer/graphene composite with a very low percolation threshold. Materials Letters
2010, 64 (20), 2226-2229.

103. Kuilla, T.; Bhadra, S.; Yao, D.; Kim, N.; Bose, S.; Lee, J. Recent advances in graphene
based polymer composites. Progress in Polymer Science 2010, 35 (11), 1350-1375.

104. Kim, H. Processing, Morphology and Properties of Graphene Reinforced Polymer


Nanocomposites. PhD Thesis, University of Minnesota, Minnesota2009.

105. Schlapbach, L.; Züttel, A. Hydrogen-storage materials for mobile applications. Nature
2001, 414 (6861), 353-358.

106. Schedin, F.; Geim, A.; Morozov, S.; Hill, E.; Blake, P.; Katsnelson, M.; Novoselov, K.
Detection of individual gas molecules adsorbed on graphene. Nature Materials 2007, 6
(9), 652-655.

157
107. Yang, K.; Gu, M. Enhanced thermal conductivity of epoxy nanocomposites filled with
hybrid filler system of triethylenetetramine-functionalized multi-walled carbon
nanotube/silane-modified nano-sized silicon carbide. Composites Part A: Applied Science
and Manufacturing 2010, 41 (2), 215-221.

108. Gaspar, H.; Pereira, C.; Rebelo, S.; Pereira, M.; Figueiredo, J.; Freire, C. Understanding
the silylation reaction of multi-walled carbon nanotubes. Carbon 2011, 49 (11), 3441-
3453.

109. Ganguli, S.; Roy, A. K.; Anderson, D. P. Improved thermal conductivity for chemically
functionalized exfoliated graphite/epoxy composites. Carbon 2008, 46 (5), 806-817.

110. Prud'Homme, R. K.; Aksay, I. A.; Adamson, D.; Abdala, A. Thermally exfoliated
graphite oxide US Patent 7,658,901, 2010.

111. Worch, M.; Engelmann, H. J.; Blum, W.; Zschech, E. Cross-sectional thin film
characterization of Si compounds in semiconductor device structures using both
elemental and ELNES mapping by EFTEM. Thin Solid Films 2002, 405, 198-204.

112. Shanmugharaj, A.; Bae, J.; Lee, K. Y.; Noh, W. H.; Lee, S. H.; Ryu, S. H. Physical and
chemical characteristics of multiwalled carbon nanotubes functionalized with
aminosilane and its influence on the properties of natural rubber composites. Composites
Science and Technology 2007, 67 (9), 1813-1822.

113. Waltman, R.; Pacansky, J.; Bates Jr, C. X-ray photoelectron spectroscopic studies on
organic photoconductors: Evaluation of atomic charges on chlorodiane blue and p-
(diethylamino) benzaldehyde diphenylhydrazone. Chemistry of Materials 1993, 5 (12),
1799-1804.

114. Jing, S. Y.; Lee, H. J.; Choi, C. K. Chemical bond structure on Si-OC composite films
with a low dielectric constant deposited by using inductively coupled plasma chemical
vapor deposition. Journal of Korean Physcial Society 2002, 41 (5), 769-773.

115. Wakeland, S.; Martinez, R.; Grey, J. K.; Luhrs, C. C. Production of graphene from
graphite oxide using urea as expansion–reduction agent. Carbon 48 (12), 3463-3470.

116. Urbonaite, S.; Wachtmeister, S.; Mirguet, C.; Coronel, E.; Zou, W. Y.; Csillag, S.;
Svensson, G. EELS studies of carbide derived carbons. Carbon 2007, 45 (10), 2047-
2053.

117. Falqui, A.; Serin, V.; Calmels, L.; Snoeck, E.; Corrias, A.; Ennas, G. EELS investigation
of FeCo/SiO2 nanocomposites. Journal of microscopy 2003, 210 (1), 80-88.

158
118. Giannakopoulos, K.; Boukos, N.; Travlos, A. Zinc oxide nanoparticles on silicon.
Superlattices and Microstructures 2006, 39 (1), 115-123.

119. Laffont, L.; Monthioux, M.; Serin, V.; Mathur, R. B.; Guimon, C.; Guimon, M. F. An
EELS study of the structural and chemical transformation of PAN polymer to solid
carbon. Carbon 2004, 42 (12), 2485-2494.

120. Skiff, W. M.; Carpenter, R. W.; Lin, S. H. Near-edge fine-structure analysis of core-shell
electronic absorption edges in silicon and its refractory compounds with the use of
electron-energy-loss microspectroscopy. Journal of applied physics 1987, 62 (6), 2439-
2449.

121. Botton, G. A.; Phaneuf, M. W. Imaging, spectroscopy and spectroscopic imaging with an
energy filtered field emission TEM. Micron 1999, 30 (2), 109-119.

122. Auchterlonie, G. J.; McKenzie, D. R.; Cockayne, D. J. H. Using ELNES with parallel
EELS for differentiating between a-Si: X thin films. Ultramicroscopy 1989, 31 (2), 217-
222.

123. BittoloáBon, S. High concentration few-layer graphene sheets obtained by liquid phase
exfoliation of graphite in ionic liquid. Journal of Materials Chemistry 2011, 21 (10),
3428-3431.

124. Pimenta, M.; Dresselhaus, G.; Dresselhaus, M. S.; Cancado, L.; Jorio, A.; Saito, R.
Studying disorder in graphite-based systems by Raman spectroscopy. Physical Chemistry
Chemical Physics 2007, 9 (11), 1276-1290.

125. Liu, Y.-T.; Yang, J.-M.; Xie, X.-M.; Ye, X.-Y. Polystyrene-grafted graphene with
improved solubility in organic solvents and its compatibility with polymers. Materials
Chemistry and Physics 2011, 130 (1), 794-799.

126. Graf, D.; Molitor, F.; Ensslin, K.; Stampfer, C.; Jungen, A.; Hierold, C.; Wirtz, L.
Spatially resolved Raman spectroscopy of single-and few-layer graphene. Nano letters
2007, 7 (2), 238-242.

127. Ferrari, A.; Meyer, J.; Scardaci, V.; Casiraghi, C.; Lazzeri, M.; Mauri, F.; Piscanec, S.;
Jiang, D.; Novoselov, K.; Roth, S. Raman spectrum of graphene and graphene layers.
Physical review letters 2006, 97 (18), 187401.

128. Kudin, K. N.; Ozbas, B.; Schniepp, H. C.; Prud'Homme, R. K.; Aksay, I. A.; Car, R.
Raman spectra of graphite oxide and functionalized graphene sheets. Nano letters 2008, 8
(1), 36-41.

159
129. Fang, M.; Wang, K.; Lu, H.; Yang, Y.; Nutt, S. Single-layer graphene nanosheets with
controlled grafting of polymer chains. Journal of Materials Chemistry 2010, 20 (10),
1982-1992.

130. He, Y.-S.; Gao, P.; Chen, J.; Yang, X.; Liao, X.-Z.; Yang, J.; Ma, Z.-F. A novel bath lily-
like graphene sheet-wrapped nano-Si composite as a high performance anode material for
Li-ion batteries. RSC Advances 2011, 1 (6), 958-960.

131. Clark, J.; Macquarrie, D. Catalysis of liquid phase organic reactions using chemically
modified mesoporous inorganic solids. Chemical Communications 1998, (8), 853-860.

132. Wight, A.; Davis, M. Design and preparation of organic-inorganic hybrid catalysts.
Chemical reviews 2002, 102 (10), 3589-3614.

133. Utracki, L. A.; Wilkie, C. A. Polymer blends handbook; Springer Reference2014.

134. Paul, D.; Newman, S. Polymer blends, vols. I and II. Academic, New York 1978.

135. Soontaranum, W.; Higgins, J.; Papathanasiou, T. Rheology and thermodynamics in


partially miscible polymer blends. Journal of non-newtonian Fluid Mechanics 1996, 67,
191-212.

136. Van Puyvelde, P.; Moldenaers, P. Rheology and morphology development in immiscible
polymer blends. Rheology Reviews 2005, 101.

137. Wagner, E.; Robeson, L. Impact polystyrene: factors controlling the rubber efficiency.
Rubber Chemistry and Technology 1970, 43 (5), 1129-1137.

138. Cancio, L. V.; Miller, G. W.; Wu, P.-C. Polyester compositions for gas and moisture
barrier materials1981.

139. Percec, E. S.; Melamud, L.; Coffey, G. P. Polymer blend films for food packages1992.

140. Heuschen, J.; Vion, J.; Jerome, R.; Teyssie, P. Polycaprolactone-based block copolymers:
4. Interfacial activity in poly (vinyl chloride) containing polyblends. Polymer 1990, 31
(8), 1473-1480.

141. Nishi, T.; Wang, T. Melting point depression and kinetic effects of cooling on
crystallization in poly (vinylidene fluoride)-poly (methyl methacrylate) mixtures.
Macromolecules 1975, 8 (6), 909-915.

142. Vleminckx, G.; Bose, S.; Leys, J.; Vermant, J.; Wubbenhorst, M.; Abdala, A. A.;
Macosko, C.; Moldenaers, P. Effect of Thermally Reduced Graphene Sheets on the Phase
160
Behavior, Morphology, and Electrical Conductivity in Poly [(α-methyl styrene)-co-
(acrylonitrile)/poly (methyl-methacrylate) Blends. ACS applied materials & interfaces
2011, 3 (8), 3172-3180.

143. Hameed, T.; Hussein, I. A. Rheological study of the influence of M w and comonomer
type on the miscibility of m-LLDPE and LDPE blends. Polymer 2002, 43 (25), 6911-
6929.

144. Utracki, L.; Schlund, B. Linear low density polyethylenes and their blends: Part 4 shear
flow of LLDPE blends with LLDPE and LDPE. Polymer Engineering & Science 1987,
27 (20), 1512-1522.

145. Hussein, I. A. Melt miscibility and mechanical properties of metallocene linear


low‐density polyethylene blends with high‐density polyethylene: influence of
comonomer type. Polymer international 2005, 54 (9), 1330-1336.

146. Moly, K.; Radusch, H.; Androsh, R.; Bhagawan, S.; Thomas, S. Nonisothermal
crystallisation, melting behavior and wide angle X-ray scattering investigations on linear
low density polyethylene (LLDPE)/ethylene vinyl acetate (EVA) blends: effects of
compatibilisation and dynamic crosslinking. European polymer journal 2005, 41 (6),
1410-1419.

147. Gonsalves, K.; Patel, S.; Trivedi, D. Development of potentially degradable materials for
marine applications. III. Polyethylene‐polyethylene oxide blends. Journal of applied
polymer science 1992, 45 (2), 217-225.

148. Naqvi, M. K.; Choudhary, M. S. Chemically modified polyolefins and their blends.
Journal of Macromolecular Science, Part C: Polymer Reviews 1996, 36 (3), 601-629.

149. Kim, H.; Kobayashi, S.; AbdurRahim, M. A.; Zhang, M. J.; Khusainova, A.; Hillmyer,
M. A.; Abdala, A. A.; Macosko, C. W. Graphene/polyethylene nanocomposites: effect of
polyethylene functionalization and blending methods. Polymer 2011, 52 (8), 1837-1846.

150. Guruvenket, S.; Rao, G. M.; Komath, M.; Raichur, A. M. Plasma surface modification of
polystyrene and polyethylene. Applied Surface Science 2004, 236 (1), 278-284.

151. Schuster, L.; Hettche, A.; Liedy, W.; Weiss, S.; Ehemann, L. Oxidized polyethylene and
use as lubricant for polyvinylchloride1991.

152. Chen, C.; Wesson, R.; Collier, J.; Lo, Y. Studies of rigid poly (vinyl chloride)(PVC)
compounds. IV. Fusion characteristics and morphology analyses. Journal of applied
polymer science 1995, 58 (7), 1107-1115.

161
153. Weiland, M.; Daro, A.; David, C. Biodegradation of thermally oxidized polyethylene.
Polymer degradation and Stability 1995, 48 (2), 275-289.

154. Iring, M. Thermal oxidation of polyethylene and polypropylene: effects of chemical


structure and reaction conditions on the oxidation process. Progress in Polymer Science
1990, 15 (2), 217-262.

155. Iwata, H.; Kishida, A.; Suzuki, M.; Hata, Y.; Ikada, Y. Oxidation of polyethylene surface
by corona discharge and the subsequent graft polymerization. Journal of Polymer Science
Part A: Polymer Chemistry 1988, 26 (12), 3309-3322.

156. Rugg, F. M.; Smith, J. J.; Bacon, R. C. Infrared spectrophotometric studies on


polyethylene. II. Oxidation. Journal of Polymer Science 1954, 13 (72), 535-547.

157. Gugumus, F. Re-examination of the thermal oxidation reactions of polymers 2. Thermal


oxidation of polyethylene. Polymer degradation and Stability 2002, 76 (2), 329-340.

158. Vandiver, M. A.; Caire, B. R.; Poskin, Z.; Li, Y.; Seifert, S.; Knauss, D. M.; Herring, A.
M.; Liberatore, M. W. Durability and performance of polystyrene‐b‐poly (vinylbenzyl
trimethylammonium) diblock copolymer and equivalent blend anion exchange
membranes. Journal of applied polymer science 2015, 132 (10).

159. Liu, Y.; Wang, J.; Yang, Y.; Brenner, T. M.; Seifert, S. e.; Yan, Y.; Liberatore, M. W.;
Herring, A. ε. Anion Transport in a Chemically Stable, Sterically Bulky α-C Modified
Imidazolium Functionalized Anion Exchange Membrane. The Journal of Physical
Chemistry C 2014, 118 (28), 15136-15145.

160. Chan, M.; Allara, D. Infrared reflection studies of the mechanism of oxidation at a
copper—polyethylene interface. Journal of colloid and interface science 1974, 47 (3),
697-704.

161. Krupa, I.; Luyt, A. Thermal properties of uncross-linked and cross-linked LLDPE/wax
blends. Polymer Degradation and Stability 2000, 70 (1), 111-117.

162. Mandelkern, L.; Fatou, J.; Denison, R.; Justin, J. A calorimetric study of the fusion of
molecular weight fractions of linear polythylene (1). Journal of Polymer Science Part B:
Polymer Letters 1965, 3 (10), 803-807.

163. Hoffman, J. D.; Weeks, J. J. X‐Ray Study of Isothermal Thickening of Lamellae in Bulk
Polyethylene at the Crystallization Temperature. The Journal of chemical physics 1965,
42 (12), 4301-4302.

162
164. Silva, M. E. S.; Mano, V.; Pacheco, R. R.; Freitas, R. F. Miscibility Behavior of
Polyacrylamides Poly (Ethylene Glycol) Blends: Flory Huggins Interaction Parameter
Determined by Thermal Analysis. Journal of Modern Physics 2013, 4 (07), 45.

165. Zhang, L.; Goh, S.; Lee, S.; Hee, G. Miscibility, melting and crystallization behavior of
two bacterial polyester/poly (epichlorohydrin-co-ethylene oxide) blend systems. Polymer
2000, 41 (4), 1429-1439.

166. Flory, P. J. Principles of polymer chemistry. 1953.

167. Sperling, L. H. Introduction to physical polymer science; John Wiley & Sons2015.

168. Ruehle, D. A.; Perbix, C.; Castañeda, M.; Dorgan, J. R.; Mittal, V.; Halley, P.; Martin, D.
Blends of biorenewable polyamide-11 and polyamide-6, 10. Polymer 2013, 54 (26),
6961-6970.

169. Rostami, S. Advances in theory of equilibrium melting point depression in miscible


polymer blends. European polymer journal 2000, 36 (10), 2285-2290.

170. Painter, P. C.; Shenoy, S. L.; Bhagwagar, D. E.; Fishburn, J.; Coleman, M. M. Effect of
hydrogen bonding on the melting point depression in polymer blends where one
component crystallizes. Macromolecules 1991, 24 (20), 5623-5629.

171. Glatter, O.; Kratky, O. Editors. Small-Angle X-ray Scattering. New York: Academic
Press, 1982.

172. Beaucage, G. Small-angle scattering from polymeric mass fractals of arbitrary mass-
fractal dimension. Journal of Applied Crystallography 1996, 29 (2), 134-146.

173. Jackson, C.; Bauer, B.; Nakatani, A.; Barnes, J. Synthesis of hybrid organic-inorganic
materials from interpenetrating polymer network chemistry. Chemistry of materials 1996,
8 (3), 727-733.

174. Canetti, M.; Bertini, F. Crystalline and supermolecular structure evolution of poly
(ethylene terephthalate) during isothermal crystallization and annealing treatment by
means of wide and small angle X-ray investigations. European polymer journal 2010, 46
(2), 270-276.

175. Rambo, R. P.; Tainer, J. A. Characterizing flexible and intrinsically unstructured


biological macromolecules by SAS using the Porod‐Debye law. Biopolymers 2011, 95
(8), 559-571.

163
176. Cross, M. M. Rheology of non-Newtonian fluids: a new flow equation for pseudoplastic
systems. Journal of colloid science 1965, 20 (5), 417-437.

177. El Achaby, M.; Arrakhiz, F. E.; Vaudreuil, S.; el Kacem Qaiss, A.; Bousmina, M.;
Fassi‐Fehri, O. Mechanical, thermal, and rheological properties of graphene‐based
polypropylene nanocomposites prepared by melt mixing. Polymer Composites 2012, 33
(5), 733-744.

178. Liu, C.; Wang, J.; He, J. Rheological and thermal properties of m-LLDPE blends with m-
HDPE and LDPE. Polymer 2002, 43 (13), 3811-3818.

179. Kyu, T.; Hu, S. R.; Stein, R. S. Characterization and properties of polyethylene blends II.
Linear low‐density with conventional low‐density polyethylene. Journal of Polymer
Science Part B: Polymer Physics 1987, 25 (1), 89-103.

180. Tucker III, C. L.; Moldenaers, P. Microstructural evolution in polymer blends. Annual
Review of Fluid Mechanics 2002, 34 (1), 177-210.

181. GRACE†, H. P. Dispersion phenomena in high viscosity immiscible fluid systems and
application of static mixers as dispersion devices in such systems. Chemical Engineering
Communications 1982, 14 (3-6), 225-277.

182. Everaert, V.; Aerts, L.; Groeninckx, G. Phase morphology development in immiscible
PP/(PS/PPE) blends influence of the melt-viscosity ratio and blend composition. Polymer
1999, 40 (24), 6627-6644.

183. Burch, H. E.; Scott, C. E. Effect of viscosity ratio on structure evolution in miscible
polymer blends. Polymer 2001, 42 (17), 7313-7325.

184. Han, C. D.; Chuang, H. K. Criteria for rheological compatibility of polymer blends.
Journal of applied polymer science 1985, 30 (11), 4431-4454.

185. Meissner, J. Basic parameters, melt rheology, processing and end-use properties of three
similar low density polyethylene samples. Pure and Applied Chemistry 1975, 42 (4), 551-
612.

186. Osaki, K.; Inoue, T.; Uematsu, T.; Yamashita, Y. Evaluation methods of the longest
Rouse relaxation time of an entangled polymer in a semidilute solution. Journal of
Polymer Science Part B: Polymer Physics 2001, 39 (14), 1704-1712.

187. σeppalli, R.; εarega, C.; εarigo, A.; Bajgai, ε. P.; Kim, H. Y.; Causin, V. Poly ( -
caprolactone) filled with electrospun nylon fibres: a model for a facile composite
fabrication. European polymer journal 2010, 46 (5), 968-976.

164
188. Pukánszky, B.; Mudra, I.; Staniek, P. Relation of crystalline structure and mechanical
properties of nucleated polypropylene. Journal of Vinyl and Additive Technology 1997, 3
(1), 53-57.

189. Jeong, S.; Kim, D.; Seo, J. Preparation and antimicrobial properties of LDPE composite
films melt-blended with polymerized urushiol powders (YPUOH) for packaging
applications. Progress in Organic Coatings 2015, 85, 76-83.

190. Layek, R.; Samanta, S.; Chatterjee, D. Physical and Mechanical Properties of Poly
(methyl methacrylate)-functionalized Graphene/Poly (vinylidine fluoride)
Nanocomposites: Piezoelectric [beta] Polymorph Formation. Polymer 2010.

191. Kim, H.; Macosko, C. W. Processing-property relationships of polycarbonate/graphene


composites. Polymer 2009, 50 (15), 3797-3809.

192. Kalaitzidou, K.; Fukushima, H.; Drzal, L. T. Mechanical properties and morphological
characterization of exfoliated graphite-polypropylene nanocomposites. Composites Part
A: Applied Science and Manufacturing 2007, 38 (7), 1675-1682.

193. Li, M. A nano-graphite/paraffin phase change material with high thermal conductivity.
Applied Energy 2013, 106, 25-30.

194. Dittrich, B.; Wartig, K.-A.; Hofmann, D.; Mülhaupt, R.; Schartel, B. Flame retardancy
through carbon nanomaterials: carbon black, multiwall nanotubes, expanded graphite,
multi-layer graphene and graphene in polypropylene. Polymer Degradation and Stability
2013, 98 (8), 1495-1505.

195. Bao, C.; Guo, Y.; Yuan, B.; Hu, Y.; Song, L. Functionalized graphene oxide for fire
safety applications of polymers: a combination of condensed phase flame retardant
strategies. Journal of Materials Chemistry 2012, 22 (43), 23057-23063.

196. Demczyk, B.; Wang, Y.; Cumings, J.; Hetman, M.; Han, W.; Zettl, A.; Ritchie, R. Direct
mechanical measurement of the tensile strength and elastic modulus of multiwalled
carbon nanotubes. Materials Science and Engineering: A 2002, 334 (1), 173-178.

197. Cheng, S.; Chen, X.; Hsuan, Y. G.; Li, C. Y. Reduced graphene oxide-induced
polyethylene crystallization in solution and nanocomposites. Macromolecules 2011, 45
(2), 993-1000.

198. Kuila, T.; Khanra, P.; Mishra, A. K.; Kim, N. H.; Lee, J. H. Functionalized-
graphene/ethylene vinyl acetate co-polymer composites for improved mechanical and
thermal properties. Polymer Testing 2012, 31 (2), 282-289.

165
199. Chatterjee, T.; Krishnamoorti, R. Rheology of polymer carbon nanotubes composites.
Soft Matter 2013, 9 (40), 9515-9529.

200. Kalaitzidou, K.; Fukushima, H.; Drzal, L. T. Multifunctional polypropylene composites


produced by incorporation of exfoliated graphite nanoplatelets. Carbon 2007, 45 (7),
1446-1452.

201. Thomassin, J.; Huynen, I.; Jérôme, R.; Detrembleur, C. Functionalized polypropylenes as
efficient dispersing agents for carbon nanotubes in a polypropylene matrix; application to
electromagnetic interference (EMI) absorber materials. Polymer 2010, 51 (1), 115-121.

202. Kim, S.; Do, I.; Drzal, L. T. Multifunctional xGnP/LLDPE nanocomposites prepared by
solution compounding using various screw rotating systems. Macromolecular Materials
and Engineering 2009, 294 (3), 196-205.

203. Steurer, P.; Wissert, R.; Thomann, R.; Mülhaupt, R. Functionalized graphenes and
thermoplastic nanocomposites based upon expanded graphite oxide. Macromolecular
rapid communications 2009, 30 (4‐5), 316-327.

204. Kuila, T.; Bose, S.; Mishra, A. K.; Khanra, P.; Kim, N. H.; Lee, J. H. Effect of
functionalized graphene on the physical properties of linear low density polyethylene
nanocomposites. Polymer Testing 2012, 31 (1), 31-38.

205. Shen, J. W.; Huang, W. Y.; Zuo, S. W.; Hou, J. Polyethylene/grafted


polyethylene/graphite nanocomposites: Preparation, structure, and electrical properties.
Journal of applied polymer science 2005, 97 (1), 51-59.

206. Chaudhry, A.; Mittal, V. High‐density polyethylene nanocomposites using masterbatches


of chlorinated polyethylene/graphene oxide. Polymer Engineering & Science 2013, 53
(1), 78-88.

207. Chaudhry, A.; Mittal, V. Blends of high‐density polyethylene with chlorinated


polyethylene: Morphology, thermal, rheological, and mechanical properties. Polymer
Engineering & Science 2014, 54 (1), 85-95.

208. Fim, F. d. C.; Basso, N. R.; Graebin, A. P.; Azambuja, D. S.; Galland, G. B. Thermal,
electrical, and mechanical properties of polyethylene–graphene nanocomposites obtained
by in situ polymerization. Journal of Applied Polymer Science 2013, 128 (5), 2630-2637.

209. Li, Y.; Zhu, J.; Wei, S.; Ryu, J.; Sun, L.; Guo, Z. Poly (propylene)/graphene nanoplatelet
nanocomposites: melt rheological behavior and thermal, electrical, and electronic
properties. Macromolecular Chemistry and Physics 2011, 212 (18), 1951-1959.

166
210. Iqbal, M. Z.; Abdala, A. A.; Liberatore, M. W. Synthesis and characterization of
polyethylene/oxidized polyethylene miscible blends and role of OPE as a viscosity
control. Journal of Applied Polymer Science 2016, n/a-n/a.

211. Marcano, D. C.; Kosynkin, D. V.; Berlin, J. M.; Sinitskii, A.; Sun, Z.; Slesarev, A.;
Alemany, L. B.; Lu, W.; Tour, J. M. Improved synthesis of graphene oxide. ACS nano
2010, 4 (8), 4806-4814.

212. Vandiver, M. A.; Caire, B. R.; Poskin, Z.; Li, Y.; Seifert, S.; Knauss, D. M.; Herring, A.
M.; Liberatore, M. W. Durability and performance of polystyrene-b-poly(vinylbenzyl
trimethylammonium) diblock copolymer and equivalent blend anion exchange
membranes. Journal of applied polymer science 2015, 132 (10), n/a-n/a.

213. Liu, Y.; Wang, J.; Yang, Y.; Brenner, T. M.; Seifert, S.; Yan, Y.; Liberatore, M. W.;
Herring, A. ε. Anion Transport in a Chemically Stable, Sterically Bulky α-C Modified
Imidazolium Functionalized Anion Exchange Membrane. The Journal of Physical
Chemistry C 2014, 118 (28), 15136-15145.

214. Huang, C.-L.; Wang, C. Rheological and conductive percolation laws for syndiotactic
polystyrene composites filled with carbon nanocapsules and carbon nanotubes. Carbon
2011, 49 (7), 2334-2344.

215. Wen, Y. H.; Lin, H. C.; Li, C. H.; Hua, C. C. An experimental appraisal of the Cox–Merz
rule and Laun's rule based on bidisperse entangled polystyrene solutions. Polymer 2004,
45 (25), 8551-8559.

216. Sabzi, M.; Jiang, L.; Liu, F.; Ghasemi, I.; Atai, M. Graphene nanoplatelets as poly (lactic
acid) modifier: linear rheological behavior and electrical conductivity. Journal of
Materials Chemistry A 2013, 1 (28), 8253-8261.

217. Sahimi, M.; Arbabi, S. Mechanics of disordered solids. II. Percolation on elastic networks
with bond-bending forces. Physical Review B 1993, 47 (2), 703.

218. Surve, M.; Pryamitsyn, V.; Ganesan, V. Universality in structure and elasticity of
polymer-nanoparticle gels. Physical review letters 2006, 96 (17), 177805.

219. Kim, H.; Macosko, C. W. Morphology and properties of polyester/exfoliated graphite


nanocomposites. Macromolecules 2008, 41 (9), 3317-3327.

220. Feng, S.; Sen, P. N. Percolation on elastic networks: new exponent and threshold.
Physical review letters 1984, 52 (3), 216.

167
221. Kanai, H.; Navarrete, R.; Macosko, C.; Scriven, L. Fragile networks and rheology of
concentrated suspensions. Rheologica acta 1992, 31 (4), 333-344.

222. Ren, J.; Silva, A. S.; Krishnamoorti, R. Linear viscoelasticity of disordered polystyrene-
polyisoprene block copolymer based layered-silicate nanocomposites. Macromolecules
2000, 33 (10), 3739-3746.

223. Shih, W.-H.; Shih, W. Y.; Kim, S.-I.; Liu, J.; Aksay, I. A. Scaling behavior of the elastic
properties of colloidal gels. Physical Review A 1990, 42 (8), 4772.

224. Kharchenko, S. B.; Douglas, J. F.; Obrzut, J.; Grulke, E. A.; Migler, K. B. Flow-induced
properties of nanotube-filled polymer materials. Nature materials 2004, 3 (8), 564-568.

225. Larson, R. The structure and rheology of complex fluids. 1999. In New York: Oxford,
1999; Vol. 688.

226. Doi, M.; Edwards, S. The theory of polymer dynamics. 1986. Claredon, Oxford. ISBN 0-
19-852033-6 1986.

227. Wierenga, A. M.; Philipse, A. P. Low-shear viscosity of isotropic dispersions of


(Brownian) rods and fibres; a review of theory and experiments. Colloids and Surfaces A:
Physicochemical and Engineering Aspects 1998, 137 (1), 355-372.

228. Cassagnau, P. Linear viscoelasticity and dynamics of suspensions and molten polymers
filled with nanoparticles of different aspect ratios. Polymer 2013, 54 (18), 4762-4775.

229. Malekani, K.; Rice, J. A.; Lin, J.-S. Comparison of techniques for determining the fractal
dimensions of clay minerals. Clays and clay minerals 1996, 44 (5), 677-685.

230. Schmidt, P. W. Use of scattering to determine the fractal dimension. The fractal
approach to heterogeneous chemistry 1989, 67-79.

231. Schaefer, D. W.; Zhao, J.; Brown, J. M.; Anderson, D. P.; Tomlin, D. W. Morphology of
dispersed carbon single-walled nanotubes. Chemical Physics Letters 2003, 375 (3), 369-
375.

232. Fornes, T.; Paul, D. Modeling properties of nylon 6/clay nanocomposites using
composite theories. Polymer 2003, 44 (17), 4993-5013.

233. Sheng, N.; Boyce, M. C.; Parks, D. M.; Rutledge, G.; Abes, J.; Cohen, R. Multiscale
micromechanical modeling of polymer/clay nanocomposites and the effective clay
particle. Polymer 2004, 45 (2), 487-506.

168
234. Halpin, J. C.; Kardos, J. The Halpin-Tsai equations: a review. Polymer engineering and
science 1976, 16 (5), 344-352.

235. Mori, T.; Tanaka, K. Average stress in matrix and average elastic energy of materials
with misfitting inclusions. Acta metallurgica 1973, 21 (5), 571-574.

236. Eshelby, J. D. In The determination of the elastic field of an ellipsoidal inclusion, and
related problems, Proceedings of the Royal Society of London A: Mathematical, Physical
and Engineering Sciences, 1957; The Royal Society, pp 376-396.

237. Tandon, G.; Weng, G. The effect of aspect ratio of inclusions on the elastic properties of
unidirectionally aligned composites. Polymer composites 1984, 5 (4), 327-333.

238. Gómez-Navarro, C.; Burghard, M.; Kern, K. Elastic properties of chemically derived
single graphene sheets. Nano letters 2008, 8 (7), 2045-2049.

239. Suk, J. W.; Piner, R. D.; An, J.; Ruoff, R. S. Mechanical properties of monolayer
graphene oxide. ACS nano 2010, 4 (11), 6557-6564.

240. Cho, J.; Luo, J.; Daniel, I. Mechanical characterization of graphite/epoxy nanocomposites
by multi-scale analysis. Composites science and technology 2007, 67 (11), 2399-2407.

241. Sjoerdsma, S.; Bleijenberg, A.; Heikens, D. The Poisson ratio of polymer blends, effects
of adhesion and correlation with the Kerner packed grains model. Polymer 1981, 22 (5),
619-623.

242. Du, J.; Zhao, L.; Zeng, Y.; Zhang, L.; Li, F.; Liu, P.; Liu, C. Comparison of electrical
properties between multi-walled carbon nanotube and graphene nanosheet/high density
polyethylene composites with a segregated network structure. Carbon 2011, 49 (4), 1094-
1100.

243. Stauffer, D.; Aharony, A. Introduction to percolation theory; CRC press1994.

244. Gao, J.-F.; Li, Z.-M.; Meng, Q.-j.; Yang, Q. CNTs/UHMWPE composites with a two-
dimensional conductive network. Materials Letters 2008, 62 (20), 3530-3532.

245. Xue, Q. The influence of particle shape and size on electric conductivity of metal–
polymer composites. European Polymer Journal 2004, 40 (2), 323-327.

246. Boudenne, A.; Ibos, L.; Fois, M.; Gehin, E.; Majeste, J. C. Thermophysical properties of
polypropylene/aluminum composites. Journal of Polymer Science Part B: Polymer
Physics 2004, 42 (4), 722-732.

169
247. Li, F.; Qi, L.; Yang, J.; Xu, M.; Luo, X.; Ma, D. Polyurethane/conducting carbon black
composites: structure, electric conductivity, strain recovery behavior, and their
relationships. Journal of Applied Polymer Science 2000, 75 (1), 68-77.

248. Yi, X. S.; Wu, G.; Ma, D. Property balancing for polyethylene‐based carbon black‐filled
conductive composites. Journal of applied polymer science 1998, 67 (1), 131-138.

249. Calixto, C. M. F.; Mendes, R. K.; Oliveira, A. C. d.; Ramos, L. A.; Cervini, P.;
Cavalheiro, É. T. G. Development of graphite-polymer composites as electrode materials.
Materials Research 2007, 10 (2), 109-114.

250. Chujo, Y.; Saegusa, T. Organic polymer hybrids with silica gel formed by means of the
sol-gel method. In Macromolecules: Synthesis, Order and Advanced Properties;
Springer, 1992, pp 11-29.

251. Novak, B. M. Hybrid nanocomposite materials—between inorganic glasses and organic


polymers. Advanced Materials 1993, 5 (6), 422-433.

252. Loy, D. A.; Shea, K. J. Bridged polysilsesquioxanes. Highly porous hybrid organic-
inorganic materials. Chemical Reviews 1995, 95 (5), 1431-1442.

253. Yu, A.; Ramesh, P.; Sun, X.; Bekyarova, E.; Itkis, M. E.; Haddon, R. C. Enhanced
thermal conductivity in a hybrid graphite nanoplatelet–carbon nanotube filler for epoxy
composites. Advanced Materials 2008, 20 (24), 4740-4744.

254. Ma, P.-C.; Liu, M.-Y.; Zhang, H.; Wang, S.-Q.; Wang, R.; Wang, K.; Wong, Y.-K.;
Tang, B.-Z.; Hong, S.-H.; Paik, K.-W. Enhanced electrical conductivity of
nanocomposites containing hybrid fillers of carbon nanotubes and carbon black. ACS
applied materials & interfaces 2009, 1 (5), 1090-1096.

255. Zhang, W. D.; Phang, I. Y.; Shen, L.; Chow, S. Y.; Liu, T. Polymer nanocomposites
using urchin‐shaped carbon nanotube‐silica hybrids as reinforcing fillers.
Macromolecular rapid communications 2004, 25 (21), 1860-1864.

256. Wu, G.; Zheng, Q. Estimation of the agglomeration structure for conductive particles and
fiber‐filled high‐density polyethylene through dynamic rheological measurements.
Journal of Polymer Science Part B: Polymer Physics 2004, 42 (7), 1199-1205.

257. Kaelble, E. F. Handbook of X-rays; McGraw-hill1967.

258. Johnson, D. W.; Dobson, B. P.; Coleman, K. S. A manufacturing perspective on graphene


dispersions. Current Opinion in Colloid & Interface Science 2015, 20 (5), 367-382.

170
259. Bharadwaj, R. K. Modeling the barrier properties of polymer-layered silicate
nanocomposites. Macromolecules 2001, 34 (26), 9189-9192.

260. Griffiths, S. Nanotechnology in the food industry.

261. Laboratory Scale Extruders by Kuhn Group, Germany.

262. Extensional Rheometer by TA Instruments.

263. Worch, M.; Engelmann, H.; Blum, W.; Zschech, E. Cross-sectional thin film
characterization of Si compounds in semiconductor device structures using both
elemental and ELNES mapping by EFTEM. Thin solid films 2002, 405 (1), 198-204.

264. Lee, S. W.; Kim, B.-S.; Chen, S.; Shao-Horn, Y.; Hammond, P. T. Layer-by-Layer
Assembly of All Carbon Nanotube Ultrathin Films for Electrochemical Applications.
Journal of the American Chemical Society 2009, 131 (2), 671-679.

171
APPENDIX-A SILANE MODIFIED GRAPHENE

Experimental details

The oxidation of graphite and the extent of exfoliation of GO is studied using X-ray

diffraction (XRDX’Pert PRτ εPD diffractometer, PAσalytical). Scans at β from 5 to 35° with

0.02 /sec step size at 40 KV voltage with an intensity of 20 Å using CuKα radiation of

wavelength 1.5406 Å was conducted.

Transmission Electron Microscopy (TEM) was performed using FEI Tecnai G20 with

0.11 nm point resolution and operated at 200 kV. Samples were prepared by dispersing

approximately 0.5 mg of graphene in 25 mL of dimethyl formamide by sonication at room

temperature for 5 minutes. Two drops of the suspension were deposited on 400 mesh copper

grids covered with thin amorphous lacey carbon film. The coupled Energy Dispersion

Spectroscopy (EDS) X-ray analysis is used for elemental composition of TRG and f-TRG.

Electron Energy Loss Spectroscopy (EELS) was acquired using post column energy filtered

camera (Gatan Quantum 963). The energy resolution was measured to be 0.9 eV FWHM at the

zero loss peak. Energy filtered TEM (EFTEM) mapping was applied to map the location of the

elements on the surface of f-TRG by measuring core, post-edge and pre-edge losses of the

respective elements using the “three-window method” technique263. EELS spectra of the

aforementioned elements were also recorded at their respective core losses and the type of bonds

were identified. Acquisition time necessary to obtain a good signal to noise ratio was 5 s for

silicon L2-3-edge, 10 s for carbon K-edge, 20 s for nitrogen K-edge and 30 s for oxygen K-edge.

172
XPS measurements were performed using SSX-100 system (Surface Science

δaboratories, Inc.) equipped with a monochromated Al Kα X-ray source, a hemispherical sector

analyzer (HSA) and a resistive anode detector. The base pressure was 4.0 x 10 -10 Torr. During

the data collection, the pressure was ca. 1 x 10-8 Torr. The X-ray spot size was 1x1 mm2, which

corresponded to an X-ray power of 200 W. Each sample was separately mounted on a sample

holder using a piece of double-sided carbon sticking tape. Care was taken to ensure the surface

was fully covered by the samples. The survey spectra were collected using 150 eV pass energy

and 1 eV/step. The atomic percentages were calculated from the survey spectrum using the

ESCA 2005 software provided with the XPS system. The high resolution spectra were collected

using 50 eV pass energy and 0.1 eV/step. For the high resolution data, the lowest binding-energy

C 1s peak (presumably, C-C/ C-H peak) was set at 285.0 eV and used as the reference for all of

the other elements. The curve fitting used a combination of Gaussian/Lorenzian function with the

Gaussian percentages being at 80% or higher.

The electrical conductivity was measured by a custom made conductivity cell at the

National Renewable Energy Laboratory (Golden, CO). A very small quantity of TRG (2-10 mg)

was dispersed in 10 mL of a 3 to 1 water to isopropanol solution via 20 minutes of bath

sonication followed by an additional 2 minutes of tip sonication to break any remaining clumps

of TRG. About 1 mL of the dispersion was placed on a copper stub at a time and dried at 40° C

in a drying oven (Despatch, model LBB1-23A-1). Subsequent layers of dispersed TRG were

developed in this way until copper stub is completely covered with TRG layers. The complete

coverage of the copper stub was obtained from ~8mL of pure TRG, and ~10mL of the f-TRG

samples. The copper stub covered with a sample is loaded carefully in the custom made

173
conductivity cell. The resistivity was measured using a multimeter (Wavetek 28XT,

Accuracy: ±1% + 0.1 Ohm).

Characterization of TRG

The evidence of the complete exfoliation of graphite oxide has been obtained by XRD

(Figure A.1a). The interlayer spacing in graphite has been observed to be 3.37Å from the single

only intrinsic 002-peak at  = 13.25. Due to the presence of polar groups on GO from oxidation

and because of the adsorbed water contents in GO, the 002-peak shifts to  = 5.7 (indicating the

interlayer d-spacing of 0.78Å). The increased d-spacing is a clear evidence of expansion of

graphite layers upon oxidation.

a) b)
TRG

GO

Graphite

2
5 10 15 20 25 30

Figure A.1 XRD spectra of graphite, graphite oxide (GO) and TRG (a), and TEM of pure TRG
(b)

In XRD of TRG, no interlayer spacing is observed indicating a complete exfoliation of

GO. This confirms the production of no stacked structure and that we are not dealing with

nanocrystalline graphite but with TRG. In addition, TEM image (Figure A.1b) evidences the

presence of overlapping layers of graphene. The transparent graphene sheets are clearly visible

174
confirming that GO has been successfully exfoliated. However, the elastic corrugations and the

scrolled or folded edges often result in different brightness in the surface of graphene 95.

XPS Spectra

Important insights into the reaction mechanism and the material structure come from the

core-level binding energies (BEs) obtained from XPS high resolution spectra. Figure A.2 shows

the overall survey data of the pure and f-TRG samples. Clear differences between pure and f-

TRG samples can be observed. The XPS spectra have been collected on dry TRG samples. In

TRG overall survey, there has been no Si and N detected. Usually, the Si 2p is detectable between

95-110 eV whereas N1s can be found between 400-407eV, depending upon the chemical

environment of the molecule. As compared with pure TRG, both the f-TRGp and f-TRGs show

the presence of Si2p and N1s originating from APTS, indicating the successful covalent

functionalization of TRG with APTS.

5000 3000 3000


C1s O1s O1s C1s
TRG C1s 2500 f-TRGs 2500
4000 f-TRGp
2000 2000
3000

Counts
Counts
Counts

1500 1500
Si2s Si2s
N1s Si2p
O1s 2000 N1s Si2p
1000 1000

1000 500 500


C12p
C12p
0 0 0
1000 900 800 700 600 500 400 300 200 100 0 1000 900 800 700 600 500 400 300 200 100 0 1000 900 800 700 600 500 400 300 200 100 0

Binding Energy, ev Binding Energy, ev Binding Energy, ev

Figure A.2 Overall XPS survey for the pure and f-TRG

The broad band in high resolution N1s spectra of f-TRGp and f-TRGs fitted with two

peaks at 399.95 eV (~400.0) and ~401.0 eV BE (Figure A.3). The lower BE peak is assigned to

aliphatic amine, imine or amide groups (R-NH2, C=N, N-C=O) and the higher energy peak is

assigned to positively charged quaternary nitrogen (R-NH3+). The latter peak might have resulted

from the amine protonation by acidic groups on the surface of TRG264.

175
1700
1450 f-TRGp N1s f-TRGs N1s

399.95 1500 399.95


1250

1300
1050
401.24 401.74
850 1100

650 900

450 700
396.5 398.5 400.5 402.5 404.5 396.5 398.5 400.5 402.5 404.5
Binding Energy, eV Binding Energy, eV

700
f-TRGs Si2p
600

500 102.9
103.5
400

300

200

100

0
95 98 101 104 107 110
Binding energy, eV

Figure A.3 N1s and S2p high resolution XPS scans for TRGp and TRGs

In Si2p region, the presence of bands confirms the grafting reaction of silane on TRG

(Figure A.3). The Si2p spectra of the two functionalized samples appear almost similar. The two

shoulders are observed exactly at the same position for the functionalized samples. One small

shoulder peak appears at 102.9 eV and the other at 103.5 eV. The peak at 103.5 eV is attributed

to the development of a strong cage-like Si-O4bonds114 or Si-O-Si structure due to lateral

polymerization, which affects the theoretical predictions of the product concentrations in

silanization reaction. However, due to anhydrous conditions during the reaction, the lateral

polymerization effects can be neglected108. The shoulder at 102.9 eV in both spectra can be

attributed to N bonding with the carbon surface. However, when deconvoluted, the two peaks

merge into one smooth peak at ~102-103 eV. This single peak contains the contributions from

Si2p3/2 and Si2p1/2 which cannot be resolved by non-monochromatic radiation as shown as inset

in Figure A.3 for Si in f-TRGp. And therefore the single deconvoluted peak around 103 eV BE is

176
assigned to Si-O bond. The relative contribution of each band in high resolution fitting is shown

in Table A.1.

Table A.1 High resolution fitting of functional groups in N1s and Si2p XPS scans

N1s peaks Si2p peaks


R-NH2 R-NH3+ N-C Si-O4
f-TRGp 399.95 401.24 102.9 103.5
f-TRGs 399.95 401.74 102.9 103.5

EDS analysis

The EDS analyses were performed on selected thin areas of graphene sheets (Table A.2).

However, EDS shows a qualitative picture of the attached moieties which can be taken as a

rough estimate for quantitative analysis.

Table A.2 Averaged EDS results

Atomic %
f-TRGp f-TRGs
C(K) 73.98 74.98
N(K) 4.53 4.10
O(K) 13.53 14.39
Si(K) 7.89 6.27
S(K) 0.04 0.14
Cl(K) 0.01 0.10

177
APPENDIX-B PE/OPE MISCIBLE BLENDS

-3
at 10 sec-1

Arrhenius Fit
-3.5

-4
lnη

-4.5

-5

-5.5
2 2.2 2.4 2.6 2.8
1000/T
Figure B.1 Arrhenius Plot of the viscosity of OPE (at shear rate = 10 1/s). Viscosity data was
obtained from performing steady state flow test on OPE after melting on the Peltier plate

130 120

120
90
∆Hc, J/g

110
Tc, C

60
100

30
90

80 0
0 10 20 30 40 50 60

OPE wt%

Figure B.2 Crystallization temperature and heat of crystallization of PE/OPE blends with
different OPE wt%.

178
16 (a)
50/50
14
12 70/30
Heat Flow, W/g 10 80/20

8
100/0
6
4
2
0
20 40 60 80 100 120 140
Temperature, C

Figure B.3 Thermograms of PE/OPE blends show the melting profiles of the blends. The inset
shows the melting profile of neat OPE

16

14

12
Heat Flow, W/g

10

2
20 40 60 80 100 120 140
Temprature, °C

Figure B.4 Melting profile of pure PE. This area has been used for calculation of melt enthalpy
and % crystallinity

179
Transmission, % (a.u.)

4000 3500 3000 2500 2000 1500 1000 500


Wavenumbver, cm-1

Figure B.5 Complete FTIR spectrum of OPE

2.5
TPeak

2.0
dW/dT, %/°C

1.5

1.0

0.5

0.0

200 300 400 500 600


Temperature, °C
Figure B.6 Derivative of weight loss of PE/OPE Blends

180
Table B.1 Thermal stability of blends with different compositions

Tpeak Tonset Total


Blend T25% T50% T75% T95%
(°C) (°C) wt loss
100/0 473 442 454 467 480 493 98.9
90/10 478 449 453 470 478 490 99.2
80/20 478 444 449 468 480 491 99.4
70/30 467 428 434 459 470 483 98.6
60/40 471 426 427 456 472 486 99.1
50/50 467 417 416 450 468 483 98.9
0/100 460 407 390 445 465 483 99.8

The solid phase miscibility of OPE in PE was studied by WAXD. PE showed two main

prominent peaks at β ~ β1° and βγ.γ° corresponding to (110) and (β00), respectively (Figure B-
155, 157,
6). The oxidation of PE to OPE is known to break down higher molecular weight chains
160
. OPE exhibited the same peaks as those in PE with extra peaks above β ~γη°. Inclusion of

τPE in PE slightly shifts the β for (110) plane to lower value, but no other change was seen

with increasing OPE concentration. No distinguishable peaks were observed in PE/OPE blends

which indicated blend immiscibility. Thus, in the solid state, blends exhibited the complete

miscibility.

50/50

60/40
Intensity [a.u.]

70/30

80/20

90/10

OPE

PE

2
10 20 30 40 50 60 70

Figure B.7 WAXD patterns of neat PE, OPE and PE/OPE blends at 25°C with Y-offset

181
Figure B.8 Cross model fitting to different blend compositions at various temperatures

Table B.2 Cross model fitting parameters for PE/OPE Blends

ηo (Pa-s) τo (s) n

180° 160° 140° 130° 180° 160° 140° 130° 180° 160° 140° 130°
PE/OPE

100/0 1.01×104 1.35×104 1.88×104 1.01×104 0.51 0.63 0.78 0.82 0.63 0.64 0.66 0.66
3 3 3 4
90/10 6.21×10 8.49×10 9.85×10 1.09×10 0.60 0.90 0.73 0.74 0.57 0.56 0.60 0.61
80/20 2.81×103 2.87×103 3.38×103 3.95×103 0.23 0.23 0.20 0.24 0.57 0.56 0.58 0.57
70/30 1.23×103 1.43×103 1.72×103 1.81×103 0.13 0.14 0.16 0.21 0.54 0.55 0.56 0.51
60/40 6.09×102 7.17×102 8.25×102 7.47×102 0.08 0.10 0.12 0.14 0.58 0.57 0.55 0.49
50/50 1.29×102 1.21×102 1.41×102 1.00×102 0.02 0.02 0.03 0.02 0.85 0.73 0.54 0.61

182
6
10
5
10
4
10
G, Pa 10
3

2
10 100/0
1 90/10
10 80/20
0 70/30
10 60/40
-1 50/50
10

G, Pa
1 2 3 4 5 6
10 10 10 10 10 10

Figure B.9 Han plots for the blends

3.00
100/0
90/10
Horizontal shift factor, a T

2.50 80/20
70/30
60/40
2.00 50/50

1.50

1.00

0.50

0.00
100 110 120 130 140 150 160 170 180 190
Temperature, C

Figure B.10 Shift factor for time-temperature master curves for PE/OPE blends

183
1000000
1000000
100/0 90/10
100000
100000

10000

G', Pa
G', Pa

10000
1000

180 100 180


1000
160 160
140 10 140
130 130
100 1
0.1 1 10 100 1000 0.1 1 10 100 1000
Frequency, Hz Frequency, Hz

1000000 1000000
80/20 70/30
100000 100000

10000 10000
G', Pa

G', Pa
1000 1000

100 100
180 180
160 160
10 10
140 140
130 130
1 1
0.1 1 10 100 1000 0.1 1 10 100 1000

Frequency, Hz Frequency, Hz

100000 100000
60/40 50/50
10000
10000

1000
G', Pa

G', Pa

1000

100
100
10
180 180
10 160 160
1
140 140
130 130
1 0.1
0.1 1 10 100 1000 0.01 0.1 1 10 100
Frequency, Hz Frequency, Hz

Figure B.11 Time-temperature master curves for all blend compositions against the tested
temperatures

184
80
90
PE/OPE 90/10

80 70
180
70 160 60

60 140
delta, [ ]

delta, [ ]
50
130
50
40
40
30
30 180

20 160
20
140
10 10
PE/OPE 100/0 130
0 0
0.01 0.1 1 10 100 1000 0.1 1 10 100 1000

Frequency, Hz Frequency, Hz

90 90

80 80

70 70

60
60
delta [ ]

50 delta [ ] 50

40
40

30
180 30

20 160 180
20
140 160
10 130 140
PE/OPE 80/20 10
PE/OPE 70/30
130
0
0.1 1 10 100 1000 0
0.1 1 10 100 1000
Frequency Frequency, Hz
90 90

80 80

70 70

60 60
delta [ ]
delta [ ]

50 50

40 40

30 30
180
180
20 20 160
160
140
10 140 10
130 PE/OPE 50/50
130 PE/OPE 60/40
0 0
0.1 1 10 100 1000 0.1 1.0 10.0 100.0
Frequency, Hz Frequency, Hz

Figure B.12 Time-temperature master curves for all blend compositions against the tested
temperatures

185
APPENDIX-C PE/OPE/TRG AND PE/OPE/CB COMPOSITES

Graphite showed an intrinsic 00β diffraction peak at β ~ βθ.η°, corresponding to a d-

spacing of 3.37Å. The oxidation of graphite to GO shifted the 00β peak to β ~11.4°, and doubled

the interlayer spacing to a d-spacing of 7.8Å. It is worth mentioning that the 002 peak in graphite

was a narrow and high intensity peak, exhibiting ordered stacked structure, whereas in GO the

peak was broad and small in intensity. The broadness in 002 peak was attributed to intercalated

oxygen functionalities and water within the stacks of graphite resulted from oxidation. On the

other hand, no diffraction peak was observed in TRG, showing that the ordered structure was

completely dissolved, and GO was successfully converted into non-stacked, and unlayered

graphene by thermal exfoliation.

2=26.5°
(a)
d ~ 0.34 nm
2=11.4°
d ~ 0.78 nm

Graphite
GO

2
TRG
5 10 15 20 25 30

Figure C.1 XRD of graphite, graphite oxide and TRG

186
TRG, produced from exfoliating GO, typically contains some residual oxy-functional

groups, such as carboxylic, hydroxyl, and epoxy groups on the surface37. The FTIR spectrum of

TRG showed these functional groups in 800-2000 cm-1 IR range. The C-O-C stretching vibration

in epoxy groups was observed at 1018 cm-1, and -OH stretching vibration in hydroxyl groups

was seen at 1259 cm-1. A weak OH deformation vibration was also observed at ~1370 cm-1.

Typical carboxylic group stretching vibrations in C=O was observed at ~1770 cm-1. The region

marked with bracket in 1550-1680 cm-1 IR range was attributed to multiple stretching vibrations

from C=C graphitic carbon in graphene37.

(a) (b)
Transmission (a.u)

Transmission (a.u)

1767

Sp2 C=C 1259 1018

2000 1800 1600 1400 1200 1000 800


4000 3500 3000 2500 2000 1500 1000 500
Wavenumber, cm-1 Wavenumber, cm-1

TRG

OPE
Transmission (%)

OPE + TRG

(c)
4000 3500 3000 2500 2000 1500 1000 500
Wavenumber (cm-1)

Figure C.2 FTIR of TRG (a, b), and OPE+TRG (c).

187
The particle size distribution of CB particles showed that CB particles were in 300-500

nm, in agreement with TEM sizing. Moreover, the DLS of dilute TRG solution showed bimodal

distribution of TRG nanosheets: smaller sheets of size 600-800 nm, and some larger sheets of

size 2-5µm.

100 (f)
Intensity [a.u.]
CB TRG
80
60
40
20
0
2 3 4
10 10 10
Particle Size [nm]
Figure C.3 DLS of TRG and CB

Figure C.4 TEM of Carbon Black (CB)

Melt rheology of polymers, and polymer nanocomposites is considered central for

understanding processing properties, and structure-property relationships. Viscoelastic (VE)

188
response can be connected to the intrinsic behavior of polymers, inter-particle, and polymer-

particle interaction. In general, rheological properties depend on the input strain, whether in linear

or non-linear VE region. The dynamic strain sweep determines the strain dependent viscoelastic

response, and the linear viscoelastic (LVE) region. The frequency sweep experiments conducted

at a strain of = 1% (in δVE region) within the frequency range of 0.01 to 100 s -1 at 160°C are

shown in Figure C.5 (a, b). Due to plasticization with increasing τPE concentration, Gˊ and G˝,

and the complex viscosity, * decreased gradually in blends compared with neat PE. In polymer

dynamics, typically, Gˊ~ƒ2 and G˝~ƒ1 behavior is expected at low frequency by relaxed polymer

chains. Here, neat PE (PE/τPE 100/0) exhibited Gˊ~ƒ1.1 behavior, which relaxed to Gˊ~ƒ1.2 for

θ0/40 blends. Similarly, G˝~ƒ behavior changed from G˝~ƒ0.8 for 100/0 to G˝~ƒ0.9 for 60/40

blends. The increase in low frequency slope for 60/40 blend indicated formation of the relaxed

polymer chains, which was attributed to the plasticization, and miscibility effects of OPE in

PE210. Similar behavior was observed for * since * also decreased with increasing τPE

concentration. The three blend compositions showed excellent fit to the Cross model176 with zero

shear viscosity ( o) of 10,330 Pa-s for PE that decreased to 3,540 Pa-s for 80/20 and 1,499 Pa-s

for 60/40 blends. The Cross exponent, decreased from 0.64 for PE to 0.56 for 80/20 blends, and

slightly increased to 0.57 for 60/10 blends.

189
10
10 (a) 4 (b)
Closed-G
10 4 10 Cross Fit
Open-G
100/0

10
8
10
G G [Pa]

* [Pa-s]
6 80/20
100/0
10
10 0 3
10
o, Pa-s
4 60/40
10 80/20
2 100/0 10,330
10 80/20 3,540
60/40
60/40 1,499
2
10
, s , s
-2 -1 0 1 2 -2 -1 0 1 2
10 10 10 -1 10 10 10 10 10 -1 10 10

Figure C.5 Rheological properties of PE/OPE Blends: (a) shear and loss moduli vs frequency,
and (b) complex viscosity profiles as a function of blend composition at 160°C. (100/0 circles;
80/20 upper triangles; 60/40 lower triangles). Pre-factors in (a) were used to shift G′~ƒ curves
vertically for clarity.
Complex Viscosity, *, Pa-s

6 5
10 10
(a) (b) PE
Shear Modulus, G , Pa

0.5
5
10 1
2.5
4.3
4 6.5
10 4 10
PE 10
0.5
3
10 1
2.5
4.3
2 6.5
10 10
3
10
1
10 -3
Frequency, , Hz Frequency, , Hz
-2 -1 0 1 2
10 10 10 10 10 10 10
-3
10
-2
10
-1
10
0
10
1
10
2

6
Complex Viscosity, *, Pa-s

5
10 10
PE/OPE 80/20
(c)
Shear Modulus, G , Pa

(d) CB 0.55 wt%


5 CB 1.16 wt%
10 CB 2.57 wt%
4
CB 4.3 wt%
4 10 CB 10 wt%
10

3
10 PE/OPE 80/20 3
CB 0.55 wt% 10
CB 1.16 wt%
2
10 CB 2.57 wt%
CB 4.3 wt%
CB 10 wt%
1 2
10 -3 10 -3
Frequency, , Hz Frequency, , Hz
-2 -1 0 1 2 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10 10 10

Figure C.6 Rheological data of CB-filler composites

190
Complex Viscosity, *, Pa-s
6 4
10 10
Shear Modulus, G , Pa
PE/OPE 60/40
(e) (f) CB 0.55 wt%
5 CB 1.16 wt%
10
CB 2.57 wt%
CB 4.3 wt%
4
10
3 3
10 10
2
10 PE/OPE 60/40
CB 0.55 wt%
1 CB 1.16 wt%
10 CB 2.57 wt%
CB 4.3 wt%
0 2
10 -3 10 -3
Frequency, , Hz Frequency, , Hz
-2 -1 0 1 2 -2 -1 0 1 2
10 10 10 10 10 10 10 10 10 10 10 10

Figure C.6 Continued

5
10
4 Pure PE
Shear Modulus, G , Pa

8x10 PE/TRG 1 wt%


4
6x10

4
4x10

4
2x10

-2 -1 0 1 2

Strain, , %
10 10 10 10 10

Figure C.7 Representative strain sweep on PE/TRG nanocomposites at 160°C. Arrows point out
the limit of linearity used for scaling argument

191
4.0 3.3
PE/TRG
80/20/TRG
3.1

decaying exponent, α
3.5 60/40/TRG
decaying exponent, α
2.9
3.0
2.7
2.5
2.5
3.0
2.0
2.5
2.3
2.0 PE/CB
1.5 1.5
80/20/CB
2.1
1.0
0 0.1 0.2 0.3 0.4 0.5 60/40/CB
1.0 1.9
0 1 2 3 4 5 6 0 2 4 6 8 10 12
TRG wt% CB wt%

Figure C.8 Power law decaying exponent for TRG-filled nanocomposites, and CB-filled
composite, obtained from SAXS I ~ q log-log plot

2
10

1
10
10

0 1
I(q) [a.u.]

10
I(q) [a.u.]

-1 0.1
10

-2 PE/CB 0.5
10 0.01
PE/CB 1.0 80/20/CB 0.55
PE/CB 2.5 80/20/CB 2.57
-3
10 PE/CB 4.3 80/20/CB 4.3
0.001
PE/CB 6.5 80/20/CB 6.58
-4 PE/CB 10 80/20/CB 10
10
9 2 3 4 5 6 7 8 9 2 3 0.0001
0.01 0.1 9 2 3 4 5 6 7 8 9 2 3
0.01 0.1
q [1/Å] q [1/Å]

60/40/CB 0.55
10 60/40/CB 1.16
60/40/CB 2.57
60/40/CB 4.3
1 60/40/CB 6.58
60/40/CB 10
I [q]

0.1

0.01

0.001
9 2 3 4 5 6 7 8 9 2 3
0.01 0.1
q [1/Å]

Figure C.9 SAXS patterns for CB-filled composites

192
I(q)xq [a.u.]

wt%
t%

1
2
5w

TRG
TRG

/4 0
PE 60
PE/O

0.05 0.10 0.15 0.20


q [1/Å]

Figure C.10 Representative Kratky plot for SAXS data

100 (a) Mass Retained, % 100 (b)


Mass Retained, %

80 80

60 60

40 40

80/20
PE 20
20
PE/TRG 5 wt % 80/20/TRG 5 wt%

0 0
0 100 200 300 400 500 600 0 100 200 300 400 500 600

Temperature, C Temperature, C

100 (c) 100 (d)


Mass Retained, %

Mass Retained, %

80 80

60 60

40 40

60/40 PE
20 20
60/40/TRG 5 wt% PE/CB 10 wt%
0 0
0 100 200 300 400 500 600 0 100 200 300 400 500 600

Temperature, C Temperature, C

100 (e) 100 (f)


Mass Retained, %

80
Mass Retained, %

80

60 60

40 40

20 80/20 20
60/40
80/20/CB 10 wt% 60/40/CB 10 wt%
0 0
0 100 200 300 400 500 600 0 100 200 300 400 500 600

Temperature, C Temperature, C

Figure C.11 Thermal stability of TRG-filled, and CB-filled composites

193
1.E+08 1.E+08
Breakage Strain, % Breakage Strain, % Breakage Stress, Pa
(a) (b)
Breakage Stress, Pa
1.E+07 1.E+07

1.E+06 1.E+06

1.E+05 1.E+05

1.E+04 1.E+04

1.E+03 1.E+03

1.E+02 1.E+02

1.E+01 1.E+01

1.E+00 1.E+00
0.03 0.07 0.16 0.5 1 3 5 0.03 0.07 0.16 0.42 1 2 3 5
1.E+08
(c) Breakage Strain, % Breakage Stress, Pa

1.E+07

1.E+06

1.E+05

1.E+04

1.E+03

1.E+02

1.E+01

1.E+00
0.03 0.07 0.16 0.42 1 3 5

Figure C.12 Strain at break and breakage stress of PE/TRG (a), 80/20/TRG, and 60/40/TRG (c)
nanocomposites

194
1.E+08 1.E+08
(b) Breakage Strain, % Breakage Stress, Pa
(a) Breakage Strain, % Breakage Stress, Pa
1.E+07 1.E+07

1.E+06 1.E+06

1.E+05 1.E+05

1.E+04 1.E+04

1.E+03 1.E+03

1.E+02 1.E+02

1.E+01 1.E+01

1.E+00 1.E+00
0.5 1 2.5 4.3 6.5 10 0.55 2.57 4.33 6.58 10

1.E+08
(c) Breakage Strain, % Breakage Stress, Pa

1.E+07

1.E+06

1.E+05

1.E+04

1.E+03

1.E+02

1.E+01

1.E+00
0.55 1.16 2.57 4.33 6.58 10

Figure C.13 Strain at break and breakage stress of PE/CB (a), 80/20/CB (b), and 60/40/CB (c)
composites

195
5.0
PE/TRG
80/20/TRG
4.0 60/40/TRG

3.0
Ec/Em

2.0

1.0

0.0
0 1 2 3 4 5
CB vol%

Figure C.14 Young's modulus of CB-filled composites, normalized by matrix moduli. Straight
lines are linear fit to experimental data with intercept 1.

0
10
-2
PE/CB
10
-4
60/40/CB
10
 [S/cm]

-6
10 10
-1

-8 -3 ~3.0
10 10

10 ~3.4
-10 -5
10 80/20/CB
-7 ~2.4
-12 10 -1
10 10 10
0
10
1

-14
W-Wp
10
0 3 6 9 12
CB wt%
Figure C.15 Electrical conductivity of CB-filled composites. Inset shows the power law fit to the
experimental data

196

You might also like