You are on page 1of 21

Accepted Manuscript

Mechanical, Electrical and Thermal Properties of In-situ Exfoliated Graphene/


Epoxy Nanocomposites

Yan Li, Han Zhang, Harshit Porwal, Zhaohui Huang, Emiliano Bilotti, Ton Peijs

PII: S1359-835X(17)30020-9
DOI: http://dx.doi.org/10.1016/j.compositesa.2017.01.007
Reference: JCOMA 4542

To appear in: Composites: Part A

Received Date: 14 October 2016


Revised Date: 5 January 2017
Accepted Date: 7 January 2017

Please cite this article as: Li, Y., Zhang, H., Porwal, H., Huang, Z., Bilotti, E., Peijs, T., Mechanical, Electrical and
Thermal Properties of In-situ Exfoliated Graphene/Epoxy Nanocomposites, Composites: Part A (2017), doi: http://
dx.doi.org/10.1016/j.compositesa.2017.01.007

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Mechanical, Electrical and Thermal Properties of In-situ
Exfoliated Graphene/Epoxy Nanocomposites

Yan Li 1, Han Zhang 1,2, Harshit Porwal 1,2 , Zhaohui Huang 3, Emiliano Bilotti 1,2*, Ton Peijs 1,2*

1
School of Engineering and Materials Science, Queen Mary University of London, Mile End Road,
E1 4NS, London, UK
2
Nanoforce Technology Ltd., Mile End Road, E1 4NS London, UK
3
School of Materials Science and Technology, China University of Geosciences (Beijing), 100083,
P. R. China.

*Corresponding authors:
E-mail addresses: t.peijs@qmul.ac.uk, e.bilotti@qmul.ac.uk

Keywords: In-situ exfoliation, few layer graphene, graphite nanoplatelets, epoxy resin

ABSTRACT

Depending on processing conditions, in-situ exfoliated graphite nanoplatelets (GNP) with low
defect content and average aspect ratios up to 300-1000 and thicknesses of 5-17 nm could be
produced by three roll milling (TRM). This paper focuses on the mechanical, electrical and thermal
properties of in-situ GNP/epoxy nanocomposites, evaluated in terms of simple analytical models.
Good mechanical reinforcement (160% increase in flexural modulus @ 4 wt.% GNP), electrical
conductivity (~10-2 S/m @ 3 wt.% GNP with a percolation threshold of 0.52 vol.%) and thermal
conductivity (0.70 W·m-1·K-1 @ 5 wt.% GNP) were obtained. The production of GNP-filled resins
using TRM technology can potentially remove important cost barriers for GNP modified plastics,
composites and coatings as compared to traditional multi-step solvent based exfoliation methods.

1 INTRODUCTION

Interest in graphene and its potential applications has grown rapidly since 2004, when the
material was first isolated [1, 2]. Graphene is a one-atom thick planar sheet and an emerging class
of nanomaterials with remarkable properties, such as, high thermal conductivity (∼5000 W·m-1·K-1)
[3], electrical conductivity (108 S·m-1) [4], high intrinsic mobility (2×105 cm2·s-1·v-1) [5], extremely
high tensile strength (~130 GPa), Young’s modulus (∼1.0 TPa) [6], and large specific surface area

1
(2360 m2·g-1) [7], all of which giving it the potential for replacing commercial fillers currently in use
for the fabrication of polymer composites and coatings. Single-layer graphene possesses the best
intrinsic properties, however, it is still difficult to obtain in large-scale and low cost. Moreover its
tendency to roll, scroll, fold or wrinkle [8], unless constrained onto a solid surface, is of great
concern as only a high aspect ratio, flat filler morphology does theoretically offer optimal physical
property (e.g. mechanical, electrical, thermal, gas barrier, corrosion resistance, etc.) enhancement
in polymer materices. Few layer graphene (FLG) and/or graphite nanoplatelets (GNP) might be a
more viable alternative to single layer graphene as a promising future nanofiller for polymer
composites and coatings, if produced with good quality, cheaply and in bulk. GNP has already been
shown to provide significant improvements in mechanical properties like stiffness, strength and
surface hardness [9-12]. However, the above mentioned properties strongly depend on the
number of layers stacked in the GNP, the degree of crystallinity in graphitic planes, their aspect
ratio (AR) and the order of stacking [13]. Fabrication of GNP/epoxy composites is also an
important aspect to take into consideration as it typically involves a multi-step process starting
from the production of GNP suspensions in various solvents followed by mixing with polymers [14]
and solvents evaporation. Even the best production methods for graphene/FLG/GNP (e.g. liquid
phase exfoliation) require large quantities of organic solvents [15, 16], resulting in difficulties in
transfering this technology to industrial scale applications due to costs and environmental impact
concerns. Coleman and coworkers [16, 17] produced graphene dispersions with concentrations up
to 63 mg/ml, by exfoliation of graphite in organic solvents such as N-methyl-pyrrolidone (NMP).
Similarly O’Neill et al. [18] were able to identify low boiling point solvents and optimize conditions
for stable graphene dispersions in chloroform, acetone and isopropanol. Depending on the
preparation conditions, which include long-term and low-energy bath sonication and
centrifugation, dispersions with a graphene concentration of 0.5 mg/ml could be obtained. All
processes described above use either high/low boiling point solvents or surfactants to successfully
exfoliate graphite into graphene sheets [15, 16], which can be detrimental to processing
conditions or final properties of the nanocomposites. The yield of GNP production using the above
described methods is very low, typically 1 to 3%. Therefore, there is a need for technologies that
are able to produce GNPs at an industrial scale. Recent developments have allowed graphene and
GNP to be produced in larger quantities, for example, XG Science Inc. commercialized expanded
graphite nanoplatelets, while Coleman’s group recently reported scalable production of large
quantities of defect-free few layer graphene by high-shear mixing of graphite in suitable stabilising
liquids [19]. However, these processes remain specialised and resource intensive and this

2
continues to prohibit the use of graphene in many industries. Moreover, it should be noted that
even ultrasonication of GNPs in organic solvent for a long period of time does not guarantee good
exfoliation and dispersion especially at higher concentrations. On the other hand, GNPs easily
restack during drying, and form a layered material consisting of graphene sheets that are bonded
together by van der Waals forces. To ensure a good dispersion of GNPs in polymer matrices,
clearly restacking of the GNPs must be avoided.

Direct in-situ exfoliation/dispersion of graphene in epoxy or other polymeric resins promises to


tackle the above drawbacks leading to a viable route to applications of GNP modified resins.
Recently our group demonstrated a high throughput, single-step, top-down process able to in-situ
exfoliate graphite to FLG/GNP in liquid epoxy via TRM. This method addresses the challenges of
creating well exfoliated and dispersed GNPs at high yield without the use of solvents and/or
compatibilisers. The objective of this work is to study in more details the mechanical, electrical
and thermal properties of these in-situ exfoliated GNP/epoxy nanocomposites, and to discuss and
interpret these properties using simple analytical models.

2 EXPERIMENTAL

2.1 Materials

Natural graphite flakes (NG) were purchased from Alfa Aesar (Product No. 43319). The MVR444
two-part epoxy (EP) resin (aeronautic grade) was kindly supplied by Cytec (UK). The epoxy resin
(EP) was an aeronautical grade epoxy resin (MVR444), based on a resin (MVR444R) and hardener
(MVR444H), kindly supplied by Cytec Ltd. (UK).

2.2 Exfoliation and Dispersion Method

GNP/epoxy mixtures were prepared in analogy to previous work [20], using a three-roll mill (TRM)
(80E EXAKT GmbH, Germany). In a typical experiment, the epoxy/NG mixture was passed 8
consecutive times through the TRM (referred to as 8 cycles), with the same rotational speed (200
rpm) of the apron roll. The first two cycles were in gap mode (fixed roll-to-roll distance) while the
following six cycles were in force mode (fixed applied force of 5 N/mm). Every cycle was
performed at the fixed speed ratio of 1:3:9 between N1 (feed roller) : N2 (central roller) : N3 (apron
roller). During the first cycle in gap mode, gaps of N 1/N2 = 120 µm and N2/N3 = 40 µm were used.
For the second cycle, N1/N2 and N2/N3 were reduced to 60 and 20 µm, respectively. Temperture

3
was varied between room temperature (RT) and 40 °C (25, 30, 35, 40 °C). By fine tuning the resin
temperature during TRM the dispersability of graphene in epoxy can be further improved as the
surface energy of epoxy can optimised to match that of graphite, which in turn assists the in-situ
exfoliation and dispersion process. For more details the reader is referred to our previous
publication [20].

2.3 Fabrication of GNP/Epoxy Composites

The hardener was added to the GNP/epoxy mixture in a 58:100 weight ratio. This mixture was
degassed in a vacuum chamber (pressure of -1 bar) at 70 oC for 60 min while mechanical stirring
was applied throughout the process. Mixtures were then poured into stainless steel moulds at
room temperature (RT) and cured in an oven. The following curing conditions were applied: i)
temperature was ramped from RT to 120 oC (3 oC·min-1) followed by a 90 min isotherm, ii)
temperature was ramped from 120 oC to 180 oC at (3 oC·min-1) followed by a 3 h isotherm and iii)
cooling down from 180 oC to RT (3 oC·min-1). The curing process followed was the one
recommended by the supplier (Cytec, UK).

2.4 Characterisation Techniques

Morphological studies were carried out using optical microscopy (OM, Olympus BX60,Japan) and
scanning electron microscopy (SEM) (FEI, Inspector-F, Netherlands) with an acceleration voltage of
20 kV. The morphology of the cured GNP/epoxy composites after gold coating was investigated by
imaging cryo-fractured cross-sectional surfaces.

Transmission electron microscopy (TEM, JOEL JEM-2010, Japan) was used to investigate the
dispersion quality and the exfoliation of the filler in the composites, 60 nm thin slices were cut
from the cured composites using an ultra-microtome, and then used for TEM imaging (for details
see Supplementary Information). TEM was also used to characterize the morphology and different
degrees of the exfoliated graphene. Please refer to our previous publication for more information
[20, 21]. In TEM characterization, effectively the edge of the GNP particles is measured which can
be positioned out-of-plane (‘curled up’) even when deposited in-plane on a flat surface. The
thickness of the GNP particles is estimated from the thickness of the edge of the particles sticking
out of the plane of the copper grid. Data from 10−15 representative particles were collected for
each specimen.

4
Atomic force microscopy (AFM, NT-MDT NTegra, Russia) was also used to study the morphology of
graphene flakes. AFM probes (Point probe-plus silicon – SPM sensor, Type: PPP-FMR-10, Force
constant 0.5-9.8 N/m) were purchased from NanosensorsTM. Specimens were prepared by
“extracting” the particles from different dispersions straight after TRM. Typically, a small amount
of uncured epoxy/particle dispersion was immersed in acetone to dissolve the epoxy, followed by
filtration using a 0.2 μm membrane. The obtained particles were then washed three times to
remove any remaining epoxy, which was confirmed by optical microscopy and FTIR analysis.
These particles were re-dispersed in acetone to a final concentration less than 0.5 mg∙ml-1. The
AFM samples were prepared by drop casting a diluted GNP/acetone dispersion on a mica
substrate. After evaporation of acetone, the specimens were scanned as prepared.

X-ray diffraction (XRD) (Philips PW 3830, Netherland) was used to characterize the thickness of
exfoliated GNPs within the cured epoxy matrix. The samples were cut into rectangular beams
(dimensions 3.2 x 10 x 30 mm3). The average out-of-plane crystallite thickness of the GNPs (t) was
estimated using the full width at half maximum (FWHM) of the (002) peak by Scherrer´s equation.
Please refer to our previous publication for more information [20]. The coefficient K was taken to
be 0.89 according to Raza et al. [22, 23]. XRD data was based on the average of three repetitions.

Raman spectroscopy (Nicolet Almega XR, Thermo Fisher Scientific, USA) was utilized to
characterise the NG and exfoliated GNP. GNP sheets were prepared by vacuum filtration of the
dispersion through a porous membrane (PVDF, pore size 0.45 μm). Raman measurements were
performed with a wavelength of 532 nm.

Flexural test specimens were prepared according to the ASTM-D 790 standard. The samples were
cut into beams of dimensions 3.2 x 12.7 x 70 mm3 using a diamond cutting wheel. The cut surfaces
were polished by hand using different grades of abrasive paper (from 1000 to 4000 grade). The
support span to depth ratio was 16:1 while the cross-head speed was calculated in accordance to
ASTM-D 790 with the strain rate in the outer fibre equal to 0.01 mm/mm/min.

The electrical conductivity of all samples was measured by a two-probes method using a
picoammeter (Keithley 6485, Textronix, USA) and a DC voltage source (Agilet 6614C, USA). Casted
panels were cut into specimens with dimensions of 3.2 x 10 x 30 mm3. Silver paint was applied to
both ends to ensure good contact between the electrodes and the sample. Voltages in the range
of 5 to 10 V were used. For specimens with resistance exceeding 1010 Ohm, electrical resistivity

5
was no longer measurable and the samples were considered as ‘non-conductive’. Five specimens
for each composition were tested and average values reported.

Thermal diffusivity (α) of GNP/epoxy and neat epoxy reference specimens with graphite coated
surfaces (φ=20 mm and t=3.0 mm) were measured by a Netzsch LFA 457 MicroFlash® Laser
Apparatus (Netzsch, Germany) between 20 and 120 °C. The front side of a plane-parallel sample is
heated by a short laser pulse (HeNe laser 633 nm, Laser pulse energy: up to 18 J). The thermal
diffusivity (α) obtained from the transient thermal pulse using Cowen analysis. Thermal diffusivity
was measured by Netzsch LFA 457 Laser and λ was calculated via the formula:

 (T )   (T )   (T )  C p (T ) (1)

where density (ρ) was measured using Archimedes method and the specific heat capacity [24]
measured with the flash diffusivity system using Vespel SP-1 polyimide as a heat capacity
reference.

3 RESULTS AND DISCUSSION

3.1 Morphological Analysis of Graphite, GNP and GNP/Epoxy

The natural graphite (NG) flakes showed lateral dimensions varying between 600 ± 150 μm and
800 ± 200 μm and thickness of ca. 40 μm (Fig. 1a,b), giving the initial NG flakes an apect ratio (AR)
of ca. 20 ± 5. After the in-situ exfoliation process in TRM, depending on the actual processing
conditions [20], nanoplatelets of different dimensions and aspect ratios can be obtained (Table 1)
The highest aspect ratios were obtained at a resin temperature of 35 °C (Table 1, Fig. 1c-e) as
discussed in our previous publication [20] (see for details Supplementary Information). AFM
images and Raman data (Fig. 1e,f) confirm that FLG/GNP was obtained after in-situ exfoliation in
epoxy resin. Moreover, Raman spectroscopy (Fig. 1f) of FLG (containing D, G and 2D peaks)
extracted from the epoxy/acetone solution confirmed a graphitic structure with low defect
content.

6
Fig. 1. (a,b) SEM images of NG; (c) representative SEM, (d) TEM and (e) AFM images of FLG/GNP
obtained by in-situ exfoliation in epoxy resin at 35 °C; (f) Raman image of FLG (containing D, G and
2D peaks) extracted from the epoxy/acetone solution, confirming a graphitic structure with low
defect content.

Table 1. Properties of initial NG and GNP particles obtained for different in-situ exfoliation epoxy
resin temperatures.

Length by SEM Thickness by XRD Thickness by TEM Aspect Ratio ID/IG


Sample
L (μm) TXRD (nm) TTEM (nm) (L/ TXRD - L/ TTEM) (-)

NG ~800 - ~ 40000* N.A. - ~ 20 0

GNP (RT) 3.5 ± 0.9 ~ 23 ± 2 14 ± 5 150 - 250 0.16 ± 0.08

GNP (25 °C) 3.5 ± 1.5 28 ± 8 14 ± 8 125 - 250 0.16 ± 0.08

GNP (30 °C) 4.0 ± 1.3 27 ± 6 12 ± 4 150 - 333 0.15 ± 0.04

GNP (35 °C) 5.2 ± 2.0 17 ± 5 5±4 306 - 1040 0.07 ± 0.05

GNP (40 °C) 4.6 ± 1.8 20 ± 10 9±5 230 - 511 0.12 ± 0.04

*estimated from SEM micrographs.

The discrepancy between the thickness values as measured by TEM and XRD was discussed in [20].
For sake of completeness, both (average) values of thickness, as measured by XRD and TEM will be
used to estimate two values of aspect ratio (L/TXRD and L/ TTEM) .

7
The ratio of the intensities of the D and G bands, ID/IG, gives an indication of the quality of the
graphene particles through its inverse relation to the flakes’ lateral dimensions. That is, the lower
the ID/IG ratio the lower the defects in the graphitic structure, suggesting a reduction of edge
defects [15]. The particles obtained at 35 °C attain the lowest ID/IG ratio (0.07 ± 0.05). This
difference can be explained by a reduction in edge defects for larger flakes processed at 35 °C, in
analogy with previous reports. For instance, Khan et al. showed a decrease of the ID/IG ratio from
0.22 to 0.08 when larger flakes were selectively separated from a solvent dispersion method by
decreasing the centrifugation rotational speed from 4000 to 500 rpm [15].

5.0 wt% GNP/epoxy mixtures were observed under the optical microscopy. After eight cycles of in-
situ exfoliation, GNP was homogenously distributed within matrix (see the optical microscopy
images in Supplementary Information). The degree of dispersion and spatial distribution of GNPs
within the epoxy matrix were observed from cryo-fractured cross-sectional surfaces of cured
composites (Fig. 2, see for more SEM images Supplementary Information).

Fig. 2. SEM images of cross sectional areas of (a) pure epoxy, (b) 1 wt.% GNP/epoxy, (c) 3 wt.%
GNP/epoxy, (d) 5 wt.% GNP/epoxy at 35 °C and (e-h) 3 wt.% GNP/epoxy at different in-situ
exfoliation temperatures (25, 30, 35 and 40°C).

Pure epoxy samples (Fig. 2a) present smooth and flat cross-sections after cryo-fracturing,
compared to GNP/epoxy nanocomposites (Fig. 2b-h). Fig. 2b-d show an overall good dispersion of
GNP in epoxy, after curing, with some evidence of (re)agglomeration for systems incorporating 5

8
wt.% GNP (Fig. 2 d, see for higher magnification images Fig.5-2 f in Supplementary Information).
The series of SEM images in Fig. 2e-h show a level of GNP dispersion invariant with the in-situ
exfoliation temperature. GNP in-situ exfoliated in epoxy resin at 35 °C appear slightly larger in
Fig.2 g (see for higher magnification images Supplementary Information), in qualitative agreement
with Table 1.

3.2 Mechanical Property and Micromechanical Model Prediction

Good dispersions and high aspect ratio GNP particles should translate in high mechanical
properties of the corresponding epoxy nanocomposites. Fig. 3a shows the flexural modulus and
flexural strength of GNP/epoxy composites, in-situ exfoliated at different resin temperatures, as a
function of nanofiller loading. It is noted that all cured composites exhibit a similar Tg (indicating
similar cross-link densities) and low porosity, which makes comparison between physical
properties easier and fairer (see Supplementary Information about the density,TGA and DSC
results). As a general trend the modulus increases with filler content, up to 4.0 wt.%, indicating
good dispersion levels for these compositions, in agreement with Fig. 2. Higher filler contents (> 4
wt.%) lead to a reduced mechanical reinforcement efficiency (relatively low or no increase in
modulus), suggesting some reagglomeration, in agreement with Fig. 2d. The decrease of viscosity
with increasing temperature is independent of filler aspect ratio. In terms of the process of
reagglomeration (dynamic percolation) there are two contrasting factors to take into account: 1)
the viscosity should increase with filler aspect ratio (percolation phenomenon), hence limiting
reagglomeration; 2) the larger the aspect ratio, the higher the specific surface area, the larger the
driving force for reagglomeration (surface energy minimization). In our case, we expect that factor
2 is more important than factor 1. That is why the higher aspect ratio GNPs (obtained at 35 °C and
40 °C) are more prone to reagglomeration. Interestingly, the modulus increase depends
significantly on the temperature at which the in-situ exfoliation/dispersion process has been
carried out. In particular the modulus increase scales with the GNP aspect ratio, achieving a
maximum in correspondence with composites in-situ exfoliated at 35 °C, which contain GNPs of
the largest AR (see Table 1). Our optimal composite (4.0 wt.% GNP/epoxy, prepared at 35 °C)
shows a 160% increase in the flexural modulus, with respect to the neat epoxy resin, which is the
largest ever reported in the scientific literature for such relatively high loadings [25-45].

9
Fig. 3. (a) Flexural modulus and flexural strength of GNP/epoxy composites processed at various in-
situ exfoliation temperatures; (b) Halpin-Tsai models with Mori–Tanaka modified shape factor of
composite modulus ratio Ec/Em as a function of filler aspect ratio together with experimental data
of GNP/epoxy composites. Model predictions assumed for platelet moduli of 200 and 300 GPa with
a 2D planar or 3D random orientation.

In order to analyse in more details the mechanical results, modulus increase (Ec/Em, where Ec and
Em are the modulus of the final composite and neat matrix, respectively) is compared to
theoretical predictions in Fig. 3b. The Halpin-Tsai semi-empirical micromechanical model (see
Supplementary Information) has been used, in analogies with previous studies [46, 47].

10
Ec/Em values plotted as a function of the average experimental GNP aspect ratio (Fig. 3b) follow a
sigmonoidal curve as predicted by the Halpin-Tsai model. The experimental data could be best
fitted assuming a 3D random orientation of platelets with an elastic modulus of 300 GPa. This
morphology of 3D random instead of 2D in-plane platelet orientations is in agreement with the
SEM images in Fig. 2. Such a GNP modulus of around 300 GPa s in agreement with previous
studies. For example, Frank et al. [48] determined a elastic modulus value of 500 GPa for a stack of
graphene sheets (n < 5), while Zhang et al. [49] measured the modulus of multiple layer graphene
sheets at 370 GPa. Based on this it can be concluded that the effective modulus of the GNP in the
composite is relatively high, suggesting a high mechanical reinforcing efficiency for composites
fabricated by the in-situ exfoliation and dispersion method.

3.3 Electrical Conductivity

Fig. 4 shows the electrical conductivity values of GNP/epoxy composites, obtained at different in-
situ exfoliation temperatures, as a function of filler content. By comparing the different data sets,
it can be observed that the maximum electrical conductivity and percolation threshold
respectively increases and decreases with GNP aspect ratio. The increase in maximum conductivity
with aspect ratio (and lateral dimension of GNP) can be explained by the reduction of density of
defects in the graphitic structure [50]. The higher aspect ratio means a larger lateral dimension
and/or a lower thickness of GNP. Good dispersion of higher aspect ratio of GNP can therfore
reduce density defects in the graphitic structure.

Percolation theory explains well the inverse proportionality between percolation threshold and
GNP aspect ratio [51, 52]. The data sets for nanocomposites processed at RT and 35 °C show
typical exponential functions, indicating percolative behaviour. Percolation theory, often describes
well the insulator to conductor transition in conductive polymer composites, and can be used to
interpret the data points in Fig. 4, using an equation of the type [53]:

(  c ) t
c   f [ ] (2)
1  c

Where  f is the conductivity of the filler,  and c are the filler volume fraction and the

percolation threshold and t is the ‘universal critical exponent’ that depends on dimensionality of
the formed network where t ≈ 2 normally refers to a three-dimensional network [54-56]. For the

GNP/epoxy composite in-situ exfoliated at 35 °C, fitting resulted in c = 0.52 vol.% and t = 3.0 (see

11
inset Fig. 4), indeed indicating a 3D network. It is noted that the value of t is sligtly higher than the
theoretical value for a 3Dl network, which can be explained by a broad distribution of the
tunnelling resistance, and hence a broad distribution of the inter-particle distance [57].

Fig. 4. Electrical conductivity of GNP/epoxy composites, in-situ exfoliated at different resin


tempratures, as a function of GNP loading, together with electrical percolation threshold (EPT)
curves (t equals 2.48 and 3.00 for in-situ exfoliation at RT and 35 °C, respectively).

An analytical model previously presented by Lu et al. [58], based on the average interparticle
distance concept , can be used to analyse the percolation threshold obtained. The model assumes
a 3D random distribution of GNPs, uniformly distributed within the matrix. GNPs are assumed to
be thin and round impenetrable rigid disks [59] with thickness t, diameter D, and aspect ratio, AR =
D/t. Lu et al.’s model [58] was found to fit relatively well the composites based on GNPs of high AR
(i.e. AR > 250) [59]. According to this model the percolation threshold ( c ) can be expressed as:

2.154
c  (3)
AR

Here we found an experimental percolation threshold of c is 0.52 vol.% for nanocomposites in-
situ exfoliated at 35 °C, which should correspond to an average GNP aspect ratio of 414 according

12
to Eq. 3. This value is well within the range of the 300-1000 experimentally found and reported in
Table 1. This analysis supports the understanding of the morphology of the nanocomposites, and
is consistent with a random 3D orientation of GNPs, homogeneously dispersed within the epoxy
matrix.

3.4 Thermal Conductivity and Model Prediction

Fig. 5 shows the thermal conductivity values of GNP/epoxy composites, obtained at an in-situ
exfoliation temperature of RT and 35 °C, as a function of filler content. The data show a typical
linear increase and no percolation behaviour, as in the case of electrical conductivity, in
agreement with previously reported literature [60, 61]. The enhancement of thermal conductivity
of GNP/epoxy nanocomposites is in general attributed and affected by the intrinsic thermal
conductivity of the filler, the filler loading, dispersion, orientation and the thermal resistance of
the interface between fillers/matrix. The most basic thermal conductivity models are based on the
standard rule of mixture (Eq. 4) and inverse rule of mixture (Eq. 5) [47], representing respectively
upper and lower bound scenarios, written as:

n
  i  i (4)
i 1

n
i

1
 (5)
 i i

where, ‘λ’ is the thermal conductivity of the composite, ‘n’ is the number of constituents in the
composite, ‘i’ is the index variable for the composite constituents, ‘  ’ is the volume fraction of
constituents.

In Fig. 5 the upper and lower theoretical bounds are drawn assuming a thermal conductivity for
GNP of 1000 W·m-1·k-1 based on literature data for GNP with the thickness of 8 nm [62]. As
expected, the experimental data lies between the lower and upper bound limits but closer to the
lower theorical limit constituted by the inverse rule of mixture (Eq. 5).

We believe that this is primarily due to the interfacial GNP/epoxy thermal resistance (also known
as Kapitza resistance [63], the transfer of thermal energy is carried out by free electron interaction
and lattice vibration between two contacted interfaces) which is ignored in our models. If the
interfacial resistance would accurately be known [64], it could simply be added into Eq. 4-5 as an
additional interface component ‘i’, for instance, in a series configuration. It should be noted that

13
the larger the aspect ratio and lateral size of the GNP, the lower the importance of the interfacial
resistance as the number of filler/polymer contact points decreases. In fact, Fig. 5 demonstrates a
slight enhancement in thermal conductivity for composites in-situ exfoliated in epoxy resin at
35 °C (i.e. higher aspect ratios) compared to RT processed composites.

In orde to benchmark our results, Fig. 6 (red and pink data points) presents the increase in thermal
conductivity of our GNP/epoxy composites and most representative results taken from the
scientific literature [26, 44, 61, 64-66]. Our data approximately follows the general trend found in
literature, with the exception of a few studies, most notably the work of Shahil et al. forhigh filler
contents (10 vol. % graphene/MLG/epoxy with n < 5) and seemingly unusually thin samples [64].

It is interesting to note that generally GNP based nanocomposites show better thermal properties
than CNT based nanocomposites [67]. Since GNPs have a higher surface contact area with the
polymer matrix than MWCNTs, thin epoxy layers on GNP can reduce the barriers for phonon
transport, while the better graphitic intergrity of GNP possesses better conductance. To obtain a
further enhancement in the thermal conductivity either the filler content or the GNP lateral size,
or both, must be increased. Unfortunately in our case, 5.0 wt.% GNP loading was the maximum
filler loading at an in-situ exfoliation temperature of 35 °C to yield high quality well dispersed
nanocomposites. Higher filler loadings result in re-agglomeration. Moreover, increasing the filler
loading has its toll on the viscosity of the resin and makes subsequent processing more difficult.
Without compromising resin properties and other physical properties of the nanocomposites,
another solution might be to combine CNTs and GNPs. Future research is now focussing on such
hybrid systems where both CNTs and GNPs are combined to form 3D conductive networks with
CNTs bridging the GNPs, resulting in synergetic effects [68].

14
Fig. 5. Thermal conductivity analytical model predictions using the input parameters shown and
experimental data for in-situ exfoliated GNP/epoxy composites processed at RT and 35 °C.

Fig. 6 Comparison of mechanical reinforcement in terms of percentage increase in flexural modulus


and thermal conductivity of the in-situ exfoliated GNP/epoxy composites together with literature
data reporting both these properties [26, 44, 61, 64-66] ("black" and "grey" data points refer to
flexural modulus increase and "red" and "pink" data points refer thermal conductivity increase).

15
Fig. 6 shows a final comparison of the flexural modulus and thermal conductivity increase (see for
more details and absolute values of flexural modulus Supplementary Information) of the current
in-situ exfoliated nanocomposites together with other GNP based epoxy composites taken from
literature. From this graph it becomes clear that the current methodology of in-situ exfoliation by
TRM results in high quality nanocomposites with mechanical reinforcement levels that out-
perform other – often more elaborately processed – systems. Also thermal conductivity values are
among the highest reported for such a GNP/epoxy system with only one study reporting a higher
value but at a four times higher filler loading.

4 CONCLUSIONS

The simple methodology based on three-roll milling (TRM) for the in-situ exfoliation of natural
graphite into GNP within the epoxy was demonstrated to bring about remarkable enhancement in
mechanical, electrical and thermal properties. TRM was able to produce in-situ GNP particles with
low defect content and aspect ratios up to 1000, while at the same time creating homogeneous
GNP dispersions in epoxy up to relatively high filler loadings of around 4-5 wt.%. GNP/epoxy
composites showed high mechanical reinforcement efficiency (160% increase in flexural modulus
at 4 wt.% GNP; the largest ever reported for epoxy based systems at this filler loading), a relatively
low percolation threshold (0.52 vol.%) combined with an electrical conductivity of ~10-2 S/m at 3
wt.% GNP), and a high thermal conductivity (100% increase in thermal conductivity @ 5 wt.%
GNP). In short, this study has shown that the production of GNP-filled resins using TRM technology
can potentially remove important cost barriers for GNP modified resins for applications in
composites, adhesives and coatings. The challenge for the future will be to in-situ exfoliate the
natural graphite to even higher aspect ratio GNP (> 1000) and homogeneously disperse and orient
them at even higher filler loadings.

ACKNOWLEDGEMENTS

This research has received funding from NanoSynth Project funded by the Technology Strategy
Board (TSB) through the Technology Inspired Innovation – NANO Competition, No. 101257. Y. L.
would also like to acknowledge the financial support through the China Scholarship Council (CSC)
scheme.

16
COMPETING INTERESTS

The authors declare that there is no conflict of interest regarding the publication of this paper.

REFERENCES

[1] Novoselov KS, Geim AK, Morozov S, Jiang D, Zhang Y, Dubonos S, Grigorieva I, Firsov A. Electric
Field Effect in Atomically Thin Carbon Films. science. 2004;306(5696):666-669.
[2] Geim AK, Novoselov KS. The Rise of Graphene. Nature materials. 2007;6(3):183-191.
[3] Nika D, Ghosh S, Pokatilov E, Balandin A. Lattice Thermal Conductivity of Graphene Flakes:
Comparison with Bulk Graphite. Applied Physics Letters. 2009;94(20):203103.
[4] Gagné M, Therriault D. Lightning Strike Protection of Composites. PrAeS. 2014;64(1-16.
[5] Morozov S, Novoselov K, Katsnelson M, Schedin F, Elias D, Jaszczak J, Geim A. Giant Intrinsic
Carrier Mobilities in Graphene and Its Bilayer. Physical Review Letters. 2008;100(1):016602.
[6] Lee C, Wei X, Kysar JW, Hone J. Measurement of the Elastic Properties and Intrinsic Strength of
Monolayer Graphene. science. 2008;321(5887):385-388.
[7] Huang X, Qi X, Boey F, Zhang H. Graphene-Based Composites. Chemical Society Reviews.
2012;41(2):666-686.
[8] Young RJ, Kinloch IA, Gong L, Novoselov KS. The Mechanics of Graphene Nanocomposites: A
Review. Composites Science and Technology. 2012;72(12):1459-1476.
[9] Das B, Prasad KE, Ramamurty U, Rao C. Nano-Indentation Studies on Polymer Matrix
Composites Reinforced by Few-Layer Graphene. Nanotechnology. 2009;20(12):125705.
[10] Maitra U, Prasad KE, Ramamurty U, Rao C. Mechanical Properties of Nanodiamond-
Reinforced Polymer-Matrix Composites. Solid State Communications. 2009;149(39):1693-1697.
[11] Kim H, Abdala AA, Macosko CW. Graphene/Polymer Nanocomposites. Macromolecules.
2010;43(16):6515-6530.
[12] Porwal H, Grasso S, Reece M. Review of Graphene–Ceramic Matrix Composites. Advances in
Applied Ceramics. 2013;112(8):443-454.
[13] Sengupta R, Bhattacharya M, Bandyopadhyay S, Bhowmick AK. A Review on the Mechanical
and Electrical Properties of Graphite and Modified Graphite Reinforced Polymer Composites.
Progress in polymer science. 2011;36(5):638-670.
[14] Khan U, May P, Porwal H, Nawaz K, Coleman JN. Improved Adhesive Strength and Toughness
of Polyvinyl Acetate Glue on Addition of Small Quantities of Graphene. ACS applied materials &
interfaces. 2013;5(4):1423-1428.
[15] Khan U, O’Neill A, Porwal H, May P, Nawaz K, Coleman JN. Size Selection of Dispersed,
Exfoliated Graphene Flakes by Controlled Centrifugation. Carbon. 2012;50(2):470-475.
[16] Khan U, Porwal H, O’Neill A, Nawaz K, May P, Coleman JN. Solvent-Exfoliated Graphene at
Extremely High Concentration. Langmuir. 2011;27(15):9077-9082.
[17] Hernandez Y, Nicolosi V, Lotya M, Blighe FM, Sun Z, De S, McGovern I, Holland B, Byrne M,
Gun'Ko YK. High-Yield Production of Graphene by Liquid-Phase Exfoliation of Graphite. Nature
nanotechnology. 2008;3(9):563-568.
[18] A. O'Neill, U. Khan, P. N. Nirmalraj, J. Boland, J. N. Coleman, J. Phys. Chem. C

17
[19] Paton KR, Varrla E, Backes C, Smith RJ, Khan U, O’Neill A, Boland C, Lotya M, Istrate OM, King
P. Scalable Production of Large Quantities of Defect-Free Few-Layer Graphene by Shear Exfoliation
in Liquids. Nature materials. 2014;13(6):624-630.
[20] Li Y, Zhang H, Crespo M, Porwal H, Picot O, Santagiuliana G, Huang Z, Barbieri E, Pugno NM,
Peijs T. In Situ Exfoliation of Graphene in Epoxy Resins: A Facile Strategy to Efficient and Large
Scale Graphene Nanocomposites. ACS Applied Materials & Interfaces. 2016;8(36):24112-24122.
[21] Zhang H, Liu Y, Huo S, Briscoe J, Tu W, Picot OT, Rezai A, Bilotti E, Peijs T. Filtration Effects of
Graphene Nanoplatelets in Resin Infusion Processes: Problems and Possible Solutions. Compos Sci
Technol. 2017;139:138-145.
[22] Raza M, Westwood A, Brown A, Stirling C. Texture, Transport and Mechanical Properties of
Graphite Nanoplatelet/Silicone Composites Produced by Three Roll Mill. Composites Science and
Technology. 2012;72(3):467-475.
[23] Raza MA, Westwood A, Brown A, Hondow N, Stirling C. Characterisation of Graphite
Nanoplatelets and the Physical Properties of Graphite Nanoplatelet/Silicone Composites for
Thermal Interface Applications. Carbon. 2011;49(13):4269-4279.
[24] Wong C, Bollampally RS. Thermal Conductivity, Elastic Modulus, and Coefficient of Thermal
Expansion of Polymer Composites Filled with Ceramic Particles for Electronic Packaging. Journal of
Applied Polymer Science. 1999;74(14):3396-3403.
[25] Song L, Lu S, Xiao X, Qi B, He Z, Xu X, Rao B, Yu J. Enhanced Thermal and Mechanical
Properties of Liquid Crystalline-Grafted Graphene Oxide-Filled Epoxy Composites. Polym Bull. 1-17.
[26] Debelak B, Lafdi K. Use of Exfoliated Graphite Filler to Enhance Polymer Physical Properties.
Carbon. 2007;45(9):1727-1734.
[27] Yasmin A, Luo J-J, Daniel IM. Processing of Expanded Graphite Reinforced Polymer
Nanocomposites. Composites Science and Technology. 2006;66(9):1182-1189.
[28] Bortz DR, Heras EG, Martin-Gullon I. Impressive Fatigue Life and Fracture Toughness
Improvements in Graphene Oxide/Epoxy Composites. Macromolecules. 2011;45(1):238-245.
[29] Yasmin A, Daniel IM. Mechanical and Thermal Properties of Graphite Platelet/Epoxy
Composites. Polymer. 2004;45(24):8211-8219.
[30] Rafiee M, Rafiee J, Yu Z-Z, Koratkar N. Buckling Resistant Graphene Nanocomposites. Applied
Physics Letters. 2009;95(22):223103.
[31] Chatterjee S, Wang J, Kuo W, Tai N, Salzmann C, Li W, Hollertz R, Nüesch F, Chu B. Mechanical
Reinforcement and Thermal Conductivity in Expanded Graphene Nanoplatelets Reinforced Epoxy
Composites. Chemical Physics Letters. 2012;531(6-10.
[32] King JA, Klimek DR, Miskioglu I, Odegard GM. Mechanical Properties of Graphene
Nanoplatelet/Epoxy Composites. Journal of Applied Polymer Science. 2013;128(6):4217-4223.
[33] Shen M-Y, Chang T-Y, Hsieh T-H, Li Y-L, Chiang C-L, Yang H, Yip M-C. Mechanical Properties
and Tensile Fatigue of Graphene Nanoplatelets Reinforced Polymer Nanocomposites. Journal of
Nanomaterials. 2013;2013(1.
[34] Yang Y, Rigdon W, Huang X, Li X. Enhancing Graphene Reinforcing Potential in Composites by
Hydrogen Passivation Induced Dispersion. Scientific reports. 2013;3(
[35] Rafiee MA, Rafiee J, Srivastava I, Wang Z, Song H, Yu ZZ, Koratkar N. Fracture and Fatigue in
Graphene Nanocomposites. small. 2010;6(2):179-183.

18
[36] Corcione CE, Freuli F, Maffezzoli A. The Aspect Ratio of Epoxy Matrix Nanocomposites
Reinforced with Graphene Stacks. Polymer Engineering & Science. 2013;53(3):531-539.
[37] Naebe M, Wang J, Amini A, Khayyam H, Hameed N, Li LH, Chen Y, Fox B. Mechanical Property
and Structure of Covalent Functionalised Graphene/Epoxy Nanocomposites. Scientific reports.
2014;4(
[38] Gudarzi MM, Sharif F. Enhancement of Dispersion and Bonding of Graphene-Polymer through
Wet Transfer of Functionalized Graphene Oxide. Express Polymer Letters. 2012;6(12):1017-1031.
[39] Miller SG, Bauer JL, Maryanski MJ, Heimann PJ, Barlow JP, Gosau J-M, Allred RE.
Characterization of Epoxy Functionalized Graphite Nanoparticles and the Physical Properties of
Epoxy Matrix Nanocomposites. Composites Science and Technology. 2010;70(7):1120-1125.
[40] Li J, Kim J-K, Sham ML. Conductive Graphite Nanoplatelet/Epoxy Nanocomposites: Effects of
Exfoliation and Uv/Ozone Treatment of Graphite. Scripta Materialia. 2005;53(2):235-240.
[41] Jana S, Zhong W-H. Graphite Particles with a “Puffed” Structure and Enhancement in
Mechanical Performance of Their Epoxy Composites. Materials Science and Engineering: A.
2009;525(1):138-146.
[42] Ramos-Galicia L, Mendez L, Martínez-Hernández AL, Espindola-Gonzalez A, Galindo-Esquivel I,
Fuentes-Ramirez R, Velasco-Santos C. Improved Performance of an Epoxy Matrix as a Result of
Combining Graphene Oxide and Reduced Graphene. International Journal of Polymer Science.
2013;2013(
[43] Rafiee MA, Lu W, Thomas AV, Zandiatashbar A, Rafiee J, Tour JM, Koratkar NA. Graphene
Nanoribbon Composites. ACS nano. 2010;4(12):7415-7420.
[44] Chatterjee S, Nafezarefi F, Tai N, Schlagenhauf L, Nüesch F, Chu B. Size and Synergy Effects of
Nanofiller Hybrids Including Graphene Nanoplatelets and Carbon Nanotubes in Mechanical
Properties of Epoxy Composites. Carbon. 2012;50(15):5380-5386.
[45] Sellam C, Zhai Z, Zahabi H, Picot OT, Deng H, Fu Q, Bilotti E, Peijs T. High Mechanical
Reinforcing Efficiency of Layered Poly (Vinyl Alcohol)–Graphene Oxide Nanocomposites.
Nanocomposites. 2015;1(2):89-95.
[46] Bilotti E, Deng H, Zhang R, Lu D, Bras W, Fischer HR, Peijs T. Synergistic Reinforcement of
Highly Oriented Poly (Propylene) Tapes by Sepiolite Nanoclay. Macromolecular Materials and
Engineering. 2010;295(1):37-47.
[47] Bilotti E, Zhang H, Deng H, Zhang R, Fu Q, Peijs T. Controlling the Dynamic Percolation of
Carbon Nanotube Based Conductive Polymer Composites by Addition of Secondary Nanofillers:
The Effect on Electrical Conductivity and Tuneable Sensing Behaviour. Composites Science and
Technology. 2013;74(0):85-90.
[48] Frank I, Tanenbaum DM, Van der Zande A, McEuen PL. Mechanical Properties of Suspended
Graphene Sheets. Journal of Vacuum Science & Technology B. 2007;25(6):2558-2561.
[49] Zhang Y, Pan C. Measurements of Mechanical Properties and Number of Layers of Graphene
from Nano-Indentation. Diamond and related materials. 2012;24(1-5.
[50] Li L, Reich S, Robertson J. Defect Energies of Graphite: Density-Functional Calculations.
Physical Review B. 2005;72(18):184109.
[51] Yi JY, Choi GM. Percolation Behavior of Conductor-Insulator Composites with Varying Aspect
Ratio of Conductive Fiber. Journal of electroceramics. 1999;3(4):361-369.

19
[52] Wu D, Wu L, Zhou W, Sun Y, Zhang M. Relations between the Aspect Ratio of Carbon
Nanotubes and the Formation of Percolation Networks in Biodegradable Polylactide/Carbon
Nanotube Composites. Journal of Polymer Science Part B: Polymer Physics. 2010;48(4):479-489.
[53] Stankovich S, Dikin DA, Dommett GH, Kohlhaas KM, Zimney EJ, Stach EA, Piner RD, Nguyen ST,
Ruoff RS. Graphene-Based Composite Materials. nature. 2006;442(7100):282-286.
[54] Zhang R, Baxendale M, Peijs T. Universal Resistivity–Strain Dependence of Carbon
Nanotube/Polymer Composites. Physical Review B. 2007;76(19):195433.
[55] Vionnet-Menot S, Grimaldi C, Maeder T, Strässler S, Ryser P. Tunneling-Percolation Origin of
Nonuniversality: Theory and Experiments. Physical Review B. 2005;71(6):064201.
[56] Zhang H, Bilotti E, Tu W, Lew CY, Peijs T. Static and Dynamic Percolation of Phenoxy/Carbon
Nanotube Nanocomposites. European Polymer Journal. 2015;68(128-138.
[57] Grimaldi C, Balberg I. Tunneling and Nonuniversality in Continuum Percolation Systems.
Physical review letters. 2006;96(6):066602.
[58] Lu C, Mai Y-W. Influence of Aspect Ratio on Barrier Properties of Polymer-Clay
Nanocomposites. Physical review letters. 2005;95(8):088303.
[59] Maxian O, Pedrazzoli D, Manas-Zloczower I. Modeling the Electrical Percolation Behavior of
Hybrid Nanocomposites Based on Carbon Nanotubes and Graphene Nanoplatelets. Materials
Research Express. 2015;2(9):095013.
[60] Kim HS, Bae HS, Yu J, Kim SY. Thermal Conductivity of Polymer Composites with the
Geometrical Characteristics of Graphene Nanoplatelets. Scientific reports. 2016;6(
[61] Chatterjee S, Wang JW, Kuo WS, Tai NH, Salzmann C, Li WL, Hollertz R, Nüesch FA, Chu BTT.
Mechanical Reinforcement and Thermal Conductivity in Expanded Graphene Nanoplatelets
Reinforced Epoxy Composites. Chemical Physics Letters. 2012;531(6-10.
[62] Balandin AA, Nika DL. Thermal Transport in Graphene and Graphene Multilayers. arXiv
preprint arXiv:150302124. 2015;
[63] Wei J, Vo T, Inam F. Epoxy/Graphene Nanocomposites - Processing and Properties: A Review.
RSC Advances. 2015;5(90):73510-73524.
[64] Shahil KM, Balandin AA. Graphene–Multilayer Graphene Nanocomposites as Highly Efficient
Thermal Interface Materials. Nano letters. 2012;12(2):861-867.
[65] Saw WS, Mariatti M. Properties of Synthetic Diamond and Graphene Nanoplatelet-Filled
Epoxy Thin Film Composites for Electronic Applications. Journal of Materials Science: Materials in
Electronics. 2012;23(4):817-824.
[66] Yang S-Y, Lin W-N, Huang Y-L, Tien H-W, Wang J-Y, Ma C-CM, Li S-M, Wang Y-S. Synergetic
Effects of Graphene Platelets and Carbon Nanotubes on the Mechanical and Thermal Properties of
Epoxy Composites. Carbon. 2011;49(3):793-803.
[67] Teng C-C, Ma C-CM, Lu C-H, Yang S-Y, Lee S-H, Hsiao M-C, Yen M-Y, Chiou K-C, Lee T-M.
Thermal Conductivity and Structure of Non-Covalent Functionalized Graphene/Epoxy Composites.
Carbon. 2011;49(15):5107-5116.
[68] Yu L, Park JS, Lim Y-S, Lee CS, Shin K, Moon HJ, Yang C-M, Lee YS, Han JH. Carbon Hybrid Fillers
Composed of Carbon Nanotubes Directly Grown on Graphene Nanoplatelets for Effective Thermal
Conductivity in Epoxy Composites. Nanotechnology. 2013;24(15):155604.

20

You might also like