Polymers: Mechanical, Thermal, and Electrical Properties of Graphene-Epoxy Nanocomposites-A Review
Polymers: Mechanical, Thermal, and Electrical Properties of Graphene-Epoxy Nanocomposites-A Review
Polymers: Mechanical, Thermal, and Electrical Properties of Graphene-Epoxy Nanocomposites-A Review
Review
Mechanical, Thermal, and Electrical Properties of
Graphene-Epoxy Nanocomposites—A Review
Rasheed Atif, Islam Shyha and Fawad Inam *
Department of Mechanical and Construction Engineering, Faculty of Engineering and Environment,
Northumbria University, Newcastle upon Tyne NE1 8ST, UK; [email protected] (R.A.);
[email protected] (I.S.)
* Correspondence: [email protected]; Tel.: +44-191-227-3741
Abstract: Monolithic epoxy, because of its brittleness, cannot prevent crack propagation and is
vulnerable to fracture. However, it is well established that when reinforced—especially by nano-fillers,
such as metallic oxides, clays, carbon nanotubes, and other carbonaceous materials—its ability to
withstand crack propagation is propitiously improved. Among various nano-fillers, graphene has
recently been employed as reinforcement in epoxy to enhance the fracture related properties of
the produced epoxy–graphene nanocomposites. In this review, mechanical, thermal, and electrical
properties of graphene reinforced epoxy nanocomposites will be correlated with the topographical
features, morphology, weight fraction, dispersion state, and surface functionalization of graphene.
The factors in which contrasting results were reported in the literature are highlighted, such as the
influence of graphene on the mechanical properties of epoxy nanocomposites. Furthermore, the
challenges to achieving the desired performance of polymer nanocomposites are also suggested
throughout the article.
1. Introduction
Polymer Matrix Composites (PMCs) have found extensive applications in aerospace, automotive,
and construction, owing to ease of processing and high strength-to-weight ratio, which is an important
property required for aerospace applications [1]. Among different polymers, epoxy is the most
commonly used thermosetting polymer matrix in PMCs [2]. The damage tolerance and fracture
toughness of epoxy can be enhanced with the incorporation of (nano-) reinforcement, such as metallic
oxides [3–5], clays [6–8], carbon nanotubes (CNTs) [9–11], and other carbonaceous materials [12–14].
After the groundbreaking experiments on the two-dimensional material graphene by Nobel Laureates
Sir Andre Geim and Konstantin Novoselov [15] from the University of Manchester, graphene came into
the limelight in the research community, mainly because of its excellent electrical [16], thermal [17], and
mechanical properties [18]. Graphene found widespread applications in electronics [19], bio-electric
sensors [20], energy technology [21], lithium batteries [22], aerospace [23], bio-engineering [24], and
various other fields of nanotechnology [25]. There is an exponential rise in the use of graphene in
different research areas, mainly because of the properties inherited in, and transferred by, graphene to
the processed graphene-based materials.
To summarize the research trends related to graphene-based nanocomposites, multiple review
articles were recently published in which various aspects of graphene-based nanocomposites were
discussed. There are numerous ways to produce and characterize graphene-based materials [26].
Graphene-based materials were studied for different properties, such as thermal properties [27],
mechanical properties [28], electrical properties [29], rheological properties [30], microwave
adsorption [31,32], environmental and toxicological impacts [33], effect of preparation [34], and
gas barrier properties [35]. These materials have found biological applications, especially related
to toxicity [36], and in other applications like electrically-conductive adhesives [37] and selective
photoredox reactions [38]. Because of their hierarchical pore structures, these materials were found
suitable for gas sorption, storage, and separation [39]. Various factors influence the mechanical
properties of graphene-based materials—e.g., γ-ray irradiation was found to have a strong influence
on the structure–property relationship [40]. Various theoretical models were developed to predict
the mechanical properties of epoxy–graphene nanocompsites and correlated with interphases and
interfacial interactions [41]. It was presented that continuum mechanics can be used to predict the
minimum graphene sheet dimensions and optimum number of layers for good reinforcement [42].
Graphene was compared with other reinforcements, such as clays [43] and CNTs [44], and was shown
to have properties superior to the other nano-fillers. Various surface modifications were employed to
improve interfacial interactions, and their influence on the performance of polymer nanocomposites
was studied [45].
To date, eclectic reviews on graphene composites are covering a broad range of graphene-related
issues; it can, however, be observed that there is an obvious gap in the lack of a review article discussing
the mechanical, thermal, and electrical properties of epoxy–graphene nanocomposites. Therefore, this
review article discusses the correlation between graphene structure, morphology, weight fraction,
dispersion, surface modifications, and the corresponding mechanical, thermal, and electrical properties
of epoxy–graphene nanocomposites.
2. Epoxy as Matrix
There are various types of epoxy which have a wide range of applications because of their
superior attributes, such as improvement in composite mechanical properties, acceptable cost, and
processing flexibility [2]. Phenolic glycidyl ethers are formed by the condensation reaction between
epichlorohydrin and a phenol group. Within this class, the structure of the phenol-containing molecule
and the number of phenol groups per molecule distinguish different types of resins and the final
properties of monolithic epoxies and nanocomposites [2]. The epoxies have found some “high-end”
applications, including aerospace, marine, automotive, high-performance sports equipment (such
as tennis rackets), electronics, and industrial applications [46]. Due to the superior properties of
carbonaceous materials, such as high strength and stiffness, they are most widely used at present as
reinforcement in advanced Epoxy Matrix Composites (EMCs) [47–50].
Epoxy resins are of particular interest to structural engineers because these resins provide a unique
balance of chemical and mechanical properties combined with extreme processing versatility [51].
When a composite is produced from epoxy-carbon using hand lay-up process, a great flexibility
in aligning the fraction of fibers in a particular direction is available, which is dependent upon
the in-service load on the composite structural member. In-plane isotropy can also be achieved by
stacking the resin-impregnated fiber layers at equal numbers of 0˝ , +45˝ , ´45˝ , and 90˝ . There are
also other stacking sequences that can be used to achieve in-plane isotropy. The specific stiffness
of quasi-isotropic epoxy–graphite laminated composite is higher than many structural metals. The
highest specific strength achieved in epoxy–graphite is higher than common structural metals, with
the exception of ultrahigh-strength steels and some β-titanium alloys. For example, the epoxy-carbon
crutch is 50% lighter and still stronger than the aluminium crutch [2].
3. Graphene as Reinforcement
Graphene—a densely packed honey-comb crystal lattice made of carbon atoms having a thickness
equal to the atomic size of one carbon atom—has revolutionized the scientific parlance due to
its exceptional physical, electrical, and chemical properties. The graphene now found in various
applications was previously considered only a research material and a theoretical model to describe
Polymers 2016, 8, 281 3 of 37
the properties of other carbonaceous materials such as fullerenes, graphite, Single-Walled Carbon
Nanotubes (SWNTs), and Multi-Walled Carbon Nanotubes (MWNTs). It was believed that the real
existence of stand-alone single layer graphene would not be possible because of thermal fluctuations,
as the stability of long-range crystalline order found in graphene was considered impossible at finite
(room) temperatures. This perception was turned into belief by experiments when the stability of
thin films was found to have direct relation with the film thickness; i.e., film stability decreases with
a decrease in film thickness [52]. However, graphene can currently be found on a silicon substrate
or suspended in a liquid and ready for processing. Although its industrial applications are not
ubiquitous, it is widely used for research purposes (e.g., as reinforcement in PMCs) and has shown
significant improvement in different (mechanical, thermal, electrical etc.) properties of produced
nanocomposites [52–56].
The ability of a material to resist the propagation of an advancing crack is vital to the prevention
of failure/fracture [57]. Graphene can significantly improve fracture toughness of epoxy at very low
volume fraction by deflecting the advancing crack in the matrix. The details of the influence of various
kinds of graphene/graphite nanoplatelets (GNPs) on the fracture toughness of epoxy nanocomposites
are listed in Table 1. In all the composite systems mentioned in Table 1, epoxy was used as matrix
and the nanocomposites were produced using solution casting technique, except [58] where the resin
infiltration method was employed. The incorporation of graphene in epoxy can increase its fracture
toughness by as much as 131% [59]. It can also be observed that graphene size, weight fraction, surface
modification, and dispersion mode have strong influence on the improvement in fracture toughness
values of the produced epoxy–graphene nanocomposites. Monolithic epoxy shows brittle fracture and
beeline crack propagates, which results in straight fracture surfaces. The advancing crack in epoxy
interacts with the graphene sheets. Initially, the crack propagates through the epoxy matrix as there
are no significant intrinsic mechanisms available in monolithic epoxy to restrict crack propagation.
However, no sooner than the crack faces strong graphene sheets ahead, it surrenders and subdues.
Nevertheless, the extent of matrix strengthening and crack bridging provided by graphene strongly
depends upon its dispersion state and interfacial interactions with the epoxy matrix [60,61].
4. Fracture Toughness
The successful employment of epoxy-based nanocomposites relies on the ability of the composite
system to meet design and service requirements. The epoxy-based nanocomposites have found
applications in aerospace, automotive, and construction due to ease of processing and high
strength-to-weight ratio. In many applications, the composite system undergoes external loadings.
The relationship between loads acting on a system and the response of the system towards the
applied loads is studied in terms of mechanical properties. Therefore, epoxy-based nanocomposites
are supposed to have superior mechanical properties. There are various tests to measure mechanical
properties, such as tensile testing, bend testing, creep testing, fatigue testing, and hardness testing,
to name a few. These tests usually take specimens of specific geometries and subject to loading at
certain rate. In general, the industrial scale samples contain porosity and notches which act as stress
concentrators and are deleterious to the mechanical properties of nanocomposites. Sometimes, it
becomes difficult to control the maximum flaw size. The shape of the flaw is another very important
parameter, as pointed notch (V-notch) is more detrimental than round notch (U-shaped) [62].
Polymers 2016, 8, 281 4 of 37
Table 1. A brief record of epoxy-based nanocomposites studied for improvement in fracture toughness values.
Dispersion % Increase in
Sr. Authors Year Reinforcement/(wt %) Remarks Ref.
method K1C (MPa¨m1/2 )
GO (0.25 wt %) 25.6 K1C drops after 0.25 wt %
1 Wan et al. 2014 Sn + BM [63]
DGEBA-f-GO (0.25 wt %) 40.7 of reinforcement
Table 1. Cont.
Dispersion % Increase in
Sr. Authors Year Reinforcement/(wt %) Remarks Ref.
method K1C (MPa¨m1/2 )
11 Jia et al. 2014 GF (0.1 wt %) (resin infiltration) None 70 K1C did not change much between 0.1 to 0.5 wt % [58]
Poorly dispersed RGO (0.2 wt %) Sn 24
12 Tang et al. 2013 Trend still increasing [72]
Highly dispersed RGO (0.2 wt %) Sn + BM 52
10.79 µm (0.5wt %) 12
13 Wang et al. 2013 GO USn K1C drops after 0.5 wt % of reinforcement [57]
1.72 µm (0.5 wt %) 61
0.70 µm (0.1 wt %) 75
Dispersion and K1C improved
14 Chandrasekaran et al. 2013 GNPs* (0.5 wt %) 3RM 43 [73]
with three roll milling
APTS-GO (0.5 wt %) 25 Trend still increasing
15 Li et al. 2013 USn [74]
GPTS-GO (0.2 wt %) 43 K1C drops after 0.2 wt % of reinforcement
ND (0.5 wt %) No effect Fracture toughness improvement is higher by
16 Shadlou et al. 2013 USn CNF and GO (high aspect ratio) compared with [75]
CNF (0.5 wt %) 4.3
that by spherical ND
GO (0.5 wt %) 39.1
GO (0.1 wt %) 31 Trend remains same after 1 wt % of reinforcement
17 Jiang et al. 2013 Sn [76]
ATS (1 wt %) 58.6 K1C drops after 0.1 wt % of reinforcement
The maximum improvement is achieved with
ATGO (1 wt %) 86.2
functionalization
p-CNFs (0.4 wt %) 41
18 Liu et al. 2013 Sn Trend still increasing [77]
m-CNFs (0.4 wt %) 80
ATP (1 wt %) 14 K1C drops after 0.1 wt %
19 Wang et al. 2013 Sn [78]
GO (0.2 wt %) 19 Trend still increasing after 0.2 wt %
K1C drops with the further increase in ATP of
ATP (1 wt %) + GO (0.2 wt %) 27
reinforcement
ND (0.5 wt %) ´26.9
20 Alishahi et al. 2013 CNF (0.5 wt %) Sn 19 Trend still increasing [79]
GO (0.5 wt %) 23
CNT (0.5 wt %) 23.8
Polymers 2016, 8, 281 6 of 37
Table 1. Cont.
Dispersion % Increase in
Sr. Authors Year Reinforcement/(wt %) Remarks Ref.
method K1C (MPa¨m1/2 )
U-GnP (0.5 wt %) 49
21 Ma et al. 2013 MgSr + USn Trend still increasing [80]
m-GnP (0.5 wt %) 109
22 Feng et al. 2013 Graphene (0.5 wt %) Sn 76 K1C decreases after 0.5 wt % of reinforcement [81]
GnPs (5 µm, 2 wt %) 60
23 Chatterjee et al. 2012 GnPs (25 µm, 2 wt %) 3RM 80 Trend still increasing [82]
CNTs (2 wt %) 80
CNT:GnP = (9:1) (2 wt %) 76
24 Chatterjee et al. 2012 EGNPs (0.1 wt %) HPH + 3RM 66 K1C drops after 0.1 wt % of reinforcement [83]
GP (2.5 wt %) 57 The surface modification significantly improved
25 Zaman et al. 2011 Sn + MS [84]
the K1C
m-GP (4 wt %) 90
26 Rana et al. 2011 CNFs Sn + MS 40 K1C is dependent upon mixing time [85]
27 Bortz et al. 2011 GO (0.5 wt %) 3RM 60 K1C drops after 0.5 wt % of reinforcement [86]
CNFs (0.5 wt %) 19.4
28 Zhang et al. 2010 3RM Trend still increasing [87]
SCFs (15 wt %) 125.8
SCF (10 wt %)/CNF (0.75 wt %) 210
29 Fang et al. 2010 GNs MS + Sn 93.8 Better results with combination of MS and Sn [88]
30 Jana et al. 2009 GP with “puffed” structure (5 wt %) Sn 28 Trend still increasing [89]
SWNT (0.1 wt %) 17 Graphene platelets have more influence on K1C
31 Rafiee et al. 2009 Sn + MS [90]
MWNT (0.1 wt %) 20 than CNTs
3RM: three roll milling; APTS-GO: amino-functionalized graphene oxide (GO); ATGO: 3-Aminopropyltriethoxysilane functionalized silica nanoparticles attached GO; ATP: attapulgite;
ATS: 3-amino functionalized silica nanoparticles; BM: ball milling; CNF: carbon nanofiber; CNT: carbon nanotube; DGEBA-f-GO: diglycidyl ether of bisphenol-A functionalized
GO; EGNP: amine functionalized expanded graphene nanoplatelets; fGnP: polybenzimidazole functionalized graphene platelets (GnPs); G-NH2: amino-functionalized GNPs;
G-Si: silane modified GNPs; GF: graphene foam; GN: amine functionalized graphene sheet; GnP: graphene platelet; GNP*: graphite nanoplatelet; GNS: graphene nanosheet; GO:
graphite; GP: graphite particles; GPL: graphene nanoplatelets; GPTS-GO: epoxy functionalized GO; HPH: high pressure homogenizer; m-clay: surface modified nano clay; m-CNF:
triazole functionalized carbon nanofiber; m-GnP: surface modified GnP; m-GnP*: surfactant modified graphene platelet; m-GP: surface modified graphene platelets; MERGO:
microwave exfoliated reduced graphene oxide; MgSr: magnetic stirring; MS: mechanical stirring; MWCNT: multi-walled carbon nanotube; MWNT: multi-walled carbon nanotubes;
ND: nanodiamond; pCNF: pristine carbon nanofibers; RGO: thermally reduced graphene oxide; SATPGO: 3-aminopropyltriethoxysilane modified silica nanoparticles attached GO;
SCF: short carbon fibers; Silane-f-GO: silane functionalized GO; Sn: Sonication; SWNT: single-walled carbon nanotubes; U-GnP: unmodified graphene platelets; UG: unmodified
graphene nanoplatelets; USn: ultrasonication.
Polymers 2016,
Polymers 2016, 8,
8, 281
281 66 of
of 35
35
Polymers 2016, 8, 281 7 of 37
Due to
Due to the
the pronounced
pronounced effect effect of
of defects
defects on on nanocomposite
nanocomposite properties,
properties, it it is
is important
important to to
understand
understand how
how a system
a system will will tolerate external loading in the presence of a flaw under operating
Due to the pronounced effecttolerate external
of defects loading in the
on nanocomposite presenceitof
properties, a flaw under
is important operating
to understand
conditions, and
conditions, and how
how aa system
system willwill resist
resist the
the propagation
propagation of of cracks
cracks from
from these
these flaws.
flaws. Therefore,
Therefore, howhow
how a system will tolerate external loading in the presence of a flaw under operating conditions, and
the material will
the will behave in in reality will
will only
only bebe determined
determined when when the the test
test specimen
specimen contains
contains possible
possible
howmaterial
a system willbehaveresist thereality
propagation of cracks from these flaws. Therefore, how the material will
flaws,
flaws, such as a notch. To deal with this issue in a pragmatic way, an intentional notch is produced
behavesuch as a notch.
in reality will onlyTo deal with this issue
be determined when inthe
a pragmatic
test specimenway,contains
an intentional
possible notch is produced
flaws, such as a
in the
in the specimen,
specimen, and and resistance
resistance to to fracture
fracture isis measured
measured and and is is termed
termed fracture
fracture toughness.
toughness. Different
Different
notch. To deal with this issue in a pragmatic way, an intentional notch is produced in the specimen,
specimens are
specimens are used
used forfor fracture
fracture toughness,
toughness, such as as notched
notched tension,
tension, three-point bending, bending, and
and resistance to fracture is measured and is such termed fracture toughness.three-point
Different specimensand are
compact
compact tension specimen, as shown in Figure 1. The toughness is usually measured in three different
used for tension
fracturespecimen,
toughness, as shown
such asinnotched
Figure 1.tension,
The toughness is usually
three-point measured
bending, in three different
and compact tension
modes namely
modes namely (1) (1) Mode-I (tensile
(tensile mode); (2) (2) Mode-II (shearing
(shearing mode);
mode); andand (3)(3) Mode-III
Mode-III (tearing
(tearing
specimen, as shownMode-Iin Figure 1. Themode); toughnessMode-II
is usually measured in three different modes namely
mode),
mode), as shown
as (tensile
shown in in Figure
Figure 2. Most of the literature on epoxy nanocomposites reported Mode-I
(1) Mode-I mode); (2) 2. Most of
Mode-II the literature
(shearing mode); onandepoxy nanocomposites
(3) Mode-III reported
(tearing mode), Mode-I
as shown in
fracture
fracture toughness. Mode-I is preferred in contrast to Mode-II, because shear yielding is the dominant
Figure 2.toughness.
Most of theMode-Iliteratureis preferred
on epoxyinnanocomposites
contrast to Mode-II, because
reported shear
Mode-I yielding
fracture is the dominant
toughness. Mode-I
mechanism of
mechanism of failure
failure that is is acting
acting under
under Mode-II, delivering
delivering higher values values thanthan in
in Mode-I.
Mode-I. Mode-
is preferred in contrastthat to Mode-II, because Mode-II,
shear yielding is the higherdominant mechanism of failureMode-
that is
III
III is never
is never practiced.
practiced.
acting under Mode-II, delivering higher values than in Mode-I. Mode-III is never practiced.
Figure 1.
Figure Various fracture
1. Various toughness test
fracture toughness
toughness specimen geometries:
test specimen
specimen geometries: (a)
geometries: (a) notched
notched tensile; (b–d) compact
tensile; (b–d) compact
tension; (e)
(e)compact
compact bend; and
bend; (f)
and single-edge
(f) notched
single-edge three-point
notched bend
three-point specimens.
bend The
specimens.
tension; (e) compact bend; and (f) single-edge notched three-point bend specimens. The arrows
The indicate
arrows
arrows
the axis
indicate of
theloading.
axis of loading.
indicate the axis of loading.
Figure 2.
Figure 2. Various
Various fracture modes:
Various fracture modes: (a)
(a) mode-I,
mode-I, (b)
(b) mode-II,
mode-II, and
and (c)
(c) mode-III.
mode-III.
Some of the
Some the fracture toughness
toughness tests include
include double torsion,
torsion, indentation, double
double cantilever tests,
tests,
Some ofof the fracture
fracture toughness tests
tests include double
double torsion, indentation,
indentation, double cantilever
cantilever tests,
and Chevron
and Chevron notch
notch method.
method. Chevron
Chevron notch
notch method
method is popular,
popular, as
as it uses
uses aa relatively
relatively small
small amount
amount
and Chevron notch method. Chevron notch method isispopular, as it it
uses a relatively small amount of
of material and no material constants are needed for the calculations. The technique is also
of material and no material constants are needed for the calculations. The technique is also suitablesuitable
Polymers 2016, 8, 281 8 of 37
material and no material constants are needed for the calculations. The technique is also suitable for
high-temperature testing, as flaw healing is not a concern. However, it requires a complex specimen
shape that incurs an extra machining cost. The most commonly used specimen is a single-edge notched
beam subjected to three or four-point bending. Unfortunately, it has been reported that the results of
this test are very sensitive to the notch width and depth. Therefore, a pre-notched or molded beam is
preferred. As polymers and polymer nanocomposites can be molded into a desired shape, a specific
kind of notch can be replicated in multiple specimens. Due to the reproducibility of notch dimensions,
the single-edge notched beam test can give reproducible values of fracture toughness in polymers and
polymer nanocomposites. These are the reasons that most of the literature published on polymers
and polymer nanocomposites used single-edge notch beams (subjected to three-point bend loading)
to determine fracture toughness values (K1C ). Impact loading methods, such as Charpy and Izod
impact tests, are also used to determine impact fracture toughness. Fracture toughness values obtained
through different techniques cannot be directly compared [91].
Fracture can be defined as the mechanical separation of a solid owing to the application of stress.
Ductile and brittle are the two broad modes of fracture, and fracture toughness is related to the
amount of energy required to create fracture surfaces. In ideally-brittle materials (such as glass), the
energy required for fracture is simply the intrinsic surface energy of the materials, as demonstrated
by Griffith [92]. For structural alloys at room temperature, considerably more energy is required for
fracture, because plastic deformation accompanies the fracture process. In polymer nanocomposites,
the fracture path becomes more tortuous as cracks detour around strong reinforcement. This increase in
crack tortuosity provides additional work to fracture and, therefore, an increase in fracture toughness.
In polymers, the fracture process is usually dominated by crazing or the nucleation of small cracks and
their subsequent growth [93].
Toughness is defined as the ability of a material to absorb energy before fracture takes place. It
is usually characterized by the area under a stress–strain curve for a smooth (un-notched) tension
specimen loaded slowly to fracture. The term fracture toughness is usually associated with the fracture
mechanics methods that deal with the effect of defects on the load-bearing capacity of structural
components. The fracture toughness of materials is of great significance in engineering design because
of the high probability of flaws being present. Defined another way, it is the critical stress intensity at
which final fracture occurs. The plane strain fracture toughness (critical stress intensity factor, K1C )
can be calculated for a single-edge notched three-point bending specimen using Equation (1), where
Pmax is the maximum load of the load–displacement curve (N), f (a/w) is a constant related to the
geometry of the sample and is calculated using Equation (2), B is sample thickness (mm), W is sample
width (mm), and a is crack length (a should be kept between 0.45 W and 0.55 W, according to ASTM
D5045) [72]. The critical strain energy release rate (G1C ) can be calculated using Equation (3), where
E is the Young’s modulus obtained from the tensile tests (MPa), and ν is the Poisson’s ratio of the
polymer. The geometric function f(a/W) strongly depends on the a/W ratio [94].
The fracture toughness is dependent on many factors, such as type of loading and environment
in which the system will be loaded [95]. However, the key defining factor is the microstructure as
summed up in Figure 3 [96]. The properties of nanocomposites are also significantly dependent on filler
shape and size [51]. The graphene size, shape, and topography can be controlled simultaneously [97].
a
Pmax f p W q
K1C “ (1)
BW 1{2
`a˘
” !
f W )ı
a ´13.32 a 2 `14.72 a 3 ´5.6 4
p2` Wa q 0.0866`4.64p W q pW q pW q p a
W q (2)
“ a 3{2
p 1´ W q
` ˘
K 2 1 ´ ν2
G1c “ 1c (3)
E
Polymers 2016, 8, 281 9 of 37
Polymers 2016, 8, 281 8 of 35
in fracture toughness, among all other nano-reinforcements, can be obtained using graphene—mainly
because of its better capability of deflecting the propagating cracks [99,100].
As graphene is a 2D structure, each carbon atom can undergo chemical reaction from the sides,
resulting in high chemical reactivity. The carbon atoms on the edge of graphene sheet have three
incomplete bonds (in single layer graphene) that impart especially high chemical reactivity to edge
carbon atoms. In addition, defects within a graphene sheet are high energy sites and preferable sites
for chemical reactants. All these factors make graphene a very highly chemical reactive entity. The
graphene oxide can be reduced by using Al particles and potassium hydroxide [101]. The graphene
structure can be studied using Transmission Electron Microscopy (TEM) and other high-resolution
tools. Wrinkles were observed in graphene flat sheet, which were due to the instability of the 2D lattice
structure [72,102].
Wrinkling is a large and out-of-plane deflection caused by compression (in-plane) or shear.
Wrinkling is usually found in thin and flexible materials, such as cloth fabric [103]. Graphene
nanosheets (GNSs) were also found to undergo a wrinkling phenomenon [104]. When wrinkling
takes place, strain energy is stored within GNSs which is not sufficient to allow the GNSs to regain
their shape. Wrinkling can be found on GNSs as well as on exfoliated graphite. The wrinkles in
GNSs are sundering apart at different locations while getting closer at other regions. As GNSs do not
store sufficient elastic strain energy, wrinkling is an irreversible phenomenon, but can be altered by
external agency [105]. The surface roughness varies depending on graphene sheets, owing to their
dissimilar topographical features, such as wrinkles’ size and shape. Therefore, the ability of sheets
to mechanically interlock with other sheets and polymer chains is dissimilar. Wang et al. showed
that a wrinkle’s wavelength and amplitude are directly proportional to sheet size (length, width, and
thickness), as is clear from Equations (4) and (5), where λ is wrinkle wavelength, ν is Poisson’s ratio, L
is graphene sheet size, t is thickness of graphene sheet, ε is edge contraction on a suspended graphene
sheet, and A is wrinkle amplitude [57].
Palmeri et al. showed that the graphene sheets have a coiled structure that helps them to store a
sufficient amount of energy [106]. The individual sheet and chunk of sheets together are subjected to
plastic deformation at the application of external load. The applied energy is utilized in undertaking
plastic work that enhances the material’s ability to absorb more energy. It is believed that large
graphene sheets have large size wrinkles [107]. These wrinkles twist, bend, and fold the graphene
sheets. The wrinkles and other induced defects remain intact while curing of polymer matrix. This
reduces the geometric continuity and regularity of graphene and lowers load transfer efficiency, and
can cause severe localized stress concentration.
4π2 νL2 t2
λ4 « ` ˘ (4)
3 1 ´ ν2 ε
16νL2 t2 ε
A4 « ` ˘ (5)
3π2 1 ´ ν2
area of graphene is calculated using the formula for a rectangular sheet. The thickness of graphene is
considered variable, so the same relation can be used for multiple layers of graphene sheets stacked
together in the form of graphene nanosheets. The length of short carbon fiber is taken to be 1 µm and
the diameter 0.1 µm. The dimensions of graphene are ` ˆ w ˆ t = 1 µm ˆ 0.1 µm ˆ 10 nm. The density
of both short carbon fiber and graphene is taken to be 2.26 to make comparison based on dimensions
only. The surface area of 1 g of carbon fibers is 19 m2 and that of graphene is 98 m2 . It can be observed
that although the lengths of both reinforcements are the same and the width of graphene is equal to
the diameter of a short carbon fiber, there is a large difference in surface areas when the thickness
of graphene is kept 10 nm. This difference will further increase if graphene dimensions are reduced.
The specific surface area of graphene is as high as 2600 m2 /g [111,112]. It shows that graphene,
having a much larger surface area, can significantly improve the fracture toughness of the epoxy
nanocomposites [113,114]. There is also improved thermal conduction among graphene–graphene
links that significantly improves the overall thermal conductivity of the nanocomposites [115,116].
The electrical conductivity also increases with graphene as graphene sheets form links and provide a
passageway for electrical conduction [117].
Zhao and Hoa used a theoretical computer simulation approach to study the improvement in
toughness when epoxy is reinforced with 2D nano-reinforcements of different particle size [118,119].
The simulation results showed that there is a direct relation between particle size and stress
concentration factor up to 1 µm, after which point the stress concentration factor was impervious to any
further size increase. However, Chatterjee et al. [82] showed that fracture toughness was improved by
increasing the graphene size, which is in negation with simulation results by Zhao and Hoa [120,121].
The relationship between graphene size and stress concentration factor can be correlated with the
facile analogy of substitutional solid solution. The extent of strain field produced by a foreign atom
depends upon the difference in atomic sizes of the foreign and parent atoms. When there is a large
difference between foreign and parent atoms, a large strain field around the atom is generated. On the
contrary, when the difference in atomic sizes of parent and foreign atoms is small, the strain field is
limited. As both atomic and GNPs sizes are in the nano-meter range, the analogy can arguably be
applied to an epoxy–graphene system where large sheet size will cause higher stress concentration
factor than that produced by small sheet size. Therefore, graphene with smaller sheet size can be more
efficient in improving the fracture toughness than the larger graphene sheets.
The increase in the fracture toughness of epoxy was found to be strongly dependent upon the
graphene sheet size [57]. For the nanocomposites, an inverse relation was found between sheet size
and fracture toughness in most cases. The increase in fracture toughness with a decrease in sheet size
can be explained on the basis of stress concentration factor, as discussed above. Although graphene
acts as reinforcement, however, it has associated stress and strain fields which arise from the distortion
of the structure of polymer matrix. When sheet size, weight fraction, or both are increased beyond a
certain value, the stress concentration factor dominates the reinforcing character. As a result, fracture
toughness and other mechanical properties—such as tensile and flexural strength and stiffness—start
decreasing, which is in accordance with Zhao and Hoa’s simulation results [118].
Wang et al. used Graphene Oxide (GO) of three different sizes, namely GO-1, GO-2, GO-3,
having average diameters 10.79, 1.72, and 0.70 µm, respectively, to produce nanocomposites using
an epoxy matrix [57]. They observed that fracture toughness was strongly dependent on GO sheet
size. The maximum increase in fracture toughness was achieved with the smallest GO sheet size.
The K1C values dropped when weight fraction was increased beyond 0.1 wt %. This decrease in K1C
with increasing weight fraction can be correlated with crack generation and dispersion state.
toughness of the nanocomposite [57]. The other reason for such behavior is due to the high probability
of agglomeration at higher weight fractions due to Van der Waals forces [57].
The weight fractions of reinforcements at which maximum K1C was achieved for different
Polymers 2016, 8, 281 nanocomposites are shown in Figure 4. All the published research articles 11
epoxy–graphene stated
of 35
that the maximum K1C values were achieved at or below 1 wt % of graphene, and K1C dropped when
the weight fraction of graphene was was raised
raised beyond
beyond 11 wt
wt %.%. The
The decrease
decrease in in KK1C
1C with a higher weight
140
120
Increase in K1C (% )
100
80
60
40
20
0
0.1 0.2 0.25 0.3 0.4 0.5 1
Reinforcement (Wt%)
Figure 4. The
Figure 4. The weight
weight fractions
fractions of
of reinforcements
reinforcements at
at which
which maximum
maximum K K1C
1C was
was achieved
achieved in different
in different
epoxy/graphene
epoxy/graphene nanocomposites
nanocompositesand andcorresponding
correspondingimprovement
improvement(%)
(%)in inK1CK1C
(See references
(See in
references
Table 1).
in Table 1).
It can be observed from Figure 4 that there is no fixed value of GNPs wt % at which a maximum
It can be observed from Figure 4 that there is no fixed value of GNPs wt % at which a maximum
increase in K1C is achieved. In addition, the increase in K1C at fixed GNP wt % is not the same. For
increase in K1C is achieved. In addition, the increase in K1C at fixed GNP wt % is not the same.
example, at 0.5 wt %, the % increase in K1C is reported to be up to 45% by Chandrasekaran et al. [67], and
For example, at 0.5 wt %, the % increase in K1C is reported to be up to 45% by Chandrasekaran et al. [67],
about 110% by Ma et al. [80]. Therefore, it can be concluded that the wt % of GNPs is not the sole
and about 110% by Ma et al. [80]. Therefore, it can be concluded that the wt % of GNPs is not the
factor defining the influence of GNPs on the mechanical properties of nanocomposites. There are
sole factor defining the influence of GNPs on the mechanical properties of nanocomposites. There are
other influential factors as well, such as dispersion method, use of dispersant, and functionalization.
other influential factors as well, such as dispersion method, use of dispersant, and functionalization.
In addition, the use of organic solvent is another important parameter in defining the improvement
In addition, the use of organic solvent is another important parameter in defining the improvement
in mechanical properties. It was observed that a lower improvement in K1C was observed when
in mechanical properties. It was observed that a lower improvement in K1C was observed when
dispersion was carried out with only sonication, and a higher improvement in K1C was observed
dispersion was carried out with only sonication, and a higher improvement in K1C was observed when
when sonication was assisted with a secondary dispersion method, especially mechanical stirring.
sonication was assisted with a secondary dispersion method, especially mechanical stirring.
8. Dispersion State and Fracture Toughness
The end product of most of the graphene synthesis methods is agglomerated graphene [33]. In
addition, graphene tends to agglomerate due to weak intermolecular Van der Waals forces [113].
Therefore, dispersing graphene in epoxy matrix is always a challenge. The relationship between
dispersion state and the nature of crack advancement is shown schematically in Figure 5. The
advancing cracks can be best barricaded by uniformly dispersed graphene. Tang et al. produced
Polymers 2016, 8, 281 13 of 37
Figure
Figure 5. Influence
5. Influence of of graphenedispersion
graphene dispersion on
on crack
crack propagation
propagation method;
method;(a)(a)poorly dispersed
poorly dispersed
graphene;
graphene; (b) (b) Ideally
Ideally uniformly dispersed
uniformly dispersed graphene.
graphene. The The
arrows indicate
arrows the paththe
indicate followed by cracks by
path followed
cracksthrough thethe
through graphene
graphenesheets.
sheets.
Several dispersion modes to disperse reinforcement into epoxy matrix were successfully
adopted (see references in Table 1). The maximum % increase in K1C as a function of dispersion mode
is shown in Figure 6. In most of these articles, sonication is the main mode of dispersing reinforcement
in epoxy matrix. It can be observed that when sonication is assisted by a supplementary dispersion
technique (such as mechanical stirring and magnetic stirring), the K1C values were significantly
Polymers 2016, 8, 281 14 of 37
Several dispersion modes to disperse reinforcement into epoxy matrix were successfully adopted
(see references in Table 1). The maximum % increase in K1C as a function of dispersion mode is shown
in Figure 6. In most of these articles, sonication is the main mode of dispersing reinforcement in epoxy
matrix. It can be observed that when sonication is assisted by a supplementary dispersion technique
Polymers 2016, 8, 281 13 of 35
(such as mechanical stirring and magnetic stirring), the K1C values were significantly increased.
The maximum improvement of 131% in K1C was achieved when a combination of sonication and
sonication and mechanical stirring was employed [59]. The second highest improvement in K1C was
mechanical stirring was employed [59]. The second highest improvement in K1C was achieved with a
achieved with a combination of sonication and magnetic stirring, an increase in K1C of 109% [80]. The
combination of sonication and magnetic stirring, an increase in K1C of 109% [80]. The minimum values
minimum values in K1C improvements are achieved when sonication is coupled with ball milling
in K1C improvements are achieved when sonication is coupled with ball milling [60,64,100]. Since both
[60,64,100]. Since both the sonication and ball milling processes reduce the sheet size and produce
the sonication and ball milling processes reduce the sheet size and produce surface defects [120–134],
surface defects [120–134], we believe that the surface defects significantly increased and sheet size
we believe that the surface defects significantly increased and sheet size was reduced below the
was reduced below the threshold value, and therefore a greater improvement in K1C was not
threshold value, and therefore a greater improvement in K1C was not achieved. Although three roll
achieved. Although three roll milling (3RM, calendering process) is an efficient way of dispersing the
milling (3RM, calendering process) is an efficient way of dispersing the reinforcement into the polymer
reinforcement into the polymer matrix due to high shear forces, the maximum improvement in K1C
matrix due to high shear forces, the maximum improvement in K1C using three roll mill was reported
using three roll mill was reported as 86% [77], which is far below that achieved with a combination
as 86% [77], which is far below that achieved with a combination of sonication and mechanical stirring
of sonication and mechanical stirring (131% [59]).
(131% [59]).
150
Increase in K1C (%)
100
50
Mode of dispersion
Figure 6.
Figure 6. The
Themaximum
maximumimprovement in Kin1CK
improvement as1Ca function of dispersion
as a function mode. mode.
of dispersion (See references in Table
(See references
1). Table 1).
in
9. Functionalization
9. Functionalizationand
andFracture
FractureToughness
Toughness
Achieving
Achieving maximum
maximum improvement
improvement in in fracture toughness of
fracture toughness polymers by
of polymers by using
using graphene
graphene
depends on the ability to optimize the dispersibility of graphene and the interfacial
depends on the ability to optimize the dispersibility of graphene and the interfacial interactions interactions with
the epoxy matrix. As described previously, graphene tends to agglomerate
with the epoxy matrix. As described previously, graphene tends to agglomerate due to the weak due to the weak Van der
Waals
Van derinteractions, and its smoother
Waals interactions, surface texture
and its smoother surfaceinhibits
texturestrong
inhibitsinterfacial interactions.
strong interfacial To tackle
interactions.
To tackle the limited dispersibility and interfacial bonding of graphene, surface modifications out
the limited dispersibility and interfacial bonding of graphene, surface modifications are carried are
[135–139]. In fact, the introduction of functional groups on the graphene surface
carried out [135–139]. In fact, the introduction of functional groups on the graphene surface can can induce novel
properties
induce [140–144].
novel Various
properties methods
[140–144]. to modify
Various the graphene
methods to modify surface
the have been employed,
graphene surface have andbeen
can
be categorized into two main groups, namely: (1) chemical functionalization;
employed, and can be categorized into two main groups, namely: (1) chemical functionalization; and and (2) physical
functionalization.
(2) physical functionalization.
In chemical functionalization,
In chemical functionalization, chemical
chemical entities
entities are
are typically
typically attached
attached covalently.
covalently. For
For example,
example,
in defect functionalization, functional groups are attached at the defect
in defect functionalization, functional groups are attached at the defect sites of graphene, such as sites of graphene,
such as –COOH
–COOH (carboxylic(carboxylic
acid) andacid)
–OHand –OH (hydroxyl)
(hydroxyl) groups.can
groups. Defects Defects
be anycan be any departure
departure from
from regularity,
regularity, including pentagons and heptagons in the hexagonal structure
including pentagons and heptagons in the hexagonal structure of graphene. Defects may also of graphene. Defects may
be
also be produced by reaction with strong acids such as HNO 3, H2SO4, or their mixture, or strong
produced by reaction with strong acids such as HNO3 , H2 SO4 , or their mixture, or strong oxidants
oxidants including
including KMnO4 , KMnOozone,4,and
ozone, and reactive
reactive plasmaThe
plasma [145]. [145]. The functional
functional groups groups
attached attached at the
at the defect
defect sites of graphene can undergo further chemical reactions, including but not limited to
silanation, thiolation, and esterification [146]. Unlike chemical functionalization, physical
functionalization has non-covalent functionalization, where the supermolecular complexes of
graphene are formed as a result of the wrapping of graphene by surrounding polymers [33].
Surfactants lower the surface tension of graphene, thereby diminishing the driving force for the
Polymers 2016, 8, 281 15 of 37
sites of graphene can undergo further chemical reactions, including but not limited to silanation,
thiolation, and esterification [146]. Unlike chemical functionalization, physical functionalization has
non-covalent functionalization, where the supermolecular complexes of graphene are formed as a
result of the wrapping of graphene by surrounding polymers [33]. Surfactants lower the surface
Polymers
tension2016, 8, 281
of graphene, thereby diminishing the driving force for the formation of aggregates. 14 of 35
The
graphene dispersion can be enhanced by non-ionic surfactants in case of water-soluble polymers [33].
The different functionalization methods adopted to study their influence on K1C values with
The different functionalization methods adopted to study their influence on K1C values with
corresponding improvements (%) in K1C values are shown in Figure 7. The minimum improvement
corresponding improvements (%) in K1C values are shown in Figure 7. The minimum improvement was
was achieved for amino-functionalized graphene oxide (APTS-GO) [74], while the maximum
achieved for amino-functionalized graphene oxide (APTS-GO) [74], while the maximum improvement
improvement was recorded for surfactant-modified graphene nanoplatelets [59]. This could be
was recorded for surfactant-modified graphene nanoplatelets [59]. This could be attributed to the
attributed to the improvement in the dispersion state of graphene in the epoxy matrix when
improvement in the dispersion state of graphene in the epoxy matrix when surfactants were used, in
surfactants were used, in addition to improving interactions without causing a reduction in graphene
addition to improving interactions without causing a reduction in graphene sheet size or imparting
sheet size or imparting surface defects on graphene sheets.
surface defects on graphene sheets.
140
120
Increase in K1C (%)
100
80
60
40
20
Type of functionalization
Themaximum
Figure7.7.The
Figure maximumimprovement inKK1C1Casasa afunction
improvementin functionofoffunctionalization
functionalizationmethod.
method.(See
(Seereferences
references
in Table 1).
in Table 1).
10.
10.Crosslink
CrosslinkDensity
Densityand
andFracture
FractureToughness
Toughness
In thermosettingmaterials,
In thermosetting materials, suchsuch as epoxy,
as epoxy, high crosslink
high crosslink density isdensity
desirableis for
desirable for the
the improvement
improvement
of mechanicalofproperties.
mechanical However,
properties.highHowever, highdensity
crosslink crosslink density
has has a detrimental
a detrimental effect oneffect on
fracture
fracture
toughness toughness [57]. Therefore,
[57]. Therefore, a crosslink
a crosslink density density
threshold threshold is required
is required in orderintoorder to achieve
achieve optimal
optimal
propertiesproperties
[147,148].[147,148].
During theDuring
curing the
ofcuring of thermoset
thermoset polymers,polymers,
while phasewhile phase transformation
transformation takes place,
takes place,
graphene graphene
sheets tend tosheets tend to in
agglomerate agglomerate in order
order to reduce to reduce configurational
configurational entropy [57].
entropy [57]. Additionally, the
Additionally, the reduces
viscosity initially viscositywhen
initially reduces when
the temperature the temperature
is increased is increased
during curing, whichduring
makescuring, which
the movement
makes the movement
of the graphene sheetsof the graphene
relatively sheets relatively
easy, supporting easy, supporting
their agglomeration. Duetheir agglomeration.
to the Due to
wrinkle-like structure
the wrinkle-like structure and high specific surface area of graphene, strong interfacial
and high specific surface area of graphene, strong interfacial interactions are possible with epoxy interactions
are possible
chains. It may with
alsoepoxy chains.
affect the It curing
overall may also affectbythe
reaction overall the
changing curing reaction
maximum by changing
exothermic the
heat flow.
maximum
Molecular exothermic heat flow.
dynamics studies Molecular
conducted dynamics
by Smith studies
et al. also conducted
showed by Smith
that there was a et al. also
change in showed
polymer
that
chainthere was acaused
mobility changeby ingeometric
polymer chain mobility
constraints caused
at the by geometric
surface constraints at
of nano-reinforcement the surface of
[149].
nano-reinforcement
The graphene affects[149]. the crosslink density of epoxy [65]. When graphene is dispersed in epoxy,
The graphene
the polymer chainsaffects the crosslink
are restricted, density of epoxy
and crosslinking [65]. When
is decreased. Thegraphene
decrease isindispersed
crosslinkingin epoxy,
lowers
the polymer chains are restricted, and crosslinking is decreased. The decrease in crosslinking lowers
the heat release rate. It was reported that both graphene platelets (GnPs) and polybenzimidazole
functionalized GnPs (fGnPs) decreased the heat release rate of the curing reaction and increased the
curing temperature [65]. It can also be attributed to the dispersion state of the reinforcement.
Uniformly dispersed reinforcement will have a more pronounced effect on heat release rate and
curing temperature than poorly dispersed reinforcement. Therefore, fGnPs have a better dispersion
Polymers 2016, 8, 281 16 of 37
the heat release rate. It was reported that both graphene platelets (GnPs) and polybenzimidazole
functionalized GnPs (fGnPs) decreased the heat release rate of the curing reaction and increased
the curing temperature [65]. It can also be attributed to the dispersion state of the reinforcement.
Uniformly dispersed reinforcement will have a more pronounced effect on heat release rate and curing
temperature than poorly dispersed reinforcement. Therefore, fGnPs have a better dispersion state
than GnPs [65]. There are two opposite effects of filler in the matrix: (1) the fillers could restrict the
polymer chains, which should increase Tg ; (2) the reactive fillers could lower the crosslinking density
of epoxy, which should decrease Tg . An increase in Tg shows that interfacial interactions dominate the
crosslinking density effect [65].
The G1C also improved with the incorporation of graphene in epoxy nanocomposites. Meng et al.
produced epoxy–graphene nanocomposites and reported an increase in G1C of up to 597% [173].
heat from electronic devices may also be barricaded when the high thermal conductivity of graphene is
conductivity of polymers
efficiently utilized. than CNTs
The graphene [175].higher
has shown It has been found
efficiency experimentally
in increasing that the Effective
the thermal conductivity of
Thermal Conductivity (K eff) of graphene-based polymer nanocomposites
polymers than CNTs [175]. It has been found experimentally that the Effective Thermal Conductivityhas a non-linear
dependence on graphene polymer
(Keff ) of graphene-based weight nanocomposites
fraction [176–178].
has a Xie et al. dependence
non-linear proposed an analyticalweight
on graphene model to
determine
fractionthe Keff of graphene-based
[176–178]. Xie et al. proposednanocomposites [179].toTheir
an analytical model modelthe
determine proposed very high thermal
Keff of graphene-based
nanocomposites
conductivity values,[179]. Their
as the modelmodel
didproposed
not take very
into high thermal
account the conductivity values, resistance.
interfacial thermal as the modelLin et
al. developed a model based on Maxwell–Garnett effective medium approximation based
did not take into account the interfacial thermal resistance. Lin et al. developed a model theory to
on Maxwell–Garnett
determine effectiveconductivity
the effective thermal medium approximation theory tonanocomposites
of graphene-based determine the effective thermal
[180,181].
conductivity of graphene-based nanocomposites [180,181]. They showed that the enhancement in
They showed that the enhancement in thermal conductivity is strongly influenced by the aspect ratio
thermal conductivity is strongly influenced by the aspect ratio and orientation of graphene.
and orientation of graphene.
120
Tensile
100
strength
% increase
80
60
40
20
0
Authors
Figure 8. The % increase in tensile properties of epoxy/graphene nanocomposites [164,165,182–191].
Figure 8. The % increase in tensile properties of epoxy/graphene nanocomposites [164,165,182–191].
Hu Hu et used
et al. al. used a molecular
a molecular dynamicsapproach
dynamics approachto
toshow
show that
that the
the agglomeration
agglomerationofofgraphene
graphene is is of
of major concern in increasing the thermal conductivity of the system [192]. The variation in thermal
major concern in increasing the thermal conductivity of the system [192]. The variation in thermal
conductivity with various forms of graphene and graphite nanocomposites is summarized in Table 2,
conductivity with various forms of graphene and graphite nanocomposites is summarized in Table
and the influence of dispersion mode on the improvement of thermal conductivity is shown in Figure 9.
2, and the influence of dispersion mode on the improvement of thermal conductivity is shown in
The maximum improvement in thermal conductivity was observed in the case of mechanical stirring.
Figure 9. The sonication
In general, maximum improvement
caused in thermalin conductivity
a lower improvement was observed
thermal conductivity. However,in the case of
maximum
mechanical
improvementstirring. In general,
in thermal sonication
conductivity caused
(not shown a lower
in Figure improvement
9) was observed in in
thethermal conductivity.
case of sonication,
However,
1.6 ˆ 10maximum
4 % [193]. improvement in thermal conductivity (not shown in Figure 9) was observed in
the case of sonication, 1.6 × 104% [193].
% increase in thermal
4000
conductivity
3000
2000
1000
0
BM
SM
M
M
S
n
major concern in increasing the thermal conductivity of the system [192]. The variation in thermal
conductivity with various forms of graphene and graphite nanocomposites is summarized in Table
2, and the influence of dispersion mode on the improvement of thermal conductivity is shown in
Figure 9. The maximum improvement in thermal conductivity was observed in the case of
mechanical stirring. In general, sonication caused a lower improvement in thermal conductivity.
However,
Polymers 2016,maximum
8, 281 improvement in thermal conductivity (not shown in Figure 9) was observed in 18 of 37
the case of sonication, 1.6 × 104% [193].
% increase in thermal
4000
conductivity
3000
2000
1000
0
BM
SM
3RM
Sn+ShM
MS
Sn
Dispersion method
Figure 9. Percent increase in thermal conductivity as a function of dispersion method (see references
Figure 9. Percent increase in thermal conductivity as a function of dispersion method (see references
in Table 2).
in Table 2).
Polymers 2016, 8, 281 17 of 35
14. Electrical Properties
14. Electrical Properties
Tailoring the electrical properties of graphene can unlock its many potential electronic
Tailoring the electrical properties of graphene can unlock its many potential electronic
applications [194,195]. For example, effective gauge fields are introduced when graphene lattice
applications [194,195]. For example, effective gauge fields are introduced when graphene lattice
deformation takestakes
deformation place. LikeLike
place. the effective magnetic
the effective field,field,
magnetic the produced
the producedeffective gauge
effective fieldsfields
gauge influence
the Dirac
influence the Dirac fermions [196]. The Fermi level in undoped graphene lies at the Dirac point, where the
fermions [196]. The Fermi level in undoped graphene lies at the Dirac point, where
minimum conductivity
the minimum valuesvalues
conductivity are achieved [197].
are achieved By By
[197]. adding
addingfree
freecharge
chargecarriers
carriers (i.e., dopants), the
(i.e., dopants),
the electrical
electrical properties
properties of graphene
of graphene cancan
be be improved,and
improved, and conductivity
conductivity increases
increases linearly withwith
linearly carrier
carrier
density
density [198,199].
[198,199]. ForFor example,
example, boronas
boron asdopant
dopant can
can contribute
contribute~0.5~0.5carriers perper
carriers dopant in a graphene
dopant in a graphene
sheetsheet [200].
[200]. Dopants
Dopants canbebeintroduced
can introduced during
during thethesynthesis
synthesisofof graphene
graphene using chemical
using vaporvapor
chemical
deposition (CVD) [201]. The variation in electrical conductivity with various forms of graphene and
deposition (CVD) [201]. The variation in electrical conductivity with various forms of graphene and
graphite nanocomposites is summarized in Table 3, and the influence of dispersion mode on the
graphite nanocomposites is summarized in Table 3, and the influence of dispersion mode on the
improvement of thermal conductivity is shown in Figure 10. The maximum improvement in electrical
improvement of thermal conductivity is shown in Figure 10. The maximum improvement in electrical
conductivity was observed in the case of a combination of ball milling and mechanical stirring.
conductivity
Therefore, bothobserved
was thermal and inelectrical
the caseconductivities
of a combination
improved of in
ball
the milling and mechanical
case of mechanical stirring. stirring.
Therefore, both thermal and electrical conductivities improved in the case of mechanical stirring.
% increase in electrical conductivity (Trillions)
1800000
1600000
1400000
1200000
1000000
800000
600000
400000
200000
0
Dispersion method
Figure 10. Percent increase in electrical conductivity as a function of dispersion method (see reference
in Table 3). 10. Percent increase in electrical conductivity as a function of dispersion method (see reference
Figure
in Table 3).
Polymers 2016, 8, 281 19 of 37
Table 2. A brief record of epoxy-based nanocomposites studied for improvement in thermal conductivity values.
% Increase
Dispersion
Sr. Authors Year Reinforcement (wt %) in thermal Remarks Ref.
method
conductivity
GnP (1.9 wt %) 9
The simultaneous inclusion of GnPs and SnP/SnW at a combined loading of 1 vol % resulted in about 40%
SnP/(0.09 wt %) 18 enhancement in the through-thickness thermal conductivity, while the inclusion of GnP at the same loading
1 Kandre et al. 2015 Sn [202]
SnW/(0.09 wt %) 8 resulted in only 9% improvement. A higher increment with simultaneous addition of GnP and SnP/SnW can
be attributed to synergistic effects.
GnP (1.9 wt %), SnP (0.09 wt %) 38
GnP (1.9 wt %), SnW (0.09 wt %) 40
Three-dimensional graphene network (3DGNs) (30 wt %) None 1,900 (Composites produced using layer-by-layer dropping method.) The filler with large size is more effective in
2 Tang et al. 2015 increasing the thermal conductivity of epoxy because of continuous transmission of acoustic phonons and [203]
Chemically reduced graphene oxide (RGO) (30 wt %) 1,650
Sn + MS minimum scattering at the interface due to reduced interfacial area. High intrinsic thermal conductivity of
Natural graphite powder (NG) (30 wt %) 1,400 graphene is the major reason for the obtained high thermal conductivity of nanocomposites.
Graphite flakes (12 wt %) (GRA-12) 237.5 As the filler/matrix interfaces increase, the thermal resistance increases due to phonon scattering. In order to
improve the thermal conductivity of a composite, it is better to structure a sample with an adapted
Graphite flakes (15 wt %) (GRA-15) 325
morphology than trying to have the best dispersion. A 3D-network was first prepared with graphite foils
3 Burger et al. 2015 Sn + MgSr [204]
Graphite flakes (14–15 wt %) (Network) 775 oriented through the thickness of the sample and then stabilized with DGEBA/DDS resin. The produced
composite sample was called as “Network”. In “fibers”, all the graphite flakes were aligned through the
Graphite flakes (11–12 wt %) (Fibers) 666.7
thickness of sample. When a DGEBA interface layer was applied in “fiber”, the sample was called
Graphite flakes (11–12 wt %) (Fiber + 1 interface) 608.3 “Fiber + 1 interface”. When two DGEBA interface layers was applied in “fiber” the sample was
Graphite flakes (11–12 wt %) (Fiber + 2 interface) 237.5 called as “Fiber + 2 interfaces”.
Liquid crystal perylene bisimides polyurethane (LCPU) Along with the increase in thermal conductivity, the impact and flexural strengths increased up to 68.8% and
4 Zeng et al. 2015 Sn 44.4 [205]
modified reduced graphene oxide (RGO) (1 wt %) 48.5%, respectively, at 0.7 wt % LCPU/RGO.
GnPs, 1 µm, (GnP-C750) 9.1
5 Wang et al. 2015 Sn + MgSr + 3RM The increase in thermal conductivity is higher in the case of larger particle size than smaller particle size. [206]
GnPs, 5 µm 115
6 Zhou et al. 2015 Multi-layer graphene oxide (MGO) (2 wt %) Sn 95.5 The thermal conductivity decreases after 2 wt % MGO. [207]
Al2 O3 nanoparticles (30 wt %) 50
7 Zeng et al. 2015 Sn The thermal conductivity can be improved by using hybrid fillers. [208]
Aminopropyltriethoxy-silane modified Al2 O3 nanoparticles
68.8
(Al2 O3 -APS) (30 wt %)
Liquid-crystal perylene-bisimide polyurethane (LCPBI)
functionalized reduced graphene oxide (RGO) and 106.2
Al2 O3 -APS (LCPBI/RGO/Al2 O3 -APS)
Al2 O3 (18.4 wt %) 59.1
8 Tang et al. 2015 Sn + MS The increase in thermal conductivity decreases with Al2 O3 coating of graphite. [209]
Graphite (18.4 wt %) 254.6
Al2 O3 -coated graphite (Al2 O3 -graphite) (18.4 wt %) 195.5
Perylene bisimide (PBI)-hyper-branched polyglycerol
9 Pan et al. 2015 (HPG) modified reduced graphene oxide (RGO), Sn 37.5 The filler was observed to be uniformly dispersed, resulting in strong interfacial thermal resistance. [210]
(PBI-HPG/RGO) (1 wt %)
SiO2 , 15 nm, (1 wt %) 14.3
SiO2 nanoparticles are more effective in increasing thermal conductivity than GO. The maximum
10 Wang et al. 2015 Sn [211]
GO (1 wt %) 4.8 improvement in thermal conductivity was observed in the case of hybrid filler.
As-prepared nanosilica/graphene
28.6
oxide hybrid (m-SGO) (1 wt %)
GNPs (3.7 wt %), Al2 O3 nanoparticles (ANPs), (65 wt %) 550.4
11 Zha et al. 2015 Sn + MS Al2 O3 nanofibers are more effective in improving thermal conductivity than Al2 O3 nanoparticles. [212]
GNPs (3.7 wt %), Al2 O3 fibers (Afs) (65 wt %) 756.7
Polymers 2016, 8, 281 20 of 37
Table 2. Cont.
% Increase
Dispersion
Sr. Authors Year Reinforcement (wt %) in thermal Remarks Ref.
method
conductivity
12 Zhou et al. 2015 Multi-layer graphene oxide (MGO) (2 wt %) Sn 104.8 The thermal conductivity decreases after 2 wt % MGO. [213]
13 Wang et al. 2015 GNPs (8 wt %) MS 627 The thermal conductivity increases with GNPs at the loss of Vickers microhardness after 1 wt % of GNP. [214]
RGO (1 wt %) 21.8 The thermal conductivity decreases after 1 wt % RGO. The silica layer on S-graphene makes electrically
14 Pu et al. 2014 Sn + MgSr conducting graphene insulating, reduces the modulus mismatch between the filler and matrix, and improves [215]
3-aminopropyl triethoxysilane (APTES) functionalized
47.1 the interfacial interactions of the nanocomposites, which results in enhanced thermal conductivity.
graphene oxide (A-graphene) (8 wt %)
Silica-coated A-graphene (S-graphene) (8 wt %) 76.5
Graphite (44.30 wt %) 888.2
The maximum improvement in thermal conductivity was observed in the case of graphene sheets with
15 Fu et al. 2014 MS [216]
Graphite nanoflakes (16.81 wt %) 982.3 thickness of 1.5 nm.
Graphene sheets (10.10 wt %) 2258.8
The alignment of MLG causes an exceptional improvement in thermal conductivity and exceeds other
16 Li et al. 2014 Aligned MLG (AG) (11.8 wt %) Sn 16670 [193]
filler-based epoxy nanocomposites.
GNPs (25 wt %) Sn 780 Ball milling is more effective in improving the thermal conductivity of GNP/epoxy than sonication. The
17 Guo and Chen 2014 [126]
thermal conductivity decreases when ball milling is carried out for more than 30 h.
GNPs (25 wt %) BM 1420
Natural graphite (NG) (1 wt %) 24.1 The thermal conductivity decreases with increasing wt % of NG after 1 wt %. The thermal conductivity
Corcione and decreases after 2 wt % of GNPs. The maximum improvement in thermal conductivity was observed with
18 2013 Sn [217]
Maffezzoli GNPs (2 wt %) 89.8
expanded graphite.
Expanded graphite (EGS) (3 wt %) 232.1
19 Chandrasekaran et al. 2013 GNP (2 wt %) 3RM 14 The thermal conductivity increases with increasing temperature. [73]
High aspect ratio of GNPs and oxygen functional groups play a significant role in improving thermal
20 Min et al. 2013 GNPs (5 wt %) Sn 240 [218]
conductivity of nanocomposites.
Silica (1 wt %) 19
The existence of the intermediate silica layer enhances the interfacial attractions between TRGO and epoxy
21 Hsiao et al. 2013 Thermally reduced graphene oxide (TRGO) (1 wt %) Sn + ShM 26.5 [219]
and improved dispersion state, which caused a significant increase in thermal conductivity.
Silica nanosheets (Silica-NS) (1 wt %) 37.5
TRGO-silica-NS (1 wt %) 61.5
Untreated GNPs (12 wt %) 139.3
22 Zhou et al. 2013 Sn + MgSr Silane functionalization can significantly improve thermal conductivity of GNP/epoxy. [220]
Silane-treated COOH-MWCNTs (6 wt %) 192.9
Silane-treated GNPs (6 wt %) 525
GNPs, 5 µm, 30 wt %, in rubbery epoxy MS 818.6
GNPs, 5 µm, 20 wt %, in rubbery epoxy ShM 332.6 The thermal conductivity increases with increasing particle size. The particle size distribution significantly
GNPs, 15 µm, 25 wt %, in rubbery epoxy MS 1228.4 influences the thermal conductivity. GNPs with a broad particle size distribution gave higher thermal
23 Raza et al. 2012 [221]
conductivity than the particles with a narrow particle size distribution, due to the availability of smaller
GNPs, 15 µm, 25 wt %, in rubbery epoxy ShM 1118.2 particles that can bridge gaps between larger particles.
GNPs, 20 µm, 20 wt %, in rubbery epoxy ShM 684.6
GNPs, 20 µm, 12 wt %, in glassy epoxy ShM 567.6
GNPs, 15 µm, 20 wt %, in glassy epoxy MS 683
GO (3 wt %) 90.4
24 Kim et al. 2012 Sn The increase in thermal conductivity decreases with Al(OH)3 coating on GO. [222]
Al(OH)3 -coated graphene oxide (Al-GO) (3 wt %) 35.1
Polymers 2016, 8, 281 21 of 37
Table 2. Cont.
% Increase
Dispersion
Sr. Authors Year Reinforcement (wt %) in thermal Remarks Ref.
method
conductivity
Chatterjee et Amine functionalized expanded graphene nanoplatelets
25 2012 Sn + 3RM 36 The EGNPs form a conductive network in the epoxy matrix allowing for increased thermal conductivity. [83]
al. (EGNPs) (2 wt %)
Thermally conductive graphene oxide (GO) (50 wt %) 111 The thermal conductivity decreases after 50 wt %, which can be attributed to residual epoxy that forms an
26 Im and Kim 2012 Sn [223]
insulting layer on reinforcement. MWCNT helps the formation of 3D network structure.
Thermally conductive graphene oxide (GO) (50 wt %),
203.4
MWCNTs (0.36 wt %)
Al2 O3 (80 wt %), GO (5 wt %) 1,650
27 Heo et al. 2012 3RM The increase in thermal conductivity decreases with Al(OH)3 coating of GO. [224]
Al(OH)3 -coated GO (5 wt %) 1,450
MWNTs (65 wt %) 1,100
GNPs are more effective in improving thermal conductivity than MWNTs. The maximum improvement in
28 Huang et al. 2012 MS [225]
GNPs (65 wt %) 2,750 thermal conductivity was observed in the case of hybrid fillers.
MWNTs (38 wt %), GNPs (38 wt %) 3,600
MWNT (4 wt %) 160 GNPs showed a significantly greater increase in thermal conductivity than MWNTs. The maximum
29 Teng et al. 2011 Sn improvement in thermal conductivity is shown by non-covalent functionalized GNS, which can be attributed [114]
GNPs(4 wt %) 700
to high surface area and uniform dispersion of GNS.
Poly(glycidyl methacrylate containing localized pyrene
860
groups (Py-PGMA) functionalized GNPs (Py-PGMA-GNS)
MWNTs (1 wt %) in nanofluids 66.7
f-MWNTs (0.6 wt %) in nanofluids 20
The layered structure of MWNTs enables an efficient phonon transport through the inner layers, while SWNTs
30 Gallego et al. 2011 SWNTs (0.6 wt %) in nanofluids ShM 20 present a higher resistance to heat flow at the interface, due to its higher surface area. The f-MWNTs have [226]
Functionalized graphene sheet (FGS) (1 wt %) in nanofluids 0 functional groups on their surface, acting as scattering points for the phonon transport.
GO (1 wt %) in nanofluids 0
MWNTs(1 wt %) in nanocomposites 72.7
Functionalized graphene sheet (FGS) (1 wt %) in
63.6
nanocomposites
31 Tien et al. 2011 Graphene flakes (12 wt %) Sn 350 The thermal conductivity increases exponentially with increasing wt % of graphene flakes. [227]
Exfoliated graphite flakes (20 wt %) 2,087.2
32 Ganguli et al. 2008 SM The thermal conductivity increases with chemical functionalization. [177]
Chemically functionalized graphite flakes (20 wt %) 2,907.2
Carbon black (CB) (10 wt %) 75
SWNTs (10 wt %) 125 The hybrid filler demonstrates a strong synergistic effect and surpasses the performance of the individual
33 Yu et al. 2008 Sn + ShM [228]
SWNT and GNP filler.
GNPs (10 wt %) 625
GNPs (7.5 wt %), SWNTs (2.5 wt %) 775
Polymers 2016, 8, 281 22 of 37
Table 3. A brief record of epoxy-based nanocomposites studied for improvement in electrical conductivity values. HSM: high speed mixing.
% Increase
Dispersion
Sr. Authors Year Reinforcement/wt % in electrical Remarks Ref.
method
conductivity
GNPs (1.5 wt %), transverse to alignment Sn + 3RM 1 ˆ 107
The maximum thermal conductivity was observed in the case of
1 Wu et al. 2015 [229]
GNPs (3 wt %), randomly oriented 1 ˆ 108 aligned GNPs.
GNPs (3 wt %), parallel to alignment 1 ˆ 1010
(Samples were produced using resin infiltration.) The average
2 Liu et al. 2015 Graphene woven fabric (GWF) (0.62 wt %) None. 1 ˆ 1013 [230]
number of graphene layers in GWFs varied between 4 and 12.
(Samples were produced using hot pressing.) The electrical
conductivity of pure graphene foam (GF) is 2.9 S-cm-1 , which is
3 Ming et al. 2015 Graphene foam (GF) (80 wt %) None. 8 ˆ 102 [231]
much lower than graphene, which can be because of the presence
of structural defects.
GNPs (1.1 wt %) 1.39 ˆ 106 GNPs are more effective in improving the thermal conductivity of
5 Ghaleb et al. 2014 Sn [159]
epoxy than MWCNTs.
MWCNTs (1.9 wt %) 1.62 ˆ 105
GO (5 wt %) 1.92 ˆ 109 The surface functionalization of GO can significantly improve the
6 Tang et al. 2014 Sn + HSM [232]
electrical conductivity of GO–epoxy.
Diamine polyetheramine functionalized
1.92 ˆ 1012
GO (GO-D230) (5 wt %)
Ag–graphene can be used in electronic applications due to its high
7 Dou et al. 2014 Silver plated graphene (Ag-G) (25 wt %) Sn + MS 4.13 ˆ 102 [233]
electrical conductivity.
GO (3.6 wt %) 1 ˆ 1018 The surface functionalization significantly improves electrical
8 Tang et al. 2014 Sn [234]
conductivity.
Polyetheramine refluxed GO (GO-D2000)
1ˆ 1017
(3.6 wt %)
GNPs (3 wt %) 2.08 ˆ 105 The samples were produced using chloroform. [235]
9 Monti et al. 2013 Sn + MS
GNPs (3 wt %) 1.16 ˆ 105 The samples were produced using tetrahydrofuran.
10 Wajid et al. 2013 GNPs (0.24 wt %) Sn + MS 2.22 ˆ 103 The samples were produced using dimethylformamide. [189]
GNPs (1 wt %) Sn + ShM 1ˆ 104 3RM is more effective in improving the electrical conductivity of
11 Chandrakekaran et al. 2013 [73]
epoxy than sonication and high speed shear mixing.
GNPs (2 wt %) 3RM 1 ˆ 108
GNPs (80 wt %), CNTs (5 wt %),
7.30 ˆ 1017
through-plane
12 Suherman et al. 2013 BM + MS The electrical conductivity significantly increases with hybrid filler. [236]
GNPs (80 wt %), CNTs (5 wt %), in-plane 1.80 ˆ 1018
GNPs (80 wt %), through-plane 4 ˆ 1017
GNPs (80 wt %) in-plane 5 ˆ 1017
Polymers 2016, 8, 281 23 of 37
Table 3. Cont.
% Increase
Dispersion
Sr. Authors Year Reinforcement/wt % in electrical Remarks Ref.
method
conductivity
GO (0.5 wt %) 240 The conductivity was measured before post-curing.
GO (0.5 wt %) Sn 730 The conductivity was measured after post-curing.
13 Mancinelli et al. 2013 [237]
Octadecylamine (ODA)-treated partially
reduced and chemically modified GO (MGO) 550 The conductivity was reduced after functionalization.
(0.5 wt %)
GO (0.5 wt %) Two phase 240 The system was not fully cured during curing process.
GO (0.5 wt %) extraction 7.80 ˆ 103 The conductivity significantly increased after post-curing.
14 Al-Ghamdi et al. 2013 Foliated graphite nanosheets (FGNs) (40 wt %) Centrifugal mixing 9.90 ˆ 103 Dielectric properties of epoxy–FGN composites decreased with an increase in frequency. [238]
Al(OH)3 functionalized GO (Al-GO) (3 wt %) 75
15 Kim et al. 2012 MS + MgSr The increase in electrical conductivity decreases with Al(OH)3 functionalization of GO. [239]
GO (3 wt %) 115
Al2 O3 (80 wt %), Al(OH)3 functionalized GO
3RM 2.90 ˆ 103 The increase in electrical conductivity with Al(OH)3 functionalization decreased. The electrically
16 Heo et al. 2012 (Al-GO) (5 wt %) insulating Al(OH)3 on the graphene oxide nanosheet can prevent electron tunneling and act as ion [224]
Al2 O3 (80 wt %), GO (5 wt %) 4.90 ˆ 103 traps which block ion mobility, resulting in a decrease in the electrical properties of nanocomposites.
17 Tien et al. 2011 Graphite flakes (14 wt %) Sn 4 ˆ 107 The percolation threshold was 8 wt %. [227]
GNPs (5 wt %) 5.50 ˆ 1010
18 Fan et al. 2009 Sn + MS The maximum electrical conductivity was observed in the case of hybrid fillers. [240]
GNPs (4.5 wt %), carbon black (CB) (0.5 wt %) 5.50 ˆ 1012
19 Jovic et al. 2008 Expanded graphite (EG) (8 wt %) Sn 5.50 ˆ 1017 The electrical conductivity further increases with the application of electric field. [241]
20 Li et al. 2007 MWCNTs (1 wt %) Sn 4.63 ˆ 107 The samples were produced using acetone. [242]
21 Pecastaings et al. 2004 MWCNTs (20 wt %) Sn + MS 4.53 ˆ 103 The samples were produced using acetone. [243]
Polymers 2016, 8, 281 24 of 37
15. Conclusions
The following are the key points related to epoxy/graphene nanocomposites:
1. Epoxy is an excellent matrix for graphene composites because of its efficient properties such as
enhancement in composite mechanical properties, processing flexibility, and acceptable cost [2].
2. Graphene can significantly enhance the fracture toughness of epoxy nanocomposites—i.e., up to
131% [59]. When epoxy is reinforced with graphene, the carbonaceous sheets shackle the crack
and restrict its advancement. This obstruction and deflection of the crack by the graphene at the
interface is the foremost mechanism of raising the fracture toughness of nanocomposites.
3. The graphene sheets with smaller length, width, and thickness are more efficient in improving
the fracture toughness than those with larger dimensions [57]. Large graphene sheets have a
high stress concentration factor, because of which crack generation becomes easy in the epoxy
matrix [118,119]. The cracks deteriorate the efficiency of graphene in enhancing the fracture
toughness of epoxy/graphene nanocomposites.
4. Uniformly dispersed graphene improves fracture toughness significantly as compared to the
poorly dispersed graphene [72]. It is evident from the published literature that the fracture
toughness dropped when graphene weight fraction was increased beyond 1 wt %. The decrease in
fracture toughness with higher weight fraction of graphene can be correlated with the dispersion
state of graphene. As graphene weight fraction increases beyond 1 wt %, the dispersion state
becomes inferior.
5. Three roll milling or calendering process is an efficient way of dispersing the reinforcement into a
polymer matrix, as it involves high shear forces [244–248]. However, the maximum enhancement
in fracture toughness was achieved with a combination of sonication and mechanical stirring [59].
6. In thermosetting materials such as epoxy, high crosslink density is desirable for improved
mechanical properties. However, fracture toughness is dropped with high crosslinking [57].
7. The literature has proved the absence of consensus of graphene’s role in improving the mechanical
properties of nanocomposites [150–154]. Generally, graphene acts as panacea and raises the
mechanical properties [116,155–158]. On the contrary, it acts as placebo and shows no effect on
mechanical properties. Even worse, it is inimical and razes the mechanical properties [160–164].
The main factors that dictate graphene’s influence on the mechanical properties of epoxy
nanocomposites include topographical features, morphology, weight fraction, dispersion state,
surface modifications, and interfacial interactions.
Acknowledgments: The authors would like to thank the Department of Mechanical and Construction Engineering,
Northumbria University, UK for the provision of research facilities, without which the analysis of relevant data
was not possible.
Author Contributions: Rasheed Atif compiled the literature and wrote the manuscript. Islam Shyha and
Fawad Inam supervised the project and proofread the manuscript.
Conflicts of Interest: The authors declare no conflict of interest.
Abbreviations
The following abbreviations are used in this manuscript:
Polymers 2016, 8, 281 25 of 37
References
1. Carlson, R.L.; Kardomateas, G.A.; Craig, J.I. Mechanics of Failure Mechanisms in Structures, 1st ed.; Springer:
Berlin, Germany, 2012.
2. Miracle, D.B., Donaldson, S.L., Eds.; ASM Handbook Volume 21: Composites; ASM International: Materials
Park, OH, USA, 2001.
3. Yao, X.F.; Zhou, D.; Yeh, H.Y. Macro/microscopic fracture characterizations of SiO2 /epoxy nanocomposites.
Aerosp. Sci. Technol. 2008, 12, 223–230. [CrossRef]
4. Wetzel, B.; Rosso, P.; Haupert, F.; Friedrich, K. Epoxy nanocomposites—Fracture and toughening mechanisms.
Eng. Fract. Mech. 2006, 73, 2375–2398. [CrossRef]
5. Naous, W.; Yu, X.Y.; Zhang, Q.X.; Naito, K.; Kagawa, Y. Morphology, tensile properties, and fracture
toughness of epoxy/Al2 O3 nanocomposites. J. Polym. Sci. Part B 2006, 44, 1466–1473. [CrossRef]
6. Kim, B.C.; Park, S.W.; Lee, D.G. Fracture toughness of the nano-particle reinforced epoxy composite.
Compos. Struct. 2008, 86, 69–77. [CrossRef]
7. Wang, K.; Chen, L.; Wu, J.; Toh, M.L.; He, C.; Yee, A.F. Epoxy nanocomposites with highly exfoliated clay:
Mechanical properties and fracture mechanisms. Macromolecules 2005, 38, 788–800. [CrossRef]
8. Liu, W.; Hoa, S.V.; Pugh, M. Fracture toughness and water uptake of high-performance epoxy/nanoclay
nanocomposites. Compos. Sci. Technol. 2005, 65, 2364–2373. [CrossRef]
9. Gojny, F.H.; Wichmann, M.H.G.; Köpke, U.; Fiedler, B.; Schulte, K. Carbon nanotube-reinforced
epoxy-composites: Enhanced stiffness and fracture toughness at low nanotube content. Compos. Sci. Technol.
2004, 64, 2363–2371. [CrossRef]
10. Yu, N.; Zhang, Z.H.; He, S.Y. Fracture toughness and fatigue life of MWCNT/epoxy composites. Mater. Sci.
Eng. A 2008, 494, 380–384. [CrossRef]
11. Srikanth, I.; Kumar, S.; Kumar, A.; Ghosal, P.; Subrahmanyam, C. Effect of amino functionalized MWCNT on
the crosslink density, fracture toughness of epoxy and mechanical properties of carbon-epoxy composites.
Compos. Part. A Appl. Sci. Manuf. 2012, 43, 2083–2086. [CrossRef]
12. Mathews, M.J.; Swanson, S.R. Characterization of the interlaminar fracture toughness of a laminated
carbon/epoxy composite. Compos. Sci. Technol. 2007, 67, 1489–1498. [CrossRef]
13. Arai, M.; Noro, Y.; Sugimoto, K.I.; Endo, M. Mode-I and mode II interlaminar fracture toughness of CFRP
laminates toughened by carbon nanofiber interlayer. Compos. Sci. Technol. 2008, 68, 516–525. [CrossRef]
14. Wong, D.W.Y.; Lin, L.; McGrail, P.T.; Peijs, T.; Hogg, P.J. Improved fracture toughness of carbon fibre/epoxy
composite laminates using dissolvable thermoplastic fibres. Compos. Part. A 2010, 41, 759–767. [CrossRef]
15. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A.
Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [CrossRef] [PubMed]
16. Pokharel, P.; Truong, Q.-T.; Lee, D.S. Multi-step microwave reduction of graphite oxide and its use in
the formation of electrically conductive graphene/epoxy composites. Compos. Part B 2014, 64, 187–193.
[CrossRef]
17. Tian, M.; Qu, L.; Zhang, X.; Zhang, K.; Zhu, S.; Guo, X.; Han, G.; Tang, X.; Sun, Y. Enhanced mechanical and
thermal properties of regenerated cellulose/graphene composite fibers. Carbohydr. Polym. 2014, 111, 456–462.
[CrossRef] [PubMed]
18. Xu, Z.; Zhang, J.; Shan, M.; Li, Y.; Li, B.; Niu, J.; Zhou, B.; Qian, X. Organosilane-functionalized graphene oxide
for enhanced antifouling and mechanical properties of polyvinylidene fluoride ultrafiltration membranes.
J. Membr. Sci. 2014, 458, 1–13. [CrossRef]
19. Bkakri, R.; Sayari, A.; Shalaan, E.; Wageh, S.; Al-Ghamdi, A.A.; Bouazizi, A. Effects of the graphene
doping level on the optical and electrical properties of ITO/P3HT:Graphene/Au organic solar cells.
Superlattices Microstruct. 2014, 76, 461–471. [CrossRef]
20. Lian, Y.; He, F.; Wang, H.; Tong, F. A new aptamer/graphene interdigitated gold electrode piezoelectric
sensor for rapid and specific detection of staphylococcus aureus. Biosens. Bioelectron. 2014, 65, 314–319.
[CrossRef] [PubMed]
21. Abdin, Z.; Alim, M.A.; Saidur, R.; Islam, M.R.; Rashmi, W.; Mekhilef, S.; Wadi, A. Solar energy harvesting
with the application of nanotechnology. Renew. Sustain. Energy Rev. 2013, 26, 837–852. [CrossRef]
Polymers 2016, 8, 281 27 of 37
22. Sun, W.; Hu, R.; Liu, H.; Zeng, M.; Yang, L.; Wang, H.; Zhu, M. Embedding nano-silicon in graphene
nanosheets by plasma assisted milling for high capacity anode materials in lithium ion batteries.
J. Power Sources 2014, 268, 610–618. [CrossRef]
23. Azeez, A.A.; Rhee, K.Y.; Park, S.J.; Hui, D. Epoxy clay nanocomposites—Processing, properties and
applications: A review. Compos. Part. B 2013, 45, 308–320. [CrossRef]
24. Aziz, A.; Lim, H.N.; Girei, S.H.; Yaacob, M.H.; Mahdi, M.A.; Huang, N.M.; Pandikumar, A. Silver/graphene
nanocomposite-modified optical fiber sensor platform for ethanol detection in water medium. Sens. Actuators
B Chem. 2015, 206, 119–125.
25. Agnihotri, N.; Chowdhury, A.D.; De, A. Non-enzymatic electrochemical detection of cholesterol using
β-cyclodextrin functionalized graphene. Biosens. Bioelectron. 2015, 63, 212–217. [CrossRef] [PubMed]
26. Galpaya, D.; Wang, M.; Liu, M.; Motta, N.; Waclawik, E.; Yan, C. Recent Advances in fabrication and
characterization of graphene-polymer nanocomposites. Sci. Res. 2012, 2012, 30–49. [CrossRef]
27. Shahil, K.M.F.; Balandin, A.A. Thermal properties of graphene and multilayer graphene: Applications in
thermal interface materials. Solid State Commun. 2012, 152, 1331–1340. [CrossRef]
28. Al-Saleh, M.H.; Sundararaj, U. Review of the mechanical properties of carbon nanofiber/polymer composites.
Compos. Part A 2011, 42, 2126–2142. [CrossRef]
29. Sanjinés, R.; Abad, M.D.; Vâju, C.; Smajda, R.; Mionić, M.; Magrez, A. Electrical properties and applications
of carbon based nanocomposite materials: An overview. Surf. Coat. Technol. 2011, 206, 727–733. [CrossRef]
30. Potts, J.R.; Dreyer, D.R.; Bielawski, C.W.; Ruoff, R.S. Graphene-based polymer nanocomposites.
Polymer (Guildf) 2011, 52, 5–25. [CrossRef]
31. Qin, F.; Brosseau, C. A review and analysis of microwave absorption in polymer composites filled with
carbonaceous particles. J. Appl. Phys. 2012. [CrossRef]
32. Lee, S.-Y.; Park, S.-J. Comprehensive review on synthesis and adsorption behaviors of graphene-based
materials. Carbon Lett. 2012, 13, 73–87. [CrossRef]
33. Singh, V.; Joung, D.; Zhai, L.; Das, S.; Khondaker, S.I.; Seal, S. Graphene-based materials: Past, present and
future. Prog. Mater. Sci. 2011, 56, 1178–1271. [CrossRef]
34. Van Rooyen, L.J.; Karger-Kocsis, J.J.; Kock, L.D.; David Kock, L. Improving the helium gas barrier properties
of epoxy coatings through the incorporation of graphene nanoplatelets and the influence of preparation
techniques. J. Appl. Polym. Sci. 2015, 42584, 1–13. [CrossRef]
35. Kim, H.; Abdala, A.A.; Macosko, C.W. Graphene/polymer nanocomposites. Macromolecules 2010, 43,
6515–6530. [CrossRef]
36. Dhand, V.; Rhee, K.Y.; Kim, H.J.; Jung, D.H. A comprehensive review of graphene nanocomposites: research
status and trends. J. Nanomater. 2015, 2013, 1–15. [CrossRef]
37. Santamaria, A.; Muñoz, M.E.; Fernández, M.; Landa, M. Electrically conductive adhesives with a focus on
adhesives that contain carbon nanotubes. J. Appl. Polym. Sci. 2013, 129, 1643–1652. [CrossRef]
38. Yang, M.-Q.; Xu, Y.-J. Selective photoredox using graphene-based composite photocatalysts. Phys. Chem.
Chem. Phys. 2013, 15, 19102–19118. [CrossRef] [PubMed]
39. Srinivas, G.; Guo, Z.X. Graphene-based materials: Synthesis and gas sorption, storage and separation.
Prog. Mater. Sci. 2014. [CrossRef]
40. Xu, Z.; Chen, L.; Zhou, B.; Li, Y.; Li, B.; Niu, J.; Shan, M.; Guo, Q.; Wang, Z.; Qian, X. Nano-structure and
property transformations of carbon systems under γ-ray irradiation: A review. RSC Adv. 2013. [CrossRef]
41. Hu, K.; Kulkarni, D.D.; Choi, I.; Tsukruk, V.V. Graphene-polymer nanocomposites for structural and
functional applications. Prog. Polym. Sci. 2014, 39, 1934–1972. [CrossRef]
42. Young, R.J.; Kinloch, I.A.; Gong, L.; Novoselov, K.S. The mechanics of graphene nanocomposites: A review.
Compos. Sci. Technol. 2012, 72, 1459–1476. [CrossRef]
43. Zaman, I.; Manshoor, B.; Khalid, A.; Araby, S. From clay to graphene for polymer nanocomposites—A survey.
J. Polym. Res. 2014, 21, 429. [CrossRef]
44. Sun, X.; Sun, H.; Li, H.; Peng, H. Developing polymer composite materials: Carbon nanotubes or graphene?
Adv. Mater. 2013, 25, 5153–5176. [CrossRef] [PubMed]
45. Kuilla, T.; Bhadra, S.; Yao, D.; Kim, N.H.; Bose, S.; Lee, J.H. Recent advances in graphene-based polymer
composites. Prog. Polym. Sci. 2010, 35, 1350–1375. [CrossRef]
46. Rasheed, A.; Khalid, F.A. Fabrication and properties of CNTs reinforced polymeric matrix nanocomposites
for sports applications. IOP Conf. Ser. Mater. Sci. Eng. 2014. [CrossRef]
Polymers 2016, 8, 281 28 of 37
47. Yue, L.; Pircheraghi, G.; Monemian, S.A.; Manas-Zloczower, I. Epoxy composites with carbon nanotubes and
graphene nanoplatelets—Dispersion and synergy effects. Carbon 2014, 78, 268–278. [CrossRef]
48. Jean-Pierre, P.; Roberto, W. Epoxy Polymers New Materials and Innovations; Wiley-VCH: Weinheim,
Germany, 2010.
49. Sanjay, M. Composites Manufacturing Materials, Product, and Process Engineering; CRC Press: Boca Raton, FL,
USA, 2002.
50. Valery, V.; Evgeny, M. Mechanics and Analysis of Composite Materials; Elsevier: Amsterdam, The Netherlands, 2001.
51. Atif, R.; Inam, F. Influence of macro-topography on damage tolerance and fracture toughness of monolithic
epoxy for tribological applications. World J. Eng. Technol. 2016, 4, 335–360. [CrossRef]
52. Wongbong, C.; Jo-Won, L. Graphene Synthesis and Applications; CRC Press: Boca Raton, FL, USA, 2012.
53. Warner, J.H.; Fransizka, S.; Mark, R.; Bachmatiuk, A. Graphene: Fundamentals and Emergent Applications;
Elsevier: Amsterdam, The Netherlands, 2013.
54. Mikhail, K.; Iosifovich, K.M. Graphene: Carbon in Two Dimensions; Cambridge University Press: Cambridge,
UK, 2012.
55. Wolf, E.L. Graphene: A New Paradigm in Condensed Matter and Device Physics; OUP: Oxford, UK, 2013.
56. Quintana, M.; Spyrou, K.; Grzelczak, M.; Browne, W.R.; Rudolf, P.; Prato, M. Functionalization of graphene.
ACS Nano 2010, 4, 3527–3533. [CrossRef] [PubMed]
57. Wang, X.; Jin, J.; Song, M. An investigation of the mechanism of graphene toughening epoxy. Carbon 2013,
65, 324–333. [CrossRef]
58. Jia, J.; Kan, C.-M.; Lin, X.; Shen, X.; Kim, J.-K. Effects of processing and material parameters on synthesis of
monolayer ultralarge graphene oxide sheets. Carbon 2014, 77, 244–254. [CrossRef]
59. Ma, J.; Meng, Q.; Zaman, I.; Zhu, S.; Michelmore, A.; Kawashima, N.; Wang, C. H.; Kuan, H.-C. Development
of polymer composites using modified, high-structural integrity graphene platelets. Compos. Sci. Technol.
2014, 91, 82–90. [CrossRef]
60. Loomis, J.; Panchapakesan, B. Dimensional dependence of photomechanical response in carbon
nanostructure composites: A case for carbon-based mixed-dimensional systems. Nanotechnology 2012,
23. [CrossRef] [PubMed]
61. Karger-Kocsis, J.; Mahmood, H.; Pegoretti, A. Recent advances in fiber/matrix interphase engineering for
polymer composites. Prog. Mater. Sci. 2015, 73, 1–43. [CrossRef]
62. Dieter, G.E. Mechanical Metallurgy, SI Metric ed.; McGraw-Hill: New York, NY, USA, 1988.
63. Wan, Y.-J.; Tang, L.-C.; Gong, L.-X.; Yan, D.; Li, Y.-B.; Wu, L.-B.; Jiang, J.-X.; Lai, G.-Q. Grafting of epoxy
chains onto graphene oxide for epoxy composites with improved mechanical and thermal properties. Carbon
2014, 69, 467–480. [CrossRef]
64. Bindu Sharmila, T.K.; Nair, A.B.; Abraham, B.T.; Beegum, P.M.S.; Thachil, E.T. Microwave exfoliated
reduced graphene oxide epoxy nanocomposites for high performance applications. Polymer (Guildf) 2014, 55,
3614–3627.
65. Zhang, Y.; Wang, Y.; Yu, J.; Chen, L.; Zhu, J.; Hu, Z. Tuning the interface of graphene platelets/epoxy
composites by the covalent grafting of polybenzimidazole. Polymer (Guildf) 2014, 55, 4990–5000. [CrossRef]
66. Ahmadi-Moghadam, B.; Sharafimasooleh, M.; Shadlou, S.; Taheri, F. Effect of functionalization of graphene
nanoplatelets on the mechanical response of graphene/ epoxy composites. Mater. Des. 2014, 66, 142–149.
[CrossRef]
67. Chandrasekaran, S.; Sato, N.; Tölle, F.; Mülhaupt, R.; Fiedler, B.; Schulte, K. Fracture toughness and failure
mechanism of graphene-based epoxy composites. Compos. Sci. Technol. 2014, 97, 90–99. [CrossRef]
68. Wan, Y.-J.; Gong, L.-X.; Tang, L.-C.; Wu, L.-B.; Jiang, J.-X. Mechanical properties of epoxy composites filled
with silane-functionalized graphene oxide. Compos. Part A 2014, 64, 79–89. [CrossRef]
69. Zaman, I.; Manshoor, B.; Khalid, A.; Meng, Q.; Araby, S. Interface modification of clay and graphene platelets
reinforced epoxy nanocomposites: A comparative study. J. Mater. Sci. 2014, 49, 5856–5865. [CrossRef]
70. Jiang, T.; Kuila, T.; Kim, N.H.; Lee, J.H. Effects of surface-modified silica nanoparticles attached graphene
oxide using isocyanate-terminated flexible polymer chains on the mechanical properties of epoxy composites.
J. Mater. Chem. A 2014. [CrossRef]
71. Shokrieh, M.M.; Ghoreishi, S.M.; Esmkhani, M.; Zhao, Z. Effects of graphene nanoplatelets and graphene
nanosheets on fracture toughness of epoxy nanocomposites. Fatigue Fract. Eng. Mater. Struct. 2014, 37,
1116–1123.
Polymers 2016, 8, 281 29 of 37
72. Tang, L.-C.; Wan, Y.-J.; Yan, D.; Pei, Y.-B.; Zhao, L.; Li, Y.-B.; Wu, L.-B.; Jiang, J.-X.; Lai, G.-Q. The effect of
graphene dispersion on the mechanical properties of graphene/epoxy composites. Carbon 2013, 60, 16–27.
[CrossRef]
73. Chandrasekaran, S.; Seidel, C.; Schulte, K. Preparation and characterization of graphite nano-platelet
(GNP)/epoxy nano-composite: Mechanical, electrical and thermal properties. Eur. Polym. J. 2013, 49,
3878–3888. [CrossRef]
74. Li, Z.; Wang, R.; Young, R.J.; Deng, L.; Yang, F.; Hao, L.; Jiao, W.; Liu, W. Control of the functionality of
graphene oxide for its application in epoxy nanocomposites. Polymer (Guildf) 2013, 54, 6437–6446. [CrossRef]
75. Shadlou, S.; Alishahi, E.; Ayatollahi, M.R. Fracture behavior of epoxy nanocomposites reinforced with
different carbon nano-reinforcements. Compos. Struct. 2013, 95, 577–581. [CrossRef]
76. Jiang, T.; Kuila, T.; Kim, N.H.; Ku, B.-C.; Lee, J.H. Enhanced mechanical properties of silanized silica
nanoparticle attached graphene oxide/epoxy composites. Compos. Sci. Technol. 2013, 79, 115–125. [CrossRef]
77. Liu, W.; Kong, J.; Toh, W.E.; Zhou, R.; Ding, G.; Huang, S.; Dong, Y.; Lu, X. Toughening of epoxies by
covalently anchoring triazole-functionalized stacked-cup carbon nanofibers. Compos. Sci. Technol. 2013, 85,
1–9. [CrossRef]
78. Wang, R.; Li, Z.; Liu, W.; Jiao, W.; Hao, L.; Yang, F. Attapulgite–graphene oxide hybrids as thermal and
mechanical reinforcements for epoxy composites. Compos. Sci. Technol. 2013, 87, 29–35. [CrossRef]
79. Alishahi, E.; Shadlou, S.; Doagou-R, S.; Ayatollahi, M.R. Effects of carbon nanoreinforcements of different
shapes on the mechanical properties of epoxy-based nanocomposites. Macromol. Mater. Eng. 2013, 298,
670–678. [CrossRef]
80. Ma, J.; Meng, Q.; Michelmore, A.; Kawashima, N.; Izzuddin, Z.; Bengtsson, C.; Kuan, H.-C. Covalently
bonded interfaces for polymer/graphene composites. J. Mater. Chem. A 2013. [CrossRef]
81. Feng, H.; Wang, X.; Wu, D. Fabrication of spirocyclic phosphazene epoxy-based nanocomposites with
graphene via exfoliation of graphite platelets and thermal curing for enhancement of mechanical and
conductive properties. Ind. Eng. Chem. Res. 2013, 52, 10160–10171. [CrossRef]
82. Chatterjee, S.; Nafezarefi, F.; Tai, N.H.; Schlagenhauf, L.; Nüesch, F.A.; Chu, B.T.T. Size and synergy effects
of nanofiller hybrids including graphene nanoplatelets and carbon nanotubes in mechanical properties of
epoxy composites. Carbon 2012, 50, 5380–5386. [CrossRef]
83. Chatterjee, S.; Wang, J.W.; Kuo, W.S.; Tai, N.H.; Salzmann, C.; Li, W.L.; Hollertz, R.; Nüesch, F.A.; Chu, B.T.T.
Mechanical reinforcement and thermal conductivity in expanded graphene nanoplatelets reinforced epoxy
composites. Chem. Phys. Lett. 2012, 531, 6–10. [CrossRef]
84. Zaman, I.; Phan, T.T.; Kuan, H.-C.; Meng, Q.; Bao La, L.T.; Luong, L.; Youssf, O.; Ma, J. Epoxy/graphene
platelets nanocomposites with two levels of interface strength. Polymer (Guildf) 2011, 52, 1603–1611.
[CrossRef]
85. Rana, S.; Alagirusamy, R.; Joshi, M. Development of carbon nanofibre incorporated three phase carbon/epoxy
composites with enhanced mechanical, electrical and thermal properties. Compos. Part. A Appl. Sci. Manuf.
2011, 42, 439–445. [CrossRef]
86. Bortz, D.R.; Merino, C.; Martin-Gullon, I. Carbon nanofibers enhance the fracture toughness and fatigue
performance of a structural epoxy system. Compos. Sci. Technol. 2011, 71, 31–38. [CrossRef]
87. Zhang, G.; Karger-Kocsis, J.; Zou, J. Synergetic effect of carbon nanofibers and short carbon fibers on the
mechanical and fracture properties of epoxy resin. Carbon 2010, 48, 4289–4300. [CrossRef]
88. Fang, M.; Zhang, Z.; Li, J.; Zhang, H.; Lu, H.; Yang, Y. Constructing hierarchically structured interphases for
strong and tough epoxy nanocomposites by amine-rich graphene surfaces. J. Mater. Chem. 2010. [CrossRef]
89. Jana, S.; Zhong, W.-H. Graphite particles with a “puffed” structure and enhancement in mechanical
performance of their epoxy composites. Mater. Sci. Eng. A 2009, 525, 138–146. [CrossRef]
90. Rafiee, M.A.; Rafiee, J.; Srivastava, I.; Wang, Z.; Song, H.; Yu, Z.-Z.; Koratkar, N. Fracture and fatigue in
graphene nanocomposites. Small 2010, 6, 179–183. [CrossRef] [PubMed]
91. Kuhn, H.; Medlin, D. ASM Handbook, Volume 8: Mechanical Testing and Evaluation; ASM International:
Materials Park, OH, USA, 2000.
92. Griffith, A.A. The Phenomena of rupture and flow in solids. Philos. Trans. R. Soc. Lond. Ser. A Contain. Pap.
Math. Phys. Character 1921, 221, 163–198. [CrossRef]
Polymers 2016, 8, 281 30 of 37
93. Zhang, W.; Srivastava, I.; Zhu, Y.F.; Picu, C.R.; Koratkar, N.A. Heterogeneity in epoxy nanocomposites
initiates crazing: Significant improvements in fatigue resistance and toughening. Small 2009, 5, 1403–1407.
[CrossRef] [PubMed]
94. ASM Handbook Volume 19: Fatigue and Fracture; ASM International: Materials Park, OH, USA, 1996.
95. Saharudin, M.S.; Atif, R.; Shyha, I.; Inam, F. The degradation of mechanical properties in polymer
nano-composites exposed to liquid media—A review. RSC Adv. 2016, 6, 1076–1089. [CrossRef]
96. Atif, R.; Shyha, I.; Inam, F. The degradation of mechanical properties due to stress concentration caused by
retained acetone in epoxy nanocomposites. RSC Adv. 2016, 6, 34188–34197. [CrossRef]
97. Chen, Q.; Liu, W.; Guo, S.; Zhu, S.; Li, Q.; Li, X.; et al. Synthesis of well-aligned millimeter-sized
tetragon-shaped graphene domains by tuning the copper substrate orientation. Carbon 2015, 93, 945–952.
[CrossRef]
98. Bhushan, B. Springer Handbook of Nanotechnology, 3rd ed.; Springer: Berlin, Germany, 2010.
99. Faber, K.T.; Evans, A.G. Crack deflection processes—I. Theory. Acta Metall. 1983, 31, 565–576. [CrossRef]
100. Faber, K.T.; Evans, A.G. Crack deflection processes—II. Experiment. Acta Metall. 1983, 31, 577–584. [CrossRef]
101. Xie, F. A facile strategy for the reduction of graphene oxide and its effect on thermal conductivity of
epoxy-based composites. Express Polym. Lett. 2016, 10, 470–478. [CrossRef]
102. Atif, R.; Inam, F. The dissimilarities between graphene and frame-like structures. Graphene 2016, 1, 55–72.
[CrossRef]
103. Fan, B.-B.; Guo, H.-H.; Zhang, R.; Jia, Y.; Shi, C.-Y. Structural evolution during the oxidation process of
graphite. Chin. Phys. Lett. 2014. [CrossRef]
104. Xu, Z.; Xue, K. Engineering graphene by oxidation: A first-principles study. Nanotechnology 2010. [CrossRef]
[PubMed]
105. Kuo, W.-S.; Tai, N.-H.; Chang, T.-W. Deformation and fracture in graphene nanosheets. Compos. Part A Appl.
Sci. Manuf. 2013, 51, 56–61. [CrossRef]
106. Palmeri, M.J.; Putz, K.W.; Brinson, L.C. Sacrificial bonds in stacked-cup carbon nanofibers: Biomimetic
toughening mechanisms for composite systems. ACS Nano 2010, 4, 4256–4264. [CrossRef] [PubMed]
107. Lee, D.; Zou, X.; Zhu, X.; Seo, J.W.; Cole, J.M.; Bondino, F.; Magnano, E.; Nair, S. K.; Su, H. Ultrafast carrier
phonon dynamics in NaOH-reacted graphite oxide film. Appl. Phys. Lett. 2012. [CrossRef]
108. Shojaee, S.A.; Zandiatashbar, A.; Koratkar, N.; Lucca, D.A. Raman spectroscopic imaging of graphene
dispersion in polymer composites. Carbon 2013, 62, 510–513. [CrossRef]
109. Tamburrano, A.; Sarasini, F.; de Bellis, G.; D’Aloia, A.G.; Sarto, M.S. The piezoresistive effect in
graphene-based polymeric composites. Nanotechnology 2013. [CrossRef] [PubMed]
110. Yang, H.; Li, F.; Shan, C.; Han, D.; Zhang, Q.; Niu, L.; Ivaska, A. Covalent functionalization of chemically
converted graphene sheets via silane and its reinforcement. J. Mater. Chem. 2009, 19, 4632–4638. [CrossRef]
111. Wang, G.; Shen, X.; Wang, B.; Yao, J.; Park, J. Synthesis and characterisation of hydrophilic and organophilic
graphene nanosheets. Carbon 2009, 47, 1359–1364. [CrossRef]
112. Samanman, S.; Numnuam, A.; Limbut, W.; Kanatharana, P.; Thavarungkul, P. Highly-sensitive label-free
electrochemical carcinoembryonic antigen immunosensor based on a novel Au nanoparticles–graphene–
chitosan nanocomposite cryogel electrode. Anal. Chim. Acta 2015, 853, 521–532. [CrossRef] [PubMed]
113. Lee, S.-Y.; Chong, M.-H.; Park, M.; Kim, H.-Y.; Park, S.-J. Effect of chemically reduced graphene oxide on
epoxy nanocomposites for flexural behaviors. Carbon Lett. 2014, 15, 67–70. [CrossRef]
114. Teng, C.-C.; Ma, C.-C.M.; 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, 5107–5116. [CrossRef]
115. Chu, K.; Li, W.; Dong, H.; Tang, F. Modeling the thermal conductivity of graphene nanoplatelets reinforced
composites. EPL Europhys. Lett. 2012, 100, 36001–36005. [CrossRef]
116. Yang, S.-Y.; Lin, W.-N.; Huang, Y.-L.; Tien, H.-W.; Wang, J.-Y.; Ma, C.-C.M.; 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, 793–803. [CrossRef]
117. Pu, N.-W.; Peng, Y.-Y.; Wang, P.-C.; Chen, C.-Y.; Shi, J.-N.; Liu, Y.-M.; Ger, M.-D.; Chang, C.-L. Application
of nitrogen-doped graphene nanosheets in electrically conductive adhesives. Carbon 2014, 67, 449–456.
[CrossRef]
Polymers 2016, 8, 281 31 of 37
118. Zhao, Q.; Hao, S. Toughening mechanism of epoxy resins with micro/nano particles. J. Compos. Mater. 2007,
41, 201–219. [CrossRef]
119. Zhao, Q.; Hoa, S.; Ouellette, P. Progressive failure of triaxial woven fabric (TWF) composites with open holes.
Compos. Struct. 2004, 65, 419–431. [CrossRef]
120. Bastwros, M.; Kim, G.-Y.; Zhu, C.; Zhang, K.; Wang, S.; Tang, X.; Wang, X. Effect of ball milling on graphene
reinforced Al6061 composite fabricated by semi-solid sintering. Compos. Part B 2014, 60, 111–118. [CrossRef]
121. Wu, H.; Rook, B.; Drzal, L.T. Dispersion optimization of exfoliated graphene nanoplatelet in polyetherimide
nanocomposites: Extrusion, precoating, and solid state ball milling. Polym. Compos. 2013, 34, 426–432.
[CrossRef]
122. Yu, M.; Shao, D.; Lu, F.; Sun, X.; Sun, H.; Hu, T.; Wang, G.; Sawyer, S.; Qiu, H.; Lian, J. ZnO/graphene
nanocomposite fabricated by high energy ball milling with greatly enhanced lithium storage capability.
Electrochem. Commun. 2013, 34, 312–315. [CrossRef]
123. Jiang, X.; Drzal, L.T. Reduction in percolation threshold of injection molded high-density
polyethylene/exfoliated graphene nanoplatelets composites by solid state ball milling and solid state
shear pulverization. J. Appl. Polym. Sci. 2011, 124, 525–535. [CrossRef]
124. Wu, H.; Zhao, W.; Chen, G. One-pot in situ ball milling preparation of polymer/graphene nanocomposites.
J. Appl. Polym. Sci. 2012, 125, 3899–3903. [CrossRef]
125. Xu, J.; Jeon, I.-Y.; Seo, J.-M.; Dou, S.; Dai, L.; Baek, J.-B. Edge-selectively halogenated graphene nanoplatelets
(XGnPs, X = Cl, Br, or I) prepared by ball-milling and used as anode materials for lithium-ion batteries.
Adv. Mater. 2014, 26, 7317–7323. [CrossRef] [PubMed]
126. Guo, W.; Chen, G. Fabrication of graphene/epoxy resin composites with much enhanced thermal
conductivity via ball milling technique. J. Appl. Polym. Sci. 2014, 131, 40565–40569. [CrossRef]
127. Rodriguez, A.M.; Prieto, P.; Prato, M.; Va, E. Exfoliation of graphite with triazine derivatives under
ball-milling conditions: Preparation of few-layer graphene via selective noncovalent interactions. ACS Nano
2014, 8, 563–571.
128. Xu, J.; Shui, J.; Wang, J.; Wang, M.; Liu, H.; Dou, S.X.; Jeon, I. Sulfur–graphene nanostructured cathodes
via ball-milling for high-performance lithium–sulfur batteries. ACS Nano 2014, 8, 10920–10930. [CrossRef]
[PubMed]
129. Cravotto, G.; Cintas, P. Sonication-assisted fabrication and post-synthetic modifications of graphene-like
materials. Chemistry 2010, 16, 5246–5259. [CrossRef] [PubMed]
130. Yi, M.; Shen, Z.; Zhang, X.; Ma, S. Vessel diameter and liquid height dependent sonication-assisted production
of few-layer graphene. J. Mater. Sci. 2012, 47, 8234–8244. [CrossRef]
131. Ciesielski, A.; Samorì, P. Graphene via sonication assisted liquid-phase exfoliation. Chem. Soc. Rev. 2014, 43,
381–398. [CrossRef] [PubMed]
132. Wang, S.; Tang, L.A.L.; Bao, Q.; Lin, M.; Deng, S.; Goh, B.M.; Loh, K. P. Room-temperature synthesis of
soluble carbon nanotubes by the sonication of graphene oxide nanosheets. J. Am. Chem. Soc. 2009, 131,
16832–16837. [CrossRef] [PubMed]
133. Akhavan, O.; Ghaderi, E.; Esfandiar, A. Wrapping bacteria by graphene nanosheets for isolation from
environment, reactivation by sonication, and inactivation by near-infrared irradiation. J. Phys. Chem. B 2011,
115, 6279–6288. [CrossRef] [PubMed]
134. Polyakova Stolyarova, E.Y.; Rim, K.T.; Eom, D.; Douglass, K.; Opila, R.L.; Heinz, T.F.; Teplyakov, A.V.;
Flynn, G.W. Scanning tunneling microscopy and X-ray photoelectron spectroscopy studies of graphene films
prepared by sonication-assisted dispersion. ACS Nano 2011, 5, 6102–6108. [CrossRef] [PubMed]
135. Xu, P.; Loomis, J.; King, B.; Panchapakesan, B. Synergy among binary (MWNT, SLG) nano-carbons in polymer
nano-composites: A Raman study. Nanotechnology 2012. [CrossRef] [PubMed]
136. Cheng, Y.C.; Kaloni, T.P.; Zhu, Z.Y.; Schwingenschlögl, U. Oxidation of graphene in ozone under ultraviolet
light. Appl. Phys. Lett. 2012. [CrossRef]
137. Gracia-espino, E.; Hu, G.; Shchukarev, A.; Wa, T. Understanding the interface of six-shell cuboctahedral
and icosahedral palladium clusters on reduced graphene oxide: Experimental and theoretical study. J. Am.
Chem. Soc. 2014, 136, 6626–6633. [CrossRef] [PubMed]
138. Velizhanin, K.A.; Dandu, N.; Solenov, D. Electromigration of bivalent functional groups on graphene.
Phys. Rev. B 2014. [CrossRef]
Polymers 2016, 8, 281 32 of 37
139. Radovic, L.R.; Suarez, A.; Vallejos-Burgos, F.; Sofo, J.O. Oxygen migration on the graphene surface.
2. Thermochemistry of basal-plane diffusion (hopping). Carbon 2011, 49, 4226–4238. [CrossRef]
140. Radovic, L.R.; Silva-Tapia, A.B.; Vallejos-Burgos, F. Oxygen migration on the graphene surface. 1. Origin of
epoxide groups. Carbon 2011, 49, 4218–4225. [CrossRef]
141. Botas, C.; Álvarez, P.; Blanco, C.; Santamaría, R.; Granda, M.; Ares, P.; Rodríguez-Reinoso, F.; Menéndez, R.
The effect of the parent graphite on the structure of graphene oxide. Carbon 2012, 50, 275–282. [CrossRef]
142. Šljivančanin, Ž.; Milošević, A.S.; Popović, Z.S.; Vukajlović, F.R. Binding of atomic oxygen on graphene from
small epoxy clusters to a fully oxidized surface. Carbon 2013, 54, 482–488. [CrossRef]
143. Ahmed, M.S.; Han, H.S.; Jeon, S. One-step chemical reduction of graphene oxide with oligothiophene for
improved electrocatalytic oxygen reduction reactions. Carbon 2013, 61, 164–172. [CrossRef]
144. Yuan, F.-Y.; Zhang, H.-B.; Li, X.; Ma, H.-L.; Li, X.-Z.; Yu, Z.-Z. In situ chemical reduction and functionalization
of graphene oxide for electrically conductive phenol formaldehyde composites. Carbon 2014, 68, 653–661.
[CrossRef]
145. Jiang, X.; Nisar, J.; Pathak, B.; Zhao, J.; Ahuja, R. Graphene oxide as a chemically tunable 2-D material for
visible-light photocatalyst applications. J. Catal. 2013, 299, 204–209. [CrossRef]
146. Park, J.S.; Yu, L.; Lee, C.S.; Shin, K.; Han, J.H. Liquid-phase exfoliation of expanded graphites into graphene
nanoplatelets using amphiphilic organic molecules. J. Colloid Interface Sci. 2014, 417, 379–384. [CrossRef]
[PubMed]
147. Karger-Kocsis, J.; Friedrich, K. Microstructure-related fracture toughness and fatigue crack growth behaviour
in toughened, anhydride-cured epoxy resins. Compos. Sci. Technol. 1993, 48, 263–272. [CrossRef]
148. Karger-Kocsis, J.; Gremmels, J. Use of hygrothermal decomposed polyester–urethane waste for the impact
modification of epoxy resins. J. Appl. Polym. Sci. 2000, 5, 1139–1151. [CrossRef]
149. Smith, G.; Bedrov, D.; Li, L.; Byutner, O. A molecular dynamics simulation study of the viscoelastic properties
of polymer nanocomposites. J. Chem. Phys. 2002, 117, 9478–9489. [CrossRef]
150. Corcione, C.E.; Freuli, F.; Maffezzoli, A. The aspect ratio of epoxy matrix nanocomposites reinforced with
graphene stacks. Polym. Eng. Sci. 2013, 53, 531–539. [CrossRef]
151. Ramos-Galicia, L.; Mendez, L.N.; Martínez-Hernández, A.L.; Espindola-Gonzalez, A.; Galindo-Esquivel, I.R.;
Fuentes-Ramirez, R.; Velasco-Santos, C. Improved performance of an epoxy matrix as a result of combining
graphene oxide and reduced graphene. Int. J. Polym. Sci. 2013, 2013, 1–7. [CrossRef]
152. Li, Z.; Young, R.J.; Wang, R.; Yang, F.; Hao, L.; Jiao, W.; Liu, W. The role of functional groups on graphene
oxide in epoxy nanocomposites. Polymer (Guildf) 2013, 54, 5821–5829. [CrossRef]
153. Liu, W.; Koh, K.L.; Lu, J.; Yang, L.; Phua, S.; Kong, J.; Chen, Z.; Lu, X. Simultaneous catalyzing and reinforcing
effects of imidazole-functionalized graphene in anhydride-cured epoxies. J. Mater. Chem. 2012. [CrossRef]
154. Yang, H.; Shan, C.; Li, F.; Zhang, Q.; Han, D.; Niu, L. Convenient preparation of tunably loaded chemically
converted graphene oxide/epoxy resin nanocomposites from graphene oxide sheets through two-phase
extraction. J. Mater. Chem. 2009, 19, 8856. [CrossRef]
155. Galpaya, D.; Wang, M.; George, G.; Motta, N.; Waclawik, E.; Yan, C. Preparation of graphene oxide/epoxy
nanocomposites with significantly improved mechanical properties. J. Appl. Phys. 2014. [CrossRef]
156. Li, W.; Dichiara, A.; Bai, J. Carbon nanotube–graphene nanoplatelet hybrids as high-performance
multifunctional reinforcements in epoxy composites. Compos. Sci. Technol. 2013, 74, 221–227. [CrossRef]
157. Cao, L.; Liu, X.; Na, H.; Wu, Y.; Zheng, W.; Zhu, J. How a bio-based epoxy monomer enhanced the properties
of diglycidyl ether of bisphenol A (DGEBA)/graphene composites. J. Mater. Chem. A 2013, 1, 5081–5088.
[CrossRef]
158. Wan, Y.-J.; Tang, L.-C.; Yan, D.; Zhao, L.; Li, Y.-B.; Wu, L.-B.; Jiang, J.-X.; Lai, G.-Q. Improved dispersion and
interface in the graphene/epoxy composites via a facile surfactant-assisted process. Compos. Sci. Technol.
2013, 82, 60–68. [CrossRef]
159. Ghaleb, Z.A.; Mariatti, M.; Ariff, Z.M. Properties of graphene nanopowder and multi-walled carbon
nanotube-filled epoxy thin-film nanocomposites for electronic applications: The effect of sonication time and
filler loading. Compos. Part A 2014, 58, 77–83. [CrossRef]
160. King, J.A.; Klimek, D.R.; Miskioglu, I.; Odegard, G.M. Mechanical properties of graphene nanoplatelet/epoxy
composites. J. Appl. Polym. Sci. 2013, 128, 4217–4223. [CrossRef]
Polymers 2016, 8, 281 33 of 37
161. Wang, X.; Song, L.; Pornwannchai, W.; Hu, Y.; Kandola, B. The effect of graphene presence in flame retarded
epoxy resin matrix on the mechanical and flammability properties of glass fiber-reinforced composites.
Compos. Part A 2013, 53, 88–96. [CrossRef]
162. Serena Saw, W.P.; Mariatti, M. Properties of synthetic diamond and graphene nanoplatelet-filled epoxy thin
film composites for electronic applications. J. Mater. Sci. Mater. Electron. 2011, 23, 817–824. [CrossRef]
163. Zaman, I.; Kuan, H.-C.; Meng, Q.; Michelmore, A.; Kawashima, N.; Pitt, T.; Zhang, L.; Gouda, S.;
Luong, L.; Ma, J. A Facile Approach to Chemically Modified Graphene and its Polymer Nanocomposites.
Adv. Funct. Mater. 2012, 22, 2735–2743. [CrossRef]
164. Hsu, C.-H.; Hsu, M.-H.; Chang, K.-C.; Lai, M.-C.; Liu, P.-J.; Chuang, T.-L.; Yeh, J.-M.; Liu, W.-R. Physical
study of room-temperature-cured epoxy/thermally reduced graphene oxides with various contents of
oxygen-containing groups. Polym. Int. 2014, 63, 1765–1770. [CrossRef]
165. Yang, Y.; Rigdon, W.; Huang, X.; Li, X. Enhancing graphene reinforcing potential in composites by hydrogen
passivation induced dispersion. Sci. Rep. 2013, 3, 2086–2093. [CrossRef] [PubMed]
166. Naebe, M.; Wang, J.; Amini, A.; Khayyam, H.; Hameed, N.; Li, L.H.; Chen, Y.; Fox, B. Mechanical property
and structure of covalent functionalised graphene/epoxy nanocomposites. Sci. Rep. 2014, 4, 4375–4382.
[CrossRef] [PubMed]
167. Qi, B.; Yuan, Z.; Lu, S.; Liu, K.; Li, S.; Yang, L.; Yu, J. Mechanical and thermal properties of epoxy composites
containing graphene oxide and liquid crystalline epoxy. Fibers Polym. 2014, 15, 326–333. [CrossRef]
168. Ren, F.; Zhu, G.; Ren, P.; Wang, Y.; Cui, X. In situ polymerization of graphene oxide and cyanate ester–epoxy
with enhanced mechanical and thermal properties. Appl. Surf. Sci. 2014, 316, 549–557. [CrossRef]
169. Qi, B. Enhanced thermal and mechanical properties of epoxy composites by mixing thermotropic liquid
crystalline epoxy grafted graphene oxide. Express Polym. Lett. 2014, 8, 467–479. [CrossRef]
170. Lu, S.; Li, S.; Yu, J.; Yuan, Z.; Qi, B. Epoxy nanocomposites filled with thermotropic liquid crystalline epoxy
grafted graphene oxide. RSC Adv. 2013, 3, 8915. [CrossRef]
171. Shen, X.-J.; Liu, Y.; Xiao, H.-M.; Feng, Q.-P.; Yu, Z.-Z.; Fu, S.-Y. The reinforcing effect of graphene nanosheets
on the cryogenic mechanical properties of epoxy resins. Compos. Sci. Technol. 2012, 72, 1581–1587. [CrossRef]
172. Bao, C.; Guo, Y.; Song, L.; Kan, Y.; Qian, X.; Hu, Y. In situ preparation of functionalized graphene oxide/epoxy
nanocomposites with effective reinforcements. J. Mater. Chem. 2011, 21, 13290–13298. [CrossRef]
173. Meng, Q.; Jin, J.; Wang, R.; Kuan, H.-C.; Ma, J.; Kawashima, N.; Michelmore, A.; Zhu, S.; Wang, C.H.
Processable 3-nm thick graphene platelets of high electrical conductivity and their epoxy composites.
Nanotechnology 2014, 25, 125707–125719. [CrossRef] [PubMed]
174. Atif, R.; Shyha, I.; Inam, F. Modeling and experimentation of multi-layered nanostructured graphene-epoxy
nanocomposites for enhanced thermal and mechanical properties. J. Compos. Mater. 2016. [CrossRef]
175. Yu, A.; Ramesh, P.; Itkis, M.E.; Bekyarova, E.; Haddon, R.C. Graphite nanoplatelet—epoxy composite thermal
interface materials. J. Phys. Chem. C 2007, 111, 7565–7569. [CrossRef]
176. Yavari, F.; Fard, H.R.; Pashayi, K.; Rafiee, M.a.; Zamiri, A.; Yu, Z.; Ozisik, R.; Borca-Tasciuc, T.; Koratkar, N.
Enhanced thermal conductivity in a nanostructured phase change composite due to low concentration
graphene additives. J. Phys. Chem. C 2011, 115, 8753–8758. [CrossRef]
177. Ganguli, S.; Roy, A.K.; Anderson, D.P. Improved thermal conductivity for chemically functionalized
exfoliated graphite/epoxy composites. Carbon 2008, 46, 806–817. [CrossRef]
178. Fukushima, H.; Drzal, L.T.; Rook, B.P.; Rich, M.J. Thermal conductivity of exfoliated graphite nanocomposites.
J. Therm. Anal. Calorim. 2006, 85, 235–238. [CrossRef]
179. Xie, S.H.; Liu, Y.Y.; Li, J.Y. Comparison of the effective conductivity between composites reinforced by
graphene nanosheets and carbon nanotubes. Appl. Phys. Lett. 2008, 92, 1–3. [CrossRef]
180. Lin, W.; Zhang, R.; Wong, C.P. Modeling of thermal conductivity of graphite nanosheet composites.
J. Electron. Mater. 2010, 39, 268–272. [CrossRef]
181. Nan, C.-W.; Birringer, R.; Clarke, D.R.; Gleiter, H. Effective thermal conductivity of particulate composites
with interfacial thermal resistance. J. Appl. Phys. 1997, 81, 6692–6699. [CrossRef]
182. Shiu, S.-C.; Tsai, J.-L. Characterizing thermal and mechanical properties of graphene/epoxy nanocomposites.
Compos. Part B 2014, 56, 691–697. [CrossRef]
183. Yu, G.; Wu, P. Effect of chemically modified graphene oxide on the phase separation behaviour and properties
of an epoxy/polyetherimide binary system. Polym. Chem. 2014, 5, 96–104. [CrossRef]
Polymers 2016, 8, 281 34 of 37
184. Liu, T.; Zhao, Z.; Tjiu, W.W.; Lv, J.; Wei, C. Preparation and characterization of epoxy nanocomposites
containing surface-modified graphene oxide. J. Appl. Polym. Sci. 2014, 131, 40236–40242. [CrossRef]
185. Liu, F.; Guo, K. Reinforcing epoxy resin through covalent integration of functionalized graphene nanosheets.
Polym. Adv. Technol. 2014, 25, 418–423. [CrossRef]
186. Guan, L.-Z.; Wan, Y.-J.; Gong, L.-X.; Yan, D.; Tang, L.-C.; Wu, L.-B.; Jiang, J.-X.; Lai, G.-Q. Toward
effective and tunable interphases in graphene oxide/epoxy composites by grafting different chain lengths of
polyetheramine onto graphene oxide. J. Mater. Chem. A 2014, 2, 15058–15069. [CrossRef]
187. Martin-Gallego, M.; Bernal, M.M.; Hernandez, M.; Verdejo, R.; Lopez-Manchado, M.A. Comparison of filler
percolation and mechanical properties in graphene and carbon nanotubes filled epoxy nanocomposites.
Eur. Polym. J. 2013, 49, 1347–1353. [CrossRef]
188. Ribeiro, H.; Silva, W.M.; Rodrigues, M.-T.F.; Neves, J.C.; Paniago, R.; Fantini, C.; Calado, H.D.R.; Seara, L.M.;
Silva, G.G. Glass transition improvement in epoxy/graphene composites. J. Mater. Sci. 2013, 48, 7883–7892.
[CrossRef]
189. Wajid, A.S.; Ahmed, H.S.T.; Das, S.; Irin, F.; Jankowski, A.F.; Green, M.J. High-Performance Pristine
Graphene/Epoxy Composites With Enhanced Mechanical and Electrical Properties. Macromol. Mater. Eng.
2013, 298, 339–347. [CrossRef]
190. Zhang, X.; Alloul, O.; He, Q.; Zhu, J.; Verde, M.J.; Li, Y.; Wei, S.; Guo, Z. Strengthened magnetic epoxy
nanocomposites with protruding nanoparticles on the graphene nanosheets. Polymer (Guildf) 2013, 54,
3594–3604. [CrossRef]
191. Wang, X.; Xing, W.; Feng, X.; Yu, B.; Song, L.; Hu, Y. Functionalization of graphene with grafted
polyphosphamide for flame retardant epoxy composites: Synthesis, flammability and mechanism.
Polym. Chem. 2014. [CrossRef]
192. Hu, L.; Desai, T.; Keblinski, P. Thermal transport in graphene-based nanocomposite. J. Appl. Phys. 2011, 110,
1–6. [CrossRef]
193. Li, Q.; Guo, Y.; Li, W.; Qiu, S.; Zhu, C.; Wei, X.; Chen, M.; Liu, C.; Liao, S.; Gong, Y.; Mishra, A. K.;
Liu, L. Ultrahigh Thermal Conductivity of Assembled Aligned Multilayer Graphene/Epoxy Composite.
Chem. Mater. 2014, 26, 4459–4465. [CrossRef]
194. Geim, A.K. Graphene: Status and Prospects. Science 2009, 324, 1530–1535. [CrossRef] [PubMed]
195. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [CrossRef] [PubMed]
196. Yan, W.; He, W.-Y.; Chu, Z.-D.; Liu, M.; Meng, L.; Dou, R.-F.; Zhang, Y.; Liu, Z.; Nie, J.-C.; He, L. Strain and
curvature induced evolution of electronic band structures in twisted graphene bilayer. Nat. Commun. 2013, 4,
1–7. [CrossRef] [PubMed]
197. Castro Neto, A.H.; Peres, N.M.R.; Novoselov, K.S.; Geim, A.K.; Guinea, F. The electronic properties of
graphene. Rev. Mod. Phys. 2009, 81, 109–162. [CrossRef]
198. Zhang, Y.; Tan, Y.-W.; Stormer, H. L.; Kim, P. Experimental observation of the quantum Hall effect and Berry’s
phase in graphene. Nature 2005, 438, 201–204. [CrossRef] [PubMed]
199. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Katsnelson, M.I.; Grigorieva, I.V.; Dubonos, S.V.;
Firsov, A.A. Two-dimensional gas of massless Dirac fermions in graphene. Nature 2005, 438, 197–200.
[CrossRef] [PubMed]
200. Zhao, L.; Levendorf, M.; Goncher, S.; Schiros, T.; Pálová, L.; Zabet-Khosousi, A.; Rim, K. T.; Gutiérrez, C.;
Nordlund, D.; Jaye, C.; Hybertsen, M.; Reichman, D.; Flynn, G. W.; Park, J.; Pasupathy, A. N. Local atomic
and electronic structure of boron chemical doping in monolayer graphene. Nano Lett. 2013, 13, 4659–4665.
[CrossRef] [PubMed]
201. Han, W.; Kawakami, R.K.; Gmitra, M.; Fabian, J. Graphene spintronics. Nat. Nanotechnol. 2014, 9, 794–807.
[CrossRef] [PubMed]
202. Kandare, E.; Khatibi, A.A.; Yoo, S.; Wang, R.; Ma, J.; Olivier, P.; Gleizes, N.; Wang, C.H. Improving the
through-thickness thermal and electrical conductivity of carbon fibre/epoxy laminates by exploiting synergy
between graphene and silver nano-inclusions. Compos. Part A 2015, 69, 72–82. [CrossRef]
203. Tang, B.; Hu, G.; Gao, H.; Hai, L. Application of graphene as filler to improve thermal transport property of
epoxy resin for thermal interface materials. Int. J. Heat Mass Transf. 2015, 85, 420–429. [CrossRef]
204. Burger, N.; Laachachi, A.; Mortazavi, B.; Ferriol, M.; Lutz, M.; Toniazzo, V.; Ruch, D. Alignments and network
of graphite fillers to improve thermal conductivity of epoxy-based composites. Int. J. Heat Mass Transf. 2015,
89, 505–513. [CrossRef]
Polymers 2016, 8, 281 35 of 37
205. Zeng, C.; Lu, S.; Xiao, X.; Gao, J.; Pan, L.; He, Z.; Yu, J. Enhanced thermal and mechanical properties of
epoxy composites by mixing noncovalently functionalized graphene sheets. Polym. Bull. 2014, 72, 453–472.
[CrossRef]
206. Wang, F.; Drzal, L.T.; Qin, Y.; Huang, Z. Mechanical properties and thermal conductivity of graphene
nanoplatelet/epoxy composites. J. Mater. Sci. 2014, 50, 1082–1093. [CrossRef]
207. Zhou, T.; Nagao, S.; Sugahara, T.; Koga, H.; Nogi, M.; Suganuma, K.; Nge, T.T.; Nishina, Y. Facile identification
of the critical content of multi-layer graphene oxide for epoxy composite with optimal thermal properties.
RSC Adv. 2015, 5, 20376–20385. [CrossRef]
208. Zeng, C.; Lu, S.; Song, L.; Xiao, X.; Gao, J.; Pan, L.; He, Z.; Yu, J. Enhanced thermal properties in a hybrid
graphene–alumina filler for epoxy composites. RSC Adv. 2015, 5, 35773–35782. [CrossRef]
209. Tang, D.; Su, J.; Yang, Q.; Kong, M.; Zhao, Z.; Huang, Y.; Liao, X.; Liu, Y. Preparation of alumina-coated
graphite for thermally conductive and electrically insulating epoxy composites. RSC Adv. 2015, 5,
55170–55178. [CrossRef]
210. Pan, L.; Ban, J.; Lu, S.; Chen, G.; Yang, J.; Luo, Q.; Wu, L.; Yu, J. Improving thermal and mechanical properties
of epoxy composites by using functionalized graphene. RSC Adv. 2015, 5, 60596–60607. [CrossRef]
211. Wang, R.; Zhuo, D.; Weng, Z.; Wu, L.; Cheng, X.; Zhou, Y.; Wang, J.; Xuan, B. A novel nanosilica/graphene
oxide hybrid and its flame retarding epoxy resin with simultaneously improved mechanical, thermal
conductivity, and dielectric properties. J. Mater. Chem. A 2015, 3, 9826–9836. [CrossRef]
212. Zha, J.-W.; Zhu, T.-X.; Wu, Y.-H.; Wang, S.-J.; Li, R.K.Y.; Dang, Z.-M. Tuning of thermal and dielectric
properties for epoxy composites filled with electrospun alumina fibers and graphene nanoplatelets through
hybridization. J. Mater. Chem. C 2015, 3, 7195–7202. [CrossRef]
213. Zhou, T. Targeted kinetic strategy for improving the thermal conductivity of epoxy composite containing
percolating multi-layer graphene oxide chains. Express Polym. Lett. 2015, 9, 608–623. [CrossRef]
214. Wang, Y.; Yu, J.; Dai, W.; Song, Y.; Wang, D.; Zeng, L.; Jiang, N. Enhanced Thermal and Electrical Properties
of Epoxy Composites Reinforced With Graphene Nanoplatelets. Polym. Compos. 2015. [CrossRef]
215. Pu, X.; Zhang, H.-B.; Li, X.; Gui, C.; Yu, Z.-Z. Thermally conductive and electrically insulating epoxy
nanocomposites with silica-coated graphene. RSC Adv. 2014, 4, 15297–15303. [CrossRef]
216. Fu, Y.-X.; He, Z.-X.; Mo, D.-C.; Lu, S.-S. Thermal conductivity enhancement of epoxy adhesive using graphene
sheets as additives. Int. J. Therm. Sci. 2014, 86, 276–283. [CrossRef]
217. Esposito Corcione, C.; Maffezzoli, A. Transport properties of graphite/epoxy composites: Thermal,
permeability and dielectric characterization. Polym. Test. 2013, 32, 880–888. [CrossRef]
218. Min, C.; Yu, D.; Cao, J.; Wang, G.; Feng, L. A graphite nanoplatelet/epoxy composite with high dielectric
constant and high thermal conductivity. Carbon 2013, 55, 116–125. [CrossRef]
219. Hsiao, M.-C.; Ma, C.-C.M.; Chiang, J.-C.; Ho, K.-K.; Chou, T.-Y.; Xie, X.; et al. Thermally conductive
and electrically insulating epoxy nanocomposites with thermally reduced graphene oxide-silica hybrid
nanosheets. Nanoscale 2013, 5, 5863–5871. [CrossRef] [PubMed]
220. Zhou, T.; Wang, X.; Cheng, P.; Wang, T.; Xiong, D.; Wang, X. Improving the thermal conductivity of
epoxy resin by the addition of a mixture of graphite nanoplatelets and silicon carbide microparticles.
Express Polym. Lett. 2013, 7, 585–594. [CrossRef]
221. Raza, M.A.; Westwood, A.V.K.; Stirling, C. Effect of processing technique on the transport and
mechanical properties of graphite nanoplatelet/rubbery epoxy composites for thermal interface applications.
Mater. Chem. Phys. 2012, 132, 63–73. [CrossRef]
222. Kim, J.; Yim, B.; Kim, J.; Kim, J. The effects of functionalized graphene nanosheets on the thermal and
mechanical properties of epoxy composites for anisotropic conductive adhesives (ACAs). Microelectron. Reliab.
2012, 52, 595–602. [CrossRef]
223. Im, H.; Kim, J. Thermal conductivity of a graphene oxide–carbon nanotube hybrid/epoxy composite. Carbon
2012, 50, 5429–5440. [CrossRef]
224. Heo, Y.; Im, H.; Kim, J.; Kim, J. The influence of Al(OH)3 -coated graphene oxide on improved thermal
conductivity and maintained electrical resistivity of Al2 O3 /epoxy composites. J. Nanopart. Res. 2012, 14,
1–10. [CrossRef]
225. Huang, X.; Zhi, C.; Jiang, P. Toward Effective Synergetic Effects from Graphene Nanoplatelets and Carbon
Nanotubes on Thermal Conductivity of Ultrahigh Volume Fraction Nanocarbon Epoxy Composites. J. Phys.
Chem. C 2012, 116, 23812–23820. [CrossRef]
Polymers 2016, 8, 281 36 of 37
226. Martin-gallego, M.; Verdejo, R.; Khayet, M.; Maria, J.; De Zarate, O.; Essalhi, M.; Lopez-manchado, M.A.
Thermal conductivity of carbon nanotubes and graphene in epoxy nanofluids and nanocomposites.
Nanoscale Res. Lett. 2011, 6, 1–7. [CrossRef] [PubMed]
227. Tien, D.H.; Joonkyu, P.; Sang, A.H.; Muneer, A.; Yongho, S.; Koo, S. Electrical and Thermal Conductivities of
Stycast 1266 Epoxy/Graphite Composites. J. Korean Phys. Soc. 2011, 59, 2760–2764.
228. Yu, A.; Ramesh, P.; Sun, X.; Bekyarova, E.; Itkis, M.E.; Haddon, R.C. Enhanced thermal conductivity in a
hybrid graphite nanoplatelet—Carbon nanotube filler for epoxy composites. Adv. Mater. 2008, 20, 4740–4744.
[CrossRef]
229. Wu, S.; Ladani, R.B.; Zhang, J.; Bafekrpour, E.; Ghorbani, K.; Mouritz, A.P.; Kinloch, A.J.; Wang, C.H. Aligning
multilayer graphene flakes with an external electric field to improve multifunctional properties of epoxy
nanocomposites. Carbon 2015, 94, 607–618. [CrossRef]
230. Liu, X.; Sun, X.; Wang, Z.; Shen, X.; Wu, Y.; Kim, J.-K. Planar Porous Graphene Woven Fabric/Epoxy
Composites with Exceptional Electrical, Mechanical Properties, and Fracture Toughness. ACS Appl.
Mater. Interfaces 2015, 7, 21455–21464. [CrossRef] [PubMed]
231. Ming, P.; Zhang, Y.; Bao, J.; Liu, G.; Li, Z.; Jiang, L.; Cheng, Q. Bioinspired highly electrically conductive
graphene–epoxy layered composites. RSC Adv. 2015, 5, 22283–22288. [CrossRef]
232. Tang, G.; Jiang, Z.-G.; Li, X.; Zhang, H.-B.; Hong, S.; Yu, Z.-Z. Electrically conductive
rubbery epoxy/diamine-functionalized graphene nanocomposites with improved mechanical properties.
Compos. Part B Eng. 2014, 67, 564–570. [CrossRef]
233. Dou, S.; Qi, J.; Guo, X.; Yu, C. Preparation and adhesive performance of electrical conductive epoxy-acrylate
resin containing silver-plated graphene. J. Adhes Sci. Technol. 2014, 28, 1556–1567. [CrossRef]
234. Tang, G.; Jiang, Z.-G.; Li, X.; Zhang, H.-B.; Yu, Z.-Z. Simultaneous functionalization and reduction of graphene
oxide with polyetheramine and its electrically conductive epoxy nanocomposites. Chin. J. Polym. Sci. 2014,
32, 975–985. [CrossRef]
235. Monti, M.; Rallini, M.; Puglia, D.; Peponi, L.; Torre, L.; Kenny, J.M. Morphology and electrical properties of
graphene–epoxy nanocomposites obtained by different solvent assisted processing methods. Compos. Part A
Appl. Sci. Manuf. 2013, 46, 166–172. [CrossRef]
236. Suherman, H.; Sulong, A.B.; Sahari, J. Effect of the compression molding parameters on the in-plane and
through-plane conductivity of carbon nanotubes/graphite/epoxy nanocomposites as bipolar plate material
for a polymer electrolyte membrane fuel cell. Ceram. Int. 2013, 39, 1277–1284. [CrossRef]
237. Mancinelli, P.; Heid, T.F.; Fabiani, D.; Saccani, A.; Toselli, M.; Frechette, M.F.; Savoie, S.; David, E. Electrical
conductivity of graphene-based epoxy nanodielectrics. In Proceedings of the 2013 Annual Report Conference
on Electrical Insulation and Dielectric Phenomena, Shenzhen, China, 20–23 October 2013; pp. 772–775.
238. Al-Ghamdi, A.A.; Al-Hartomy, O.A.; Al-Solamy, F.; Al-Ghamdi, A.A.; El-Tantawy, F. Electromagnetic wave
shielding and microwave absorbing properties of hybrid epoxy resin/foliated graphite nanocomposites.
J. Appl. Polym. Sci. 2013, 127, 2227–2234. [CrossRef]
239. Kim, J.; Im, H.; Kim, J.; Kim, J. Thermal and electrical conductivity of Al(OH)3 covered graphene oxide
nanosheet/epoxy composites. J. Mater. Sci. 2011, 47, 1418–1426. [CrossRef]
240. Fan, Z.; Zheng, C.; Wei, T.; Zhang, Y.; Lu, G. Effect of Carbon Black on Electrical Property of Graphite
Nanoplatelets/Epoxy Resin Composites. Polym. Eng. Sci. 2009, 49, 2041–2045. [CrossRef]
241. Jović, N.; Dudić, D.; Montone, A.; Antisari, M.V.; Mitrić, M.; Djoković, V. Temperature dependence of the
electrical conductivity of epoxy/expanded graphite nanosheet composites. Scr. Mater. 2008, 58, 846–849.
[CrossRef]
242. Li, J.; Ma, P.C.; Chow, W.S.; To, C.K.; Tang, B.Z.; Kim, J.-K. Correlations between Percolation Threshold,
Dispersion State, and Aspect Ratio of Carbon Nanotubes. Adv. Funct. Mater. 2007, 17, 3207–3215. [CrossRef]
243. Sandler, J.; Shaffer, M.S.; Prasse, T.; Bauhofer, W.; Schulte, K.; Windle, A. Development of a dispersion process
for carbon nanotubes in an epoxy matrix and the resulting electrical properties. Polymer (Guildf) 1999, 40,
5967–5971. [CrossRef]
244. Raza, M.A.; Westwood, A.; Stirling, C. Effect of processing technique on the transport and mechanical
properties of vapour grown carbon nanofibre/rubbery epoxy composites for electronic packaging
applications. Carbon 2012, 50, 84–97. [CrossRef]
245. Mas, B.; Fernández-Blázquez, J.P.; Duval, J.; Bunyan, H.; Vilatela, J.J. Thermoset curing through Joule heating
of nanocarbons for composite manufacture, repair and soldering. Carbon 2013, 63, 523–529. [CrossRef]
Polymers 2016, 8, 281 37 of 37
246. Chang, K.-C.; Hsu, M.-H.; Lu, H.-I.; Lai, M.-C.; Liu, P.-J.; Hsu, C.-H.; Ji, W.-F.; Chuang, T.-L.; Wei, Y.; Yeh, J.-M.;
Liu, W.-R. Room-temperature cured hydrophobic epoxy/graphene composites as corrosion inhibitor for
cold-rolled steel. Carbon 2014, 66, 144–153. [CrossRef]
247. Prolongo, S.G.; Jiménez-Suárez, A.; Moriche, R.; Ureña, A. Graphene nanoplatelets thickness and lateral size
influence on the morphology and behavior of epoxy composites. Eur. Polym. J. 2014, 53, 292–301. [CrossRef]
248. Prolongo, S.G.; Moriche, R.; Jiménez-Suárez, A.; Sanchez, M.; Ureña, A. Advantages and disadvantages of
the addition of graphene nanoplatelets to epoxy resins. Eur. Polym. J. 2014, 61, 206–214. [CrossRef]
© 2016 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC-BY) license (http://creativecommons.org/licenses/by/4.0/).