You are on page 1of 176

Graphene Synthesis & Graphene/Polymer Nanocomposites

A DISSETRTATION
SUBMITTED TO THE FACULTY OF
UNIVERSITY OF MINNESOTA
BY

Ken-Hsuan Liao

IN FULFILLMENT OF THE REQUIREMENTS


FOR THE DEGREE OF
DOCTOR OF PHILOSOPHY

Christopher W. Macosko

November 2012
© Ken-Hsuan Liao 2012
ALL RIGHTS RESERVED
Acknowledgements

The completion of this thesis would not have been possible without assistance from a
number of people. First, I would like to thank my advisors, Professor Christopher W.
Macosko and Professor K. Andre Mkhoyan, for giving me the opportunity to be a part of
their groups. I am much indebted to my advisors’ patience and the insightful discussions
we had during the process. The reason I have turned out to be a creative researcher must be
credited to Professor Macosko who instructed me to make my research ideas become real
results. I would like to thank Dr. Greg Haugstad for teaching me many things on AFM. I
would like to thank Professor Andreas Stein and Professor David Morse for being my
thesis defense committee members. Second, I would like to take this opportunity to thank
Dr. Hyunwoo Kim, who helped me to start my project in graphene/polymer
nanocomposites, and I am deeply indebted to him for his guidance and discussions. I
appreciate Susan Viker for helping me on purchasing orders and many other things which
facilitated my PhD research. I would like to thank Yuqiang Qian and Wei Xie for many
helps on sample property measurements. I want to thank Karen Haman for helping me with
English writing. I wish to dedicate many thanks to every member of the Macosko's group
and my undergraduate researchers for experimental assistance: Steven Maslo and Jerry Yeh.
This thesis would not be possible without the financial supports from
Abu-Dhabi-Minnesota Institute for Research Excellence (ADMIRE) program. I appreciate
Professor Amed Abdala for helpful discussion and providing thermally reduced graphene
for my research.
I am also sincerely grateful to my wife, Hui Sun, for her constant support for the
encouragement to work hard, for the interesting discussions about politics and
entertainment, for all the small things that happened in our lives, and for sharing a
promising picture of future and planning every detail in it. I want to thank my parents for
cheering me up and encouraging me through graduate school. Last, but not least, I would
like to thank all my friends who shared happiness and difficulties with me in these 4 years
of graduate school.

i
Abstract

Graphene, a two-dimensional carbon sheet with single-atom thickness, has recently

attracted significant interest due to its unique mechanical and electrical properties. It has

been reported that incorporation of graphene in polymers can efficiently improve the

materials’ electrical and mechanical properties. For reliable integration of graphene into

practical graphene/polymer nanocomposites, it is essential to have a simple, reproducible

and controllable technique to produce graphene on a large scale. We successfully

developed a novel, fast, hydrazine-free, high-yield method for producing single-layered

graphene. Graphene sheets were formed from graphite oxide by reduction with de-ionized

water at 130 ºC. Over 65% of the sheets are single graphene layers. A dehydration reaction

of exfoliated graphene oxide was utilized to reduce oxygen and transform C-C bonds from

sp3 to sp2. The reduction appears to occur in large uniform interconnected oxygen-free

patches so that despite the presence of residual oxygen the sp2 carbon bonds formed on the

sheets are sufficient to provide electronic properties comparable to reduced graphene

sheets obtained using other methods.

Cytotoxicity of aqueous graphene was investigated with Dr. Yu-Shen Lin by

measuring mitochondrial activity in adherent human skin fibroblasts using two assays. The

methyl-thiazolyl-diphenyl-tetrazolium bromide (MTT) assay, a typical nanotoxicity assay,

fails to predict the toxicity of graphene oxide and graphene toxicity because of the

spontaneous reduction of MTT by graphene and graphene oxide, resulting in a false

positive signal. An appropriate alternate assessment, using the water soluble tetrazolium

salt (WST-8) assay, reveals that the compacted graphene sheets are more damaging to

mammalian fibroblasts than the less densely packed graphene oxide. Clearly, the toxicity
ii
of graphene and graphene oxide depends on the exposure environment (i.e. whether or not

aggregation occurs) and mode of interaction with cells (i.e. suspension versus adherent cell

types).

Ultralow percolation concentration of 0.15 wt% graphene, as determined by surface

resistance and modulus, was observed from in situ polymerized thermally reduced

graphene (TRG)/ poly-urethane-acrylate (PUA) nanocomposite. A homogeneous

dispersion of TRG in PUA was revealed by TEM images. The aspect ratio of dispersed

TRG, calculated from percolation concentration and modulus, was found to be equivalent

to the reported aspect ratio of single-layered free standing TRG. This indicates TRG is

mono-layer-dispersed in the matrix polymer.

How graphene/polymer nanocomposite glass transition temperatures (Tg) vary was

investigated in this study. First we surveyed the literature. No changes in Tg were observed

for graphene/polymer nanocomposites synthesized via physical blending processes such as

solvent or melt blending, except aqueous blending. In contrast, chemical blending

processes such as in situ polymerization or chemically modified fillers yielded significant

Tg increases in graphene/polymer nanocomposites. We attribute these results to bonding

interactions at the interfaces between matrix polymers and fillers. Physical blending

processes cannot provide enough interaction at the interfaces, whereas chemical blending

processes can yield strong interaction such as covalent bonds. Aqueous blending of

graphene or graphene oxide nanocomposites with water soluble matrix polymers also

cause Tg increases, even though the blending processes involve no chemical reactions. The

reason for this exception is that hydrogen bonding forms between fillers (graphene oxide or

reduced graphene) and water soluble matrix polymers. We then measured Tg in PMMA.
iii
We used isotactic PMMA (i-PMMA) and syndiotactic-rich atactic PMMA (a-PMMA) to

make TRG/PMMA nanocomposites using solvent blending and in situ polymerization in

order to investigate the stereo-regularity and processing effects on the Tg. A Tg increase

was found in i-PMMA and in situ PMMA but not in a-PMMA. The results can be

explained by the thin film confinement effect of polymer. We attribute the Tg increase to

both a higher interaction density and a stronger hydrogen bonding at the interfaces.

We have studied the elastic modulus of graphene oxide with various oxygen content.

We used in situ AFM nano-indentation to measure the influence of oxygen on the elastic

modulus of graphene oxide with various carbon/oxygen (C/O) ratios. The results show that

chemical reduction (lower oxygen contents) decreases the elastic modulus of graphene

oxide. We speculate that chemical reduction of oxygen atoms of epoxy groups on graphene

oxide surface removes the bridging effect between carbon atoms, which leads to more

flexible sheets.

iv
Table of Contents

List of Tables ................................................................................................ ix


List of Figures ................................................................................................x
Chapter 1. Introduction ................................................................................1
1-1 Graphene & Graphene Oxide ................................................................................................ 1
1-2 Graphene/Polymer Nanocomposites ..................................................................................... 2
1.3 Overview of Thesis ................................................................................................................ 3

Chapter 2. Graphene Synthesis ....................................................................6


2-1 Introduction ......................................................................................................................... 6
2-2 Experiments & Characterization ......................................................................................... 7
2-2.1 Graphene & Graphene Oxide Synthesis ........................................................................... 7
2-2.2.1 Atomic Force Microscopy (AFM)................................................................................. 9
2-2.2.2 X-Ray Photoelectron Spectroscopy (XPS) .................................................................. 10
2-2.2.3 Elemental Analysis (EA) ............................................................................................. 10
2-2.2.4 Electrical Resistivity Measurements. .......................................................................... 11
2-2.2.5 Transmission Electron Microscopy (TEM) ................................................................. 12
2-2.2.6 Fourier Transform Infrared Spectroscopy (FTIR) ....................................................... 12
2-3 Results and Discussion ...................................................................................................... 12
2-4 Conclusions ....................................................................................................................... 22

Chapter 3. .....................................................................................................24
Cytotoxicity of Graphene Oxide and Graphene .......................................24
3-1 Introduction ......................................................................................................................... 24
3-2 Experiments & Characterization. ........................................................................................ 27
3-2.1 Synthesis of Graphene and Graphene Oxide .................................................................... 27
3-2.2 Characterization ................................................................................................................ 28
3-2.2.1 Wide-angle Powder X-ray Diffraction (WXRD)........................................................... 28
3-2.2.2 Atomic Force Microscopy (AFM) ................................................................................. 28
3-2.2.3 X-ray Photoelectron Spectroscopy (XPS) ..................................................................... 28

v
3-2.2.4 Dynamic Light Scattering (DLS)................................................................................... 29
3-2.2.5 ζ-Potential Measurements .............................................................................................. 29
3-2.2.6 Optical Microscopy ....................................................................................................... 29
3-2.2.7 Hemolysis Assay ........................................................................................................... 30
3-2.2.8 MTT Viability Assay ..................................................................................................... 30
3-2.2.9 WST-8 Viability Assay ................................................................................................. 32
3-3 Results and Discussion ........................................................................................................ 33
3-3.1 Synthesis and Characterization of GO and GS. ................................................................ 33
3-3.2 In Vitro Hemolytic Activity of GO and GS...................................................................... 41
3-3.3 In Vitro Fibroblast Cytotoxicity of GO and GS ............................................................... 47
3-4 Conclusions ......................................................................................................................... 52

Chapter 4......................................................................................................54
Reducing Graphene Aggregation in Preparing Nanocomposites ....54
4-1 Introduction ..................................................................................................................... 54
4-2 Experimental Section ...................................................................................................... 55
4-2.1 Materials .......................................................................................................................... 55
4-2.2 Fabrication of Graphene/TPU Nanocomposites .......................................................... 58
4-2.3 Characterization of Nanocomposites ............................................................................ 62
4-3 Results and Discussion .................................................................................................... 63
4-4 Conclusions ...................................................................................................................... 69

Chapter 5. .....................................................................................................70
Graphene/Polyurethane Acrylate Nanocomposites .................................70
5-1. Introduction ...................................................................................................................... 70
5-2 Experiments & Characterization ....................................................................................... 71
5-2.1 Materials ......................................................................................................................... 71
5-2.2 Synthesis of UAO and TPGDA solution (UA) .............................................................. 74
5-2.3 Fabrication of TRG/PUA Nanocomposites .................................................................... 74
5-2.4 Rheology and Polymerization of TRG/UA Uncured Liquids ........................................ 75
5-2.5 Characterization of TRG/PUA Nanocomposites ............................................................ 75
5-3 Results and Discussion ...................................................................................................... 76
5-4 Conclusions ....................................................................................................................... 85
vi
Chapter 6. .....................................................................................................87
Stereo-regularity and Processing Effect on Glass Transition
Temperature of Graphene/PMMA Nanocomposites ...............................87
6-1. Introduction ........................................................................................................................ 87
6-2. Experimental Section .......................................................................................................... 89
6-2.1 Materials ........................................................................................................................... 89
6-2.2 Preparation of Nanocomposites via Solvent Blending: .................................................... 91
6-2.3 Preparation of Nanocomposites via in situ Polymerization:............................................. 92
6-2.4 Molecular and Nanocomposite Characterization:............................................................. 93
6-3. Results and Discussion ....................................................................................................... 94
6-4. Conclusions ...................................................................................................................... 110

Chapter 7. ...................................................................................................111
Glass Transition Temperature of Graphene/Polymer Nanocomposites
.....................................................................................................................111
7-1. Introduction ...................................................................................................................... 111
7-2. Literature Summary .......................................................................................................... 111
7-3. Conclusions ...................................................................................................................... 114

Chapter 8. Oxygen Effect of Stiffness of Graphene Oxide ....................119


8-1 Introduction ..................................................................................................................... 119
8-2 Experiments & Results .................................................................................................... 121
8-2.1 Phase I: Silicon Wafer Substrate .................................................................................. 121
8-2.2 Phase II: TEM Grid Substrate ...................................................................................... 123
8-3 Future works .................................................................................................................... 129

Chapter 9. Results & Future Works ........................................................131


9-1 Summary of Results ........................................................................................................ 131
9-2 Future Directions ............................................................................................................. 134
9-2.1 Dispersion of Aqueous Reduced Graphene in Polymers ................................................ 134
9-2.2 Dispersion of Graphene in Solvents ............................................................................... 134
9-2.3 Glass Transition of Graphene/Polymer Nanocomposites ............................................... 135
9-2.4 Graphene/Polymer Nanocomposites .............................................................................. 135
9-2.5 Solvothermal Graphene Synthesis .................................................................................. 136
vii
References ...................................................................................................137
Appendix A .................................................................................................153
Glass Transition Temperature of PMMA with Incorporation of
Graphene via Solvent Blending ................................................................153
A-1 Introduction ...................................................................................................................... 153
A-2 Results and Discussion ..................................................................................................... 154
A-3 Conclusions ...................................................................................................................... 156

viii
List of Tables

Table 2-1. Compositions of XGO and GS ........................................................ 10


Table 3-1. GO and GS characteristics, values presented mean ± standard
deviation from triplicate measurements .................................................... 39
Table 2. Concentrations of GO and GS Leading to a 50% Lysis of RBCs ...... 43
Table 5-1. DSC polymerization results: ........................................................... 83
Table 6-1. Summary of properties of all PMMA nanocomposites. ................ 106
Table 7-1. Physical blending processes, including solvent and melt blending.
..................................................................................................................115
Table 7-2. Chemical blending processes (in situ polymerization or chemically
modified fillers) .......................................................................................116
Table 7-2. Cont' Chemical blending processes (in situ polymerization or
chemically modified fillers) .....................................................................117
Table 7-3. Aqueous blended processes (solvent blending but using water as
solvent for water-soluble matrix polymers) .............................................118

ix
List of Figures
Figure 1-1. Graphene can be considered as the mother material of buckycalls,
carbon-nanotubes and graphite.[1] (The figure was modified from
Novoselov et al.'s article[1]) ....................................................................... 2
Figure 2-1. Dispersions during reduction (XGO/DI-water, pH3): (a) before
reduction; (b) after 12 h (inset is a magnification of the top surface); (c)
after 24 h; (d) after 36 h; (e) after 48 h of heating. ..................................... 8
Figure 2-2. (a) Regular camera photo of the produced GS film with
highly-reflected "metallic-like" appearance, and (b) SEM image of the top
surface. ........................................................................................................ 9
Figure 2-3. AFM images of several GS on mica obtained in noncontact mode.
................................................................................................................... 10
Figure 2-4. Electrical resistance of GS films of different thicknesses as a
function of temperature measured using Van der Pauw method. ...............11
Figure 2-5. WXRD measurements from graphite, graphite oxide (GO),
exfoliated graphite oxide (XGO), and graphene sheets (GS) obtained at
different stages of graphite to graphene transition. ................................... 13
Figure 2-6. C1s XPS spectra and a simple molecular model describing the
structures (inset in left-hand side) of (a) XGO and (b) GS showing
reduction of C-O bonds from XGO to GS. The insets in upper-right
corners are suspensions of XGO and GS in DI water (pH3)
correspondingly. ........................................................................................ 15
Figure 2-7. Transformation of XGO to GS in DI water (pH3) at 95ºC as a
function of time. Initial XGO suspension (left vial). After 48h black and
“fluffy” precipitates with non-polar C-C bonds are observed (right vial). 16
Figure 2-8. GS reduced from XGO after 48 h at 95 ºC in DI water (pH=3) (a)
SEM image of GS film of many GS sheets. (b) AFM image of a single GS
sheet (top) with thickness profile across the flake (bottom). (c) Histogram
of number of layers per particle and (d) sheet sizes of GS. ...................... 17
Figure 2-9. (a) Low magnification TEM image of GS suspended over
holey-carbon film showing typical wrinkles occuring in most sheets. (b)
High magnification TEM image of GS showing amorphous and crystalline
regions. (c) FFT from crystalline regions of the image in (b) showing
hexagonal symmetry. (d) Schemetic of (b) where hexagons represent
crystalline regions and gray represents amorphous. (e) Electron diffraction
pattern from GS. (f) Change in intensity of -1010 spot as the GS is tilted
relative to incident beam. .......................................................................... 19
Figure 2-10. Schematic diagrams for reduction of hydrxyl and epoxy groups for
XGO to GS transformation. ...................................................................... 20
Figure 2-11. Fourier transform infrared spectra from XGO and GS. The spectra
were recorded in transmission mode......................................................... 22
Figure 3-1. WXRD spectra of (from top) graphite (precursor of GO; gray) as
x
received, graphite oxide (GO; red) as synthesized; graphene oxide (bGO;
blue) following 5-hour bath-sonication from GO; graphene oxide (pGO-5;
green) following 5-min probe-sonication from GO; graphene oxide
(pGO-30; purple) following 30-min probe-sonication from GO; and
graphene sheets (GS; black) following hydrothermal processing of pGO-5
in D.I. H2O for 20 h at 130 °C and pH=3. ................................................ 34
Figure 3-2. AFM topography images of (a) GO, (b) bGO, (c) pGO-5, (d)
pGO-30, (e) GS, and (f) cross-section topography of the white line in (c).
The cross-sectional topography shows step features of ~1 nm thickness
indicating that pGO-5 is mostly single-layered. ....................................... 35
Figure 3-3. AFM images of GO, bGO, pGO-30, and GS with cross-sectional
topography................................................................................................. 36
Figure 3-4. XPS C1s spectra and fitted curves of (a) GO and (b) GS. GS show
a significant drop in C-O (285.6 eV; red) character relative to C-C (283.5 eV;
blue). The XPS spectra show that there is significant reduction of oxygen
from GO (C/O=2/1) to GS (C/O=7/1). The purple curve is the fitted
baseline. .................................................................................................... 38
Figure 3-5. Photographs of GO, bGO, pGO-5, pGO-30, and GS sheets dispersed
at 50 μg mL-1 in D.I. water (a, b, c, d) and in PBS (e, f, g, h) after various
time points: samples right after preparation (0 hour) (a, e), after 3-hour
mixing agitation, (b,f), after 3-hour static aging (c, g), and after 24-hour
static aging (d, h). All samples were held at 37 oC. .................................. 41
Figure 3-6. (a) Percent hemolysis of RBCs incubated with different
concentrations (3.125 to 200 μg mL-1) of GO (red), bGO (blue), pGO-5
(green), pGO-30 (purple), and GS (black) for 3 hours at 37 oC with agitation.
Data represent mean ± SD from at least five independent experiments.
Percent hemolysis of RBCs incubated with pGO-30/chitosan at 100 μg
mL-1 for 3 hours at 37 oC with agitation. (b) Photographs of RBCs after
3-hour exposure to GO, bGO, pGO-5, pGO-30, and GS at different
concentrations (3.125 to 200 μg mL-1). The presence of red hemoglobin in
the supernatant indicates RBCs with membrance damage. (+) and (-)
symbols represent positive control and negative control, respectively. .... 43
Figure 3-7. Percent hemolysis of RBCs incubated with different concentrations
(3.125 to 200 μg mL-1) of graphite for 3 hours at 37 oC with agitation. Data
represent mean ± SD, n=3. Inset: photographs of RBCs after 3-hour
exposure to graphite with negative and positive controls. (+) and (-)
symbols represent positive control and negative control, respectively. .... 44
Figure 3-8. None of the five GO and GS samples adsorb significant amounts of
hemoglobin. Hemoglobin was mixed with GO, bGO, pGO-5, pGO-30, and
GS (100 μg mL-1 in PBS) and incubated for 3 hours at 37 oC. Each data
point is the average of three measurements of optical density of the
supernatant. ............................................................................................... 45
Figure 3-9. Optical microscopy images of RBCs in the presence of (a) PBS
(control), (b) pGO-5, (c) pGO-30, and (d) GS at 25 μg mL-1 for 3 hours at 37
o
C with agitation. Inset images are magnified RBCs. Scale bars in the inset
xi
images represent 10 μm. The arrows in (b) and (c) show lysed RBCs. The
arrows in (d) represent the hemagglutination caused by GS aggregates. . 46
Figure 3-10. Percent hemolysis of RBCs incubated with 100 μg mL-1 of pGO-30
and pGO-30/chitosan for 3 hours at 37 oC with agitation. Data represent
mean ± SD from three independent experiments. Inset: photographs of
RBCs after 3-hour exposure to pGO-30 and pGO-30/chitosan with negative
and positive controls. (+) and (-) symbols represent positive control and
negative control, respectively.................................................................... 47
Figure 3-11. Cell viability of human skin fibroblast cells determined from MTT
and WST-8 assay after 24-hour exposure to different concentrations of
pGO-5 and GS. Data represent mean ± SD. ............................................. 49
Figure 3-12. Photographs of MTT formazan (in the DMSO supernatant)
produced by pGO-5 and GS particles. [MTT]+ will react with the
conjugated electrons from GS to form the radical intermediate [MTT]●.
Then [MTT]● will react with the protons to form a protonated cation
[MTTH]+ under at least pH<13 environment. [MTTH]+ will then react
with another electrons from GS to form [FORMH], followed by a further
protonated reaction with another proton mainly from GS to form
[FORMH2]+. ............................................................................................. 49
Figure 3-13. Cell viability of human skin fibroblast cells determined from trypan
blue exclusion assay after exposure to different concentrations of pGO-5
and GS. Data represent mean ± SD, n=3. ................................................. 50
Figure 3-14. Optical microscopy images of human skin fibroblasts after 24-hour
incubation at 37 oC (5% CO2) in the presence of (a) medium (control), (b)
50 μg mL-1 of pGO-5, and (c) 50 μg mL-1 of GS particles. ...................... 50
Figure 3-15. Effect of pGO-5 and GS on the generation of ROS in human skin
fibroblast cells. .......................................................................................... 52
Scheme 5-1. Synthesis of urethane-acrylate oligomer (UAO): (i) Reaction of
HEMA and MDI (1:1.05 by mole) to form HEMA-isocyanate; (ii) reaction
of tri-functional polyol and HEMA-isocyanate (1:3.15 by mol) to name
UAO; (iii) UAO plus TPGDA diluent monomer were polymerized at 70 °C
with AIBN initiator to form polyurethane-acrylate, PUA. ....................... 72
Figure 5-1. Current-Voltage (I-V) curve for cured 2 wt% TRG/PUA
nanocomposite. ......................................................................................... 76
Figure 5-2. Rheology of un-cured TRG/UA uncured liquids. Dynamic strain
sweeps were run at room temperature. ..................................................... 77
Figure 5-3. Rheology of un-cured TRG/UA uncured liquids. Dynamic
frequency sweeps were run at room temperature. .................................... 77
Figure 5-4. TEM images of polymerized PUA with 0.25 wt% TRG at lower and
high resolution. ......................................................................................... 79
Figure 5-5. XRD patterns for cured TRG/PUA nanocomposites. .................... 79
Figure 5-6. (a) Normalized electrical resistance (R/Ro) of un-cured TRG/UA
liquids (●) and polymerized TRG/PUA nanocomposites (■) with various
TRG concentrations, measured using the 11-probe electrical resistance
measurement. The percolation concentration before polymerization is 0.5
xii
wt% and ~0.15 wt% after polymerization. (b) Electrical conductivity of
cured TRG/PUA nanocomposites measured by DC measurement; log c
plotted against Log(-perc), where perc is the percolation concentration
(0.07 vol%), c the electrical conductivity, and  the volume fraction (vol%)
of cured nanocomposites. The solid line was calculated based on Equation
(1), gives f = 0.23 S/cm, and t = 5.17. .................................................... 81
Figure 5-7. DSC scans of TRG/UA uncured liquids: (a) first scan: monitoring
polymerization reaction for polymerization heat (Hp) and peak
polymerization temperature (Tp); (b) second scan: glass transition
temperature (Tg; indicated by arrows) of polymerized TRG/PUA
nanocomposites. Both Hp and Tg show no dependence on TRG
concentrations. Tp increases with incorporation of TRG but no change with
various TRG concentrations (see Table 3-1). ............................................ 82
Figure 5-8. (a) Thermomechanical properties of TRG/PUA nanocomposites by
DMA at several TRG concentrations. (b) Mori-Tanaka model fit assuming
that the aspect ratio of TRG is 750 and Ef=250 GPa. TRG/PUA
nanocomposites show almost no reinforcement in glassy region due to the
low Ef/Ep ~70 but show large reinforcement in rubbery region due to the
high Ef/Ep ~104. ......................................................................................... 84
Figure 5-9. Loss modulus (E") of cured TRG/PUA nanocomposites by DMA.
................................................................................................................... 85
Figure 5-10. Summary of reported electrical percolation concentrations of TRG
polymers nanocomposites with different dispersion process.[78], [79], [84],
[94–98] TRG/PUA nanocomposites shows the lowest percolation
concentration. ............................................................................................ 86
Figure 6-1. 1H NMR spectra of syndiotactic-rich atactic PMMA (a-PMMA;
mm:mr:rr=6:38:56) and isotactic PMMA (i-PMMA; mm:mr:rr=96:3:1). 90
Figure 6-2. XPS (a) TRG, and (b) PG. Carbon/oxygen atomic ratios (C/O) are
9/1 for TRG and 49/1 for PG, fitting by Shirley-Gaussian distribution. ... 91
Figure 6-3. Electrical resistance of TRG/i-PMMA, TRG/a-PMMA and in situ
TRG/PMMA nanocomposites. The resistances of all systems drop
significantly at 1.0 wt% of TRG loading, indicating a similar dispersion
level (Af ~ 200) of TRG was achieved. ..................................................... 95
Figure 6-4. Normalized storage modulus of TRG/i-PMMA, TRG/a-PMMA and
in situ TRG/PMMA nanocomposites measured by DMA at several TRG
concentrations. Assuming filler modulus Ef = 250 GPa, Mori-Tanaka model
fitting (lines) gives an aspect ratios of ~200 for all. The results show a
similar dispersion level of TRG in i-PMMA and a-PMMA. .................... 96
Figure 6-5. DMA results (tan) of (a) TRG/a-PMMA, (b) TRG/i-PMMA, (c)
PG/i-PMMA and (d) in situ TRG/PMMA nanocomposites. ..................... 97
Figure 6-6. DMA results (loss modulus, E") of (a) TRG/a-PMMA, (b)
TRG/i-PMMA, (c) PG/i-PMMA and (d) in situ TRG/PMMA
nanocomposites. ...................................................................................... 101
Figure 6-7. DMA results (storage modulus, E') of (a) TRG/a-PMMA, (b)

xiii
TRG/i-PMMA, (c) PG/i-PMMA and (d) in situ TRG/PMMA
nanocomposites. ...................................................................................... 102
Figure 6-8. DSC curves of TRG/a-PMMA, TRG/i-PMMA, PG/i-PMMA and in
situ TRG/PMMA nanocomposites. ......................................................... 103
Figure 6-10. Three possible reactions during in situ polymerization with
presence of reduced graphene or GO. Hydrogen atom abstraction from GO
or reduced graphene followed by (1) coupling with radicals of propagating
polymer chains, or (2) initiation of polymerization. (3) A free radical of a
propagating polymer chain attacks the C=C on GO or reduced graphene
sheets. ...................................................................................................... 105
Figure 6-11. Tg measured by DSC of in situ TRG/PMMA nanocomposites and
their extracted matrix PMMA. ................................................................ 106
Figure 6-12. (a) DSC curves of extracted TRG, matrix in situ PMMA and in situ
TRG/PMMA nanocomposites (1.0 wt%). (b) AFM height images of the
extracted TRG. (The inset is the DSC curve of Extracted TRG.) ............ 107
Figure 8-1. "In situ" GO modulus measurement flowchart via AFM
nano-indentation. .................................................................................... 120
Figure 8-2. AFM images of patterned holes covered by broken single-layer GO
sheet. All the holes have the same diameter of 0.5 m and less than 1 m
thick......................................................................................................... 122
Figure 8-3. AFM image of a multi-layer GO covered hole (left) poked through
by the cantilever (right). .......................................................................... 123
Figure 8-4. The AFM sample grid was mounted on a TEM holder. .............. 124
Figure 8-5. AFM images of patterned holey carbon film covered by GO sheet.
All the holes have the same diameter of 2.5 m. .................................... 124
Figure 8-6. An example of locating the same GO sheet under TEM and AFM.
The arrows point out the feature to locate the GO sheet. The upper left insert
picture in the lower resolution TEM image is ED to check that the GO sheet
which covers the hole is single-layer. As mentioned in Chapter 2, six
electron diffraction patterns in hexagonal manner indicates single-layer
structure of GO. ...................................................................................... 125
Figure 8-7. Illustration how an AFM work. By moving the sample in Z-direction,
the sensor can detect the light position change to report the sample surface.
................................................................................................................. 126
Figure 8-8.

different runs. .......................................................................................... 127


Figure 8-9. Deflection-indentation depth curves of the same GO sheet with
different reduction times. ........................................................................ 129
Figure 8-10. Epoxy oxygen atoms form bridges between two neighboring
carbon atoms. The bridge is removed after reduction, leading to a lower
stiffness of GO. (The figure is modified from Mkhoyan et al.[16]) ....... 129
Figure A-1. (a) Glass transition temperature (Tg) and (b) DSC traces of
as-received PMMA (R-PMMA), as-precipitated (PMMA) and all
TRG/PMMA nanocomposites................................................................. 157
xiv
Figure A-2. Surface resistance and tensile modulus of as-received PMMA
(R-PMMA) as-precipitated PMMA (PMMA) and TRG/PMMA
nanocomposites. ...................................................................................... 158
Figure A-3. GPC chromatograms of as-received PMMA (R-PMMA),
as-precipitated (PMMA) and extractive content (EXC). Inset: optical
micrograph of R-PMMA......................................................................... 158
Figure A-4. 1H NMR spectra of (a) as-received PMMA (R-PMMA), (b)
as-precipitated (PMMA), and (c) extract (EXC). Signals from MMA (blue)
and EXC (red) are arrowed in figures (a) and (b). Expanded spectra with
the portion of EXC signals are also added as insets to clarify the removal
of EXC. ................................................................................................... 159

xv
Chapter 1. Introduction

1-1 Graphene & Graphene Oxide

Graphite is a naturally abundant material composed by stacked graphene layers.

Graphene, a single atomic layer of sp2-hybridized carbon covalently bonded in a

hexagonal lattice, is the thinnest material in the world. Graphene can be considered as

the mother material of buckyballs, carbon nanotubes and graphite[1] as shown in

Figure 1-1. Due to its single-atom-layer sp2 structure, graphene has attracted much

interests for its potential applications in electronics,[2], [3] and nanocomposites.[4]

Geim and Novoselov, the receivers of the Nobel Prize in Physics 2010, first discovered

a top-down method to obtain graphene by exfoliation of graphite using Scotch tape in

2004.[3] Meyer and coworkers[5] in 2007 demonstrated that a single atomic layer

graphene 2D crystal is thermodynamically stable. For reliable integration of graphene

into practical electronic devices it is essential to have a simple, reproducible and

controllable technique to produce large-area graphene sheets on a large scale.

Micromechanical cleaving[3], exfoliation of individual graphene sheets via sonication

in a solvent,[2], [6] and growth of graphene on SiC or nickel substrates[7–9] have been

reported as possible techniques to produce graphene.

One method to massively produce graphene is to use graphite oxide (GO) as a

precursor to synthesize graphene.[10–13] GO was successfully synthesized in 1958 by

Hummers[14] and it was also reported by Jeong and co-workers[15] that the oxidation

reaction to synthesize GO requires at least 60 min under acid environment to complete.

Oxygen atoms mainly exist on GO as hydroxyl, epoxy and carbonyl groups.[16], [17]

Due to oxygen functional groups, layered-structure graphite is intercalated and

1
becomes hydrophilic after Hummers' method.[18] The inter-layer distance of GO is

approximately 1 nm, whereas for graphite it is 0.34 nm.[17–19] Therefore, it is easier

to exfoliate GO into single-atom-layer sheet via simple colloidal dispersion

process.[18], [19]

Figure 1-1. Graphene can be considered as the mother material of buckycalls,


carbon-nanotubes and graphite.[1] (The figure was modified from Novoselov et
al.'s article[1])

1-2 Graphene/Polymer Nanocomposites

Clay reinforced polymer nanocomposites have been used in the automobile


2
industry by Toyota since 1991.[20] This class of materials offers large improvements

in mechanical properties with low filler loading.[21] The reinforcement of polymers

with stiff filler material depends on three criteria: (1) high aspect ratio, (2) high

mechanical properties, and (3) good adhesion between the polymer and filler.

Exfoliated clay individual layers fit two of the three criteria (high aspect ratio and

mechanical properties), whereas the adhesion between polymer and clay depends not

only on the clay itself but also on polymer properties. Oblate graphene can be used in

place of clay to reinforce polymers due to its high aspect ratio[22] and high elastic

modulus (~1TPa).[23–27] Furthermore, graphene is an especially unique conductor

with single atomic layer thickness and the high electrical conductivity of graphene

(~106 S/cm),[3], [28], [29] so a significant change in electrical conductivity of

polymers can be achieved with little incorporation of graphene. Graphite/polymer

composites are also electrically conductive; however, the low aspect ratio of graphite

makes this material not as desirable as graphene/polymer nanocomposites. This is

because the low aspect ratio of graphite provides a smaller number of effective

particles than graphene for the reinforcement of polymers.[21] From the reasons above,

production of graphene/polymer nanocomposite materials is our ultimate goal.

1.3 Overview of Thesis

This thesis addresses (1) synthesis and properties of graphene and (2) processing

and properties of graphene/polymer nanocomposites.

In Chapter 2, a novel graphene synthesis process is reported. It is important to

produce graphene using a low-cost, environmentally friendly process for use in

graphene/polymer nanocomposites. The graphene synthesis process utilizes a

3
dehydration reaction with no hazardous chemicals, which is expected to be

economical and environmentally friendly.

In Chapter 3, cytotoxicity of graphene oxide with different particle sizes is

studied. Toxicity is a crucial property for certain nanocomposite applications, for

example, food packaging materials. Both water-soluble tetrazolium (WST-8) and

methylthiazolyldiphenyl-tetrazolium bromide (MTT) assays are used to investigate

cytotoxicity of graphene & graphene oxide.

In Chapter 4, dispersion of aqueous reduced graphene (ARG; Chapter 2) in

thermoplastic polyurethane (TPU) is investigated. ARG/TPU nanocomposites are

prepared by a novel co-solvent blending process and are compared to thermally

reduced graphene (TRG)/TPU nanocomposites prepared by the typical solvent

blending process.

In Chapter 5, processing, structure, and properties of thermoset

graphene/polyurethane-acrylate (PUA) nanocomposites are presented. TRG is

dispersed into un-polymerized urethane-acrylate oligomers, followed by in situ

polymerization to synthesize TRG/PUA nanocomposites. Percolation concentration of

dispersed TRG is determined by several techniques.

In Chapter 6, the effects of stereo-regularity and processing on the glass transition

temperature (Tg) of TRG/PMMA nanocomposites will be reported. TRG/PMMA

nanocomposites are prepared by the solvent blending process and in situ

polymerization. Differing stereo-regularities of PMMA are also used as the matrix

polymer to investigate the effect of tacticity on nanocomposite Tg.

In Chapter 7, Tg of various kinds of graphene/polymer nanocomposites is studied.

We survey the literature to summarize how nanocomposite Tgs change with different

processing and matrix polymers.


4
In Chapter 8, the oxygen effect on elastic modulus of graphene oxide is studied.

We use "in situ" AFM nano-indentation to measure the modulus change of the same

single-layer graphene oxide sheet before and after chemical reduction reaction.

Several difficulties encountered are discussed.

Finally, Chapter 9 summarizes all key results of the graphene &

graphene/polymer nanocomposite and suggests future research directions.

5
Chapter 2. Graphene Synthesis1
2-1 Introduction

Graphene has zero band gap and high carrier concentrations and mobilities and,

as a result, exhibits remarkable electronic properties.[28], [30] For reliable integration

of graphene into practical electronic devices it is essential to have a simple,

reproducible and controllable technique to produce large-area graphene sheets on a

large scale. Micromechanical cleaving,[31] exfoliation of individual graphene sheets

via sonication in solvent,[2], [6] and growth of graphene on SiC or nickel

substrates[7–9] have been reported as possible techniques to produce graphene.

However, reduction of graphene from exfoliated graphite oxide/graphene oxide (XGO)

appears to be the most promising approach for large-scale production.[4], [32–35]

XGO can be obtained with a range of O:C stoichiometries depending on the

details of the particular synthesis. In XGO sheets the oxygen is bound to the carbon in

the form of hydroxyl and epoxy functional group[36], [37] and some carbonyl.[38]

Several reports suggest that XGO can be reduced to graphene either chemically by

exposing XGO to hydrazine (as a hydrate, as anhydrous hydrazine, or as

dimethylhydrazine vapor)[32], [39–41], or by rapid heating to high temperature[34] or,

alternatively, a combination of both.[33], [35], [38], [41] Toxicity of hydrazine and the

necessity of extreme temperatures, ~1000C for the thermal route, make it desirable to

seek other reduction methods. Here we describe a new hydrazine-free and aqueous

only route for reduction of graphene from XGO at atmospheric pressure and moderate

temperature. Other water-based reduction of graphene has been reported previously by

Zhou et al.,[42] Chen et al.[43] and Chen et al.[43]

1
This chapter appears in: Liao K-H, Mittal A, Bose S, Leighton C, Mkhoyan KA, Macosko CW ACS Nano 2011;5:1253.
6
2-2 Experiments & Characterization

2-2.1 Graphene & Graphene Oxide Synthesis

Graphite oxide was synthesized using modified Hummer’s method.[14] A

solution of 10 g of graphite (SP-1 graphite, from Bay Carbon) with average particle

size of 30 μm, 200 mL of concentrated sulfuric acid (H2SO4 98%), and 5 g sodium

nitrate (NaNO3) was prepared. 40 g of potassium permanganate (KMnO4) was then

slowly added to the solution, while keeping the mixture in an ice bath to maintain a

temperature of 35C. The solution was continuously stirred using a standard

magnetic stirrer. After an hour the solution became quite viscous and could not be

stirred any longer. The solution was held in ice bath till it reached 10C. The solution

was then transferred into a bigger container and was diluted with 2L deionized (DI)

water. The solution was dark brown at this point. Hydrogen peroxide (H2O2 30%)

was then slowly added until the solution turned green.

To aid filtration, 100 mL concentrated hydrochloric acid (HCl 37.5%) was

added to the mixture, and the whole solution was stirred for 30 min. Graphite oxide

was recovered from the solution after a four-step-wash procedure: (i) filtering

through a polytetrafluoroethylene (PTFE) membrane with 5 μm pore size, and after

filtration washing in 2 L DI-water with 30 min stirring followed by sonicating for 15

min, (ii) filtering through a PTFE membrane with 2 μm pore size, washing in 2 L

ethanol/100 mL HCl with 30 min stirring followed by sonicating for 15 min, (iii)

filtering through a PTFE membrane with 2 μm pore size, washing in 500 mL ether

with 30 min with stirring and again sonicating for 15 min, and finally, (iv) filtering

through a PTFE membrane with 2 μm pore size followed by washing with 500 mL

7
ether. The final product (graphite oxide) was air dried after the final filtration for 30

min. The order of ether was still present.

To obtain exfoliated graphite oxide/graphene oxide (XGO), 0.45 g graphite

oxide was dispersed in 500 mL DI water (pH~3) and sonicated (Branson 2510; 100

W) for 60 min at room temperature. Graphene sheets (GS) were obtained from XGO

by heating at 95C in the lightly capped jar shown in Figure 2-1. DI water was added

periodically to replace that lost by evaporation. A sealed pressure cooker (PRESTO,

6-Quart) was used to heat DI water to 120C. The transformation of XGO into GS is

illustrated in Figure 2-1 by series of images of the vial with solution at different

times.

Figure 2-1. Dispersions during reduction (XGO/DI-water, pH3): (a) before


reduction; (b) after 12 h (inset is a magnification of the top surface); (c) after 24 h;
(d) after 36 h; (e) after 48 h of heating.

GS film was prepared as following steps. After the reduction process described

above, the clear water (Figure 2-1(e)) was removed and the black precipitation was

re-dispersed in 500 mL dimethylformamide (DMF). The solution was then sonicated

for 5 min (Branson 2510; 100 W) and was slowly filtrated through a PTFE

8
membrane. The black filtrate cake was dried in a vacuum oven (<0.05 bar) at 80 oC

for 24 h. The produced dry filtrate cake (GS film; Figure 2-2) with metallic

appearance was the GS film used in the following electrical resistivity measurement.

Figure 2-2. (a) Regular camera photo of the produced GS film with
highly-reflected "metallic-like" appearance, and (b) SEM image of the top
surface.

2-2.2 Characterization

2-2.2.1 Atomic Force Microscopy (AFM)

A drop of dispersed GS in dimethylformamide (DMF) was dropped on

silicon wafer substrate and air-dried at room temperature. A Digital Instrument,

Nanoscope III Multimode AFM at the Institute of Technology Characterization

Facility of the University of Minnesota was used in noncontact mode for

measurements. Random lengths of GS sheets (N=300) were measured from several

AFM images. Examples of such AFM images are presented in Figure 2-3.

9
Figure 2-3. AFM images of several GS on mica obtained in noncontact mode.

2-2.2.2 X-Ray Photoelectron Spectroscopy (XPS)


Standard XPS measurements were conducted on XGO and GS powders using

Surface Science SSX-100 spectrometer housed at Characterization Facility of the

University of Minnesota. The compositions obtained from these XPS measurements

are listed in Table 2-1.

2-2.2.3 Elemental Analysis (EA)


EA was conducted at Atlantic Microlab, Inc. The samples of XGO and GS

were initially dried in vacuum oven at 25C for 5 h. The results of analysis are listed

in Table 2-1.

Table 2-1. Compositions of XGO and GS

XPS (mol.%) EA (mol.%)


O: 30%
O: 39%
XGO C: 59%
C: 61%
H: 8%

O: 13%
O: 16%
GS C: 81%
C: 84%
H: 5%

10
2-2.2.4 Electrical Resistivity Measurements.
We measured electrical resistivity of XGO and GS films with different

thicknesses using Van der Pauw method. All samples were measured in a 4-wire Van

der Pauw configuration with indium metal contacts. Measurements were made using

an LR-700 AC resistance bridge at an excitation frequency of 20 Hz. All contacts

were found to be Ohmic down to the lowest temperature of measurement, and

anisotropies were small. The results of the measurement for GS films of different

thicknesses are presented in Figure 2-4. These measurements indicate non-bulk

behavior and therefore cannot be used as a direct measure of resistivity of single GS.

However, they can be used to estimate the changes in resistivity from XGO to GS.

The ratio of the resistivity measured at room temperature from similarly prepared

thin films of GS and XGO (both with thickness of about 40m) was found to be
GS
 4 105 .
XGO



3 100 m
10
80 m
50 m
2 40 m
10
 (-cm )
2

1
10

0
10

-1
10
0 50 100 150 200 250 300
T (K)

Figure 2-4. Electrical resistance of GS films of different thicknesses as a function


of temperature measured using Van der Pauw method.

11
2-2.2.5 Transmission Electron Microscopy (TEM)
Samples for TEM study were prepared by a dropping a solution of GS

dissolved in DMF onto standard holey-carbon-film-covered Cu grid, which was left

for air-dry before loading into microscope. For these experiments, FEI Tecnai F30

300 keV (S)TEM at the University of Minnesota was used. The microscope is

equipped with Schottky field emission gun, a high-resolution pole piece with

spherical aberration coefficient of Cs=1.3 mm. For quantitative analysis of the

diffraction patterns obtained from single GS, each pattern was calibrated to same

incident beam by using the maximum intensity point in each pattern as a reference.

The intensity of diffracted spot is computed by averaging the intensities of 25 pixels

forming the spot. A slight decrease in intensities (as a function of time) due to beam

damage was observed and corresponding linear correction was applied.

2-2.2.6 Fourier Transform Infrared Spectroscopy (FTIR)


Nicolet Macna - IR760 Spectrometer was used to measure FTIR spectra from

XGO and GS. Both XGO and GS were dried in vacuum at 45C for 24 h and were

individually ground together with potassium bromide (KBr) at a mass ratio of

XGO(GS):KBr = 1:10. Ground powders were compressed into transparent pellets for

the FTIR measurements.

2-3 Results and Discussion

Large exfoliated individual graphene sheets (GS) were obtained following a

three-step process: (i) transition from graphite into graphite oxide (GO), (ii)

intercalation of GO with water to form XGO and, finally, (iii) exfoliation with
12
conversion of XGO into GS. In the first step, GO was synthesized from graphite, with

an average particle size of 30 μm, using a modified Hummer’s method[14] (the details

can be found in Supporting Information). Wide-angle X-ray diffraction (WXRD)

measurements were performed on the original graphite and the obtained GO, Figure

2-5. The measurements confirm the expected shift in the main peak: the 2θ Bragg

angle moved from 26.5, corresponding to d0001= 0.34 nm spacing between atomic

planes in graphite, to 10.1 in GO, which corresponds to an interplane distance of

d0001= 0.94 nm.[44], [45]

Figure 2-5. WXRD measurements from graphite, graphite oxide (GO),


exfoliated graphite oxide (XGO), and graphene sheets (GS) obtained at different
stages of graphite to graphene transition.

In the second step, GO particles were dispersed in DI water at a concentration

of about 0.9 mg/mL (pH3) then sonicated for about 1 h at room temperature[46] (at

40 kHz with power of 100W). The WXRD measurement from the XGO in Figure 2-5

shows further position reduction and broadening of the Bragg peak to 2θ=8 (d = 1.12
13
nm). We speculate that the 0.22 nm increase in distance between exfoliated layers is

due to water swelling of GO during sonication. The increase in d-spacing is slightly

less than the size of a water molecule, ~0.28 nm. XGO layers are not perfectly flat. A

roughness of ~0.6 nm[16] has been measured which may provide space to

accommodate water molecules.

X-ray photoelectron spectroscopy (XPS) and elemental analysis (EA) were

performed on these XGO sheets. Figure 2-6(a) shows that the C1s spectrum is a

superposition of two strong peaks at 286.2 eV and 284.4 eV that are fingerprints of

C-O (including epoxy and hydroxyl groups) and C-C bonds.[38], [47], [48]

Decomposition of the C1s XPS spectrum into these main components indicates that the

fractions of C-O and C-C bonds in XGO are 70% and 30% respectively. Although

some C=O and C(=O)-(OH) bonds (with corresponding peaks at 287.5 and 289.2 eV)

are also expected to be present in XGO,[39] no visible amounts are detected in these

samples. Analysis of the XPS measurements indicates that the oxygen to carbon ratio

in XGO is close to O:C=1:2. The same ratio is also found from the standard EA

measurements (see SI). Since it is energetically favorable for oxygen to be bonded to

two carbon atoms,[16] the XPS and EA measurements indicate that at this stage the

XGO sheets are heavily oxidized, which is in good agreement with previous

reports.[28], [38], [47]

To obtain GS from these XGO, the sample was heated in DI water at 95ºC for

more than 70 h. XPS measurements performed on the final material show a

significant drop in C-O relative to C-C bonds (see Figure 2-6(b)). Decomposition of

the C1s XPS spectrum obtained from GS into components indicates that the fractions

of C-O and C-C bonds now are 31% and 69% respectively. The ratio of oxygen to

carbon is reduced to about O:C=1:6, a value confirmed by EA.


14
Figure 2-6. C1s XPS spectra and a simple molecular model describing the
structures (inset in left-hand side) of (a) XGO and (b) GS showing reduction of
C-O bonds from XGO to GS. The insets in upper-right corners are suspensions
of XGO and GS in DI water (pH3) correspondingly.

15
Figure 2-7 shows that at 95ºC the XGO to GS transformation is completed

within 48 h. The appearance of black “fluffy” precipitates at the end of

transformation, indicative of a dramatic increase of non-polar C-C bonds and thus

reduction. Using the same procedure at 80ºC did not produce the XGO to GS

transformation even after 5 days. At 120ºC (See SI) GS with the same characteristics

as at 95ºC was obtained after 12 h.

Figure 2-7. Transformation of XGO to GS in DI water (pH3) at 95ºC as a


function of time. Initial XGO suspension (left vial). After 48h black and “fluffy”
precipitates with non-polar C-C bonds are observed (right vial).

Despite this considerable amount of oxygen still present in the GS, the sheets

have very low electrical resistance. The ratio of the resistance measured with the van

der Pauw method at room temperature from similarly prepared thin films of GS as
RGS
XGO (both with thickness of about 40m) was found to be  4 105 (Figure
RXGO
2-4). An SEM image of a typical GS film is presented in Figure 2-8(a). Such five

order of magnitude reduction in resistance going from


 XGO to GS is in agreement

with previously reported data.[33], [38], [49] The resistance of XGO and GS are

16
highly sensitive to the fractions of the sp2 and sp3 bonds of carbon atoms in the

sheets and concentration of oxygen atoms, which act as scattering centers. While

reduction of oxygen in XGO is essential for minimizing the density of scattering

centers, its effects on sp3 to sp2 transitions of carbon bonds are not clear. Our

measurements indicate that, despite the relatively large fraction of oxygen still

remaining in the GS, the oxygen reduction method described here provides sheets

that contain large interconnected patches of oxygen-free graphene with sp2 bonds.

Figure 2-8. GS reduced from XGO after 48 h at 95 ºC in DI water (pH=3) (a)


SEM image of GS film of many GS sheets. (b) AFM image of a single GS sheet
(top) with thickness profile across the flake (bottom). (c) Histogram of number
of layers per particle and (d) sheet sizes of GS.

Thicknesses of 300 individual GS were measured using atomic force

microscopy (AFM) (Figure 2-8). The thickness of individual GS is about 1 nm,

which is smaller than 1.6 nm, the reported thickness of XGO,[47] consistent with

thickness measurements of graphene.[50], [51] The statistical analysis of AFM

measurements indicates that approximately 65% of the sheets are single layers (with

a thicknesses ranging from 0.6 nm to 1.2 nm), 27% are double layered (1.2-2.4 nm),

and the remaining 8% are multi-layered (>2.4 nm; Figure 2-8(c)). The average

diameter of GS is ~300nm (see Figure 2-8(d)).

17
Transmission electron microscopy (TEM) was used to further analyze oxygen

reduced GS. A FEI Tecnai F30 300 keV (S)TEM at the University of Minnesota

equipped with a Schottky field emission gun and a high-resolution pole piece with

spherical aberration coefficient of Cs=1.3 mm was used. Figure 2-9(a) shows a low

magnification image of a typical GS suspended over the holey-carbon-film covering a

standard TEM copper grid. All expected characteristics of the single sheet, including

numerous wrinkles, can be seen. Detailed examination of high-resolution images of

individual sheets (see Figure 2-9(b)) suggest that the sheet consists of two regions: (i)

large uniform regions of minimum contrast variation and hexagonal symmetry, which

can be seen in the fast fourier transform (FFT) pattern shown in Figure 2-9(c), and, (ii)

amorphous-like regions in between with strong contrast variation. The uniform areas

in the sheets are characteristic of oxygen free regions34 and the fact that they are

connected is consistent with resistance measurements. These regions make up ~50%

of the area in the high resolution image. If we assume that the area appearing

amorphous is due to oxidation, and that the O:C ratio in these areas is 1:2, then based

on the area ratios the total sheets should have O:C ratio of 1:6. This is in agreement

with our XPS and EA result.

18
Figure 2-9. (a) Low magnification TEM image of GS suspended over
holey-carbon film showing typical wrinkles occuring in most sheets. (b) High
magnification TEM image of GS showing amorphous and crystalline regions. (c)
FFT from crystalline regions of the image in (b) showing hexagonal symmetry. (d)
Schemetic of (b) where hexagons represent crystalline regions and gray
represents amorphous. (e) Electron diffraction pattern from GS. (f) Change in
intensity of -1010 spot as the GS is tilted relative to incident beam.

To confirm that the layers studied here are indeed single layers of GS, we

analyzed the change in intensity of the (-1010) diffraction spot as the sample was tilted

with respect to the incident beam over a range of 30 degrees. For an infinitely thin

sample, the intensity should remain constant as the sample is tilted. A

single-atomic-layer-thin graphene, to a good approximation, satisfies this condition

and the intensity of the (-1010) diffraction spot should remain constant over a large
19
range of tilt. Meyer et al.[5] have shown that this approach can be used to distinguish

between single and bilayer graphene (in the case of bilayer graphene four order of

magnitude changes in diffracted spot intensity are expected). Figure 2-9(e) shows a

diffraction pattern from GS supported by an amorphous carbon film. Figure 2-9(f)

shows the changes in intensity for the (-1010) diffraction spot as a function of tilt angle,

confirming that the GS is indeed a single layer. The fact that the intensity of the (-2110)

diffracted spot is considerably lower than that of the (-1010) spot provides additional

confirmation of a single layer.[2]

Figure 2-10. Schematic diagrams for reduction of hydrxyl and epoxy groups for
XGO to GS transformation.

The oxygen reduction and simultaneous transformation of the carbon sp3

bonds into sp2 can be explained by dehydration of XGO in DI water. If hydroxyl


20
groups and hydrogen atoms are attached to two neighboring carbons, in an acidic

environment they can combine through dehydration reaction resulting in H2O and GS

with sp2 bonded carbon atoms. For epoxy groups, the reduction is a two-step process.

If an epoxy group is attached to carbon atoms of GS with two hydrogen atoms

attached to the neighboring carbons, in an acidic environment the system will first

hydrate transforming the epoxy group (-O-) to two hydroxyl groups (-OH), which

then reduce to H2O and GS with sp2 bonded carbon atoms. Schematic diagrams for

both reactions are presented in Figure 2-10. Since presence of hydrogen atoms next

to hydroxyl groups is needed for oxygen reduction, their availability will set the limit

of oxygen reduction in the XGO to GS transformation, which can explain the

presence of residual oxygen in the GS. Fourier transform infrared spectroscopy

(FTIR) performed on XGO and GS specimens before and after transformation

(Figure 2-11, for details see SI) supports dehydration as the main mechanism for

oxygen reduction. Figure 2-11 shows significant reduction of the hydroxyl group[52]

(2900-3600 cm-1) and some reduction of epoxy groups.

To achieve reduction of XGO sheets in water a low pH environment and

moderately high temperature are needed. H+ catalyst is essential for a dehydration

reaction, where the kinetics of the reaction are governed by temperature (we

observed a four-fold reaction rate increase going from 95ºC to 120ºC). It should be

noted that the reduction of XGO sheets can also take place in a basic environment

with HO- acting as a catalyst. Fan et al.,[53] have reported deoxydation of graphene

oxide sheets in alkaline water. In either case, the amount of hydrogen atoms bonded

to carbon next to hydroxyl groups sets the limit of oxygen reduction. The role of H+

catalyst in dehydration of XGO is the critical difference between oxygen removal

processes in water and in air. Unlike reduction in water, extreme high temperatures
21
(> 600ºC) are needed to observe any considerable oxygen reduction in air due to

direct thermal excitations.[34], [38] In addition thermal reduction generates CO2,

removing carbon from the sheets resulting in defect formation and size reduction .

Figure 2-11. Fourier transform infrared spectra from XGO and GS. The spectra
were recorded in transmission mode.

2-4 Conclusions

In conclusion, we successfully developed a hydrazine-free, high-yield method

to produce single-layer graphene with high sp2 fraction and conductivity at

atmospheric pressure. Graphene sheets are reduced from exfoliated graphite

oxide/graphene oxide in DI water at pH3 within 48 hours at temperature of 95ºC.

The graphene sheets produced according to this method have a resistivity comparable

to graphene sheets produced using other methods. These measurements, combined

with TEM observations, indicate that while these sheets are not fully reduced, the

22
presence of large uniform interconnected oxygen-free patches with sp2 carbon bonds

is sufficient to provide sheets with electronic properties comparable to other reduced

graphene sheets. We speculate that heating XGO for extended time in water provides

enough energy for dehydration and transformation of C-C bonds from sp3 to sp2.

23
Chapter 3.

Cytotoxicity of Graphene Oxide and Graphene2


3-1 Introduction

Graphene, a two-dimensional carbon sheet with single-atom thickness, has

recently attracted significant interest due to its unique mechanical and electrical

properties.[1] This newly discovered material has a wide range of potential

applications including transistors,[2] transparent conductors,[3] surfactants,[54]

polymer reinforcement,[4] and biodevices.[55] Recently, similar to carbon nanotubes

(CNTs), biological applications of graphene sheets (GS) and graphene oxide (GO)

have attracted attention in the scientific community based on their great potential for

bacterial inhibition,[56] drug delivery,[57] and photothermal therapy.[10] Graphene

morphology is distinct from that of CNTs; for example, the length of CNTs influence

their toxicity but GS and GO don't have a "length". One similarity between the

materials is that both GO/GS[4] and CNT[58] structures vary according to the

synthetic processes employed, which can also change their physical properties, such as

dispersity, surface functionality, and their toxicity.[59] Several methods have been

reported to produce graphene economically.[18], [51] The method employed herein

starts with exfoliation of graphite oxide to single-layered GO followed by reduction of

the GO to GS. GO can be chemically reduced without changing the sheet size because

hydroxyl and epoxy groups are converted into carbon-carbon double bonds by

dehydration or other reactions.[18], [60]

2
This chapter appears in: Liao K-H, Lin Y-S, Macosko CW, Haynes C. ACS Applied Materials & Interfaces 2011;3:2607.
24
Even with the aforementioned promise of graphene-based materials, there are

only a few studies investigating the in vitro cytotoxicity of GO and GS to baterial or

mammalian cells, all published in 2010 and 2011.[11–13], [57], [60], [61] For toxicity

of graphene and its derivatives to bateria, Hu et al. reported that GO and reduced GO

(rGO) inhibit bacterial growth with minimal toxicity to human alveolar epithelial

A549 cells.[57] Akhaven et al. compared the toxicity of GO and rGO on Escherichia

and Staphylococcus bacteria and found (1) that GO and rGO caused bacterial

membrane damage and (2) that the hydrazine-reduced GO was more toxic than

untreated GO.[60] They attributed the toxicity of their reduced GO to sharper

nanowalls. In the case of cytotoxicity of GO and GS to adherent mammalian cells,

Biris and co-workers demonstrated that both GS and CNTs induce cytotoxic effects on

phaeochromocytoma (PC-12) cells; they also concluded that the CNTs are more toxic

than graphene and that the shape of these carbon-based nanomaterials plays an

important role in their cytotoxicity.[61] In very recent work, Wang et al. demonstrated

that GO has dose-dependent toxicity to human fibroblast cells with the GO causing

obvious toxicity when the dose is higher than 50 µg/mL.[11] On the contrary, two

other very recent reports show high biocompatibility of GO or GS.[12], [13] Ryoo et al.

reported that GO and GS substrates were highly biocompatible and improved gene

transfection efficacy in NIH-3T3 fibroblasts.[12] Chang et al. suggested GO will not

enter A549 cells and showed no obvious toxicity to A549 cells, regardless of the size

or dose of GO.[13] To summarize, the limited published work indicates that both GO

and GS have high bacterial toxicity but there is no consensus on cellular toxicity. One

likely source of this apparent disagreement is that the physicochemical properties of

GO or GS, such as size, surface charge, particulate state, surface functional groups,

and residual precursors, are not always well controlled though they likely have
25
significant influence on biological/toxicological activity.[62] Another possibility is

that the most commonly used viability assay, the MTT assay, is not appropriate for

work with carbon-based nanomaterials like carbon black and CNTs.[63], [64]

To date, the reported literature only probes the in vitro toxicity of GO and GS in

bacterial, adherent mammalian, or cancerous cells. Likely due to the recent discovery

and progress on graphene, none of this work has investigated and compared the effect

of particulate state of GO and GS on the response of suspended cells like red blood

cells and adherent cells like human skin fibroblast cells. In this work, we aim to

systematically study the effects of GO exfoliation, size, oxygen content, and

particulate state on red blood cells (RBCs), a likely site of interaction for biomedical

applications that require intravenous injection, as well as human skin fibroblasts, a

likely target upon dermal interaction. RBC toxicity is assessed by tracking hemolysis,

or the release of hemoglobin upon cell lysis, under various nanomaterial exposure

conditions while fibroblast toxicity is assessed with a comparison between the

well-established methylthiazolyldiphenyl-tetrazolium bromide (MTT) and water

soluble tetrazolium salt (WST-8) assays. Together, these data reveal some critical

chemical and physical parameters that determine the biocompatibility and promise of

these exciting new materials. Based on likely broad future use of GO and GS in a

variety of products, it is critical to consider cytotoxicity of well-characterized

graphene materials on cell types that are most populous in the body and likely to

interact with foreign materials.

26
3-2 Experiments & Characterization.

3-2.1 Synthesis of Graphene and Graphene Oxide

A graphite oxide particle suspension was synthesized using Hummers’

method;[14] and a detailed process has been reported in our previous paper.[65] We

chose this simple, hydrazine-free method to synthesize graphene oxide and graphene

sheets to avoid any unintentional toxicity from the highly-toxic hydrazine species.

Graphite oxide was dialyzed (Spectrum Laboratories; 5 nm pore size) against

deionized (D.I.) water for 48 hours. The D.I. water was changed every 4 hours. To

obtain graphene oxide (GO), graphite oxide was redispersed in D.I. water at 1 wt%

concentration and sonicated with a bath-sonicator (Branson 2510; 100 W) or

probe-sonicator (Cole Parmer; CPX750, 750W) for various sonication times as

described in the Characterization section. GS was produced from pGO-5 using the

dehydration process reported in our previous paper.[65] To provide an acid

environment for dehydration of hydroxyl groups8 on pGO-5, HCl was added into the

pGO-5/D.I. water solution (1 wt%; yellow solution) until the pH reached ~3. A sealed

pressure cooker (PRESTO, 12-Quart) was used to heat the solution to 130 C for 20

hours. The solution turned from homogeneous yellow to a layered suspension with

clear water on the top and fluffy black precipitate on the bottom. All the GO and GS

samples used for cytotoxicity experiments were dialyzed in regenerated cellulose

tubing (with a molecular weight cut off, of 12,000-14,000, Fisherbrand, Pittsburgh, PA)

against two liters of D.I. water for at least 6 days. The pH of GO and GS suspensions

after dialysis was around 5.

27
3-2.2 Characterization

3-2.2.1 Wide-angle Powder X-ray Diffraction (WXRD)

GO and GS powders were obtained after drying the aqueous GO and GS

suspensions using a rotary evaporator (IKA RV10, Wilmington, NC) equipped with a

self-cleaning dry vacuum system (WELCH, Niles, IL). The WXRD measurements

for GO and GS were performed on a Bruker-AXS D-5005 (Siemens) with filtered Cu

Kα source (2.2 kW) at 45 kV and 40 mA. Data were recorded by step scanning with a

step size of 0.040° and a step time of 1.0 second.

3-2.2.2 Atomic Force Microscopy (AFM)

A single drop of GO, bGO, pGO-5, and pGO-30 aqueous suspensions were

spin-coated onto separate mica substrates (Ted Pella, Redding, CA) at 4000 rpm for 30

seconds and air-dried at 45 °C for 3 hours. A drop of dispersed GS in co-solvent (10

wt% D.I. water and 90 wt% dimethylformamide) was dropped onto a mica substrate

and air-dried at 65 °C for 3 hours. Non-contact mode AFM (Digital Instrument,

Nanoscope III Multimode AFM) was used for measurements.

3-2.2.3 X-ray Photoelectron Spectroscopy (XPS)

GO, bGO, pGO-5, pGO-30 and GS were dried in a vacuum oven at 65 °C for 3

hours before the XPS measurements. The XPS spectra were recorded on a Surface

Science SSX-100 spectrometer equipped with a monochromatic Al Kα radiation

source.

28
3-2.2.4 Dynamic Light Scattering (DLS)

The hydrodynamic diameter of GO, bGO, pGO-5, pGO-30 and GS was

determined using a 90Plus/BI-MAS particle size analyzer (Brookhaven Instruments

Corporation, Holtsville, NY). All the particles were suspended in D.I. water and PBS

at a concentration of 50 μg mL-1. Three runs and one minute run duration were set for

each measurement. Each measurement was repeated on three cuvettes taken from the

same solution.

3-2.2.5 ζ-Potential Measurements

All GO and GS samples were prepared in D.I. water, PBS, and cell culture

medium at a concentration of 50 μg mL-1. The ζ-potential was determined using a

ZetaPALS Zeta-Potential Analyzer (Brookhaven Instruments Corporation, Holtsville,

NY) equipped with a 35 mW red diode laser (660 nm). Five runs and ten cycles were

set for each measurement. Each sample was measured at least three times.

3-2.2.6 Optical Microscopy

Following 3-hour exposure to pGO-5, pGO-30, and GS sheets (25 μg mL-1) at 37


o
C with agitation, washed RBCs were transferred to a 35 x 10 mm Petri dish (Falcon,

Franklin Lakes, NJ) and observed under a Nikon Eclipse TE 2000-U inverted

microscope (Nikon USA, Melville, NY). The optical images were recorded using a

QUANTEM: 512SC camera (Photometrics, Tucson, AZ) with MetaMorph imaging

software (Molecular Devices, Downingtown, PA).

29
3-2.2.7 Hemolysis Assay

Fresh ethylenediamine tetraacetic acid (EDTA)-stabilized human whole blood

samples were purchased from Memorial Blood Center (St. Paul, MN). Typically, 5 mL

of whole blood was added to 10 mL of calcium- and magnesium-free Dulbecco’s

phosphate buffered saline (PBS, Grand Island, NY) and centrifuged at 500 g for 10

minutes to isolate RBCs from serum. This purification step was repeated five times,

and then the washed RBCs were diluted to 50 mL in PBS. To test the hemolytic

activity of GO and GS samples, 0.2 mL of diluted RBC suspension (around 4.5 x 108

cells mL-1) was added to 0.8 mL of GO and GS suspension solutions in PBS at

different concentrations. The 250 μg mL-1 of GO or GS stock solutions in PBS were

prepared by adding 0.8 mL of 500 μg mL-1 of GO or GS particle solutions in D.I.

water to 0.8 mL of 2X PBS solutions. The final concentration of GO and GS ranges

from 3.125 to 200 μg mL-1. D.I. water (+RBCs) and PBS (+RBCs) were used as the

positive control and negative control, respectively. All the samples were placed on a

rocking shaker in an incubator at 37 oC for 3 hours. After incubation, the samples were

centrifuged at 10,016 g for 3 minutes. The hemoglobin absorbance in the supernatant

was measured at 540 nm, with 655 nm as a reference, using an iMark microplate

reader (Bio-Rad, Hercules, CA). Percent hemolysis was calculated using equation 3-1.

(3-1)

3-2.2.8 MTT Viability Assay

Human skin fibroblast cells (CRL-2522) were purchased from American Type

Culture Center (Manassas, VA). Typically, 6×104 cells were seeded in 96-well plates
30
and cultured in Minimum Essential Medium Eagle (MEM, Hyclone, Logan, UT) with

Earle’s balance salt and L-glutamine, supplemented with 10% fetal bovine serum

(FBS, Invitrogen, Grand Island, NY) and 1% penicillin/streptomycin (PS, Invitrogen,

Grand Island, NY) at 37 oC under 5% CO2. After 24 hours, the cells were washed with

100 μL of serum-free MEM (1% PS) twice and incubated with 100 μL of different

concentrations of pGO-5 and GS suspensions in serum-free MEM (1% PS). The

pGO-5 and GS particles used for viability assays were washed with serum-free MEM

(1% PS) five times. After 24-hour exposure, the cells were washed twice with 100 μL

of serum-free MEM and incubated with 100 μL of 0.5 mg/mL

methylthiazolyldiphenyl-tetrazolium bromide (MTT, Invitrogen, Eugene, OR)

containing media for 2 hours at 37 oC under 5% CO2. Finally, the MTT containing

media was removed and the insoluble purple formazan crystals produced by live cells

were dissolved in 100 μL of dimethyl sulfoxide (DMSO, Sigma-Aldrich, Milwaukee,

WI). The plate was placed on a rocking shaker for at least 20 minutes and then 80 μL of

the purple DMSO solution in each well was transferred to a new 96-well plate, because

residual pGO-5 or GS can affect the absorbance values at 570 nm. Optical density of

the produced stain was monitored at 570 nm, with 655 nm as a reference, using an

iMark microplate reader (Bio-Rad, Hercules, CA). The cell viability was determined

by mitochondrial activity which was calculated using equation 2. Cells without

particle exposure were used as control. In addition, cell-free control experiments were

performed to see if the GO and GS react directly with the MTT reagents. Typically,

pGO-5 and GS particles with different concentration (3.125 -200 µg/mL) were

suspended in 1 mL of 0.5 mg/mL of MTT solution (in MEM). After two-hour

incubation at 37 oC under 5% CO2, the pGO-5 and GS particles were centrifuged and

washed with PBS one time. To see if any insoluble formazan formed during incubation,
31
1 mL of DMSO was added to redisperse the pelleted particles. The suspended pGO-5

and GS particles were centrifuged again and the optical density of DMSO supernatant

at 570 nm (655 nm as reference) was used to see if the MTT reagent reacted with

pGO-5 and GS particles.

(3-2)

3-2.2.9 WST-8 Viability Assay

In addition to the MTT assay, cell viability was measured using a cell counting

kit-8 (CCK-8, Dojindo, Rockville, MD). Typically, 6×104 cells were seeded in a

96-well plate and cultured in MEM supplemented with 10% FBS and 1% PS at 37 oC

under 5% CO2. After 24 hours, the cells were washed with 100 μL of serum-free MEM

(1% PS) two times and incubated with 100 μL of different concentrations of pGO-5

and GS suspensions in serum-free MEM (1% PS). After 24-hour exposure, the cells

were washed twice with serum-free MEM and 15 μL of CCK-8 solution was added to

each well containing 100 μL of serum-free MEM. After one hour incubation at 37 oC

under 5% CO2, 80 μL of the mixture was transferred to another 96-well plate, because

residual pGO-5 or GS can affect the absorbance values at 450 nm. The absorbance of

the mixture solutions was measured at 450 nm with 655 nm as a reference, using an

iMark microplate reader. The viability was calculated using equation 3. The cell-free

control experiments were performed to see if the GO and GS react directly with the

WST-8 reagents. Typically, 100 µL of pGO-5 and GS particles with different

concentration (3.125-200 µg/mL) were added to a 96-well plate and 10 µL of WST-8

reagent solution was added to each well; the mixture solution was incubated at 37 oC

under 5% CO2 for one hour. After incubation, the pGO-5 and GS particles were

32
centrifuged and 50 µL of supernatant was transferred to another 96-well plate. The

optical density at 450 nm (655 nm as reference) of each control was used to see if the

pGO-5 and GS particles react with WST-8 reagents.

(3-3)

3-3 Results and Discussion

3-3.1 Synthesis and Characterization of GO and GS.

High yields of GO (by Hummers’ method[14]) and GS with various sizes, extent

of exfoliation and oxygen content were synthesized using a simple aqueous and

hydrazine-free hydrothermal route accompanied by different sonication treatments.[18]

The sample preparation flowchart is summarized in Scheme 3-1. The following

samples were characterized by WXRD as dried powders: (1) graphite oxide colloidal

suspension obtained directly from Hummers’ method[14] (GO); (2) graphene oxide

obtained after 5-hour bath-sonication of GO (bGO), (3) graphene oxide obtained after

5-minute probe-sonication of GO (pGO-5), (4) graphene oxide obtained after

30-minute probe-sonication of GO (pGO-30), (5) and graphene sheets obtained after

hydrothermal processing of pGO-5 in D.I. water for 20 hours at 130 °C and pH~3 (GS).

Both bath- and probe-sonication were employed because the differing sonication

intensities allowed access to a broad range of material sizes, since the size of GO

sheets can be decreased by sonication[4] without influencing the chemical reduction

process. To investigate the morphology of the graphene-related materials, wide angle

X-ray diffraction (WXRD) was used. The WXRD (Figure 3-1) gives the spacing

between atomic planes in the main peak: 2θ (Bragg angle) = 10.1 in GO, which
33
corresponds to an interplane distance of d0001= 0.94 nm.[45] In contrast to the WXRD

spectrum of GO, the WXRD spectra of bGO, pGO-5 and pGO-30 show much weaker

peaks at 2θ = 10.1, due to the exfoliation of the sheets, a result of long bath-sonication

times and powerful probe-sonication. For GS, no peak is observed at 2θ = 10.1 in

the WXRD spectrum. The weak broad peak at 23° indicates some stacking of GS.

Scheme 3-1. Sample preparation flowchart of graphene oxide and graphene sheets:
GO, bGO, pGO-5, pGO-30, and GS.

Figure 3-1. WXRD spectra of (from top) graphite (precursor of GO; gray) as
received, graphite oxide (GO; red) as synthesized; graphene oxide (bGO; blue)
following 5-hour bath-sonication from GO; graphene oxide (pGO-5; green)
following 5-min probe-sonication from GO; graphene oxide (pGO-30; purple)
following 30-min probe-sonication from GO; and graphene sheets (GS; black)
following hydrothermal processing of pGO-5 in D.I. H2O for 20 h at 130 °C and
pH=3.

34
Figure 3-2. AFM topography images of (a) GO, (b) bGO, (c) pGO-5, (d) pGO-30,
(e) GS, and (f) cross-section topography of the white line in (c). The
cross-sectional topography shows step features of ~1 nm thickness indicating that
pGO-5 is mostly single-layered.

35
Figure 3-3. AFM images of GO, bGO, pGO-30, and GS with cross-sectional
topography.

36
Atomic force microscopy (AFM) was used to measure the particle size of GO,

bGO, pGO-5, pGO-30, and GS. Figure 3-2a shows GO can reach 10 µm in lateral

size and tens of nanometers in thickness (see Figure 3-3). In contrast to GO,

micrographs of bGO show that the 5-hour bath-sonication significantly exfoliates GO

and decreases the particle size (Figure 3-2b), but multi-layer structures are still

observed in the AFM images and cross-sectional topography (see Figure 3-3). Figure

3-2c shows that pGO-5 sheets are smaller than either bGO or GO. The white dotted

line in Figure 3-2f is a line scan showing that most of the pGO-5 sheets are

approximately 1 nm in thickness, indicating single-layer structures.[16] The AFM

images of pGO-30, shown in Figure 3-2d, reveal that 30-min probe-sonication has

broken GO into even smaller pieces than pGO-5, as expected with the longer

sonication time. GS was produced via dehydration in an acidic environment[18]

from pGO-5 (Scheme 3-1), and Figure 3-2e (see Figure 3-3) shows that most of the GS

sheets are single-layer with similar size to pGO-5, demonstrating that the

hydrothermal process does not significantly affect the particle size. The images in

Figure 3-2 (also see Figure 3-3) reveal that GS is mostly single-layer after the

dehydration reaction; these results are in agreement with previously characterized

morphology. In addition, the hydrodynamic diameters of GO, bGO, pGO-5, pGO-30,

and GS in D.I. water and PBS as determined by dynamic light scattering (DLS) are

listed in Table 3-1. Although the DLS characterization can not reveal the exact size of

these GO and GS particles in aqueous solution because of the anisotropic morphology

of GO and GS, the results still show that the hydrodynamic size of GO decreases in

either D.I. water or PBS after intense probe-sonication. Compared to the

hydrodynamic diameter of pGO-5, the GS has much larger hydrodynamic diameter,

indicating the formation of irreversible aggregates in aqueous solution.


37
Figure 3-4. XPS C1s spectra and fitted curves of (a) GO and (b) GS. GS show a
significant drop in C-O (285.6 eV; red) character relative to C-C (283.5 eV; blue).
The XPS spectra show that there is significant reduction of oxygen from GO
(C/O=2/1) to GS (C/O=7/1). The purple curve is the fitted baseline.

X-ray photoelectron spectroscopy (XPS) was used to measure the chemical

composition of GO and GS. Figure 3-4a shows that the C1s spectrum of GO is an

overlap of three peaks at 287.3, 285.6 and 283.5 eV, fingerprints of C=O, C-O and C-C

bonds, respectively.[38] XPS measurements performed on GS show a significant drop


38
in C-O character relative to C-C (see Figure 3-4b). The atomic ratio of oxygen to

carbon is reduced from C/O = 2/1 to about C/O = 7/1, which explains the hydrophobic

nature of GS and the hydrophilicity of GO sheets. In addition, the oxygen amount in

GO and GS also affects their surface charge and dispersity in aqueous solutions.

As mentioned previously, the surface charge[66] and aggregation state[67] of

nanomaterials critically influence their in vitro cytotoxicity. The surface charge of

nanoparticles plays an especially important role in cell-nanoparticle interactions

because cell membranes themselves are charged. In fact, prior to cytotoxicity

experiments, characterization of the surface charge and particulate state of the GO and

GS samples in biological media is paramount. As shown in Table 3-1, the zeta

(ζ)-potentials of all GO and GS used in this work were negative. The GO samples have

similar ζ-potentials in both D.I. water, PBS, and cell culture medium indicating similar

oxygen content. Compared to GO, GS has a lower ζ-potential, which is also consistent

with the decreased oxygen amount in the XPS data and poor aqueous dispersity.

Table 3-1. GO and GS characteristics, values presented mean ± standard deviation


from triplicate measurements
samples hydrodynamic hydrodynamic ζ-potential ζ-potential ζ-potential
diameter in diameter in in D.I. in PBS in MEM
D.I. water PBS (nm) water (mV) (mV)
(nm) (mV)
GO 765±19 1678±190 - 40.6±2.9 - 33.1±1.6 - 22.7±0.5
bGO 748±13 1574±160 - 40.6±2.2 - 31.1±0.8 - 23.9±1.7
pGO-5 672±13 1254±143 - 41.2±1.3 - 31.1±1.9 - 23.9±3.3
pGO-30 342±17 861±115 - 42.4±2.7 - 34.6±1.7 - 22.6±0.7
GS 3018±36 4312±206 - 37.2±1.6 - 24.7±2.6 - 17.1±1.4

39
Although ζ-potential plays a key role in colloidal stability, it does not show the

true particulate state in various environments. To estimate the aggregation behavior of

GO and GS samples under different aging conditions, we simply used visual evidence

of settling. The concentration of GO and GS samples used for particle stability in D.I.

water and PBS was 50 μg/mL. All the GO samples show excellent colloidal dispersity

and stability in D.I. water (see Figure 3-5 a, b, c, and d) even after 24 hours at 37 oC.

We attribute this to their high negative surface charge (electrostatic stabilization).

However, even though the surface charge of GS was negative in D.I. water, GS formed

observable aggregates and came completely out of suspension after 24-hour static

aging at 37 oC (see Figure 3-5d), probably due to more π-π stacking interactions

between the deoxygenated surfaces.[68] To study the cytotoxicity of nanomaterials, in

most cases, particles will be suspended in highly salted solutions, like phosphate

buffered saline (PBS). Accordingly, the colloidal stability of GO and GS was also

evaluated in PBS because the results will more accurately model their actual

particulate state in toxicity assays and biological environments. In PBS, the GO

samples are still homogenously dispersed after 3-hour agitation at 37 oC (see Figure

3-5f). However, all the GO samples aggregated after 3-hour static aging and

completely settled out after 24-hour static aging at 37 oC (see Figure 3-5g and h)

because of charge neutralization of surface oxygen groups by ionic salt species. In the

case of GS, the particles aggregate and settle down to the vial bottom faster in the PBS

solution than in D.I. water (see Figure 3-5c and g). In short, GO is highly stable in D.I.

water but aggregates in highly salted environments. Importantly, the aggregation of

GO can be reversed using external agitation such as mixing and sonication. Compared

to the irreversible aggregation of GS, the particulate behavior of GO in highly salted

solution is called reversible aggregation in this study.


40
Figure 3-5. Photographs of GO, bGO, pGO-5, pGO-30, and GS sheets dispersed
at 50 μg mL-1 in D.I. water (a, b, c, d) and in PBS (e, f, g, h) after various time
points: samples right after preparation (0 hour) (a, e), after 3-hour mixing
agitation, (b,f), after 3-hour static aging (c, g), and after 24-hour static aging (d, h).
All samples were held at 37 oC.

3-3.2 In Vitro Hemolytic Activity of GO and GS

First, the hemolysis assay was employed to evaluate the in vitro blood

compatibility of GO and GS because these materials have recently been used for

biomedical applications, including injectable graphene-related particles.[10], [57]

Here, a universal method for testing in vitro nanoparticle hemolysis proposed by

41
McNeil et al.[69] was employed to investigate the hemolytic activity of GO and GS.

As shown in Figure 3-6a and b, the membrane of RBCs was compromised by GO and

GS in a dose-dependent manner, leading to observable free hemoglobin in the

supernatant. The concentrations of GO and GS leading to 50% lysis of RBCs (the TC50)

are listed in Table 3-2. It is apparent that the hemolytic activity of GO increases after

sonication, especially for intensely (probe) sonicated GO, like the pGO-5 and pGO-30

samples. The higher hemolytic activity of sonicated GO may be due to the greater

extent of exfoliation and smaller GO particle size. The size-dependent cytotoxicity in

human RBCs and mammalian cells has also been demonstrated using other types of

nanoparticles, such as silica[70], [71] and latex.[72] Because GS, compared to all the

GO samples, has the lowest oxygen content, it has the lowest hemolytic activity and is

also more likely to form aqueous aggregations, yielding fewer cell-contactable

reactive oxygen groups on the GS surface. This result is similar to work recently

published by Chen and co-workers where the authors found individually dispersed

carbon nanotubes were more toxic than aggregated carbon nanotubes.[67] To further

study the possible reasons why GO and GS cause hemolysis, hemolysis experiments

were performed using graphite (the precursor material for GO and GS). Compared to

the high hemolytic activity of GO particles, graphite induces a very low percent

hemolysis, even at 200 μg/mL (~3%, see Figure 3-7), most likely due to much lower

surface area and hydrophobic surface. This result indicates that the disruption of the

RBC membrane is likely attributed to the strong electrostatic interactions between

negatively charged oxygen groups on the GO/GS surface and positively charged

phosphatidylcholine lipids which are present on the RBC outer membrane.

42
Table 2. Concentrations of GO and GS Leading to a 50% Lysis of RBCs
GO bGO pGO-5 pGO-30 GS
TC50 (μg mL-1)a 142 49.6 30.5 20.2 >200
a
TC50 was determined using ED50plus v1.0 free software to fit the data in Figure 3-6a.

Figure 3-6. (a) Percent hemolysis of RBCs incubated with different


concentrations (3.125 to 200 μg mL-1) of GO (red), bGO (blue), pGO-5 (green),
pGO-30 (purple), and GS (black) for 3 hours at 37 oC with agitation. Data
represent mean ± SD from at least five independent experiments. Percent
hemolysis of RBCs incubated with pGO-30/chitosan at 100 μg mL-1 for 3 hours at
37 oC with agitation. (b) Photographs of RBCs after 3-hour exposure to GO, bGO,
pGO-5, pGO-30, and GS at different concentrations (3.125 to 200 μg mL-1). The
presence of red hemoglobin in the supernatant indicates RBCs with membrance
damage. (+) and (-) symbols represent positive control and negative control,
respectively.

43
Figure 3-7. Percent hemolysis of RBCs incubated with different concentrations
(3.125 to 200 μg mL-1) of graphite for 3 hours at 37 oC with agitation. Data
represent mean ± SD, n=3. Inset: photographs of RBCs after 3-hour exposure to
graphite with negative and positive controls. (+) and (-) symbols represent
positive control and negative control, respectively.

To confirm that the presence of GO and GS do not interfere the hemolysis

measurement method, all particle samples were incubated directly with cell-free

hemoglobin (absorbance ~0.88 at 540 nm) for 3 hours at 37 oC. The absorbance of

supernatant was compared to a control prepared by high speed centrifugation of pure

hemoglobin (see Figure 3-8). The result showed no significant amount of adsorbed

hemoglobin on the GO and GS. Additionally, to ensure the Uv-vis absorption by GO

and GS do not cause overestimation of hemoglobin concentration in the supernatant,

all particles suspended in PBS at 100 μg/mL were centrifuged, and the optical density

of the supernatant was measured at 540 nm, using 655 nm as reference. The optical

densities of all supernatants were very close to background signal, revealing that light

absorption by GO and GS is not a significant issue in the hemolysis assay.

44
Figure 3-8. None of the five GO and GS samples adsorb significant amounts of
hemoglobin. Hemoglobin was mixed with GO, bGO, pGO-5, pGO-30, and GS
(100 μg mL-1 in PBS) and incubated for 3 hours at 37 oC. Each data point is the
average of three measurements of optical density of the supernatant.

Morphological changes and significant lysis of RBCs after GO and GS exposure

can be observed by optical microscopy. Compared to the normal biconcave shape of

untreated RBCs in PBS (Figure 3-9a), pGO-5 and pGO-30 treated RBCs (at 25 μg

mL-1) appeared in much lower numbers and demonstrate both aberrant morphology

and recently lysed RBCs (arrows and circles in the insets of Figure 3-9b and c,

respectively). While the GS-treated RBCs did not result in significant lysis of RBCs,

hemagglutination was observed (arrows in Figure 3-9d) surrounding the GS

aggregates. This data is the first of its kind, systematically studying the hemolytic

activity of graphene oxide and graphene sheets toward human RBCs, and it is clear

that both particle size and surface charge/oxygen content influence apparent blood

compatibility.

45
Figure 3-9. Optical microscopy images of RBCs in the presence of (a) PBS
(control), (b) pGO-5, (c) pGO-30, and (d) GS at 25 μg mL-1 for 3 hours at 37 oC
with agitation. Inset images are magnified RBCs. Scale bars in the inset images
represent 10 μm. The arrows in (b) and (c) show lysed RBCs. The arrows in (d)
represent the hemagglutination caused by GS aggregates.

To further confirm that the hemolytic activity of these graphene-related particles

is attributed to an interaction between RBCs and GO/GS particle surface, the GO

particle with greatest hemolytic activity, pGO-30, was coated with a biocompatible

polymer, chitosan, using electrostatic adsorption.[73] The zeta potential of

pGO-30/chitosan in acidic water (pH~4.8) was 24.8±1.8 mV, compared to the zeta

potential of pGO-5 in acidic water (-38.3±2.3 mV), indicating that chitosan was

successfully coated on pGO-30 surface. However, upon dispersion in PBS (pH~7.4),

pGO-30/chitosan aggregates form rapidly due to a pH-dependent chitosan

conformational change.[73] Compared to the high hemolytic activity of pGO-30, no


46
apparent hemolysis was observed from pGO-30/chitosan (see Figure 3-6a and Figure

3-10), revealing that chitosan either serves as a protective layer, masking the

electrostatic interactions between RBCs and oxygen groups on the pGO-30 surface, or

aggregates the particles to decrease the cell-contactable surface area.

Figure 3-10. Percent hemolysis of RBCs incubated with 100 μg mL-1 of pGO-30
and pGO-30/chitosan for 3 hours at 37 oC with agitation. Data represent mean ±
SD from three independent experiments. Inset: photographs of RBCs after 3-hour
exposure to pGO-30 and pGO-30/chitosan with negative and positive controls. (+)
and (-) symbols represent positive control and negative control, respectively.

3-3.3 In Vitro Fibroblast Cytotoxicity of GO and GS

To further explore the cytotoxicity of GO and GS, two methods, the MTT and

WST-8 assay, were employed to investigate how these carbon-based nanomaterials

interact with adherent cells. In this case, we compare only the cytotoxicity of pGO-5

and GS on mitochondrial activity of human skin fibroblast cells, to focus on the role of

oxygen content. The MTT results for both pGO-5 and GS showed no dose-dependent

47
effects on the mitochondrial activity of human skin fibroblast cells (Figure 3-11),

indicating that the adherent fibroblast cells were unaffected by either nanomaterial at

any of the employed concentrations. A second mitochondrial activity-based assay,

WST-8, was employed to verify the MTT result. The WST-8 assay operates on the

same principle but uses a negatively charged, water soluble tetrazolium dye.

Compared to the MTT results, WST-8 data showed that both pGO-5 and GS have a

dose-dependent effect on the viability of human skin fibroblast cells. There has been

previous work done by Monteiro-Riviere et al.[63] and Wörle-Knirsch et al.[64]

demonstrating interference by carbon-based materials with the viability marker, MTT

reagent, yielding inflated viability results. To rationalize the contradictory results of

these two mitochondrial assays with GO and GS particles, control experiments were

performed to see if the GO and GS react directly with the MTT or WST-8 reagents,

regardless of the state of any cells in the dish. We found that both GO and GS react

with MTT to form purple formazan (see Figure 3-12), a result that would indicate

viable cells even though there were no cells in this control sample. Marques et al.[74]

reported a detailed reaction mechanism whereby the [MTT]+ cation can be reduced by

electrons and protons, which are present in GS or GO particles (see Figure 3-12).

Accordingly, we conclude that the MTT assay is not appropriate for cytotoxicity tests

of GO or GS particles. GO and GS particles cause a false positive measure of viability,

generating an overestimation of the viability of human skin fibroblasts, especially at

high GS and GO doses.

48
Figure 3-11. Cell viability of human skin fibroblast cells determined from MTT
and WST-8 assay after 24-hour exposure to different concentrations of pGO-5
and GS. Data represent mean ± SD.

Figure 3-12. Photographs of MTT formazan (in the DMSO supernatant)


produced by pGO-5 and GS particles. [MTT]+ will react with the conjugated

49
electrons from GS to form the radical intermediate [MTT]●. Then [MTT]● will
react with the protons to form a protonated cation [MTTH]+ under at least pH<13
environment. [MTTH]+ will then react with another electrons from GS to form
[FORMH], followed by a further protonated reaction with another proton mainly
from GS to form [FORMH2]+.

Figure 3-13. Cell viability of human skin fibroblast cells determined from trypan
blue exclusion assay after exposure to different concentrations of pGO-5 and GS.
Data represent mean ± SD, n=3.

Figure 3-14. Optical microscopy images of human skin fibroblasts after 24-hour
incubation at 37 oC (5% CO2) in the presence of (a) medium (control), (b) 50 μg
mL-1 of pGO-5, and (c) 50 μg mL-1 of GS particles.

50
Compared to the MTT control experiment, the WST-8 was not reactive either

with GO or GS particles. No detectable reduction of WST-8 (WST-8 formazan

formation) occurred after one hour incubation with cell-free GO or GS particles within

a wide concentration range (3.125-200 μg mL-1). In addition to the failure of MTT on

determining the cytotoxicity of GO and GS, we further found the GO and GS particles

adsorbed lactate dehydrogenase (LDH) and generated underestimation of the LDH

results. To validate the results of the WST-8 assay, another common viability assay

based on trypan blue exclusion was performed. The trypan blue dye exclusion assay

shows similar results to those from the WST-8 assay (see Figure 3-13), which confirms

the validity of WST-8 assay. In addition, it is clear to see from optical microscopy

images (see Figure 3-14) that the fibroblast cell density decreases after 24-hour GS

exposure at 50 μg mL-1, compared to control and pGO-5 treated cells. These data

further confirm the WST-8 results. Accordingly, adherent cell toxicity conclusions will

be based on the WST-8 assay results only, indicating that GS is more toxic to adherent

cells than pGO-5. This is likely due to the faster sedimentation and formation of more

compact aggregates of GS, as compared pGO-5, during 24-hour static aging (see

Figure 3-5 g and h), which greatly inhibits nutrient availability to and growth of human

skin fibroblasts. Unlike the hemolysis assay, the human skin fibroblast cells were

grown on the bottom of the assay wells, making factors such as the sedimentation rate,

thickness and compactness of GO or GS aggregates on the top of adherent cells more

likely to affect the viability of fibroblasts. Additionally, the generation of reactive

oxygen species (ROS) is a common toxic mechanism of carbon-based and other

nanoscale materials.[75] To investigate this possibility, we performed a reactive

oxygen species (ROS) assay to measure the oxidative stress generated by pGO-5 or

GS particles in or near the human skin fibroblast cells. The results show that the
51
generation of ROS in cells is concentration-dependent after pGO-5 or GS exposure

(see Figure 3-15). After 24-hour exposure, the aggregated GS at 25 μg mL-1 induced

approximately a 9-fold increase in ROS in fibroblast cells compared to untreated cells

(control). The aggregated GS induced an even higher level of ROS (indicating more

oxidative stress) in human skin fibroblast cells compared to reversibly aggregated

pGO-5 (see Figure 3-15), revealing a likely mechanism for larger GS toxic effects

compared to pGO-5. This work clearly demonstrates that the particulate state (extent

of aggregation, irreversible or reversible aggregation) of GO and GS and the cell types

(membrane composition and suspended/adherent nature of cells) used for cytotoxicity

evaluation have a great impact on the biological/toxicological responses.

Figure 3-15. Effect of pGO-5 and GS on the generation of ROS in human skin
fibroblast cells.

3-4 Conclusions

In this first study of the blood compatibility of graphene-based materials, the

blood compatibility and cytotoxicity of graphene oxide (GO) and graphene sheets (GS)
52
of various sizes and oxygen content have been investigated in suspended human RBCs

and adherent skin fibroblasts using in vitro hemolysis and WST-8 viability assays. All

the GO and GS show dose-dependent hemolytic activity on RBCs. In the case of GO

samples, extent of exfoliation and particle size play a critical role in extent of

hemolysis. Sonicated (smaller) GO exhibited higher hemolytic activity than untreated

(larger) GO. Compared to individually dispersed GO sheets having higher surface

oxygen content, the aggregated GS showed lower hemolytic activity. Covering the GO

sheets with chitosan eliminated their hemolytic activity.

The spontaneous formation of MTT formazan by cell-free GO and GS indicates

the failure of MTT assay in predicting the cytotoxicity of graphene-related materials.

The MTT data indicate a false high biocompatibility of GO and GS with adherent

fibroblasts. However, the valid WST-8, trypan blue exclusion, and ROS data

demonstrate that aggregated GS particles are more cytotoxic than reversibly

aggregated GO on human skin fibroblast cells. Moreover, compared to reversibly

aggregated GO, the aggregated GS generated more reactive oxygen species in human

skin fibroblast cells and strongly associated to the cell surface. Based on the hemolysis

and WST-8 viability assay results, this is the first work to show that the particulate

state of graphene-based particles has a profound impact on their toxicity to suspended

erythrocytes and adherent human skin fibroblasts.

53
Chapter 4.

Reducing Graphene Aggregation in Preparing

Nanocomposites

4-1 Introduction

As described in Chapter 2, the aqueous reduction process for synthesis of

graphene was motivated by producing graphene at a large scale with lower cost and by

a more environmentally friendly approach for graphene/polymer nanocomposites. In

this chapter, dispersion of the aqueous reduced graphene (ARG) in polymers is studied

for further applications in nanocomposites. Different from thermally reduced graphene

(TRG)[34] which is exfoliated in inert gas, ARG is exfoliated and reduced in water.

Removing water from ARG can cause aggregation and stacking of single sheets,

making dried ARG powder difficult to re-disperse into solvents or molten polymers.

Liquid phase reduced graphene are usually blended into polymers via in situ

polymerization to avoid aggregation of graphene caused by drying processes.[76], [77]

Not only liquid phase reduced graphene, but TRG also has the same aggregation

problem after vacuum drying if dispersed in solvents. Kim et al. reported that TRG can

be densified by dispersing TRG into acetone followed by filtration and drying by

vacuum drying. [78] A poorer dispersion of the densified TRG in polyethylene was

observed compared to the normal non-densified TRG.

In this chapter, we report a dispersion method to resolve the problem of

aggregation during the drying process of ARG. Since the majority of aggregation
54
occurs during the drying process, we must avoid complete drying. We replace >90% of

water from ARG/water suspension and leave 10% of water in order to avoid

aggregation. Then we add organic aqueous miscible organic solvent, for example

dimethylforamide (DMF), to make ARG/water/DMF solution. Because the sample

was not totally dried, there was no significant aggregation. Then we can dissolve

polymers into ARG/water/DMF solutions to make nanocomposites. Thermoplastic

polyurethane (TPU) was used as the matrix polymer in this chapter since TRG/TPU

nanocomposites were studied previously in our group. Kim et al.[79] reported that

solvent blending process gives the highest TRG dispersion level in TPU compared to

melt blended or in situ polymerized samples. Hence, in this chapter, ARG/TPU

nanocomposites prepared by two different solvent blending processes are compared

with the TRG/TPU nanocomposites prepared by the same procedure as reported by

Kim et al.'s. [79]

4-2 Experimental Section

4-2.1 Materials

The composites were prepared using Avalon 70AE thermoplastic polyurethane

(TPU) provided by Huntsman. ARG was prepared as described in Chapter 2, and was

kept in water after reduction. TRGs were provided by Vorbeck Materials (TRG-V)

and by Professor Abdahla, Petroleum Institute, UAE (TRG-P). TRG-V and TRG-P

were prepared from Hummers' graphite oxide[14] by rapid heating process.[34] TPU,

TRG-V and TRG-P were vacuum dried at 70 °C overnight before using. Carbon

oxygen ratios of TRG-V, TRG-P and ARG, were measured by XPS measurements

using Surface Science SSX-100 spectrometer housed at the Characterization Facility


55
of the University of Minnesota and are shown in Figure 4-1. The results shows that

both TRG-V and TRG-P have a similar C/O ratio (C/O = 9/1), whereas ARG has a

lower value (C/O = 7/1). The BET results are 512 m2/g for TRG-V, 500 m2/g for

TRG-P and 400 m2/g for ARG. The dried TRG-V and TRG-P was also tested by XRD

(Siemens; X-ray Diffraction & Scattering Wide Angle Bruker-AXS D5005) Using

AFM on similar samples McAllister et al reported that lateral size of TRG is ~750, and

80 % of TRG particles are single-layer. [81] Give dimensions of ARG ref Liao

56
(a)

(b)

(c)

Figure 4-4. C1s XPS spectra of (a) TRG-V, (b) TRG-P and (c) ARG. The results
shows that both TRG-V and TRG-P have a C/O ratio (C/O) of 9/1, whereas ARG
has a lower C/O of 7/1.

57
4-2.2 Fabrication of Graphene/TPU Nanocomposites

Nanocomposites were prepared from three types of graphene's: ARG, TRG-V,

and TRG-P. The nanocomposites were prepared by solvent blending. Graphene (1.3 –

40 mg) was added to DMF (5 mL) and stirred for 2 days using a magnetic stirrer.

Dried TPU pellets (~500 mg) were then added to the solution at room temperature and

the solution was stirred for an additional 5 days. The composites were then cast as a

film (with thickness 50~100 m) on a glass plate at 55 °C for 24 h. The film was then

dried in a vacuum oven at room temperature for 2 days. ARG/water suspension was

frozen and freeze dried at -70 °C under vacuum for 5 h to produce a low-density clump

of ARG. The rest steps of blending process are the same as TRG-V and TRG-P. The

process is illustrated in Figure 4-2a. Both TRG-V and TRG-P were received as dried

powder, so the preparation process of TRG/TPU systems were started after "Freeze

Drying" in Figure 4-2a. The TRG-V/TPU nanocomposites prepared for this work are

compared with the results in Kim et al.'s[79] report in Figure 4-3, which shows similar

mechanical properties improvement and percolation concentration (~0.5 wt%).

58
(a)

(b)

(c)

Figure 4-2. Illustration of (a) the single solvent blending process to synthesize
and (b) the co-solvent blending process to make ARG/TPU nanocomposites. (c)
The slime phase separation of ARG/TPU nanocomposites.

59
The single solvent blended ARG/TPU nanocomposites shows a poor dispersion

of ARG. (More details are in the Results and Discussion part.) Hence, ARG/TPU

nanocomposites were also fabricated using a co-solvent dispersion process as

illustrated in Figure 4-2b. The water left over from the aqueous reduction process (see

Chapter 2) was decanted until the water level reached black ARG precipitation, and

then DMF was added with a DMF/water weight ratio ~ 8/1. The ARG/DMF/water

suspensions were magnetically stirred for 24 h, and TPU pellets with certain weight

were then added into the suspension. The ARG/DMF/water/TPU suspensions

(DMF/water/TPU solution as the control group) were then magnetically stirred for 2

days at room temperature. Phase separation occurred after 2 days of stirring. All the

ARG went into the slime-like (black; high viscosity) phase, while the other phase only

contains a slightly yellow liquid with water-like viscosity (Figure 4-2c). Phase

separation also occurred in the control group (without ARG), but the "slime" phase

was white and the liquid phase appeared light yellow. The liquid phase was decanted

and the slime phase was picked up from the vials and was dried in a vacuum oven at 70

°C for 4 days. The dried products were heat compressed into 50~100 m thick films

for further characterizations.

60
Figure 4-3. Comparison of TRG-V/TPU nanocomposites prepared in this work
and the results in Kim et al.'s report. [79] The results show similar mechanical
properties improvement and percolation concentration (~0.5 wt%).

61
4-2.3 Characterization of Nanocomposites

ARG/TPU nanocomposites were heat pressed to films (50~100 μm) at 70 °C for

10 min, while TRG/TPU nanocomposites were made as film. Surface resistance was

measured using an 11-point DC surface resistance meter (PRS-801, Prostat) and was

measured at three points on top of the film and three points on the bottom (glass side)

of the film and the average is reported.

Modulus measurement and dynamic mechanical analysis were performed on a

Rheometrics Solids Analyzer (RSA-II). The samples were cut in regtangular-shape.

Modulus was measured at a temperature of 15 °C and strain rate of 10-3 s-1. Dynamic

mechanical analysis (DMA) was carried out from -100 °C to 130 °C with a 0.25 %

strain.

Differential scanning calorimetry (DSC) was used to investigate thermal

properties of nanocomposites. The samples were tested over a temperature range from

-100 °C to 200 °C with a heating rate of 10 °C/min. The film-shape samples were cut

into small pieces for the test.

62
4-3 Results and Discussion

ARG is hydrophobic (see Chapter 2&3), so all ARG merges into the "slime-like"

DMF/TPU phase but not water/DMF phase as shown in Figure 4-2c. This “slime-like”

black phase is a soft, deformable solid. The electrical resistance curves with various

graphene loads for graphene/TPU nanocomposites synthesized from the same solvent

blending process are shown in Figure 4-4a. Electrical resistance of the nanocomposites

provides a measure of the percolation concentration. We define electrical percolation

when the surface resistance of the nanocomposites drops ten folds. With the same

solvent blending process, ARG/TPU nanocomposite shows the highest percolation

concentration and the highest surface resistance at 3 wt% graphene. TRG-V/TPU and

TRG-P/TPU nanocomposites have lower percolation concentrations and surface

resistances than ARG/TPU nanocomposites. The results can be explained by the

process to produce ARG creates aggregation after drying the water left over from the

aqueous reduction process. TRG is exfoliated and reduced in the air, but ARG is

exfoliated and reduced in water. Normal solvent blending process disperses dried

ARG powder into TPU. However, after removing the water left over from the aqueous

reduction process, significant aggregation of ARG forms no matter whether normal or

freeze drying is used. Similar results were also observed by Kim et al.[78] They

dispersed dried TRG in water and removed water to make

densified-TRG/polyethylene nanocomposites, and the results show a poor dispersion

of densified-TRG.

In order to avoid aggregation of ARG, the co-solvent blending process (Figure

4-2b) which does not completely dry out ARG in the water before blending was used.

Figure 4-4a shows that the percolation concentration of dispersed ARG reaches 0.5
63
wt% if dispersed via the co-solvent process, compared to 1.0 wt% via the normal

solvent blending. The results indicate that the dispersion of ARG in TPU has been

significantly improved by the co-solvent blending process.

64
(a)

(b)

Figure 4-4. (a) Normalized tensile modulus of nanocomposites. (b) Surface


resistance of ARG/TPU nanocomposites.

The elastic modulus of the nanocomposites (Figure 4-4b) exhibit similar trends to

the surface resistance. TRG-V/TPU nanocomposite has the highest elastic modulus,

with the modulus increasing by 815 % at 3 wt%, and TRG-P/TPU nanocomposite

elastic modulus increased by 515 % at 3 wt%. As for ARG/TPU nanocomposite, the

co-solvent blended samples shows a modulus increase of 668%, whereas the normal

solvent blended sample increased by 354 % at 3 wt%. The results supports the
65
electrical resistance data saying that dispersion of ARG can be improved to the TRG

dispersion level via the new co-solvent blending process.

TRG-V/TPU nanocomposite system shows a higher modulus improvement and

lower electrical resistance than TRG-P/TPU nanocomposite system. The results can be

explained by the exfoliation level of the two TRGs measured by XRD as shown in

Figure 4-5. TRG-P shows a sharp peak at 2 = 25° but not TRG-V, indicating

restacking of TRG-P occurs during the rapid heating process.[18] The reason for the

difference is likely because that the precursor GO was not completely dried.

Figure 4-5. XRD spectra of ARG, TRG-V and TRG-P. TRG-P shows a sharp
peak at 2 = 25° but not TRG-V.

Both DMA and DSC were used to identify the glass transition temperature of the

composites, Figure 4-6, 4-7 and Table 4-1. No change in the glass transition

temperature was observed by either method. The results support the conclusion that

Tg is not increased via solvent blended graphene. The effect of graphene on Tg is

discussed in Chapter 6 & 7. DMA, in Figures 4-6, showed an increase in elastic

modulus with increasing graphene load. The elastic modulus increased less in the

glassy region. This can be explained by the lower filler/matrix polymer modulus

66
ratios.[93]

Figure 4-6. DMA for regular solvent blended ARG/TPU nanocomposites


composites.

Figure 4-7. DSC traces for co-solvent blended ARG/TPU nanocomposites


composites.

67
Table 4-1. Physical properties of graphene/TPU nanocomposites.

Tg by DMA Tg by DSC Elastic Resistance


(°C) (°C) Modulus (MPa) ()
Co-solvent blended ARG/TPU nanocomposites
Neat TPU -31 -35 10.06±1.6 6.6±2.3×1010
0.25 wt% -31 -35 -- 6.4±2.1×1010
0.5 wt% -32 -34 12.59±2.1 1.1±0.3×1010
1.0 wt% -31 -34 17.64±1.8 5.0±0.3×107
3.0 wt% -30 -35 66.50±8.2 3.0±0.2×104
TRG-P/TPU nanocomposites
Neat TPU -31 -36 9.96±2.3 6.6±2.6×1010
0.25 wt% -31 -36 9.00±1.6 5.8±0.6×1010
0.5 wt% -29 -38 9.35±2.0 2.3±0.2×1010
1.0 wt% -31 -35 18.64±1.9 1.1±1.4×108
3.0 wt% -34 -34 51.32±1.9 1.8±0.1×105
TRG-V/TPU
Neat TPU -31 -36 9.96±2.3 6.6±2.6×1010
0.25 wt% -- -- -- 6.7±1.0×1010
0.5 wt% -31 -39 14.44±2.3 9.0±2.0×109
1.0 wt% -33 -38 15.50±1.6 1.4±0.4×106
3.0 wt% -31 -32 81.15±9.3 1.4±0.8×102

68
4-4 Conclusions

It was shown that ARG can be dispersed in TPU as well as TRG using a

modified co-solvent blending process. ARG is exfoliated and reduced in aqueous

solution, so the product is suspended in water. Severe aggregation of ARG occurs

when removing the water to obtain dry ARG powder as demonstrated by XRD. The

co-solvent blending process prevents aggregation of ARG caused by the removal of

the water by avoiding completely drying the ARG. Dispersion of graphene fillers

improved significantly by changing the blending process with the same ARG filler.

The results showed that not only the intrinsic properties of graphene fillers, but also

blending process can affect graphene/polymer nanocomposites. Similar results were

observed by Kim et al.[78]: with the same TRG, densified TRG cannot be dispersed in

PE as well as a non-densified one. The electrical percolation concentration of

ARG/TPU nanocomposites synthesized using the co-solvent blending process is 0.5

wt%, which is as good as TRG/TPU nanocomposites prepared via the single solvent

blending process. Elastic modulus improvement of ARG/TPU nanocomposites can

reach 668% compare to 354% obtained from the normal solvent blended systems. Tg

of all nanocomposites stay constant for both TRG and ARG systems.

69
Chapter 5.
Graphene/Polyurethane Acrylate Nanocomposites3

5-1. Introduction

Graphene has recently attracted wide interest[1] because of its 2-D structure[30]

and outstanding electrical[28] and mechanical properties.[80] Several methods to

produce graphene in large amounts have been reported, such as epitaxial growth[7]

from carbon atom and thermally reduced graphene (TRG)[34], [81] or chemically

reduced graphene (CRG)[18], [32], [51], [65], [82], [83] with graphite oxide (GO) as

precursor. TRG is produced by rapidly heating dry GO under an argon gas purge at a

high temperature (1000 °C).[34] The process exfoliates and reduces GO to

single-layer graphene in ~ 60 s. Reported average dimensions of TRG are ~ 750 nm in

diameter.[81] Due to the high surface area and aspect ratio of graphene,[34], [81]

incorporation of graphene into polymers for graphene/polymer nanocomposites has

been reported to improve mechanical properties, decrease gas permeability and

increase electrical conductivity.[4], [78], [79], [84], [85] In this work, we use TRG as

graphene filler to make TRG/polymer nanocomposites.

Dispersion of nano-layered disk-like structured fillers, such as clay and graphene,

is the primary challenge of preparing nanocomposites.[85], [86] Homogeneous

dispersion in nanocomposites with GO filler prepared by in-situ solution

polymerization have been previously reported.[76], [77] Jang et al. suggested that GO

can be intercalated with a macroazo-initiator (oligomer initiator), and that free radical

polymerization of PMMA monomers in the presence of GO can exfoliate the

3
This chapter appears in: Liao K-H, Qian Y, Macosko CW Polymer 2012;53:3756.
70
platelets.[76] Potts et al. reported good dispersion of CRG in PMMA using in-situ

free-radical polymerization.[77] Motivated by these results, we sought to make

nanocomposites using in-situ polymerization to achieve homogeneous dispersion of

TRG in thermoset polyurethane acrylate (PUA).

PUA was selected as the polymer matrix in this study since

urethane-acrylate-based materials have a wide variety of applications in the coatings

industries[87], [88] and can be polymerized by light or thermal energy. PUA was

polymerized from our synthesized urethane acrylate oligomer (UAO) with addition of

diluent monomer, tripropyleneglycol diacrylate (TPGDA). The mixture of UAO and

TPGDA will be referred to as UA henceforth. We blended TRG into UA and prepared

nanocomposites of PUA by in-situ polymerization. Percolation of both

un-polymerized TRG/UA uncured liquids and polymerized TRG/PUA

nanocomposites were studied by rheology and surface resistance. The modulus of

TRG/PUA nanocomposites at low (glassy) and high (rubbery) temperature was

measured via dynamic mechanical analysis (DMA).

5-2 Experiments & Characterization

5-2.1 Materials

TRG[34], [81] (functionalized graphene; Vorbeck) was used as received. This

TRG is from the same batch used by Kim and coworkers.[78], [79], [84] Propylene

oxide tri-functional polyol (PPO; Mn = 700; FX31-240, Huntsman), was vacuum dried

at 100 °C for 5 h before use. Methylene diphenyl diisocyanate (MDI; Aldrich,

powder), tripropyleneglycol diacrylate (TPGDA; Aldrich),

hydroxyl-ethyl-methacrylate (HEMA; Aldrich, 97%), tetrahydrofuran (THF; Fisher

71
Scientific), azobisisobutyronitrile (AIBN; Aldrich, powder), and

N,N-dimethylcyclohexylamine (Polycat®8, AirProducts, CAS 98-94-2) were used as

received.
Scheme 5-1. Synthesis of urethane-acrylate oligomer (UAO): (i) Reaction of
HEMA and MDI (1:1.05 by mole) to form HEMA-isocyanate; (ii) reaction of
tri-functional polyol and HEMA-isocyanate (1:3.15 by mol) to name UAO; (iii)
UAO plus TPGDA diluent monomer were polymerized at 70 °C with AIBN
initiator to form polyurethane-acrylate, PUA.

72
73
5-2.2 Synthesis of UAO and TPGDA solution (UA)

In the synthetic reaction shown in Scheme 5-1, HEMA was slowly added (one

drop every 10 min) into vigorously stirred melted MDI at 70 °C to attain a molar ratio

of 1:1.05 (HEMA:MDI) and stirred an additional 48 h to form HEMA-isocyanate,[89]

as shown in Scheme 3-1 (i). Next, the HEMA-isocyanate was added into 70 °C PPO at

a molar ratio of 3.15:1 (HEMA-isocyanate:PPO), with 1 wt% of Polycat®8 to catalyze

the urethane reaction. The reactants became viscous after 2 h, at which time the

TPGDA at 70 °C was added to obtain a weight ratio of 4:6

(TPGDA:[HEMA-isocyanate + PPO]). The mixture was stirred for another 48 h to

form UA as shown in Scheme 3-1 (ii). TPGDA was used as a diluent monomer. No

reaction occurred with TPGDA in the UA synthesis process. All reactions were purged

with N2 gas.

5-2.3 Fabrication of TRG/PUA Nanocomposites

First, UA/THF solutions were prepared by dissolving UA in THF at an 1:10

(UA:THF) weight ratio. Next, the desired amounts of TRG were added to UA/THF

solutions, and mixtures were bath-sonicated (Model 2510, Branson) for 5 min and

magnetically stirred for 5 days. Next, solution vials were opened to air and

magnetically stirred in a fume hood for 2 days to remove ~ 99% of the THF. During

the last 2 hours in the fume hood, 4 wt% AIBN was added. Finally, solutions were

cast on a TEFLON plate and placed in a vacuum oven at 25 °C for 24 h to remove all

remaining THF. The black, viscous liquid residue was then covered by another

TEFLON plate and cured for three hours in an oven at 105 °C for 3 h. The solid films

(~200 m thick) were TRG/PUA nanocomposites. Neat PUA was prepared by the

same procedure without incorporation of TRG.


74
5-2.4 Rheology and Polymerization of TRG/UA Uncured Liquids

The TRG/UA uncured liquid samples for rheological testing were prepared as

described previously, but without the addition of AIBN in order to avoid slow curing

of UA during tests. Viscoelastic properties of TRG/UA uncured liquids were

investigated with a 25 mm parallel plate rheometer (ARES). Critical strain (γc),

defined as the strain when G' drops to 90 % of its limiting low strain value, was

determined by dynamic strain sweep at a frequency of 1 rad/s. Percolation

concentration of TRG was determined by dynamic frequency sweep at a strain of 0.1

%, which was within the linear viscoelastic range for all samples. Tests were

conducted at 25 °C.

Differential scanning calorimetry (DSC) at a temperature ramp of 10 °C/min was

used to monitor the polymerization of TRG/UA uncured liquid. Polymerization

temperature (Tp) and heat (Hp) were determined in the first scan, and glass transition

temperature was found in a subsequent scan.

5-2.5 Characterization of TRG/PUA Nanocomposites

Surface resistance of the TRG/PUA nanocomposites was measured using an

11-probe DC resistance meter (Prostat-801), which was previously applied[4], [78],

[79] to find the electrical percolation. The probe was applied to the surfaces of the

polymerized TRG/PUA nanocomposites or submerged in TRG/UA uncured liquid to

measure the electrical resistance. TRG/UA uncured liquids appeared wax-like when

the TRG load reached 1.0 wt% or higher, making it difficult to prepare TRG/PUA

nanocomposite thin films at 1.0 wt% or higher TRG loading. Electrical conductivity of

cured TRG/PUA nanocomposites was measured using Desert Cryogenics (Lakeshore,

75
Inc.) vacuum probe station in a nitrogen filled glove box with Keithley 236 and 6517

electrometers. Cured TRG/PUA nanocomposites of known dimensions were coated

with a layer of nano silver particle (Leitsilber 200 silver paint, Ted Pella, Inc.) on each

end for DC measurement. A linear current-voltage (I-V) curve for 2 wt% TRG/PUA

nanocomposites sample is shown in Figure 5-1. Transmission electron microscopy

(TEM) images were obtained on a FEI Tecnai T12 microscope using an accelerating

voltage of 120 kV. Samples for TEM were microtomed (Leica) with a glass knife at

room temperature into 100 nm sections before being picked up on carbon-coated Cu

grids. Dynamic mechanical analysis (DMA) was conducted on a Rheometrics Solid

Analyzer, RSA-2. Samples for DMA were cut into strips measuring 3 mm x 30 mm x

0.2 mm. Dynamic measurements were performed from -100 °C to 150 °C at a

frequency of 1 rad/s under a dynamic strain of 0.5 % with a pretension of 0.01 %

strain.

Figure 5-1. Current-Voltage (I-V) curve for cured 2 wt% TRG/PUA


nanocomposite.

5-3 Results and Discussion


Linear viscoelastic properties of TRG/UA uncured liquids with various TRG

loadings were studied at room temperature. The critical strain (Figure 5-2) decreased

76
from 5% to 0.5% at higher TRG loadings, in agreement with the previously reported

results based on melt rheology of TRG/linear-polymer nanocomposites.[78], [84]

Linear rheological behavior of TRG/UA uncured (un-polymerized) liquids was

measured by dynamic frequency sweeps as shown in Figure 5-3. As shown by Figure

5-3, samples with 0.5 wt% or higher TRG loadings reach a plateau in G' at lower

frequencies, indicating that the percolation concentration for un-polymerized

TRG/UA uncured liquids is ~0.5 wt%.

Figure 5-2. Rheology of un-cured TRG/UA uncured liquids. Dynamic strain


sweeps were run at room temperature.

Figure 5-3. Rheology of un-cured TRG/UA uncured liquids. Dynamic frequency


sweeps were run at room temperature.
77
TEM images of polymerized TRG/PUA nanocomposites with 0.25 wt% TRG are

shown in Figure 5-4 at low and high magnification. These TEM images indicate that

we achieved dispersion similar to Potts et al., who reported an elastic modulus

improvement of PMMA composites by 28% at 1 wt% CRG loading.[85] We also used

X-ray diffraction (XRD) patterns to check the dispersity of TRG. As shown in Figure

5-5, all the XRD results shows no aggregation curves. However, XRD alone cannot

prove good dispersion of TRG in PUA since the TRG loads are low. Percolation

concentration is a better indicator of dispersion. It can be measured directly by

electrical resistance.[4] Normalized electrical resistances R/Ro (resistance of TRG

loaded samples divided by resistance of pure PUA or UA) of TRG/PUA

nanocomposites and TRG/UA uncured liquids are given in Figure 5-6(a). Kim et al.

report that percolation occurs at R/Ro drops 10 fold.[79] The electrical conductivity of

composites, c, above percolation concentration follows a simple power law:

c = f (-perc)t (5-1)

where f is electrical conductivity of the filler, perc the percolation concentration

(vol%),  the volume fraction of the filler, and t is the universal critical exponent.[39]

Electrical conductivities of cured TRG/graphene nanocomposites are shown in Figure

5-6(b).The line calculated using Equation (5-1), gives the f as 0.23 S/cm, and the t as

5.17. The percolation concentration, perc, is 0.07 vol% (0.15 wt%). For the TRG/UA

uncured liquid percolation occurs at 0.23 vol% (0.5 wt%). The results show that

electrical percolation of TRG was improved via radical polymerization of UA, which

may due to the volume shrinkage during polymerization. The TRG/UA uncured liquid

results are in agreement with the G' results in Figure 5-3.

78
Figure 5-4. TEM images of polymerized PUA with 0.25 wt% TRG at lower and
high resolution.

Figure 5-5. XRD patterns for cured TRG/PUA nanocomposites.

The aspect ratio of dispersed TRG, which serves as an indicator for dispersion
quality, can be estimated from the percolation concentration using Equation (5-2):[78],
[84], [90]

(5-2)

where d and h are diameter and thickness of particles, respectively, ϕsphere = 0.29

(interpenetrating randomly packed spheres),[78], [84] and Af is aspect ratio of the

dispersed graphene in nanocomposites. The percolation concentration of TRG/UA


79
uncured liquid (0.5 wt%) yields an aspect ratio of ~ 210, which is lower than the

reported aspect ratio of 750 of free-standing TRG.[81] This lower aspect ratio may

indicate aggregation of TRG in uncured liquid UA. For polymerized TRG/PUA

nanocomposites, an Af of 750 was obtained from 0.15 wt% percolation concentration

via eq. (5-2). This result indicates TRG is nearly single-layer-dispersed in PUA. The

results are also supported by the TEM image in Figure 5-4b.

AIBN was selected as the thermal initiator for free-radical polymerization of the

acrylate functional groups in UAO and TPGDA. Benzyl-peroxide was also

considered for use as an initiator, but on addition to the TRG/UA uncured liquids and

removal of THF, the mixture polymerized at room temperature. The thermal

polymerization reaction with AIBN and thermal properties of the polymerized

TRG/PUA nanocomposites were monitored by DSC. The first DSC heating scan of

the un-polymerized TRG/UA uncured liquids gave the peak temperature (TP) and heat

of polymerization (HP), shown in Figure 5-7(a). The DSC traces from the first scan in

Figure 5-7(a) indicate that TP increases by approximately 10 °C with incorporation of

TRG, but no significant trend emerges upon varying TRG loadings. We speculated the

increase in TP is due to adsorption of AIBN by high surface area TRG sheets. The

extent of polymerization, HP, is independent of TRG concentration. The second DSC

heating scan of the cured TRG/PUA nanocomposites yielded the glass transition

temperature (Tg) in Figure 5-7(b). Similar to the Hp results, no significant change in Tg

was observed for PUA incorporated with TRG. Measured values of Tp, Hp, and Tg are

listed in Table 5-1. The Tg of TRG/PUA nanocomposites is independant from

incorporation of TRG, which will be discussed in Chapter 6 & 7.

80
(a) (b)

Figure 5-6. (a) Normalized electrical resistance (R/Ro) of un-cured TRG/UA


liquids (●) and polymerized TRG/PUA nanocomposites (■) with various TRG
concentrations, measured using the 11-probe electrical resistance measurement.
The percolation concentration before polymerization is 0.5 wt% and ~0.15 wt%
after polymerization. (b) Electrical conductivity of cured TRG/PUA
nanocomposites measured by DC measurement; log c plotted against
Log(-perc), where perc is the percolation concentration (0.07 vol%), c the
electrical conductivity, and  the volume fraction (vol%) of cured
nanocomposites. The solid line was calculated based on Equation (1), gives f =
0.23 S/cm, and t = 5.17.

81
Figure 5-7. DSC scans of TRG/UA uncured liquids: (a) first scan: monitoring
polymerization reaction for polymerization heat (Hp) and peak polymerization
temperature (Tp); (b) second scan: glass transition temperature (Tg; indicated by
arrows) of polymerized TRG/PUA nanocomposites. Both Hp and Tg show no
dependence on TRG concentrations. Tp increases with incorporation of TRG but
no change with various TRG concentrations (see Table 3-1).

82
Table 5-1. DSC polymerization results:

TRG Load (wt%) Tp (°C)a Hp (J/g)b Tg (°C)c


0 77 245 14±4
0.10 94 244 14±4
0.25 89 256 14±4
0.50 89 243 14±4
a. peak temperature
b. heat of polymerization
c. glass transition after polymerization

DMA was used to probe the effects of TRG on the mechanical properties of

TRG/PUA nanocomposites. Figure 5-8(a) shows that TRG does not reinforce the

nanocomposites in glassy regions, but it does significantly reinforce PUA in the

rubbery region. Within the glassy region, TRG does not effectively improve the

nanocomposite modulus since the elastic modulus of the polymer matrix is 3.5 GPa,

corresponding to a ratio Ef/Ep on the order ~ 70 (TRG modulus was assumed as

reported 250 GPa for chemically reduced graphene.)[91], [92] However, in the

rubbery region, the elastic modulus of PUA decreases to 25 MPa, which increases

Ef/Ep to 104.

The results can be compared to prediction of the Mori-Tanaka model:[93]

(5-3)

where ϕ is the concentration of filler, Ec, Ef, and Ep are the Young's modulus of the

nanocomposite, filler, and polymer, respectively, and vm is Poisson's ratio of the

polymer matrix. A and Ai are functions of Eshelby's tensor.[78], [89] Here we used

0.006 and 0.5 as the Poisson’s ratio of TRG and PUA, respectively.[91] The

reinforcement curves with Ef/Ep ~70 and ~104 estimated by the Mori-Tanaka model
83
were plotted as dashed lines in Figure 5-8(b). By fitting the Mori-Tanaka model to the

experimental data with Ef/Ep ~70 in the glassy region and Ef/Ep ~104 in the rubbery

region, we obtained an approximate aspect ratio of 750. Loss modulus E" curves of

TRG/UA uncured liquids are shown in Figure 5-9. Glass transitions based on the peak

in E" (shown in Figure 5-9) for all cured TRG/PUA nanocomposites were in

agreement with the second run DSC results in Figure 5-7(b).

(a)

(b)

Figure 5-8. (a) Thermomechanical properties of TRG/PUA nanocomposites by


DMA at several TRG concentrations. (b) Mori-Tanaka model fit assuming that
the aspect ratio of TRG is 750 and Ef=250 GPa. TRG/PUA nanocomposites show
almost no reinforcement in glassy region due to the low Ef/Ep ~70 but show large
reinforcement in rubbery region due to the high Ef/Ep ~104.

84
Figure 5-9. Loss modulus (E") of cured TRG/PUA nanocomposites by DMA.

5-4 Conclusions

We have successfully produced TRG/PUA nanocomposites by in-situ radical

polymerization with an ultralow percolation concentration of 0.15 wt% (0.07 vol%),

which is the lowest electrical percolation concentration among all the TRG/polymer

nanocomposites reported previously (Figure 5-10).[78], [79], [84], [94–98] The

calculations of Af , the aspect ratio, of dispersed TRG nanoplatelets in polymerized

composites from the modulus data using the Mori-Tanaka model agree with Af

calculated from resistance percolation. The percolation concentration of cured

nanocomposites is lower than the 0.5 wt% obtained for TRG/UA uncured liquids by

both resistance and rheology. We speculate that volume shrinkage during free radical

polymerization of UA is a possible reason to improve the dispersion of TRG in PUA,

and further work in this area is needed. TEM images taken of the cross sections of

cured TRG/PUA nanocomposites showed a homogeneous distribution of TRG. The

polymerization temperature (90 °C), polymerization heat (240 J/g) and glass transition

temperature (14 °C) of the nanocomposites were found to be independent of TRG

loading, but incorporation of TRG caused polymerization temperature of UA to


85
increase by 10 °C. Modulus results suggest an effective mechanical reinforcement of

TRG/PUA nanocomposites in the rubbery state. Although a good dispersion of

graphene in polymer was achieved, no significant reinforcement was observed in the

glassy state. This result can be explained by Mori-Tanaka model since the modulus

ratio of filler and matrix polymer is low.

Figure 5-10. Summary of reported electrical percolation concentrations of TRG


polymers nanocomposites with different dispersion process.[78], [79], [84],
[94–98] TRG/PUA nanocomposites shows the lowest percolation concentration.

86
Chapter 6.

Stereo-regularity and Processing Effect on Glass Transition


Temperature of Graphene/PMMA Nanocomposites

6-1. Introduction

Graphene has recently attracted much interest[1] because of its 2-D structure,[30]

outstanding electrical[28] and mechanical properties.[80] Due to its high surface area

and aspect ratio,[99] graphene is capable of improving the mechanical properties, gas

impermeability and electrical conductivity of polymers.[4], [100–102] Several studies

reported that incorporation of graphene into polymers showed an increase over the

matrix polymer glass transition temperature (Tg).[77], [103–106] In contrast, many

other reports did not observe significant Tg change.[107–112] We review these studies

in Chapter 7. Unfortunately, there is still no conclusion for the Tg of graphene/polymer

nanocomposites. These various Tg results leave an open question to researchers: how

does graphene affect the Tg of matrix polymers?

Several studies in the literature have sought to predict the Tg of polymer

nanocomposites by applying concepts from polymer nanoconfinement.[113–116]

Ellison et al. [114] reported that Tg was distributed near the interfaces between fillers

and matrix polymers. Rittigstein and coworkers successfully used a silica/poly-methyl

methacrylate (PMMA)-thin-film/silica sandwich "model structure" to predict the Tg

change of silica/PMMA nanocomposites.[116] Grohens et al. reported that (1)

polymer stereo-regularity, (2) polymer-substrate interfacial interactions, and (3)

polymer conformations determine the Tg change of nano-confined polymer thin films

on substrates.[117] Grohens et al.[118] reported that thin film confinement effects in

87
PMMA depends on stereo-regularity. The Tg of syndiotactic-rich PMMA decreased by

confinement on Al2O3 and SiO2 substrates, in contrast to a significant Tg increase for

isotactic PMMA on the same substrates.[118] Keddie et al.[119] reported that the Tg of

atactic PMMA thin films increased on an attractive surface but decreased on a

low-adhesive surface, indicating that a strong interaction between polymer and

substrate can cause Tg to increase locally. Rittigstein et al.[116] suggested that the

depression of Tg is due to the voids created at the poorly adhesive polymer/filler

interfaces, and the polymer can be considered as free standing thin film. Motivated by

this previous work, we investigated two factors: (1) matrix polymer stereo-regularity,

and (2) polymer-graphene interaction effects on Tg of graphene/polymer

nanocomposites. As far as we know, no literature exists that discusses the effect of

matrix polymer stereo-regularity effect on graphene/polymer nanocomposites.

To study the effect of matrix polymer stereo-regularity, we used isotactic-PMMA

(i-PMMA) and syndiotactic-rich atactic-PMMA (a-PMMA) as matrix polymers to

prepare nanocomposites by two methods: solvent blending and in situ polymerization.

Thermally reduced graphene[34] (TRG; also referred to in the literature as

functionalized graphene sheets; FGS) has been widely used to reinforce polymers.[4],

[78], [79], [84], [98], [99], [120–124] Recent studies have reported that TRG can

enhance the tensile modulus of glassy polymers by 32~92% with a 1.9~3.1 vol%

loading using solvent-blending.[98], [99] Therefore, we used TRG as one of the

graphenes for this study.

Kan et al.[125] have shown that free radical polymerization of vinyl monomers in

the presence of graphene oxide (GO) can cause covalent bonds to form between

carbon atoms in the polymers and graphene. The phenomenon was explained by the

free radicals on the propagating chain attacking carbon-carbon double bonds (C=C)
88
from the graphene sheets. These C=C are more reactive than those in pristine graphite

due to the loss of conjugacy caused by oxidation and reduction reactions from

graphene oxide to graphene. Hence, we prepared TRG/PMMA nanocomposites via

solvent blending and in situ polymerization in order to study the process effect on Tg.

We expected that Tg of in situ polymerized TRG/PMMA nanocomposites will be

increased by TRG incorporation due to the strong covalent bonding between TRG and

PMMA. Also, to further study the effect of interfacial interactions between the matrix

polymer and the filler, we dispersed nearly pristine graphene (PG) with much lower

oxygen levels than the TRG in i-PMMA for comparison to the TRG system. PG with

significantly lower oxygen level should provide weaker interactions at the interfaces

with PMMA due to the lack of oxygen functional groups such as hydroxyl or epoxy

groups.

6-2. Experimental Section

6-2.1 Materials
The syndiotactic-rich atactic PMMA (a-PMMA) used here was purchased from

Polysciences (atactic beads, Mw = 350,000, PDI = 2.7, lot. 528266). Additives in the

as-received a-PMMA were removed by solvent extraction as described in Appendix

A. Methyl-methacrylate monomer (MMA; Sigma-Aldrich, 99%), isotactic PMMA

(i-PMMA, Sigma Aldrich; Mw = 64,000 PDI= 2.9), N,N-dimethylformamide (DMF;

Aldrich), tetrahydrofuran (THF; J.T.Baker), methanol (Macron) and

2,2'-azobis(2-methylpropionitrile (AIBN; Sigma-Aldrich) were used as received.

Stereo-regularity of syndiotactic-rich a-PMMA and i-PMMA were confirmed

using H1NMR (Figure 6-1), which shows that a-PMMA has mm:mr:rr = 6:38:56 and

89
i-PMMA has mm:mr:rr = 96:3:1. Stereo-regularity of the in situ polymerized

(H1NMR not shown) control PMMA and TRG/PMMA nanocomposites are

approximately mm:mr:rr = 7:37:58, which is similar to a-PMMA.

Thermally reduced graphene sheets (TRG; Vorbeck Materials) and pristine

graphene (PG; N002-PDR, Angstron Materials) were dried at 70 °C under vacuum

for 24 h before use. Figure 6-2 shows the XPS results of TRG and PG. Carbon/oxygen

atomic ratios (C/O) are 9/1 for TRG and 49/1 for PG. Zhamu and Jang[126] exfoliated

graphite to PG via direct ultra-sonication in organic solvents without oxidation or

acid-involved processes, which leads to no oxygen being modified onto PG

(Angstron Materials). Standard XPS measurements were conducted on TRG and PG

powders using Surface Science SSX-100 spectrometer housed at Characterization

Facility of the University of Minnesota. The powders were dried at 70 °C overnight

before the test.

Figure 6-1. 1H NMR spectra of syndiotactic-rich atactic PMMA (a-PMMA;


mm:mr:rr=6:38:56) and isotactic PMMA (i-PMMA; mm:mr:rr=96:3:1).

90
(a)

(b)

Figure 6-2. XPS (a) TRG, and (b) PG. Carbon/oxygen atomic ratios (C/O) are 9/1
for TRG and 49/1 for PG, fitting by Shirley-Gaussian distribution.

6-2.2 Preparation of Nanocomposites via Solvent Blending:

The solvent blended TRG/a-PMMA and TRG/i-PMMA nanocomposites were

prepared using the procedure reported by Ramanathan et al.[123] Briefly, TRG was

dispersed in 10 mL THF using a bath sonicator (Branson 2510, 100W) at room

temperature for 5 h to form TRG/THF solution. The a-PMMA and i-PMMA were

dried under vacuum at 80 °C overnight, dissolved in THF at a concentration of 10 wt%

91
using a magnetic stirrer, and then the solution was stirred until a clear PMMA/THF

solution formed (>2 h). The TRG/THF solutions were mixed with the PMMA/THF

solution in varying ratios to yield approximately 1 g of total mass of TRG/PMMA at

compositions of 0.05, 0.1, 0.25, 0.5 and 1.0 wt% in THF. All the TRG/PMMA/THF

solutions were magnetically stirred for one more hour in an ice water bath. Each

solution was precipitated in 300 mL of methanol and then the precipitate was

collected by suction filtration through a polytetrafluoroethylene (PTFE) membrane.

The filter cakes were then dried in a vacuum oven at 80 °C for 10 h. The samples for

thermomechanical and surface resistance tests were prepared from the dried filter

cakes by pressing at 210 °C at 2 MPa for 5 min. PG/i-PMMA nanocomposites were

produced via the same process as the TRG system. As a control, the as-received

PMMAs were subjected to the same THF solution processing. PMMA/THF solutions

(10 g of PMMA in 100 mL THF) were directly poured into 300 mL of methanol to

precipitate the polymer.

6-2.3 Preparation of Nanocomposites via in situ Polymerization:

MMA/THF solutions was prepared with a weight ratio of MMA/THF = 2/3. A

certain amount of TRG was then added into the 10 mL MMA/THF solutions and was

probe-sonicated (Cole Parmer; CPX750, 750W) for 5 min (1 second of sonication and

1 second of pause; totally 10 minute) followed by magnetically stirring for 24 hours at

room temperature. The TRG/MMA/THF suspensions were then purged by N2 for 20

min prior to polymerization. AIBN was added into the purged suspensions with a

weight ratio of MMA/AIBN = 1000/3. Then polymerization was carried out at 65 °C

for 24 hours under N2 purging. The product was dried after polymerization, and the
92
remaining monomer (< 1 wt% of total nanocomposite) was removed in hood at 80 °C

for 24 hours and then heat pressed into films. The films were further dried in a vacuum

oven at 120 °C for another 24 h for further tests. The control group without

incorporation of TRG was prepared via the same procedure.

The matrix PMMA of in situ polymerized TRG/PMMA samples was extracted.

First the TRG/PMMA nanocomposites were dissolved in THF with a concentration of

10 wt%, and the black TRG/PMMA/THF solutions were filtered through PTFE filter

(0.2 m; Acrodisc CR13) to remove the dispersed TRG. The filtered liquids appear

clear and were dried in fume hood at 80 °C for 4 hours and in a vacuum oven at 120 °C

for 24 h. More than 98 wt% by weight of matrix PMMA was collected. A clear

polymer film was obtained for tests. The filter cake was also collected for further

studies. The TRG loaded filter cake was washed with THF three times and was

re-dispersed in THF for further tests. The filtered TRG tends to stick on the filter, so

we cut the filter into small pieces followed by dipping them in vigorous stirred THF.

Atomic force microscopy (AFM) was used to investigate the morphology of the

extracted TRG. A drop of the extracted TRG/THF solution was spin coated on a mica

substrate and air-dried at room temperature for 2 h before test. A Digital Instrument,

Nanoscope III Multimode AFM at the Institute of Technology Characterization

Facility of the University of Minnesota was used in noncontact mode for

measurements. (refer to Ch 2)

6-2.4 Molecular and Nanocomposite Characterization:

Mw was determined on a Hewlett-Packard 1100 series liquid chromatograph

fitted with a Hewlett-Packard 1047A refractive index detector and three PL gel

93
columns (Jordi Gel columns of 500, 103, and 104 Å pore sizes) with chloroform as

the eluent at a flow rate of 1 mL/min at 35 °C. All the columns were calibrated

with polystyrene standards (Polymer Laboratories). 1H NMR was recorded on a

Varian INOVA-500 spectrometer at room temperature with CDCl3 as the solvent.

DSC measurements were performed using a TA Instruments Q1000 at a rate of 10

°C/min. Surface resistance of nanocomposite films was measured using an 11-probe

DC resistance meter (Prostat-801). DMA and tensile moduli were tested using the

RSA-G2, TA Instruments.

6-3. Results and Discussion

Conductivity and modulus were used to verify that all the TRG/PMMA

nanocomposite systems are at the same dispersion level. Surface electrical resistance

are shown in Figure 6-3. The resistances of all three systems drops by 2 orders of

magnitude at 1.0 wt% of TRG loading, indicating a similar dispersion level of TRG

was achieved. An aspect ratio of ~200 for the dispersed TRG can be calculated using

Equation 5-1 (in Chapter 5).

(5-1)
We can use the Mori-Tanaka model[93] to estimate the aspect ratio of the dispersed

TRG, another indicator of the dispersion level besides electrical resistance. Figure 6-4

shows the normalized storage modulus (E') at 25 °C measured by DMA of all samples

at various TRG loadings. E' is one order of magnitude larger than loss modulus (E"),

so E' approximately equals to tensile modulus (E*). A summary of E' in the glassy

state and Mori-Tanaka calculated results (solid lines) were plotted in Figure 6-4, all of
94
which give aspect ratios of approximately 200, similar to the values estimated from

conductivity percolation. These results indicate that all TRG/i-PMMA,

TRG/a-PMMA and in situ TRG/PMMA nanocomposites are at similar dispersion

levels. Dispersion of TRG and PG in i-PMMA were not compared here since the

dimensions and polarity of the two fillers are different.

Figure 6-3. Electrical resistance of TRG/i-PMMA, TRG/a-PMMA and in situ


TRG/PMMA nanocomposites. The resistances of all systems drop significantly at
1.0 wt% of TRG loading, indicating a similar dispersion level (Af ~ 200) of TRG
was achieved.

95
Figure 6-4. Normalized storage modulus of TRG/i-PMMA, TRG/a-PMMA and
in situ TRG/PMMA nanocomposites measured by DMA at several TRG
concentrations. Assuming filler modulus Ef = 250 GPa, Mori-Tanaka model
fitting (lines) gives an aspect ratios of ~200 for all. The results show a similar
dispersion level of TRG in i-PMMA and a-PMMA.

96
(a) (b)

(c) (d)

Figure 6-5. DMA results (tan) of (a) TRG/a-PMMA, (b) TRG/i-PMMA, (c)
PG/i-PMMA and (d) in situ TRG/PMMA nanocomposites.

DMA results (tan) of the three PMMA nanocomposites prepared via solvent

blending: (1) TRG/a-PMMA, (2) TRG/i-PMMA, and (3) PG/i-PMMA are shown in

Figure 6-5a,b and c. Tg of TRG/a-PMMA nanocomposites changed little with

incorporation of TRG, whereas TRG/i-PMMA increased by ~ 5 °C, PG/i-PMMA

increased by ~3 °C, and in situ TRG/PMMA increased by ~ 6 °C. The detailed values

are listed in Table 6-1. Loss modulus (E") and storage modulus (E') of all the samples

are shown in Figure 6-6 and Figure 6-7. DSC curves of all the nanocomposites are

shown in Figure 6-8, which are consistent to our DMA conclusions. Summary of the

Tg results measured by DMA and DSC are as shown in Figure 6-9.

Grohens et al.[117] reported that stereo-regularity of PMMA has a significant

97
effect on thin film nanoconfinement. In Grohens' report, the Tg of isotactic-rich

PMMA thin films can be increased significantly when supported by an aluminum or

silica substrate with hydrogen bonding at the interfaces, whereas a Tg change for

syndiotactic-rich PMMA was not observed. Grohen and coworkers speculated that the

reason for this is that i-PMMA has a higher interaction density (i.e., higher number of

active interaction sites per chain) at the interfaces between PMMA and substrates

compared to a-PMMA.[118] These nanoconfinement results predict that the Tg of

i-PMMA can be increased by incorporation of TRG but not a-PMMA.

Besides TRG/a-PMMA and TRG/i-PMMA nanocomposites, the Tg of

PG/i-PMMA nanocomposites were also studied in order to verify that interfacial

interactions are important for Tg increase. PG has a much higher C/O ratio (49/1) than

TRG (9/1) as shown in Figure 6-2, indicating less oxygen functional groups for

i-PMMA molecules to interact with. Here we propose a possible reason to explain the

stereo-regularity effect on Tg. It has been reported that PMMA can form strong

hydrogen bonds with TRG.[123] It is well known that hydrogen-bonds are highly

temperature-dependent. Noro and Lodge reported[127] that Poly(2-vinylpyridine) has

only ~40 % of number of active hydrogen bonds remained at 125 °C compared to that

of 65 °C.[127] In our case, since that i-PMMA has a Tg at 64 °C, which is a

temperature that number of hydrogen bonds is higher than a-PMMA's Tg of 125 °C.

Hence the different PMMA stereo-regularity not only increases the interaction density

as Grohen et al. described[117], but also increases the hydrogen bonding interactions

between graphene and PMMA.

DMA results of in situ TRG/PMMA nanocomposites (Figure 6-5d) shows a

larger Tg increase than the solvent blended TRG/i-PMMA system. The results can be

explained by interfacial covalent bonding for in situ polymerized system than the
98
solvent blended i-PMMA nanocomposites' hydrogen bonding. Similar results were

also observed by other in situ polymerized graphene/PMMA or GO/PMMA

nanocomposites.[77],[128],[129],[76] Kan et al. showed that free radical polymerization

of vinyl monomers in the presence of GO causes polymer chains covalently bonded on

GO surfaces.[125] They speculate that the loss of conjugate structure of graphene by

oxidation can actually improve the chemical reactivity of the C=C on the sheets, hence

the free radical can attack GO or reduced GO more easily than pristine

(dislocation-free) graphene with perfectly conjugated structure. As shown

schematically in Figure 6-10, besides the reaction of free radicals on propagating

chains and C=C on GO sheets (#3 in Figure 6-10) that Kan et al. proposed,[125]

radicals might also be generated on the surface of GO or reduced grapheme by

hydrogen abstraction (e.g., allyl position, olefin hydrogen). This would then be

followed by (1) coupling with radicals of propagating polymer chains, or (2) initiation

of monomer polymerization.

All the three reactions proposed above can provide strong covalent bonds to the

surface. Besides strong covalent bonds, we speculate that a polymer chain initiated

from the free radical on GO or reduced graphene wets the interface better than by the

solvent blending process. Rittigstein and Torkelson[130] reported that a well wetted

matrix polymer/filler interface can lead to a stronger interaction to increase Tg,

whereas a poor or non-wetted interface creates voids at the interface and leads to a Tg

depression. They showed that Tg of clay/polymer nanocomposites increase or decrease

can be controlled by choice of solvent for the nanocomposite solvent blending

process.[130]

In Figure 6-11 We compared Tg by DSC of the PMMA extracted from the 1.0

wt% in situ polymerized nanocomposites to the in situ TG/PMMA. The results show
99
that after removing the dispersed TRG, Tg decreased by ~ 7 °C, i.e. back to the Tg value

of the control group. As described in 6-2.3, the extracted matrix polymer occupies > 98

wt% of the original nanocomposites, indicating only small amount of PMMA is

covalently bonded to TRG. The dispersed TRG of 1.0 wt% in situ TRG/PMMA

nanocomposites was also extracted, and two Tg’s were observed using DSC. Figure

6-12a shows that the higher Tg (142.8 °C) is ~32 °C more than the lower Tg (110.6 °C),

which is same within standard deviation of the Tg of 1.0 wt% in situ TRG/PMMA

nanocomposites (110.9 °C). The results can be explained by that the constrained

chains mobility of the PMMA covalently bonded to TRG, as shown in Figure 6-10.

The lower Tg is the free in situ matrix PMMA left over after washing. From the AFM

images in Figure 6-12b it appears that the surface of the extracted TRG is covered by

small spheres, which are the covalently bonded PMMA chains. Similar AFM or TEM

results were also reported by Kan et al. [125] and Fang et al.[131] Kan et al. [125]

chemically grafted poly-glycidylmethacrylat (PGMA) on GO surface and found

spheres on GO which increased in size with longer polymerization time. Fang et

al.[131] grafted polystyrene (PS) on GO and graphene using ATRP and observed

similar spheres on GO sheets.

100
(a) (b)

(c) (d)

Figure 6-6. DMA results (loss modulus, E") of (a) TRG/a-PMMA, (b)
TRG/i-PMMA, (c) PG/i-PMMA and (d) in situ TRG/PMMA nanocomposites.

101
(a) (b)

(c) (d)

Figure 6-7. DMA results (storage modulus, E') of (a) TRG/a-PMMA, (b)
TRG/i-PMMA, (c) PG/i-PMMA and (d) in situ TRG/PMMA nanocomposites.

102
Figure 6-8. DSC curves of TRG/a-PMMA, TRG/i-PMMA, PG/i-PMMA and in
situ TRG/PMMA nanocomposites.

103
(a)

(b)

Figure 6-9. Summary of Tg change (Tg) measured by (a) DMA tan and (b) DSC.
Significant Tg changes were observed in TRG/i-PMMA, PG/i-PMMA and in situ
TRG/PMMA nanocomposites, while Tg of a-PMMA shows no significant change with
incorporation of TRG. Tg increase.

104
(1) (2)

(3)

Figure 6-10. Three possible reactions during in situ polymerization with


presence of reduced graphene or GO. Hydrogen atom abstraction from GO or
reduced graphene followed by (1) coupling with radicals of propagating polymer
chains, or (2) initiation of polymerization. (3) A free radical of a propagating
polymer chain attacks the C=C on GO or reduced graphene sheets.

105
Figure 6-11. Tg measured by DSC of in situ TRG/PMMA nanocomposites and
their extracted matrix PMMA.

Table 6-1. Summary of properties of all PMMA nanocomposites.


Filler Tg DSC Tg DMA Tg DMA Modulus Resistance Mn PDI
Load (°C) by E" by tan (GPa) (G) (kg/mol
(wt%) (°C) (°C) )
TRG/ 0 124.6±1.1 119.6±0.8 130.1±1.4 1.93±0.06 460±13 186 2.66
a-PMMA 0.25 124.5±0.1 117.6±0.9 129.4±1.9 2.26±0.05 530±21
0.5 125.3±1.1 120.6±0.9 129.9±1.7 2.26±0.08 470±15
1.0 124.5±0.4 117.6±0.3 128.8±1.5 2.62±0.07 0.6±0.01
TRG/ 0 55.7±1.1 45.4±0.6 62.2±0.8 2.98±0.08 490±14 6.4 2.9
i-PMMA 0.25 57.7±1.6 55.1±0.4 65.8±1.6 3.28±0.10 380±11
0.5 57.8±1.7 57.0±0.9 67.5±1.9 3.43±0.09 400±12
1.0 58.4±1.9 56.3±0.7 67.6±1.0 4.16±0.05 0.3±0.01
PG/ 0 55.7±1.1 45.4±0.6 62.2±0.8 2.98±0.08 490±14
i-PMMA 0.25 57.5±0.2 52.7±0.6 65.0±0.7 3.84±0.06 380±31
0.5 57.1±0.1 55.7±1.2 65.0±0.2 3.82±0.03 430±12
1.0 57.5±0.2 54.2±0.4 64.6±0.9 3.82±0.06 0.06±0.001
in situ 0 110.9±2.4 95.3±1.3 111.7±0.2 2.80±0.3 510±23 50 1.97
TRG/ 0.25 117.2±2.6 100.6±0.5 116.3±0.5 2.93±0.7 430±21 52 1.86
PMMA 0.5 118.3±0.5 101.8±0.8 117.7±0.3 3.54±0.04 480±13 47 1.91
1.0 120.1±0.4 103.8±1.4 117.9±0.5 3.78±0.03 2.1±1.1 53 1.81

106
(a)

(b)

Figure 6-12. (a) DSC curves of extracted TRG, matrix in situ PMMA and in situ
TRG/PMMA nanocomposites (1.0 wt%). (b) AFM height images of the extracted
TRG. (The inset is the DSC curve of Extracted TRG.)

All the measured properties of TRG/a-PMMA, TRG/i-PMMA, PG/i-PMMA and

in situ TRG/PMMA nanocomposites are summarized in Table 6-1. It is necessary to

address how the Tg changes with various loadings of graphene. In Figure 6-10, we
107
have showssubstantial Tg increases relative to neat matrix polymers were found at very

low filler loadings (0.25 wt%), but only small increases were found in Tg beyond this

level. The phenomenon can be explained by Grohen's report,[117] which stated that

the nanoconfinement effect on Tg can extend as far as 8 times of Rg from the interface

between substrates and polymers. Assuming 1 g of single-layer graphene dispersed in

PMMA with an obtained aspect ratio of 200 in round disk shape with a thickness of ~1

nm, then the surface area for the dispersed graphene is ~64,000 nm2/sheet. With 8 Rg

(~50 nm; assuming Rg of a matrix polymer is 6.3 nm)[78] of Tg affected distance, the

average immobilized volume is ~3×106 nm3 per graphene sheet. Assuming the density

of graphene sheet is 2.28 g/cm3,[78] the average mass of one graphene sheet is

~2×10-17 g. For a 0.05 wt% graphene /PMMA nanocomposites in Potts' report,[77] 1 g

of nanocomposite contains 0.5 mg of graphene, which is ~1015 graphene sheets. The

number of graphene sheets leads to a 1 cm3 immobilized volume, which is 3 times of

the volume of 1 g nanocomposites, assuming a density of 1 g/cm3. The rough

calculation above indicates that even at a graphene loading as low as 0.015 wt% with

Af ~200, the Tg affected volume can be as large as 1 cm3. Hence, only a small Tg

increase can be observed beyond this concentration, which is confirmed with Potts'

report.[77] This trend is also observed in our TRG/i-PMMA nanocomposites results,

which is discussed previously, hence we speculate that it can be a general case for a

variety of graphene/polymer nanocomposites.

As for Ramanathan et al.'s report[123], our TRG/a-PMMA nanocomposite

systems were synthesized by the same process and used the same materials as given in

their report. Tg for our TRG/a-PMMA nanocomposites were consistent with

Ramanathan et al.[123] and the PMMA nanocomposites Tg results in Table 7-1. Thus,

the 30 °C increase in Tg reported by Ramanathan et al.[123] is suspect since they


108
obtained the same reported Tg results for nanocomposites (~125 °C) as other published

results.[77], [118], [128], [129], [132] Their increase can be explained by the unfilled

PMMA they used as the control. Our control PMMA and all the others in the literature

show a Tg ~125 °C, 30 °C higher than Ramanathan et al.’s reported number was ~95

°C. The Tg and mechanical properties improvements[4] in Ramanathan et al.’s

report[123] were systematically studied and reported in detail in the Appendix A.

109
6-4. Conclusions

The influence of graphene/GO on Tg was investigated in this chapter. We used

isotactic PMMA (i-PMMA) and syndiotactic-rich atactic PMMA (a-PMMA) to make

thermally reduced graphene (TRG)/PMMA nanocomposites using solvent blending in

order to investigate the stereo-regularity and in situ polymerization process effect on

the Tg. A Tg increase was found in i-PMMA and in situ PMMA but not in a-PMMA.

The results can be explained by the thin film confinement effect of polymer. We

attribute the Tg increase of TRG/i-PMMA nanocomposites to both a higher interaction

density and a stronger hydrogen bonding at the interfaces. Nearly pristine graphene

(PG) with extremely low oxygen level was also used as filler to make PG/i-PMMA

nanocomposites. The results show that the Tg increase of PG/i-PMMA

nanocomposites was observed but not as significant as TRG/i-PMMA system. We

attribute the results to less active site (oxygen functional groups) on PG to provide

interfacial interactions with i-PMMA.

We also prepared thermally reduced graphene (TRG)/PMMA nanocomposites

using in situ polymerization. In contrast to no Tg change in TRG/a-PMMA system, in

situ polymerized TRG/PMMA nanocomposites show a significant Tg increase. We

attribute the results to covalent bonding formation on GO surface at the interfaces

between matrix polymers and fillers combined with possibly better wetting by the

combination of MMA monomer and THF.

110
Chapter 7.

Glass Transition Temperature of Graphene/Polymer

Nanocomposites

7-1. Introduction

In this chapter, we review graphene/polymer nanocomposites literature

mentioning thermal properties and discovered that Tg is highly related to the (1)

preparation processes of nanocomposites and the (2) hydrophobicity of matrix

polymers. Moreover, the letter by Ramanathan et al.[123] reported an unprecedented

shift in glass transition temperature of ~30 C for a-PMMA at 0.05 wt% of TRG. We

have not been able to find such large increases in Tg elsewhere in the literature or in our

own results. Ramanathan et al. attributed their unusually high Tg increase to strong

hydrogen bonding interactions between hydroxyl groups on the TRG surface and

carbonyl groups of PMMA, as well as the restriction of PMMA segment mobilities by

mechanical interlocking of the wrinkled surface of TRG.

7-2. Literature Summary


In order to determine rules for how Tg changes with respect to various factors, we

did a broad literature review for thermal properties of graphene/polymer

nanocomposites. The results are classified as (1) Table 7-1: physical blending

processes (no chemical reaction involved, such as solvent or melt blending,) (2) Table

7-2: chemical blending process (chemical reaction involved), such as in situ


111
polymerization or chemically modified filler and (3) Table 7-3: aqueous blending

processes (solvent blending but using water as solvent for water-soluble matrix

polymers). In Table 7-1, all graphene/polymer nanocomposites synthesized via

solvent or melt blending show no significant Tg change as measured by DMA or DSC.

It indicates that solvent and melt blending processes do not provide enough interaction

between matrix polymers and graphene or graphene oxide (GO) fillers, leading to a no

Tg change for nanocomposites. Generally, a fixed supporting substrate with strong

interaction at interfaces (such as covalent or hydrogen bonds) can immobilize the

polymer chains, which leads to Tg increase within 8Rg from interfaces.[117] Table 7-2

shows all the graphene/polymer nanocomposites synthesized via in situ

polymerization or blending with chemically modified graphene or graphene oxide

(CRG). In contrast to the solvent/melt blended nanocomposites in Table 7-1, most of

the reports in Table 7-2 exhibit significant Tg increases. Several studies in Table 7-2

involve covalently bonding short chains on the fillers' surfaces prior to the blending

process to enhance compatibility, hence even solvent-blending modified graphene or

GO in polymers can cause significant Tg increase. The literature confirms our

expectation that such strong interactions from covalent bonds can actually raise the Tg

of nanocomposites, but not the weak van der Waals forces from

non-chemical-involved blending processes. Keddie et al.[119] found that van der

Waals interactions between a polymer thin film, and substrate could only lead to a Tg

change as large as 0.003K. In contrast to in situ polymerization in GO or reduced

graphene, if the nanocomposite filler is instead pristine graphene, it will still contain

graphite’s stable conjugate structures. We speculate that the pristine graphene’s

conjugate C=C is too stable to react with the propagating free radical, hence providing

no covalent bonding between graphene and the matrix polymer. This can explain why
112
pristine graphene /PBI[133] or NIPAA[134] nanocomposites have no Tg increases, as

observed in Table 2. Furthermore, even though the Tg has no change with graphene

incorporation in Table 7-1, significant mechanical properties improvements of the

nanocomposites were commonly observed.

Within the same conclusion, however, Table 7-3 shows that aqueous blended

(solvent blending but using water or acid aqueous solution as solvents)

GO/water-soluble polymers do show a significant Tg increase even if the

nanocomposites were synthesized via non-chemical-reaction-blending processes. It is

well known that water-soluble polymers can form strong hydrogen bonding with water.

Hence, we speculate that strong hydrogen bonding was formed between polymers and

GO (or TRG, CRG) after the simple aqueous blending process. This interesting

phenomenon further proves our previous conclusion that interaction between matrix

polymers and graphene is the primary cause of Tg increase. A critical point though is

that hydrogen bonds will be weakened at temperatures higher than 100 °C, as such,

any reports with a matrix polymer's Tg much higher than 100 °C should not attribute

the Tg increase to hydrogen bonds. For example, as for GO/chitosan nanocomposites

with a matrix polymer Tg at 175.4 °C, not only weak hydrogen bonding but also

electrostatic attraction immobilizes the chitosan polymer chains, resulting in a small Tg

increase of only 5 °C.[135–137] Another possible reason for Tg increase in Table 3

will be wetting ability of water for GO. Water is a good dispersion solvent for GO

compared to other organic solvents for graphene. Hence, we speculate that water

soluble polymers can be wetted onto GO more easily than other organic solvent

blending processes, leading to strong interfacial interactions. With all the results we

reviewed and found, however, no report has ever addressed the molecular weight of

the covalently bonded chains, which might significantly affect the entanglement
113
density or morphology of the nearby free matrix polymers. What is the morphology of

the polymer chains at the interfaces? More work needs to be addressed in the future.

7-3. Conclusions

Combine the experimental results (in Chapter 6) and literature summary have

proved that the Tg of graphene or GO/polymer nanocomposites can be increased only

if strong interactions are formed between matrix polymers and fillers.

Non-chemical-reaction involved blending processes, such as solvent and melt

blending, are incapable of providing enough interactions between the matrix

polymers and fillers. Hence, the Tg of nanocomposites are independent of graphene

or GO incorporation. However, aqueous-soluble polymers can easily form strong

hydrogen bonding with TRG, CRG and GO, and the Tg increase of such

nanocomposites were commonly observed. Chemical-reaction involved blending

processes, such as chemically modified graphene or in situ polymerization of

monomers in the presence of TRG, CRG, or GO, are capable of providing strong

covalent bonds between matrix polymers and fillers, which cause an increase of the

Tg. One exception is to in situ polymerize monomers in the presence of pristine

graphene, which causes no Tg change of nanocomposites. It is because no oxidized

functional groups or reactive C=C exist on pristine graphene surface for monomer to

chemically react with.

114
Table 7-1. Physical blending processes, including solvent and melt blending.
Solvent Blending

Matrix Polymer ∆Tg (°C) Filler Technique Filler Load

PVDF[98] 0 TRG DSC 4 wt%

PBS[138] 0 CRG DMA 2 wt%

LLDPE[139] 0 EG DMA 20 wt%

PVDF[140] 3.5 TRG DMA 0.5 wt%

TPU[122] -2 TRG DSC 7 wt%

PαMSAN/PMMA[141 0 TRG DMA 1 wt%


]
Rubber[142] 0 GO DSC 10 wt%

TPU[143] -5 EG DSC 10 vol%

TPU[144] 0 (soft segment) TRG DMA 6 wt%

PS[108] 0 CRG DMA 1.94 vol%

PI[107] 2.6 -ABA-GO DMA 0.4 wt%

PVA[132] -4.5 1 wt%

PMMA[132] 0.3 0.2 wt%


TRG DSC
PEI[132] 0 0.1 wt%

TPU[145] 0 TRG DMA 4.4 wt%

Melt Blending

PE[146] 0 CRG DMA 2 wt%

PA12[147] 0 GS (XG-Science) DSC 2 wt%

PET[148] <2 TRG DMA 7 wt%

PTT[149] 0 TRG DSC 7 wt%

PC[111] 0 GO DMA 3 wt%

PP[110] 0 GO DMA 5 wt%

115
Table 7-2. Chemical blending processes (in situ polymerization or chemically
modified fillers)
In situ Polymerization

Matrix Polymer ∆Tg (°C) Filler Technique Filler load

PBI[133] 1.8 Pristine graphene DMA 0.2 wt%

NIPAA[134] 0 Pristine graphene DSC 0.13 wt%

PEO[150] 0 GO DSC 2 wt%

Epoxy[151] <2 TRG DSC 0.1 wt%

Epoxy[112] 3 GO DSC 0.5 wt%

Epoxy[152] 7 TRG DSC 0.05 wt%

40 DSC
Epoxy[153] TRG 1.5 wt%
37 DMA

10 (DDS)
Epoxy[154] Expanded graphite DMA 1.988 vol%
5 (J230)

PMMA[77] 20 CRG DMA 4 wt%

PMMA[128] 9 CRG DMA 1 wt%

3 GO
PMMA[129] DMA 1 wt%
8 CRG

14 DSC
PMMA[76] 6 wt%
GO
19 DMA

7 Micro-AIBN-initiator DSC & DMA

Vinyl Ester[155] 7 & GO


TRG -- 0.2 phr

4 GO 0.5 wt%
PUA[156] DSC
3.4 Isocyanate-GO 1 wt%

PUA[157] 0 TRG DSC & DMA 0.5 wt%

PUA[158] 7 CRG DMA 1 wt%

PS[159] 8 PS-modified CRG DSC 2.5 wt%

TPU[160] -4 TRG DMA 0.3 wt%

TPU[161] 16 CRG DMA 0.3 wt%


PBI: polybenzimidazole; NIPAA: poly(N-isopropylacrylamide)
116
Table 7-2. Cont' Chemical blending processes (in situ polymerization or chemically
modified fillers)

Modified GS

Matrix
Polymer ∆Tg (°C) Filler Technique Filler load Process

-- 0.7 PDMAEMA-graft-TRG[162] DSC -- --

PVDF[163] 8 PMMA-graft-CRG DSC 5 wt% SB

Epoxy[164] 8 PEO/PPO-modified GS DMA 0.489 vol% In Situ

PMMA[165] 12 methacryloyl chloride-GO DMA 0.5 wt% In Situ

-- 15 PS-grafted CRG[166] DSC -- --

ABS[167] 9 Octadecyl-amine-graft-GO DSC 1 wt% SB

17 PMMA-graft-TRG[168]

-- 0 PMMA-adsorb-TRG [168] DSC -- --

Phenyl isocyanate-modified
PS[109] 6 GO/CRG end DMA 0.5 wt% SB

25 PVC-graft-CRG 1.4 wt%


PVC[169] DSC SB
CRG (no modification, pure
3 SB) 5 wt%

-- 15 PS-graft-CRG[131] DSC -- --

-- 7 PGMA-graft-GO[125] DSC -- --

117
Table 7-3. Aqueous blended processes (solvent blending but using water as solvent
for water-soluble matrix polymers)

Aqueous Blending (solvent blending with water as the solvent)

Matrix Polymer ∆Tg (°C) Filler Technique Filler load

PVA[104] 12 CRG DSC 7.5 wt%

PVA[170] 4 CRG DSC 0.5 wt%

PVA[105] 14 CRG DSC 3.5 wt%

PVA[171] 9 GO DSC 0.72 vol%

PVA[172] 4 GO DSC 0.7 wt%

Chitosan[135] 5 GO DSC 1 wt%

Gelatin[173] 10 GO DSC 2 wt%

PEO[174] 9 TRG DMA 4 wt%

118
Chapter 8. Oxygen Effect of Stiffness of Graphene Oxide

8-1 Introduction

Improvement of mechanical properties is one of the main reasons for creating

graphene/polymer nanocomposites. As mentioned in Chapter 1, a filler with high

mechanical properties is one of the three criteria for reinforcement of polymers. The

high elastic modulus of graphene makes it very attractive as a filler. A value of 1 TPa

is often given[80], however, lower values have been reported in the literature. Ruoff

and co-workers reported that the elastic modulus of graphene oxide (GO) paper is

approximately ~32 GPa.[175] In 2007, Frank et al.[176] reported an elastic modulus of

0.5 TPa of multi-layer pristine graphene with a 2-8 nm thickness using AFM

nano-indentation. Gomez-Navarro et al.[91] reported that chemically reduced

graphene (CRG) has an elastic modulus of 0.25 TPa. Suk et al.[92] reported that the

elastic modulus of single-layer graphene oxide is 0.21 TPa. Moreover, nearly all

graphene/polymer nanocomposites reported to date have been reinforced with

graphene reduced from GO. According to the diverse results above, how the elastic

modulus changes with various oxygen levels on reduced GO remains an open

question.

As described in Chapter 2, we realized that the large interconnected oxygen-free

patches on GO can increase the electrical conductivity dramatically even with low C/O

ratio (6/1). Hence, there might be a specific C/O ratio of GS for oxygen-free patches to

reach a 2-D percolation. If our assumption is true, the elastic modulus and electrical

conductivity of our GS increases significantly at the C/O ratio of percolation of

oxygen-free patches. We will study the elastic modulus of single layer GO and GS

119
with different C/O ratios.

Figure 8-1. "In situ" GO modulus measurement flowchart via AFM


nano-indentation.

120
8-2 Experiments & Results

The experiment process is illustrated in Figure 8-1. Single-layer GO sheet is the

same batch of GO used in Chapter 3 & 4. A silicon wafer with patterned holes for

Phase I was provided by the Nano Fabrication Center at the University of Minnesota.

TEM grids (Quantifoil; R2/2) with patterned holes were used for Phase II. We

measured the modulus of a single-layer GO sheet suspended on holes followed by

chemical reduction by hydrazine. Chemical reduction was conducted using

dimethyl-hydrazine (Aldrich; used as received). Dimethyl-hydrazine aqueous solution

with a concentration of 1 wt% was prepared. Next, the solution was contained in an

opened small glass container, which was left in a sealed bigger glass container for

reduction. GO suspended holey silicon wafer or TEM grids were also left in the same

sealed container. Next the sealed container with dimethyl-hydrazine aqueous solution

and GO grid or wafer was left in the oven at 80 °C for certain time periods. We then

measured the same suspended GO sheet after a total reduction time of 1 hour and 4

hours. The "in situ" AFM nano-indentation allows us to understand how the GO elastic

modulus changes with various oxygen levels.

A drop of dilute GO aqueous solution was dropped on silicon wafer substrate or

the grid with patterned holes and air-dried at room temperature. A Digital Instrument,

Nanoscope III Multimode AFM at the Institute of Technology Characterization

Facility of the University of Minnesota was used in tapping mode for measurements.

8-2.1 Phase I: Silicon Wafer Substrate

We measured the same GO (or GS) sheet before and after chemical reduction. A

silicon wafer with patterned micro-holes 0.5 m in diameter was prepared as the

121
substrate by the Nano Fabrication Center of University of Minnesota. Aqueous GO

solution was obtained from the same lot (pGO-5; 5 min probe sonicated) as in Chapter

3 and was spin-coated onto the holey silicon wafer for further AFM nano-indentation

measurements. We expected that a silicon wafer with patterned holes is an ideal

substrate because it is strong enough to support suspended GO sheets and very flat.

However, as shown in Figure 8-2, it was observed that all the suspended GO sheets

were broken after the drying process followed by spin coating. We speculate that a

reason for the frustrating results is that the holes are blind (the holes were not drilled

through the wafer), which might cause the trapped water molecules to break the

covered GO sheet during evaporation. This phenomenon was observed during every

experiment. Hence, we decided that the patterned blind holey substrate is not

appropriate for our AFM measurement. Figure 8-3 is a multi-layer GO covered hole

poked through by the cantilever, the shape of cantilever tip is clearly observed.

Figure 8-2. AFM images of patterned holes covered by broken single-layer GO


sheet. All the holes have the same diameter of 0.5 m and less than 1 m thick.

122
Figure 8-3. AFM image of a multi-layer GO covered hole (left) poked through by
the cantilever (right).

8-2.2 Phase II: TEM Grid Substrate

In order to resolve the problem described in 8-2-1, we used a TEM grid carbon

thin film with patterned holes for AFM nano-indentation instead of a silicon wafer.

The grid was purchased from Quantifoil Inc, and the diameter of carbon thin film holes

was 2.5 m. The same grid was used by Suk et al.'s[92] report on AFM

nano-indentation measurement on the elastic modulus of GO. The AFM sample grid

was mounted on a TEM holder as shown in Figure 8-4. Although the holey carbon film

worked well for Suk et al.'s report, it did not work for us. The reason is the poor

thermal stability of the holey carbon film. After 1 h of hydrazine vapor reduction at 80

°C, we found that the stretched holey carbon film became loose from the copper grid

and was too soft to support the GO. In order to improve the thermal stability of the

holey carbon film, a 20 nm thick layer of chromium was coated on the both side of the

film to prevent the carbon thin film from being softened during chemical reduction

process. Aqueous GO solution described in 8-2-1 was dropped onto the TEM grid to

deposit GO on the holes. Unlike the silicon wafer, the chromium coated TEM grid
123
holey carbon film is not able to measure the thickness via height imaging as shown in

Figure 8-5.

Figure 8-4. The AFM sample grid was mounted on a TEM holder.

Figure 8-5. AFM images of patterned holey carbon film covered by GO sheet.
All the holes have the same diameter of 2.5 m.

124
TEM:

AFM:

Figure 8-6. An example of locating the same GO sheet under TEM and AFM.
The arrows point out the feature to locate the GO sheet. The upper left insert
picture in the lower resolution TEM image is ED to check that the GO sheet
which covers the hole is single-layer. As mentioned in Chapter 2, six electron
diffraction patterns in hexagonal manner indicates single-layer structure of GO.

Due to the difficulty of measuring the thickness of covered GO sheets, TEM was

used to check if the covered GO sheets are single-layer or multi-layer. We used TEM

electron diffraction (ED) to locate single-layer GO sheets covering holes of the holey

carbon film; the same GO sheet was then located again under AFM to measure the

modulus. In order to locate the same GO sheet from TEM, small features to point the

GO sheet out are needed. Figure 8-6 is an example how to locate the same GO sheets

under TEM and AFM. The red arrows point out the feature used to locate the GO sheet.

The upper left insert image in the lower resolution TEM image is ED to show that the

GO sheet which covers the hole is single-layer. As mentioned in Chapter 2, six

125
electron diffraction patterns in a hexagonal manner indicates single-layer structure of

GO.

Tapping mode AFM nano-indentation was operated after successful location of

the appropriate GO sheet. AFM works by moving the sample along Z-directions to let

the sample interact with the cantilever. The CCD sensor will then detect the position of

the light reflected by cantilever, which is illustrated in Figure 8-7:

Figure 8-7. Illustration how an AFM work. By moving the sample in Z-direction,
the sensor can detect the light position change to report the sample surface.

As illustrated in Figure 8-7, the AFM is not capable of direct measuring the force the

cantilever applies to the sample since there is no force sensor. Raw AFM data provides

only (1) Sample stage displacement and (2) cantilever deflection, as shown in Figure

8-8: (Cantilever deflection is calculated from a voltage change in sensor.)

126
(a)

(b)

(c)

Figure 8-8. (a) Illustration of deflection and stage displacement (


raw traction and (c) retraction AFM deflection as curves for two different
runs.

127
From Figure 8-8, we can tell that the force applied to the sample equals the force to

bend the cantilever, which causes cantilever deflection. The applied force will then be

the deflection multiplied by the spring constant of cantilever. Also, Figure 8-8 shows

that the AFM sample stage displacement Z does not equal the cantilever real

indented depth. The real indented depth should be Z subtracted by the cantilever

deflection.

The calibrated AFM deflection-indented depth curves of the same GO sheet with

different reduction times are shown in Figure 8-9. It can be observed that the stiffness

decreases with longer reduction time, which indicates higher C/O ratios. The results

can be explained by two possible reasons: (1) oxygen atom bridge removal after

reduction,[173] and (2) dislocation formation after reduction. As mentioned in

Chapter 2, oxygen atoms on GO mainly exist as epoxy and hydroxyl groups. As shown

in Figure 8-10, one oxygen atom of the epoxy group forms a bridge between two

neighbor carbon atoms, which can eliminate the rotation ability of the carbon-carbon

bond. According to Incze et al.'s[177] report, reduction of the epoxy oxygen removes

the oxygen bridges, which leads to a lower stiffness of GO.

128
Figure 8-9. Deflection-indentation depth curves of the same GO sheet with
different reduction times.

Figure 8-10. Epoxy oxygen atoms form bridges between two neighboring carbon
atoms. The bridge is removed after reduction, leading to a lower stiffness of GO.
(The figure is modified from Mkhoyan et al.[16])

8-3 Future works

According to the results in 8-2, the stiffness of GO will decrease with lower

oxygen levels. However, more studies need to be done. It is important to understand if


129
oxygen bridge or dislocation or both are the reason for lower stiffness. Furthermore,

absolute numbers of elastic modulus need to be reported. Lee et al.[80] reported a

cubic parabolic equation to fit force-displacement curves to obtain a 2D modulus:

(8-1)

where F is the applied force, o2D is the 2D pretension, a is the membrane diameter, 

is the indented depth at the center point, E2D is the 2D elastic modulus, q is a dimension

less constant. The elastic modulus will be E2D divided by the thickness of the GO sheet.

However, comparison of Lee et al.'s[80] report to our results in Figure 8-9 shows that

our indentation depth is too low to fit equation 8-1. Deeper indentation will be

necessary for the next measurement for calculation of the elastic modulus.

Polymer molecular nanoconfinement at the interfaces between filler and matrix

polymers is still unclear to date.[4] We propose to investigate modifications of

polymer matrix morphology at interfaces between matrix polymer and graphene using

AFM. As described in Chapter 6, DSC of bulk PMMA-graphene nanocomposites

suggest that the PMMA Tg is elevated due to the inclusion of graphene flakes. We wish

to locally measure Tg on variable-size patches of PMMA adsorbed or covalently

bonded to graphene surfaces to model the interfacial system.

130
Chapter 9. Results & Future Works
9-1 Summary of Results

We have discovered a new, simple, hydrazine-free, high-yield method for

producing single-layer graphene sheets. Graphene sheets were formed from graphite

oxide by reduction with simple D.I. water at 95 ºC under atmospheric pressure. Over

65 % of the sheets are single graphene layers; the average sheet diameter is 300 nm.

We speculate that dehydration of GO is the main mechanism for oxygen reduction and

transformation of C-C bonds from sp3 to sp2. The reduction appears to occur in large

uniform interconnected oxygen-free patches so that despite the presence of residual

oxygen the sp2 carbon bonds formed on the sheets are sufficient to provide electronic

properties comparable to reduced graphene sheets obtained using other methods. The

size and extent of exfoliation of graphene oxide sheets was varied by sonication

intensity and time. Graphene sheets were obtained from graphene oxide by a simple

(hydrazine-free) hydrothermal route. The particle size, morphology, exfoliation extent,

oxygen content, and surface charge of graphene oxide and graphene were

characterized by wide-angle powder X-ray diffraction, atomic force microscopy,

X-ray photoelectron spectroscopy, dynamic light scattering, and zeta-potential.

One method of toxicity assessment was based on measurement of the efflux of

hemoglobin from suspended red blood cells. At the smallest size, graphene oxide

showed the greatest hemolytic activity while aggregated graphene sheets exhibited the

lowest hemolytic activity. Coating graphene oxide with chitosan nearly eliminated

hemolytic activity. Together, these results demonstrate that particle size, particulate

state, and oxygen content/surface charge of graphene have a strong impact on

biological/toxicological responses to red blood cells. In addition, the cytotoxicity of


131
graphene oxide and graphene sheets was investigated by measuring mitochondrial

activity in adherent human skin fibroblasts using two assays. The

methylthiazolyldiphenyl-tetrazolium bromide (MTT) assay, a typical nanotoxicity

assay, fails to predict the toxicity of graphene oxide and graphene toxicity due to the

spontaneous reduction of MTT by graphene and graphene oxide, resulting in a false

positive signal. However, an appropriate alternate assessment, using the water soluble

tetrazolium salt (WST-8) assay, reveals that the compacted graphene sheets are more

damaging to mammalian fibroblasts than the less densely packed graphene oxide.

Clearly, the toxicity of graphene and graphene oxide depends on whether or not

aggregation occurs and mode of interaction with cells (i.e. suspension versus adherent

cell types).

Using thermally reduced graphene (TRG) and in-situ polymerization we have

created PUA nanocomposites with an ultralow percolation concentration of 0.15 wt%

(0.07 vol%) graphene. Urethane-acrylate oligomer (UAO) was synthesized and

diluted by tripropyleneglycol diacrylate (TPGDA) to form flowable UAO/TPGDA

mixture (UA). TRG was solvent-blended in UA to form uncured TRG/UA liquids and

were polymerized by free radical polymerization with azobisisobutyronitrile (AIBN)

initiator. Percolation concentrations of polymerized TRG/PUA nanocomposites

occurred at 0.15 wt% (0.07 vol%), as determined by surface resistance measurements,

bulk electrical conductivity, and modulus. TEM images revealed a homogeneous

dispersion of TRG in PUA. Differential scanning calorimetry (DSC) was used to

monitor the polymerization of TRG/UA uncured liquids and thermal properties of

polymerized TRG/PUA nanocomposites. Polymerization heat, glass transition

temperature, and polymerization temperature are independent of TRG loading, though

polymerization temperature is ~ 10 °C lower in the absence of TRG.


132
How graphene/polymer nanocomposite glass transition temperatures (Tg) vary

was investigated in this study. We first surveyed the literature to summarize the rule of

Tg change with graphene incorporation. No changes in Tg were observed for

graphene/polymer nanocomposites synthesized via physical blending processes such

as solvent or melt blending, except aqueous blending. In contrast, chemical blending

processes such as in situ polymerization or chemically modified fillers yielded

significant Tg increases in graphene/polymer nanocomposites. We attribute these

results to bonding interactions at the interfaces between matrix polymers and fillers.

Physical blending processes cannot provide enough interaction at the interfaces,

whereas chemical blending processes can yield strong interaction such as covalent

bonds. Aqueous blending of oxygen barriers graphene nanocomposites with water

soluble matrix polymers also cause Tg increases, even though the blending processes

involve no chemical reactions. The reason for this exception is that hydrogen bonding

forms between fillers (graphene oxide or reduced graphene) and water soluble matrix

polymers. Besides the literature review, we used isotactic PMMA (i-PMMA) and

syndiotactic-rich atactic PMMA (a-PMMA) to make thermally reduced graphene

(TRG)/PMMA nanocomposites using solvent blending in order to investigate the

stereo-regularity effect on the Tg. A Tg increase was found in i-PMMA and in situ

PMMA but not in a-PMMA. The results can be explained by the thin film confinement

effect of polymer. We attribute the Tg increase to both a higher interaction density and

a stronger hydrogen bonding at the interfaces.

133
9-2 Future Directions
9-2.1 Dispersion of Aqueous Reduced Graphene in Polymers

As described in Chapter 2, the aqueous reduction method for graphene was

motivated by mass production of graphene in a low cost, environmental-friendly

manners. However, dispersion of aqueous reduced graphene (ARG) in polymers has

not been fully developed. In Chapter 4, we have already successfully dispersed ARG

in TPU via the novel co-solvent blending process. In contrast to thermally reduced

graphene (TRG), which is exfoliated in the air, ARG is exfoliated and reduced in water.

Hence, removing water from ARG might cause aggregation and make dried ARG

difficult to redispersed into solvents or melt polymers. Besides the co-solvent blending

process reported in Chapter 4, other graphene dispersion processes should be studied

to improve dispersion of ARG in various matrix polymers. In Chapter 6, we have

successfully dispersed TRG in PMMA using in situ polymerization. ARG/PMMA

nanocomposites prepared via in situ polymerization should be studied in the future.

Water soluble polymers, for example poly-vinylalcohol or poly-ethyleneoxide, can be

used as matrix polymer. These water soluble polymers can be dissolved in water prior

to the aqueous reduction process for better dispersion of ARG.

9-2.2 Dispersion of Graphene in Solvents

Dispersion of graphene sheets in solvents is important for applications such as

graphene/polymer composites. As reported previously, solvent blending can facilitate

dispersion of graphene. If we plan to use solvent blending for graphene/polymer

composite materials, we will need to study the dispersion of graphene in various

solvents. Dispersion of graphene particles in solvents in dilute solutions can be

evaluated by measuring the relative viscosity, which will increase with better
134
dispersion. Competition between Brownian motion and shear stress (orientation) of

graphene oblate particles will occur. At low shear stress, Brownian motion will

dominate the system and increase the viscosity. On the other hand, when shear stress

dominates, the particles become oriented, and shear thinning will occur. Shear

thinning results from oriented non-spherical particles which decreases the viscosity.

Hence, we can disperse like amounts of graphene in different solvents and measure the

viscosities to understand the dispersion. Higher viscosities will correspond to better

dispersion.

9-2.3 Glass Transition of Graphene/Polymer Nanocomposites

According to 9-1, we understand that the interfacial interaction between graphene

and matrix polymers is the key of Tg change. However, there are still several open

questions for glass transition of nanocomposites: (1) Why are some Tg changes of

graphene/polymer nanocomposites observed using DSC not as obvious as measured

using DMA? (2) What is the morphology of polymer coils at the interfaces between

fillers and matrix polymers? (3) What is the detailed chemical reaction mechanism of

monomers that are undergo free radical polymerization in the presence of graphene or

graphene oxide? These fundamental questions need to be resolved for a deeper

understanding of graphene/polymer nanocomposites.

9-2.4 Graphene/Polymer Nanocomposites

One of the ultimate goals of graphene/polymer nanocomposites is to make

flexible transparent electrical conductive film (FTCF). However, even well dispersed

0.1 wt% of TRG/TPU nanocomposite reaching percolation concentration, as described


135
in Chapter 4, appears to be opaque. This result tells us that a heterogeneous dispersion

system is required for FTCF graphene/polymer nanocomposites. The main issue with

homogeneous graphene/polymer nanocomposites is that they require a certain

graphene loading to form an interconnected graphene network (percolation), at the

cost of transparency of the materials. Hence, heterogeneous dispersed graphene is

necessary for graphene/polymer FTCF. More details are in Appendix B.

9-2.5 Solvothermal Graphene Synthesis

Solvothermal reduction of GO for ARG has been reported in Chapter 2 with

water as the solvent media. Here we proposed to use other solvents as media to reduce

GO. Water is not a perfect solvent media because it cannot well disperse ARG after

aqueous reduction. A solvent which is capable to well disperse both GO and reduced

GO can be an ideal candidate. N,N-dimethylforamide (DMF) or N-methylpyrrolidone

(NMP) are desirable to prevent aggregation of reduced GO after solvothermal

reduction since they are good solvent for both GO and reduced GO. Furthermore,

DMF and NMP have high boiling points, a higher reduction temperature can be

applied to accelerate the reaction without increasing pressure. Reduced GO/DMF or

NMP can be used directly for solvent blending process for making graphene/polymer

nanocomposites.

136
References

[1] A. K. Geim and K. S. Novoselov, “The rise of graphene,” Nature Materials, vol. 6, no. 3,
pp. 183–191, Mar. 2007.
[2] Y. Hernandez, V. Nicolosi, M. Lotya, F. M. Blighe, Z. Sun, S. De, I. T. McGovern, B.
Holland, M. Byrne, Y. K. Gun’Ko, J. J. Boland, P. Niraj, G. Duesberg, S. Krishnamurthy, R.
Goodhue, J. Hutchison, V. Scardaci, A. C. Ferrari, and J. N. Coleman, “High-yield
production of graphene by liquid-phase exfoliation of graphite,” Nature
Nanotechnology, vol. 3, no. 9, pp. 563–568, Aug. 2008.
[3] K. S. Novoselov, “Electric Field Effect in Atomically Thin Carbon Films,” Science, vol.
306, no. 5696, pp. 666–669, Oct. 2004.
[4] H. Kim, A. A. Abdala, and C. W. Macosko, “Graphene/Polymer Nanocomposites,”
Macromolecules, vol. 43, no. 16, pp. 6515–6530, Aug. 2010.
[5] J. C. Meyer, A. K. Geim, M. I. Katsnelson, K. S. Novoselov, T. J. Booth, and S. Roth, “The
structure of suspended graphene sheets,” Nature, vol. 446, no. 7131, pp. 60–63, Mar.
2007.
[6] V. C. Tung, M. J. Allen, Y. Yang, and R. B. Kaner, “High-throughput solution processing
of large-scale graphene,” Nature Nanotechnology, vol. 4, no. 1, pp. 25–29, Nov. 2008.
[7] C. Berger, Z. Song, T. Li, X. Li, A. Y. Ogbazghi, R. Feng, Z. Dai, A. N. Marchenkov, E. H.
Conrad, P. N. First, and W. A. de Heer, “Ultrathin epitaxial graphite: 2D electron gas
properties and a route toward graphene-based nanoelectronics,” The Journal of
Physical Chemistry B, vol. 108, no. 52, pp. 19912–19916, Dec. 2004.
[8] S. Y. Zhou, G.-H. Gweon, A. V. Fedorov, P. N. First, W. A. de Heer, D.-H. Lee, F. Guinea,
A. H. Castro Neto, and A. Lanzara, “Substrate-induced bandgap opening in epitaxial
graphene,” Nature Materials, vol. 6, no. 10, pp. 770–775, Sep. 2007.
[9] K. S. Kim, Y. Zhao, H. Jang, S. Y. Lee, J. M. Kim, K. S. Kim, J.-H. Ahn, P. Kim, J.-Y. Choi,
and B. H. Hong, “Large-scale pattern growth of graphene films for stretchable
transparent electrodes,” Nature, vol. 457, no. 7230, pp. 706–710, Jan. 2009.
[10] K. Yang, S. Zhang, G. Zhang, X. Sun, S.-T. Lee, and Z. Liu, “Graphene in mice: ultrahigh
in vivo tumor uptake and efficient photothermal therapy,” Nano Letters, vol. 10, no. 9,
pp. 3318–3323, Sep. 2010.
[11] Y. Chang, S.-T. Yang, J.-H. Liu, E. Dong, Y. Wang, A. Cao, Y. Liu, and H. Wang, “In vitro
toxicity evaluation of graphene oxide on A549 cells,” Toxicology Letters, vol. 200, no. 3,
pp. 201–210, Feb. 2011.

137
[12] S.-R. Ryoo, Y.-K. Kim, M.-H. Kim, and D.-H. Min, “Behaviors of NIH-3T3 Fibroblasts on
Graphene/Carbon Nanotubes: Proliferation, Focal Adhesion, and Gene Transfection
Studies,” ACS Nano, vol. 4, no. 11, pp. 6587–6598, Nov. 2010.
[13] K. Wang, J. Ruan, H. Song, J. Zhang, Y. Wo, S. Guo, and D. Cui, “Biocompatibility of
Graphene Oxide,” Nanoscale Research Letters, vol. 6, p. 8, Aug. 2010.
[14] W. S. Hummers and R. E. Offemann, “Preparation of Graphite Oxide,” Journal of
American Chemical Society, vol. 80, p. 1339, 1958.
[15] H. K. Jeong, M. H. Jin, K. P. So, S. C. Lim, and Y. H. Lee, “Tailoring the characteristics of
graphite oxides by different oxidation times,” Journal of Physics D: Applied Physics, vol.
42, no. 6, p. 065418, Mar. 2009.
[16] K. A. Mkhoyan, A. W. Contryman, J. Silcox, D. A. Stewart, G. Eda, C. Mattevi, S. Miller,
and M. Chhowalla, “Atomic and Electronic Structure of Graphene-Oxide,” Nano Letters,
vol. 9, no. 3, pp. 1058–1063, Mar. 2009.
[17] D. R. Dreyer, S. Park, C. W. Bielawski, and R. S. Ruoff, “The chemistry of graphene
oxide,” Chemical Society Reviews, vol. 39, no. 1, p. 228, 2010.
[18] K.-H. Liao, A. Mittal, S. Bose, C. Leighton, K. A. Mkhoyan, and C. W. Macosko,
“Aqueous Only Route toward Graphene from Graphite Oxide,” ACS Nano, vol. 5, no. 2,
pp. 1253–1258, Feb. 2011.
[19] K.-H. Liao, Y.-S. Lin, C. W. Macosko, and C. L. Haynes, “Cytotoxicity of Graphene Oxide
and Graphene in Human Erythrocytes and Skin Fibroblasts,” ACS Applied Materials &
Interfaces, vol. 3, no. 7, pp. 2607–2615, Jul. 2011.
[20] Auto application of drive commercialization of nanocomposties, vol. 4. Plastic Additives
& Compounding, 2002.
[21] Y. Kojima, A. Usuki, M. Kawasumi, A. Okada, T. Kurauchi, and O. Kamigaito, “One-pot
synthesis of nylon 6–clay hybrid,” Journal of Polymer Science Part A: Polymer Chemistry,
vol. 31, no. 7, pp. 1755–1758, Jun. 1993.
[22] K. Kalaitzidou, H. Fukushima, and L. T. Drzal, “Multifunctional polypropylene
composites produced by incorporation of exfoliated graphite nanoplatelets,” Carbon,
vol. 45, no. 7, pp. 1446–1452, Jun. 2007.
[23] B. Kelly, U. Stredulinsky, J. Hoek, and J. Levy, “Antibodies in the sera of patients with
bronchogenic carcinoma that react with antigen from a tumour cell line,” Cancer
Immunology Immunotherapy, vol. 12, no. 1, Dec. 1981.
[24] L. S. Schadler, S. C. Giannaris, and P. M. Ajayan, “Load transfer in carbon nanotube
epoxy composites,” Applied Physics Letters, vol. 73, no. 26, p. 3842, 1998.
[25] O. L. Blakslee, “Elastic Constants of Compression-Annealed Pyrolytic Graphite,”
Journal of Applied Physics, vol. 41, no. 8, p. 3373, 1970.

138
[26] R. H. Knibbs, “The young’s modulus of pyrolytic graphite,” Carbon, vol. 7, no. 2, pp.
225–228, Apr. 1969.
[27] T. Tsuzuku, “Anisotropic electrical conduction in relation to the stacking disorder in
graphite,” Carbon, vol. 17, no. 3, pp. 293–299, Jan. 1979.
[28] Y. Zhang, Y.-W. Tan, H. L. Stormer, and P. Kim, “Experimental observation of the
quantum Hall effect and Berry’s phase in graphene,” Nature, vol. 438, no. 7065, pp.
201–204, Nov. 2005.
[29] E. I. Asinovskii, A. V. Kirillin, and A. V. Kostanovskii, “Experimental investigation of the
thermal properties of carbon at high temperatures and moderate pressures,”
Physics-Uspekhi, vol. 45, no. 8, pp. 869–882, Aug. 2002.
[30] K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Katsnelson, I. V. Grigorieva, S.
V. Dubonos, and A. A. Firsov, “Two-dimensional gas of massless Dirac fermions in
graphene,” Nature, vol. 438, no. 7065, pp. 197–200, Nov. 2005.
[31] K. S. Novoselov, “Two-dimensional atomic crystals,” Proceedings of the National
Academy of Sciences, vol. 102, no. 30, pp. 10451–10453, Jul. 2005.
[32] S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas, A. Kleinhammes, Y. Jia, Y. Wu, S.
T. Nguyen, and R. S. Ruoff, “Synthesis of graphene-based nanosheets via chemical
reduction of exfoliated graphite oxide,” Carbon, vol. 45, no. 7, pp. 1558–1565, Jun.
2007.
[33] G. Eda, G. Fanchini, and M. Chhowalla, “Large-area ultrathin films of reduced
graphene oxide as a transparent and flexible electronic material,” Nature
Nanotechnology, vol. 3, no. 5, pp. 270–274, Apr. 2008.
[34] H. C. Schniepp, J.-L. Li, M. J. McAllister, H. Sai, M. Herrera-Alonso, D. H. Adamson, R. K.
Prud’homme, R. Car, D. A. Saville, and I. A. Aksay, “Functionalized Single Graphene
Sheets Derived from Splitting Graphite Oxide,” The Journal of Physical Chemistry B, vol.
110, no. 17, pp. 8535–8539, May 2006.
[35] W. Lv, D.-M. Tang, Y.-B. He, C.-H. You, Z.-Q. Shi, X.-C. Chen, C.-M. Chen, P.-X. Hou, C.
Liu, and Q.-H. Yang, “Low-Temperature Exfoliated Graphenes: Vacuum-Promoted
Exfoliation and Electrochemical Energy Storage,” ACS Nano, vol. 3, no. 11, pp.
3730–3736, Nov. 2009.
[36] J. T. Paci, T. Belytschko, and G. C. Schatz, “Computational Studies of the Structure,
Behavior upon Heating, and Mechanical Properties of Graphite Oxide,” Journal of
Physical Chemistry C, vol. 111, no. 49, pp. 18099–18111, Dec. 2007.
[37] J. Chattopadhyay, A. Mukherjee, C. E. Hamilton, J. Kang, S. Chakraborty, W. Guo, K. F.
Kelly, A. R. Barron, and W. E. Billups, “Graphite Epoxide,” Journal of the American
Chemical Society, vol. 130, no. 16, pp. 5414–5415, Apr. 2008.

139
[38] C. Mattevi, G. Eda, S. Agnoli, S. Miller, K. A. Mkhoyan, O. Celik, D. Mastrogiovanni, G.
Granozzi, E. Garfunkel, and M. Chhowalla, “Evolution of Electrical, Chemical, and
Structural Properties of Transparent and Conducting Chemically Derived Graphene Thin
Films,” Advanced Functional Materials, vol. 19, no. 16, pp. 2577–2583, Aug. 2009.
[39] S. Stankovich, D. A. Dikin, G. H. B. Dommett, K. M. Kohlhaas, E. J. Zimney, E. A. Stach, R.
D. Piner, S. T. Nguyen, and R. S. Ruoff, “Graphene-based composite materials,” Nature,
vol. 442, no. 7100, pp. 282–286, Jul. 2006.
[40] S. Gilje, S. Han, M. Wang, K. L. Wang, and R. B. Kaner, “A Chemical Route to Graphene
for Device Applications,” Nano Letters, vol. 7, no. 11, pp. 3394–3398, Nov. 2007.
[41] C. Gómez-Navarro, R. T. Weitz, A. M. Bittner, M. Scolari, A. Mews, M. Burghard, and K.
Kern, “Electronic Transport Properties of Individual Chemically Reduced Graphene
Oxide Sheets,” Nano Letters, vol. 7, no. 11, pp. 3499–3503, Nov. 2007.
[42] Y. Zhou, Q. Bao, L. A. L. Tang, Y. Zhong, and K. P. Loh, “Hydrothermal Dehydration for
the ‘Green’ Reduction of Exfoliated Graphene Oxide to Graphene and Demonstration
of Tunable Optical Limiting Properties,” Chemistry of Materials, vol. 21, no. 13, pp.
2950–2956, Jul. 2009.
[43] C. Chen, Q.-H. Yang, Y. Yang, W. Lv, Y. Wen, P.-X. Hou, M. Wang, and H.-M. Cheng,
“Self-Assembled Free-Standing Graphite Oxide Membrane,” Advanced Materials, vol.
21, no. 29, pp. 3007–3011, Aug. 2009.
[44] T. Szabó, O. Berkesi, and I. Dékány, “DRIFT study of deuterium-exchanged graphite
oxide,” Carbon, vol. 43, no. 15, pp. 3186–3189, Dec. 2005.
[45] D. W. Boukhvalov and M. I. Katsnelson, “Modeling of Graphite Oxide,” Journal of the
American Chemical Society, vol. 130, no. 32, pp. 10697–10701, Aug. 2008.
[46] S. Stankovich, R. D. Piner, X. Chen, N. Wu, S. T. Nguyen, and R. S. Ruoff, “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, vol. 16, no. 2, p. 155, 2006.
[47] C. Hontoria-Lucas, A. J. López-Peinado, J. d. D. López-González, M. L. Rojas-Cervantes,
and R. M. Martín-Aranda, “Study of oxygen-containing groups in a series of graphite
oxides: Physical and chemical characterization,” Carbon, vol. 33, no. 11, pp. 1585–1592,
Jan. 1995.
[48] D.-Q. Yang and E. Sacher, “Carbon 1s X-ray Photoemission Line Shape Analysis of
Highly Oriented Pyrolytic Graphite: The Influence of Structural Damage on Peak
Asymmetry,” Langmuir, vol. 22, no. 3, pp. 860–862, Jan. 2006.
[49] G. Eda and M. Chhowalla, “Chemically Derived Graphene Oxide: Towards Large-Area
Thin-Film Electronics and Optoelectronics,” Advanced Materials, vol. 22, no. 22, pp.
2392–2415, Jun. 2010.

140
[50] X. Du, I. Skachko, A. Barker, and E. Y. Andrei, “Approaching ballistic transport in
suspended graphene,” Nature Nanotechnology, vol. 3, no. 8, pp. 491–495, Jul. 2008.
[51] X. Li, G. Zhang, X. Bai, X. Sun, X. Wang, E. Wang, and H. Dai, “Highly conducting
graphene sheets and Langmuir–Blodgett films,” Nature Nanotechnology, vol. 3, no. 9,
pp. 538–542, Aug. 2008.
[52] D. V. Kosynkin, A. L. Higginbotham, A. Sinitskii, J. R. Lomeda, A. Dimiev, B. K. Price, and
J. M. Tour, “Longitudinal unzipping of carbon nanotubes to form graphene
nanoribbons,” Nature, vol. 458, no. 7240, pp. 872–876, Apr. 2009.
[53] X. Fan, W. Peng, Y. Li, X. Li, S. Wang, G. Zhang, and F. Zhang, “Deoxygenation of
Exfoliated Graphite Oxide under Alkaline Conditions: A Green Route to Graphene
Preparation,” Advanced Materials, vol. 20, no. 23, pp. 4490–4493, Dec. 2008.
[54] J. Kim, L. J. Cote, F. Kim, W. Yuan, K. R. Shull, and J. Huang, “Graphene Oxide Sheets at
Interfaces,” Journal of the American Chemical Society, vol. 132, no. 23, pp. 8180–8186,
Jun. 2010.
[55] N. Mohanty and V. Berry, “Graphene-Based Single-Bacterium Resolution Biodevice
and DNA Transistor: Interfacing Graphene Derivatives with Nanoscale and Microscale
Biocomponents,” Nano Letters, vol. 8, no. 12, pp. 4469–4476, Dec. 2008.
[56] W. Hu, C. Peng, W. Luo, M. Lv, X. Li, D. Li, Q. Huang, and C. Fan, “Graphene-Based
Antibacterial Paper,” ACS Nano, vol. 4, no. 7, pp. 4317–4323, Jul. 2010.
[57] X. Sun, Z. Liu, K. Welsher, J. T. Robinson, A. Goodwin, S. Zaric, and H. Dai,
“Nano-graphene oxide for cellular imaging and drug delivery,” Nano Research, vol. 1,
no. 3, pp. 203–212, Aug. 2008.
[58] F. Liang and B. Chen, “A Review on Biomedical Applications of Single-Walled Carbon
Nanotubes,” Current Medicinal Chemistry, vol. 17, no. 1, pp. 10–24, Jan. 2010.
[59] M. Pumera, “Nanotoxicology: The Molecular Science Point of View,” Chemistry - An
Asian Journal, vol. 6, no. 2, pp. 340–348, Feb. 2011.
[60] O. Akhavan and E. Ghaderi, “Toxicity of Graphene and Graphene Oxide Nanowalls
Against Bacteria,” ACS Nano, vol. 4, no. 10, pp. 5731–5736, Oct. 2010.
[61] Y. Zhang, S. F. Ali, E. Dervishi, Y. Xu, Z. Li, D. Casciano, and A. S. Biris, “Cytotoxicity
Effects of Graphene and Single-Wall Carbon Nanotubes in Neural
Phaeochromocytoma-Derived PC12 Cells,” ACS Nano, vol. 4, no. 6, pp. 3181–3186, Jun.
2010.
[62] P. Rivera Gil, G. Oberd rster, A. Elder, V. Puntes, and W. J. Parak, “Correlating
Physico-Chemical with Toxicological Properties of Nanoparticles: The Present and the
Future,” ACS Nano, vol. 4, no. 10, pp. 5527–5531, Oct. 2010.
[63] N. A. Monteiro-Riviere and A. O. Inman, “Challenges for assessing carbon
nanomaterial toxicity to the skin,” Carbon, vol. 44, no. 6, pp. 1070–1078, May 2006.

141
[64] J. M. Wörle-Knirsch, K. Pulskamp, and H. F. Krug, “Oops They Did It Again! Carbon
Nanotubes Hoax Scientists in Viability Assays,” Nano Letters, vol. 6, no. 6, pp.
1261–1268, Jun. 2006.
[65] M. C. Kim, G. S. Hwang, and R. S. Ruoff, “Epoxide reduction with hydrazine on
graphene: A first principles study,” The Journal of Chemical Physics, vol. 131, no. 6, p.
064704, 2009.
[66] R. R. Arvizo, O. R. Miranda, M. A. Thompson, C. M. Pabelick, R. Bhattacharya, J. D.
Robertson, V. M. Rotello, Y. S. Prakash, and P. Mukherjee, “Effect of Nanoparticle
Surface Charge at the Plasma Membrane and Beyond,” Nano Letters, vol. 10, no. 7, pp.
2543–2548, Jul. 2010.
[67] S. Liu, L. Wei, L. Hao, N. Fang, M. W. Chang, R. Xu, Y. Yang, and Y. Chen, “Sharper and
Faster ‘Nano Darts’ Kill More Bacteria: A Study of Antibacterial Activity of Individually
Dispersed Pristine Single-Walled Carbon Nanotube,” ACS Nano, vol. 3, no. 12, pp.
3891–3902, Dec. 2009.
[68] D. Li, M. B. Müller, S. Gilje, R. B. Kaner, and G. G. Wallace, “Processable aqueous
dispersions of graphene nanosheets,” Nature Nanotechnology, vol. 3, no. 2, pp.
101–105, Jan. 2008.
[69] M. A. Dobrovolskaia, J. D. Clogston, B. W. Neun, J. B. Hall, A. K. Patri, and S. E. McNeil,
“Method for Analysis of Nanoparticle Hemolytic Properties in Vitro,” Nano Letters, vol.
8, no. 8, pp. 2180–2187, Aug. 2008.
[70] Y.-S. Lin and C. L. Haynes, “Impacts of Mesoporous Silica Nanoparticle Size, Pore
Ordering, and Pore Integrity on Hemolytic Activity,” Journal of the American Chemical
Society, vol. 132, no. 13, pp. 4834–4842, Apr. 2010.
[71] D. Napierska, L. C. J. Thomassen, V. Rabolli, D. Lison, L. Gonzalez, M. Kirsch-Volders, J.
A. Martens, and P. H. Hoet, “Size-Dependent Cytotoxicity of Monodisperse Silica
Nanoparticles in Human Endothelial Cells,” Small, vol. 5, no. 7, pp. 846–853, Apr. 2009.
[72] A. Mayer, M. Vadon, B. Rinner, A. Novak, R. Wintersteiger, and E. Fr hlich, “The role
of nanoparticle size in hemocompatibility,” Toxicology, vol. 258, no. 2–3, pp. 139–147,
Apr. 2009.
[73] M. Fang, J. Long, W. Zhao, L. Wang, and G. Chen, “pH-Responsive Chitosan-Mediated
Graphene Dispersions,” Langmuir, vol. 26, no. 22, pp. 16771–16774, Nov. 2010.
[74] E. P. Marques, J. Zhang, Y.-H. Tse, R. A. Metcalfe, W. J. Pietro, and A. B. P. Lever,
“Surface electrochemistry of 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyl-2H-tetrazolium
bromide ([MTT]Br) adsorbed on a graphite electrode,” Journal of Electroanalytical
Chemistry, vol. 395, no. 1–2, pp. 133–142, Oct. 1995.

142
[75] K. M. Garza, K. F. Soto, and L. E. Murr, “Cytotoxicity and reactive oxygen species
generation from aggregated carbon and carbonaceous nanoparticulate materials,” Int J
Nanomedicine, vol. 3, no. 1, pp. 83–94, 2008.
[76] J. Jang, M. Kim, H. Jeong, and C. Shin, “Graphite oxide/poly(methyl methacrylate)
nanocomposites prepared by a novel method utilizing macroazoinitiator,” Composites
Science and Technology, vol. 69, no. 2, pp. 186–191, Feb. 2009.
[77] J. R. Potts, S. H. Lee, T. M. Alam, J. An, M. D. Stoller, R. D. Piner, and R. S. Ruoff,
“Thermomechanical properties of chemically modified graphene/poly(methyl
methacrylate) composites made by in situ polymerization,” Carbon, vol. 49, no. 8, pp.
2615–2623, Jul. 2011.
[78] H. Kim and C. W. Macosko, “Processing-property relationships of
polycarbonate/graphene composites,” Polymer, vol. 50, no. 15, pp. 3797–3809, Jul.
2009.
[79] H. Kim, Y. Miura, and C. W. Macosko, “Graphene/Polyurethane Nanocomposites for
Improved Gas Barrier and Electrical Conductivity,” Chemistry of Materials, vol. 22, no.
11, pp. 3441–3450, Jun. 2010.
[80] C. Lee, X. Wei, J. W. Kysar, and J. Hone, “Measurement of the Elastic Properties and
Intrinsic Strength of Monolayer Graphene,” Science, vol. 321, no. 5887, pp. 385–388,
Jul. 2008.
[81] M. J. McAllister, J.-L. Li, D. H. Adamson, H. C. Schniepp, A. A. Abdala, J. Liu, M.
Herrera-Alonso, D. L. Milius, R. Car, R. K. Prud’homme, and I. A. Aksay, “Single Sheet
Functionalized Graphene by Oxidation and Thermal Expansion of Graphite,” Chemistry
of Materials, vol. 19, no. 18, pp. 4396–4404, Sep. 2007.
[82] Y. Si and E. T. Samulski, “Synthesis of Water Soluble Graphene,” Nano Letters, vol. 8,
no. 6, pp. 1679–1682, Jun. 2008.
[83] G. Wang, J. Yang, J. Park, X. Gou, B. Wang, H. Liu, and J. Yao, “Facile Synthesis and
Characterization of Graphene Nanosheets,” Journal of Physical Chemistry C, vol. 112,
no. 22, pp. 8192–8195, Jun. 2008.
[84] H. Kim and C. W. Macosko, “Morphology and Properties of Polyester/Exfoliated
Graphite Nanocomposites,” Macromolecules, vol. 41, no. 9, pp. 3317–3327, May 2008.
[85] J. R. Potts, D. R. Dreyer, C. W. Bielawski, and R. S. Ruoff, “Graphene-based polymer
nanocomposites,” Polymer, vol. 52, no. 1, pp. 5–25, Jan. 2011.
[86] D. Paul, E. Morris, R. Kensler, and J. Squire, “Three-dimensional reconstruction of the
muscle thin filament. Single particle analysis of an asymmetric filament,” Microscopy
and Microanalysis, vol. 14, no. S2, Aug. 2008.
[87] “BASF Global - BASF - The Chemical Company - Corporate Website.” [Online]. Available:
http://www.basf.com/group/corporate/en/. [Accessed: 24-Aug-2012].

143
[88] K. Hsieh, J. Ueng, C. Lin, and M. Chang, “Experimental study of tuned mass damper in
reducing floor vertical vibration due to machinery,” in World Forum on Smart Materials
and Smart Structures Technology, M. Tomizuka, R. Chen, C. Yun, B. Spencer, and W.
Chen, Eds. CRC Press, 2008.
[89] B.-S. Lee, E. H.-H. Lai, K.-H. Liao, C.-Y. Lee, K.-H. Hsieh, and C.-P. Lin, “A Novel
Polyurethane-based Root Canal–obturation Material and Urethane-Acrylate–based
Root Canal Sealer—Part 2: Evaluation of Push-out Bond Strengths,” Journal of
Endodontics, vol. 34, no. 5, pp. 594–598, May 2008.
[90] J. Vermant, S. Ceccia, M. K. Dolgovskij, P. L. Maffettone, and C. W. Macosko,
“Quantifying dispersion of layered nanocomposites via melt rheology,” Journal of
Rheology, vol. 51, no. 3, p. 429, 2007.
[91] C. Gómez-Navarro, M. Burghard, and K. Kern, “Elastic Properties of Chemically
Derived Single Graphene Sheets,” Nano Letters, vol. 8, no. 7, pp. 2045–2049, Jul. 2008.
[92] J. W. Suk, R. D. Piner, J. An, and R. S. Ruoff, “Mechanical Properties of Monolayer
Graphene Oxide,” ACS Nano, vol. 4, no. 11, pp. 6557–6564, Nov. 2010.
[93] G. P. Tandon and G. J. Weng, “The effect of aspect ratio of inclusions on the elastic
properties of unidirectionally aligned composites,” Polymer Composites, vol. 5, no. 4,
pp. 327–333, Oct. 1984.
[94] M. Yoonessi and J. R. Gaier, “Highly Conductive Multifunctional Graphene
Polycarbonate Nanocomposites,” ACS Nano, vol. 4, no. 12, pp. 7211–7220, Dec. 2010.
[95] R. Prud’homme, B. Ozbas, and I. Aksay, “Patent2008045778 AI.,” 2008.
[96] H. C. Kim, D. M. Yang, S. W. Kim, and S. J. Park, “Reassessment of CT images to
improve diagnostic accuracy in patients with suspected acute appendicitis and an
equivocal preoperative CT interpretation,” European Radiology, vol. 22, no. 6, pp.
1178–1185, Dec. 2011.
[97] H. Zhang, Q. Li, Y. Zang, Y. Yang, and G. Zheng, “Centralized Genomic Control: A
Simple Approach Correcting for Population Structures,” in Encyclopedia of Statistical
Sciences, S. Kotz, C. B. Read, N. Balakrishnan, B. Vidakovic, and N. L. Johnson, Eds.
Hoboken, NJ, USA: John Wiley & Sons, Inc., 2010.
[98] S. Ansari and E. P. Giannelis, “Functionalized graphene sheet-Poly(vinylidene fluoride)
conductive nanocomposites,” Journal of Polymer Science Part B: Polymer Physics, vol.
47, no. 9, pp. 888–897, May 2009.
[99] P. Steurer, R. Wissert, R. Thomann, and R. Mülhaupt, “Functionalized Graphenes and
Thermoplastic Nanocomposites Based upon Expanded Graphite Oxide,”
Macromolecular Rapid Communications, vol. 30, no. 4–5, pp. 316–327, Feb. 2009.
[100] J.-C. Huang, “Carbon black filled conducting polymers and polymer blends,”
Advances in Polymer Technology, vol. 21, no. 4, pp. 299–313, 2002.

144
[101] M. Moniruzzaman and K. I. Winey, “Polymer Nanocomposites Containing Carbon
Nanotubes,” Macromolecules, vol. 39, no. 16, pp. 5194–5205, Aug. 2006.
[102] M. Okamoto, “Polymer/Clay Nanocomposites,” in Encyclopedia of Nanoscience and
Nanotechnology, vol. 8, Stevenson Ranch, CA: American Scientific Publishers, 2004, pp.
791–843.
[103] S. Ganguli, A. K. Roy, and D. P. Anderson, “Improved thermal conductivity for
chemically functionalized exfoliated graphite/epoxy composites,” Carbon, vol. 46, no. 5,
pp. 806–817, Apr. 2008.
[104] H. J. Salavagione, G. Martínez, and M. A. Gómez, “Synthesis of poly(vinyl
alcohol)/reduced graphite oxide nanocomposites with improved thermal and electrical
properties,” Journal of Materials Chemistry, vol. 19, no. 28, p. 5027, 2009.
[105] X. Yang, L. Li, S. Shang, and X. Tao, “Synthesis and characterization of layer-aligned
poly(vinyl alcohol)/graphene nanocomposites,” Polymer, vol. 51, no. 15, pp. 3431–3435,
Jul. 2010.
[106] I. Zaman, T. T. Phan, H.-C. Kuan, Q. Meng, L. T. Bao La, L. Luong, O. Youssf, and J. Ma,
“Epoxy/graphene platelets nanocomposites with two levels of interface strength,”
Polymer, vol. 52, no. 7, pp. 1603–1611, Mar. 2011.
[107] G. Y. Kim, M.-C. Choi, D. Lee, and C.-S. Ha, “2D-Aligned Graphene Sheets in
Transparent Polyimide/Graphene Nanocomposite Films Based on Noncovalent
Interactions Between Poly(amic acid) and Graphene Carboxylic Acid,” Macromolecular
Materials and Engineering, vol. 297, no. 4, pp. 303–311, Apr. 2012.
[108] V. H. Pham, T. V. Cuong, T. T. Dang, S. H. Hur, B.-S. Kong, E. J. Kim, E. W. Shin, and J. S.
Chung, “Superior conductive polystyrene – chemically converted graphene
nanocomposite,” Journal of Materials Chemistry, vol. 21, no. 30, p. 11312, 2011.
[109] P.-G. Ren, D.-X. Yan, T. Chen, B.-Q. Zeng, and Z.-M. Li, “Improved properties of highly
oriented graphene/polymer nanocomposites,” Journal of Applied Polymer Science, vol.
121, no. 6, pp. 3167–3174, Sep. 2011.
[110] P. Song, Z. Cao, Y. Cai, L. Zhao, Z. Fang, and S. Fu, “Fabrication of exfoliated
graphene-based polypropylene nanocomposites with enhanced mechanical and
thermal properties,” Polymer, vol. 52, no. 18, pp. 4001–4010, Aug. 2011.
[111] J. R. Potts, S. Murali, Y. Zhu, X. Zhao, and R. S. Ruoff, “Microwave-Exfoliated Graphite
Oxide/Polycarbonate Composites,” Macromolecules, vol. 44, no. 16, pp. 6488–6495,
Aug. 2011.
[112] D. R. Bortz, E. G. Heras, and I. Martin-Gullon, “Impressive Fatigue Life and Fracture
Toughness Improvements in Graphene Oxide/Epoxy Composites,” Macromolecules, vol.
45, no. 1, pp. 238–245, Jan. 2012.

145
[113] A. Bansal, H. Yang, C. Li, K. Cho, B. C. Benicewicz, S. K. Kumar, and L. S. Schadler,
“Quantitative equivalence between polymer nanocomposites and thin polymer films,”
Nature Materials, vol. 4, no. 9, pp. 693–698, Aug. 2005.
[114] C. J. Ellison and J. M. Torkelson, “The distribution of glass-transition temperatures in
nanoscopically confined glass formers,” Nature Materials, vol. 2, no. 10, pp. 695–700,
Sep. 2003.
[115] L. Schadler, “Nanocomposites: Model interfaces,” Nature Materials, vol. 6, no. 4, pp.
257–258, Apr. 2007.
[116] P. Rittigstein, R. D. Priestley, L. J. Broadbelt, and J. M. Torkelson, “Model polymer
nanocomposites provide an understanding of confinement effects in real
nanocomposites,” Nature Materials, vol. 6, no. 4, pp. 278–282, Mar. 2007.
[117] Y. Grohens, L. Hamon, G. Reiter, A. Soldera, and Y. Holl, “Some relevant parameters
affecting the glass transition of supported ultra-thin polymer films,” The European
Physical Journal E - Soft Matter, vol. 8, no. 2, pp. 217–224, May 2002.
[118] Y. Grohens, M. Brogly, C. Labbe, M.-O. David, and J. Schultz, “Glass Transition of
Stereoregular Poly(methyl methacrylate) at Interfaces,” Langmuir, vol. 14, no. 11, pp.
2929–2932, May 1998.
[119] J. L. Keddie, R. A. L. Jones, and R. A. Cory, “Interface and surface effects on the
glass-transition temperature in thin polymer films,” Faraday Discussions, vol. 98, p. 219,
1994.
[120] B. Das, K. Eswar Prasad, U. Ramamurty, and C. N. R. Rao, “Nano-indentation studies
on polymer matrix composites reinforced by few-layer graphene,” Nanotechnology, vol.
20, no. 12, p. 125705, Mar. 2009.
[121] R. Verdejo, F. Barroso-Bujans, M. A. Rodriguez-Perez, J. Antonio de Saja, and M. A.
Lopez-Manchado, “Functionalized graphene sheet filled silicone foam
nanocomposites,” Journal of Materials Chemistry, vol. 18, no. 19, p. 2221, 2008.
[122] D. A. Nguyen, Y. R. Lee, A. V. Raghu, H. M. Jeong, C. M. Shin, and B. K. Kim,
“Morphological and physical properties of a thermoplastic polyurethane reinforced
with functionalized graphene sheet,” Polymer International, vol. 58, no. 4, pp. 412–417,
Apr. 2009.
[123] T. Ramanathan, A. A. Abdala, S. Stankovich, D. A. Dikin, M. Herrera-Alonso, R. D.
Piner, D. H. Adamson, H. C. Schniepp, X. Chen, R. S. Ruoff, S. T. Nguyen, I. A. Aksay, R. K.
Prud’Homme, and L. C. Brinson, “Functionalized graphene sheets for polymer
nanocomposites,” Nature Nanotechnology, vol. 3, no. 6, pp. 327–331, May 2008.
[124] M. A. Rafiee, J. Rafiee, Z. Wang, H. Song, Z.-Z. Yu, and N. Koratkar, “Enhanced
Mechanical Properties of Nanocomposites at Low Graphene Content,” ACS Nano, vol. 3,
no. 12, pp. 3884–3890, Dec. 2009.

146
[125] L. Kan, Z. Xu, and C. Gao, “General Avenue to Individually Dispersed Graphene
Oxide-Based Two-Dimensional Molecular Brushes by Free Radical Polymerization,”
Macromolecules, vol. 44, no. 3, pp. 444–452, Feb. 2011.
[126] A. Zhamu and B. Jang, “Mass production of pristine nano graphene materials - Patent
US8226801,” 2012.
[127] A. Noro, Y. Matsushita, and T. P. Lodge, “Thermoreversible Supramacromolecular Ion
Gels via Hydrogen Bonding,” Macromolecules, vol. 41, no. 15, pp. 5839–5844, Aug.
2008.
[128] J. Wang, H. Hu, X. Wang, C. Xu, M. Zhang, and X. Shang, “Preparation and mechanical
and electrical properties of graphene nanosheets-poly(methyl methacrylate)
nanocomposites via in situ suspension polymerization,” Journal of Applied Polymer
Science, vol. 122, no. 3, pp. 1866–1871, Nov. 2011.
[129] J. Y. Jang, H. M. Jeong, and B. K. Kim, “Compatibilizing effect of graphite oxide in
graphene/PMMA nanocomposites,” Macromolecular Research, vol. 17, no. 8, pp.
626–629, Aug. 2009.
[130] P. Rittigstein and J. M. Torkelson, “Polymer–nanoparticle interfacial interactions in
polymer nanocomposites: Confinement effects on glass transition temperature and
suppression of physical aging,” Journal of Polymer Science Part B: Polymer Physics, vol.
44, no. 20, pp. 2935–2943, Oct. 2006.
[131] M. Fang, K. Wang, H. Lu, Y. Yang, and S. Nutt, “Single-layer graphene nanosheets
with controlled grafting of polymer chains,” Journal of Materials Chemistry, vol. 20, no.
10, p. 1982, 2010.
[132] T. Wallin, “Mechanical Properties of POlymer Nanocomposites Based on
Functionalized Graphene Sheets,” College of William and Mary, Virginia, 2010.
[133] Y. Wang, Z. Shi, J. Fang, H. Xu, X. Ma, and J. Yin, “Direct exfoliation of graphene in
methanesulfonic acid and facile synthesis of graphene/polybenzimidazole
nanocomposites,” Journal of Materials Chemistry, vol. 21, no. 2, p. 505, 2011.
[134] V. Alzari, D. Nuvoli, S. Scognamillo, M. Piccinini, E. Gioffredi, G. Malucelli, S.
Marceddu, M. Sechi, V. Sanna, and A. Mariani, “Graphene-containing
thermoresponsive nanocomposite hydrogels of poly(N-isopropylacrylamide) prepared
by frontal polymerization,” Journal of Materials Chemistry, vol. 21, no. 24, p. 8727,
2011.
[135] X. Yang, Y. Tu, L. Li, S. Shang, and X. Tao, “Well-Dispersed Chitosan/Graphene Oxide
Nanocomposites,” ACS Applied Materials & Interfaces, vol. 2, no. 6, pp. 1707–1713, Jun.
2010.

147
[136] A. Leszczynska and K. Pielichowski, “Application of thermal analysis methods for
characterization of polymer/montmorillonite nanocomposites,” Journal of Thermal
Analysis and Calorimetry, vol. 93, no. 3, pp. 677–687, Oct. 2008.
[137] Z. Yao, N. Braidy, G. A. Botton, and A. Adronov, “Polymerization from the Surface of
Single-Walled Carbon Nanotubes − Preparation and Characterization of
Nanocomposites,” Journal of the American Chemical Society, vol. 125, no. 51, pp.
16015–16024, Dec. 2003.
[138] X. Wang, H. Yang, L. Song, Y. Hu, W. Xing, and H. Lu, “Morphology, mechanical and
thermal properties of graphene-reinforced poly(butylene succinate) nanocomposites,”
Composites Science and Technology, vol. 72, no. 1, pp. 1–6, Dec. 2011.
[139] S. Kim, I. Do, and L. T. Drzal, “Thermal stability and dynamic mechanical behavior of
exfoliated graphite nanoplatelets-LLDPE nanocomposites,” Polymer Composites, vol. 31,
no. 5, pp. 755–761, Jan. 2009.
[140] J. Yu, P. Jiang, C. Wu, L. Wang, and X. Wu, “Graphene nanocomposites based on
poly(vinylidene fluoride): Structure and properties,” Polymer Composites, vol. 32, no.
10, pp. 1483–1491, Oct. 2011.
[141] G. Vleminckx, S. Bose, J. Leys, J. Vermant, M. Wübbenhorst, A. A. Abdala, C.
Macosko, and P. Moldenaers, “Effect of Thermally Reduced Graphene Sheets on the
Phase Behavior, Morphology, and Electrical Conductivity in Poly[(α-methyl
styrene)-co-(acrylonitrile)/poly(methyl-methacrylate) Blends,” ACS Applied Materials &
Interfaces, vol. 3, no. 8, pp. 3172–3180, Aug. 2011.
[142] H. Lian, S. Li, K. Liu, L. Xu, K. Wang, and W. Guo, “Study on modified graphene/butyl
rubber nanocomposites. I. Preparation and characterization,” Polymer Engineering &
Science, vol. 51, no. 11, pp. 2254–2260, Nov. 2011.
[143] L. Seveyrat, A. Chalkha, D. Guyomar, and L. Lebrun, “Preparation of graphene
nanoflakes/polymer composites and their performances for actuation and energy
harvesting applications,” Journal of Applied Physics, vol. 111, no. 10, p. 104904, 2012.
[144] A. V. Raghu, Y. R. Lee, H. M. Jeong, and C. M. Shin, “Preparation and Physical
Properties of Waterborne Polyurethane/Functionalized Graphene Sheet
Nanocomposites,” Macromolecular Chemistry and Physics, vol. 209, no. 24, pp.
2487–2493, Dec. 2008.
[145] D. Cai, K. Yusoh, and M. Song, “The mechanical properties and morphology of a
graphite oxide nanoplatelet/polyurethane composite,” Nanotechnology, vol. 20, no. 8,
p. 085712, Feb. 2009.
[146] Z. Cao, P. Song, Z. Fang, Y. Xu, Y. Zhang, and Z. Guo, “Physical wrapping of reduced
graphene oxide sheets by polyethylene wax and its modification on the mechanical
properties of polyethylene,” Journal of Applied Polymer Science, p. n/a–n/a, 2012.

148
[147] S. Chatterjee, F. A. Nüesch, and B. T. T. Chu, “Comparing carbon nanotubes and
graphene nanoplatelets as reinforcements in polyamide 12 composites,”
Nanotechnology, vol. 22, no. 27, p. 275714, Jul. 2011.
[148] M. Li and Y. G. Jeong, “Poly(ethylene terephthalate)/exfoliated graphite
nanocomposites with improved thermal stability, mechanical and electrical properties,”
Composites Part A: Applied Science and Manufacturing, vol. 42, no. 5, pp. 560–566,
May 2011.
[149] M. Li and Y. G. Jeong, “Preparation and Characterization of High-Performance
Poly(trimethylene terephthalate) Nanocomposites Reinforced with Exfoliated
Graphite,” Macromolecular Materials and Engineering, vol. 296, no. 2, pp. 159–167,
Feb. 2011.
[150] M. Sangermano, S. Marchi, L. Valentini, S. B. Bon, and P. Fabbri, “Transparent and
Conductive Graphene Oxide/Poly(ethylene glycol) diacrylate Coatings Obtained by
Photopolymerization,” Macromolecular Materials and Engineering, vol. 296, no. 5, pp.
401–407, May 2011.
[151] A. Zandiatashbar, R. C. Picu, and N. Koratkar, “Mechanical Behavior of
Epoxy-Graphene Platelets Nanocomposites,” Journal of Engineering Materials and
Technology, vol. 134, no. 3, p. 031011, 2012.
[152] M. A. Rafiee, J. Rafiee, I. Srivastava, Z. Wang, H. Song, Z.-Z. Yu, and N. Koratkar,
“Fracture and Fatigue in Graphene Nanocomposites,” Small, vol. 6, no. 2, pp. 179–183,
Jan. 2010.
[153] M. Martin-Gallego, R. Verdejo, M. A. Lopez-Manchado, and M. Sangermano,
“Epoxy-Graphene UV-cured nanocomposites,” Polymer, vol. 52, no. 21, pp. 4664–4669,
Sep. 2011.
[154] I. Zaman, H.-C. Kuan, J. Dai, N. Kawashima, A. Michelmore, A. Sovi, S. Dong, L. Luong,
and J. Ma, “From carbon nanotubes and silicate layers to graphene platelets for
polymer nanocomposites,” Nanoscale, 2012.
[155] M.-C. Hsiao, S.-H. Liao, M.-Y. Yen, C.-C. Teng, S.-H. Lee, N.-W. Pu, C.-A. Wang, Y. Sung,
M.-D. Ger, C.-C. M. Ma, and M.-H. Hsiao, “Preparation and properties of a graphene
reinforced nanocomposite conducting plate,” Journal of Materials Chemistry, vol. 20,
no. 39, p. 8496, 2010.
[156] S. H. Yoon, J. H. Park, E. Y. Kim, and B. K. Kim, “Preparations and properties of
waterborne polyurethane/allyl isocyanated-modified graphene oxide
nanocomposites,” Colloid and Polymer Science, vol. 289, no. 17–18, pp. 1809–1814, Sep.
2011.

149
[157] K.-H. Liao, Y. Qian, and C. W. Macosko, “Ultralow percolation
graphene/polyurethane acrylate nanocomposites,” Polymer, vol. 53, no. 17, pp.
3756–3761, Aug. 2012.
[158] Z. Cai, L. Liu, Z. Zheng, and X. Wang, “Poly(urethane-co-vinyl imidazole)/graphene
nanocomposites,” Polymer Composites, vol. 33, no. 4, pp. 459–466, Apr. 2012.
[159] H. Hu, X. Wang, J. Wang, L. Wan, F. Liu, H. Zheng, R. Chen, and C. Xu, “Preparation
and properties of graphene nanosheets–polystyrene nanocomposites via in situ
emulsion polymerization,” Chemical Physics Letters, vol. 484, no. 4–6, pp. 247–253, Jan.
2010.
[160] M. M. Bernal, I. Molenberg, S. Estravis, M. A. Rodriguez-Perez, I. Huynen, M. A.
Lopez-Manchado, and R. Verdejo, “Comparing the effect of carbon-based nanofillers on
the physical properties of flexible polyurethane foams,” Journal of Materials Science,
vol. 47, no. 15, pp. 5673–5679, Mar. 2012.
[161] D. Yan, L. Xu, C. Chen, J. Tang, X. Ji, and Z. Li, “Enhanced mechanical and thermal
properties of rigid polyurethane foam composites containing graphene nanosheets and
carbon nanotubes,” Polymer International, vol. 61, no. 7, pp. 1107–1114, Jul. 2012.
[162] J. M. Bak, T. Lee, E. Seo, Y. Lee, H. M. Jeong, B.-S. Kim, and H. Lee,
“Thermoresponsive graphene nanosheets by functionalization with polymer brushes,”
Polymer, vol. 53, no. 2, pp. 316–323, Jan. 2012.
[163] R. K. Layek, S. Samanta, D. P. Chatterjee, and A. K. Nandi, “Physical and mechanical
properties of poly(methyl methacrylate) -functionalized graphene/poly(vinylidine
fluoride) nanocomposites: Piezoelectric β polymorph formation,” Polymer, vol. 51, no.
24, pp. 5846–5856, Nov. 2010.
[164] I. Zaman, H.-C. Kuan, Q. Meng, A. Michelmore, N. Kawashima, T. Pitt, L. Zhang, S.
Gouda, L. Luong, and J. Ma, “A Facile Approach to Chemically Modified Graphene and
its Polymer Nanocomposites,” Advanced Functional Materials, p. n/a–n/a, 2012.
[165] K. P. Pramoda, H. Hussain, H. M. Koh, H. R. Tan, and C. B. He, “Covalent bonded
polymer-graphene nanocomposites,” Journal of Polymer Science Part A: Polymer
Chemistry, vol. 48, no. 19, pp. 4262–4267, Aug. 2010.
[166] M. Fang, K. Wang, H. Lu, Y. Yang, and S. Nutt, “Covalent polymer functionalization of
graphene nanosheets and mechanical properties of composites,” Journal of Materials
Chemistry, vol. 19, no. 38, p. 7098, 2009.
[167] C. Heo, H.-G. Moon, C. S. Yoon, and J.-H. Chang, “ABS nanocomposite films based on
functionalized-graphene sheets,” Journal of Applied Polymer Science, p. n/a–n/a, 2011.
[168] S. Mohamadi, N. Sharifi-Sanjani, and H. Mahdavi, “Functionalization of Graphene
Sheets via Chemically Grafting of PMMA Chains Through in-situ Polymerization,” -

150
Journal of Macromolecular Science, Part A: Pure and Applied Chemistry, vol. 48, p. 577,
2010.
[169] H. J. Salavagione and G. Martínez, “Importance of Covalent Linkages in the
Preparation of Effective Reduced Graphene Oxide−Poly(vinyl chloride)
Nanocomposites,” Macromolecules, vol. 44, no. 8, pp. 2685–2692, Apr. 2011.
[170] L. Jiang, X.-P. Shen, J.-L. Wu, and K.-C. Shen, “Preparation and characterization of
graphene/poly(vinyl alcohol) nanocomposites,” Journal of Applied Polymer Science, p.
n/a–n/a, 2010.
[171] H.-D. Huang, P.-G. Ren, J. Chen, W.-Q. Zhang, X. Ji, and Z.-M. Li, “High barrier
graphene oxide nanosheet/poly(vinyl alcohol) nanocomposite films,” Journal of
Membrane Science, vol. 409–410, pp. 156–163, Aug. 2012.
[172] J. Liang, Y. Huang, L. Zhang, Y. Wang, Y. Ma, T. Guo, and Y. Chen, “Molecular-Level
Dispersion of Graphene into Poly(vinyl alcohol) and Effective Reinforcement of their
Nanocomposites,” Advanced Functional Materials, vol. 19, no. 14, pp. 2297–2302, Jul.
2009.
[173] C. Wan, M. Frydrych, and B. Chen, “Strong and bioactive gelatin–graphene oxide
nanocomposites,” Soft Matter, vol. 7, no. 13, p. 6159, 2011.
[174] H. B. Lee, A. Raghu, K. S. Yoon, and H. M. Jeong, “Preparation and Characterization of
Poly(ethylene oxide)/Graphene Nanocomposites from an Aqueous Medium,” Journal of
Macromolecular Science, Part B, vol. 49, no. 4, pp. 802–809, Jul. 2010.
[175] D. A. Dikin, S. Stankovich, E. J. Zimney, R. D. Piner, G. H. B. Dommett, G. Evmenenko,
S. T. Nguyen, and R. S. Ruoff, “Preparation and characterization of graphene oxide
paper,” Nature, vol. 448, no. 7152, pp. 457–460, Jul. 2007.
[176] I. W. Frank, D. M. Tanenbaum, A. M. van der Zande, and P. L. McEuen, “Mechanical
properties of suspended graphene sheets,” Journal of Vacuum Science & Technology B:
Microelectronics and Nanometer Structures, vol. 25, no. 6, p. 2558, 2007.
[177] A. Incze, A. Pasturel, and P. Peyla, “Mechanical properties of graphite oxides: Ab
initio simulations and continuum theory,” Physical Review B, vol. 70, no. 21, Dec. 2004.
[178] R. K. Prud’homme, I. A. Aksay, D. Adamson, and A. Abdala, “Thermally Exfoliated
Graphite Oxide. U.S. Patent 2010, 7, 658,901.” [Online]. Available:
http://scholar.google.com/scholar?q=U.S.+Patent+2010%2C+7%2C+658%2C901.&btnG
=&hl=en&as_sdt=0%2C24. [Accessed: 06-Jul-2012].
[179] B. A. Miller-Chou and J. L. Koenig, “A review of polymer dissolution,” Progress in
Polymer Science, vol. 28, no. 8, pp. 1223–1270, Aug. 2003.
[180] M. Dion, A. B. Larson, and B. D. Vogt, “Impact of low-molecular mass components
(oligomers) on the glass transition in thin films of poly(methyl methacrylate),” Journal
of Polymer Science Part B: Polymer Physics, vol. 48, no. 22, pp. 2366–2370, Nov. 2010.

151
[181] C.-S. Chern, Principles and Applications of Emulsion Polymerization. Hoboken, NJ,
USA: John Wiley & Sons, Inc., 2008.
[182] L. M. Gan, C. H. Chew, K. C. Lee, and S. C. Ng, “Polymerization of methyl methacrylate
in ternary oil-in-water microemulsions,” Polymer, vol. 34, no. 14, pp. 3064–3069, Jan.
1993.
[183] S. C. Pilcher and W. T. Ford, “Structures and Properties of Poly(methyl methacrylate)
Latexes Formed in Microemulsions,” Macromolecules, vol. 31, no. 11, pp. 3454–3460,
Jun. 1998.
[184] R. Tang, W. Yang, L. Zha, and S. Fu, “The Relationship of Reaction Temperature, Tg
and Rich-Syndiotacticity of Poly(alkyl methacrylate)s in Modified Microemulsion
Polymerization,” Journal of Macromolecular Science, Part A, vol. 45, no. 5, pp. 345–352,
Mar. 2008.
[185] I. M. Kalogeras, F. Pallikari, A. Vassilikou-Dova, and E. R. Neagu, “Thermal analysis
studies of doping effects on the conformational motions of polymer chains in solid
solutions with lasing molecules,” Journal of Applied Physics, vol. 101, no. 9, p. 094108,
2007.
[186] J. Brandrup, E. H. Immergut, E. A. Grulke, A. Abe, and D. R. Bloch, Polymer Handbook,
4th edition. New York: Wiley, 1999.
[187] T. Kitayama, Y. Zhang, and K. Hatada, “Preparation of heterotactic-rich poly(methyl
methacrylate) with narrow molecular weight distribution
bytert-butyllithium/bis(2,6-di-tert-butylphenoxy)methylaluminum,” Polymer Bulletin,
vol. 32, no. 4, pp. 439–446, Apr. 1994.
[188] E. V. Thompson, Journal of polymer science. Part A-2: Polymer physics, vol. 4. New
York: Interscience Publishers., 1996.

152
Appendix A

Glass Transition Temperature of PMMA with

Incorporation of Graphene via Solvent Blending4

A-1 Introduction

Recent studies reported that TRG can enhance 32~92% of tensile modulus of

glassy polymers with 1.9~3.1 vol% load using solvent-blending.[98], [99] As

described in Chapter 6, the letter by Ramanathan et al.[123] reported an unprecedented

shift in glass transition temperature (Tg) of 29 C for PMMA at 0.05 wt% of TRG. A

33% increase in modulus with 0.01 wt% TRG loading was also reported. We have not

been able to find such large increases in modulus in other published graphene/polymer

nanocomposites (Chapter 6).[4] This value is higher than the Voigt upper bound

prediction, [4] which is estimated by using fully aligned particles with infinite aspect

ratio. The authors attributed their very high modulus value to strong hydrogen bonding

interactions of hydroxyl groups on the TRG surface with the carbonyl groups of

PMMA and the restriction of PMMA segment mobilities by mechanical interlocking

of the wrinkled surface of TRG. In order to understand the reasons for these surprising

results, we repeated the experiments reported in the letter. The PMMA used here was

purchased from the same source, Polysciences (atactic beads, Mw = 350,000, PDI =

2.7, lot. 528266). TRG was obtained from Vorbeck Materials, which uses the same

thermal method to make TRG.[34], [81], [178]

4
This chapter was submitted as a Correspondence to Nature Nanotechnology 2011 as "Functionalized Graphene Sheets have a

Small Influence on the Glass Transition of Polymer Nanocomposites." with authors: Liao K-H, Kobayashi S, Kim H, Abdala A,

Macosko C.
153
A-2 Results and Discussion

Glass transition temperatures (Tg) of PMMA as-received (R-PMMA), PMMA as

precipitated (PMMA; the same material as a-PMMA in the main article) and

TRG/PMMA nanocomposites (TRG/PMMA-NC) with various TRG loadings were

determined using DSC and DMA [Figure A-1(a)]. Different from the results in the

letter, the Tg of PMMA does not change with TRG loading ~125 C. However, the Tg

of R-PMMA is ~105 C by DSC and ~95 C by DMA (Figure A-1) which is consistent

with the values reported for the control sample in the letter. These results suggest

that the Tg value of PMMA in Ramanathan et al.'s[123] report are from R-PMMA

instead of PMMA. The tensile modulus of R-PMMA, PMMA and TRG/PMMA-NC

in Figure S2 did increase with TRG loading as Ramanathan et al. reported. [123]

However, the modulus of TRG/PMMA-NC with 1 wt% TRG (~2.2 GPa) increased

only ~30% from PMMA (~1.8 GPa), significantly less than the 80% increment stated

in Ramanathan et al.'s report. [123]

Ramanathan et al. [123] state that the increase in Tg resulted from: (1)

interlocking of the polymer chains by the nanoscale surface roughness of TRG, and (2)

hydrogen bonds between the carbonyl group in PMMA and the hydroxyl group on the

TRG surface. If the first explanation of the Tg increase stated above is true, we should

expect the DSC curves will show another Tg at higher temperature since the volume

fraction of TRG is so low. PMMA coils close to the TRG surface could have another

higher Tg, while coils further from TRG should have the Tg of neat PMMA. However,

the DSC traces in Figure A-1(b) have no second Tg from 130 C to 200 C.

According to the results above, we found the Tg of PMMA is ~20 C higher than

R-PMMA. In order to understand the reason for the discrepancy between the two Tgs,

we analyzed the remaining filtrate from the precipitation process described in


154
Ramanathan et al.[123] and obtained a solid extract (EXC) by concentration of the

filtrate. The EXC is a pale yellowish solid which foamed on shaking in water and is

~4 wt% of the total R-PMMA. GPC chromatograms of EXC, PMMA and R-PMMA

in Figure A-3 indicate that the weight average molecular weight (MW) of PMMA

(MW=436,000 g/mol; PDI=2.92) is slightly higher than R-PMMA (MW=398,000 g/mol;

PDI=2.93). GPC indicates that EXC is mainly composed of (1) structures with very

low MW (~500 g/mol), and (2) polymer (or oligomer) also with much lower MW

(MW=7,000 g/mol; PDI=1.25) than both PMMA and R-PMMA. It was reported that

PMMA with a molecular weight lower than 10 kg/mol is soluble in methanol.[179],

[180]

A significant amount of additives are generally used in suspension and

emulsion polymerization. Surfactants are used to stabilize MMA droplets in the

emulsion during the polymerization process.[181–184] To obtain pure PMMA the

surfactant has to be removed by washing the product (typically using methanol).

[182–184] The inset in Figure A-3 shows that the as-received PMMA is small

spherical beads indicating an emulsion polymerization product. We can reasonably

speculate that the very low molecular weight structures in EXC are (1) surfactant for

emulsion polymerization of MMA and (2) PMMA oligomer formed during the

polymerization process. These two compounds were extracted using THF/methanol

procedure described in the letter. [123] The low molecular weight additives can act

as plasticizers and decrease the Tg of PMMA.[180], [185] Hence, the removal of

PMMA oligomer and additives should be the reason for the ~20 C increase in Tg

between R-PMMA and PMMA.

To further characterize R-PMMA and the precipitation process we used NMR.


1
H NMR results (Figure A-4) show that the spectrum of R-PMMA [Figure A-4(a)] has
155
nonnegligible minor signals compared to PMMA [Figure A-4(b)], and part of these

peaks are from the remaining MMA monomer [blue arrowed in Figure A-4(a)].

Comparison of 1H NMR spectrum of EXC [Figure A-4(c)] with R-PMMA [Figure

A-4(a)] shows that the extra signals in R-PMMA are from EXC. The chemical

structure of EXC contains −O−CnH2n+1 [n=18-19, e.g., stearyl alcohol (n=18), is a

surfactant], mono-substituted benzene and small amount of PMMA oligomer. These


1
H NMR results agree with the speculation of the EXC composition based on GPC

data. The tacticity of PMMA was estimated from 1H NMR [Figure A-4(b)]. The diad

sequences, mm(isotactic):mr(heterotactic):rr(syndiotactic) = 6:38:56, show that

PMMA is highly syndiotactic, a result expected for emulsion polymerization. This

explains why the Tg of PMMA is ~20 C higher than the commonly reported value for

atactic PMMA (Tg≈105 C).[186] The Tg for syndiotactic rich PMMA is ≈ 115~125

C.[183], [184], [186–188] The difference in tensile modulus may also due to the

plasticizing effect of the small molecules.

According to the Kim et al.'s review,[4] all the studies of graphene/polymer

nanocomposites reported percolation concentrations of graphene from 0.2 to 6.2 wt%,

which are at least three times higher than the reported 0.05 wt% in Ramanathan et al.'s

report.[123] Our surface resistance measurements in Figure A-2 indicate the

percolation concentration is ~0.8 wt%.

A-3 Conclusions

This study indicates that the significant increment of Tg reported by Ramanathan

et al.[123]. did not resulted from the addition of TRG. Instead, the improvement of Tg

is because the ~4 wt% additive (emulsifier) in the original PMMA had been extracted

during the precipitation process. An 80% increase of tensile modulus of PMMA is


156
contributed by both addition of TRG and removal of emulsifier, and only ~30%

improvement of tensile modulus can be obtained by loading 1 wt% TRG into PMMA.

Figure A-1. (a) Glass transition temperature (Tg) and (b) DSC traces of
as-received PMMA (R-PMMA), as-precipitated (PMMA) and all TRG/PMMA
nanocomposites.

157
Figure A-2. Surface resistance and tensile modulus of as-received PMMA
(R-PMMA) as-precipitated PMMA (PMMA) and TRG/PMMA nanocomposites.

Figure A-3. GPC chromatograms of as-received PMMA (R-PMMA),


as-precipitated (PMMA) and extractive content (EXC). Inset: optical micrograph
of R-PMMA.

158
Figure A-4. 1H NMR spectra of (a) as-received PMMA (R-PMMA), (b)
as-precipitated (PMMA), and (c) extract (EXC). Signals from MMA (blue) and
EXC (red) are arrowed in figures (a) and (b). Expanded spectra with the portion
of EXC signals are also added as insets to clarify the removal of EXC.

159

You might also like