Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

J Polym Eng 2015; 35(3): 257–266

Rahim Eqra, Kamal Janghorban* and Habib Daneshmanesh

Mechanical properties and toughening


mechanisms of epoxy/graphene nanocomposites
Abstract: Because of extraordinary physical, chemical and properties [1, 2], which are dependent on the morphol-
mechanical properties, graphene nanosheets (GNS) are ogy, the homogenous dispersion, the types of fillers
suitable fillers for optimizing the properties of different and matrices, and also the interface bonding between
polymers. In this research, the effect of GNS content (up to the matrix and the reinforcements [3, 4]. These types of
1 wt.%) on tensile and flexural properties, morphology of materials have been utilized in various fields of science
fracture surface, and toughening mechanism of epoxy were and engineering such as photovoltaic cells, electronics,
investigated. Results of mechanical tests showed a peak for sensors, coatings and mechanical structures for many dif-
tensile and flexural strength of samples with 0.1 wt.% GNS ferent applications such as aerospace industries, defense
such that the tensile and flexural strength improved by 13% systems, medical equipment and even in daily life.
and 3.3%, respectively. The Young’s modulus and flexural Graphene has been used as a prominent nanofiller
modulus increased linearly with GNS content, although since its discovery in 2004 by Navoselov et  al. [5]. It
the behavior of the Young’s modulus was more remark- has a two-dimensional honeycomb structure of carbon
able. Morphological investigations confirmed this behavior atoms with sp2 bonds and, owing to its extraordinary
because the GNS dispersion in the epoxy matrix was uni- mechanical, chemical and electrical properties, is con-
form at lower contents and agglomerated at higher con- sidered to be the most suitable nanofiller for optimizing
tents. Finally, microscopical observation showed that the and improving the properties of polymers. Graphene can
major toughening mechanism of graphene-epoxy nano- be produced by four main methods: (1) chemical vapor
composites was crack path deflection, which changed the deposition (CVD), (2) epitaxial growth of graphene films
mirror fracture surface of the pure epoxy to rough surface. on electrically insulating substrates, (3) mechanical
exfoliation of graphene from bulk graphite (e.g., using
Keywords: epoxy; graphene; mechanical properties; Scotch tape5) and (4) thermal or chemical reduction of
nanocomposite; toughening mechanism. graphene derivatives such as graphene oxide. The fourth
method has the potential for producing graphene in bulk
DOI 10.1515/polyeng-2014-0134
quantities that are suitable for use in composite materi-
Received May 26, 2014; accepted September 10, 2014; previously als [1, 6–11].
published online October 9, 2014 Highly cross-linked thermoset epoxy provides high
modulus and strength, good resistance to creep and cor-
rosion, and good performance at elevated temperatures.

1 Introduction Because of its excellent properties, it is widely used as


the main component for coating, adhesives and matrices
in various engineering fields, from structural composites
Compared to the properties of conventional composites,
to microelectronics, such as spacecraft, and as impreg-
the properties of polymeric materials are enhanced signif-
nating materials in the cryogenic environment and also
icantly by the addition of nanofillers, without compromis-
in superconducting cable technology. However, the high
ing their process ability, inherent mechanical properties
cross-link density makes these materials inherently
and light weights. Polymer nanocomposites have a unique
brittle, leading to poor resistance to crack propagation
combination of mechanical, thermal and electrical
[12–14]. Therefore, epoxy-graphene nanocomposites have
attracted many researchers in recent years to improve
*Corresponding author: Kamal Janghorban, Department of their mechanical, thermal and electrical properties
Materials Science and Engineering, Engineering School, Shiraz [3, 4, 11–17]. In these studies, the uniform distribu-
University, P. O. Box 71348-15939, Shiraz, Iran,
tion of graphene in epoxy matrix and the improvement
e-mail: janghor@shirazu.ac.ir
Rahim Eqra and Habib Daneshmanesh: Department of Materials
of its interface to achieve the effect of graphene filler
Science and Engineering, Engineering School, Shiraz University, have been investigated [4, 18–20]. The results of these
P. O. Box 71348-15939, Shiraz, Iran research studies show that the mechanical properties

Brought to you by | New York University Bobst Library Technical Services


Authenticated
Download Date | 7/1/15 1:07 PM
258      R. Eqra et al.: Polymer nanocomposites

of graphene-epoxy nanocomposites enhanced when the thermal shock condition at 1050°C in nitrogen atmos-
graphene content reached a peak and then decreased [4, phere. In this step, the thermal shock caused a very large
11–14]. Also, the thermal, electrical and especially dielec- expansion (up to 200%) and the graphite oxide sheets
tric characteristics and electromagnetic shielding (EMS) were simultaneously exfoliated, reduced and changed to
of epoxy improved with graphene [3, 15–17]. Acceptable GNS. Finally, the GNS were prepared for nanocomposite
interface bonding and uniform dispersion were obtained production by sonication of these sheets in acetone at
by using appropriate functional and dispersing agents [4, 400 W of power for 4 h
18–20]. Little information, however, has been reported on
the bending characteristics, fracture surface morphology
and toughening mechanism of graphene-epoxy nano- 2.3 Graphene-nanocomposite production
composites. In contrast, because of the many types of
thermoset epoxy resins with different properties and the The produced GNS with acetone were agitated using an
dependency of graphene characteristics on the produc- ultrasonic instrument under 100  W for 2 h. Then epoxy
tion method, further research is necessary to optimize the resin was added to the solution and sonicated under
properties of new graphene-epoxy nanocomposites. In the aforementioned condition for 90 min. Acetone was
this work, graphene nanosheets (GNS) were synthesized removed under vacuum at 60°C for 18 h. Afterward, a
by thermal expansion and reduction of graphene oxide, hardener was added and mixed by a magnetic stirrer.
then graphene epoxy nanocomposites based on bysphe- This mixture was degassed in a vacuum chamber and
nol-F resin were fabricated with different compositions. was injected into the tensile and bending sample mold.
Subsequently, the tensile and bending characteristic, Curing was finished after 7 days. A reference sample of
fracture surface morphology and toughening mechanism pure epoxy was also produced. The processing sequence
of these nanocomposites were investigated. is shown in Figure 1.

2.4 Characterization
2 Materials and methods
The purity of the graphite, graphite oxide and GNS was
2.1 Materials studied using a GNR Explorer diffractometer with Cu Kα
radiation (λ = 1.54 Å) at 40 kV. The morphology of GNS and
Graphite powder (art. no. 1.04206.2500), acetone and other the fracture surface of the nanocomposites and epoxy
materials for the synthesis of GNS such as sulfuric acid, sample were examined by a scanning electron microscope
nitric acid, potassium chlorate, hydrochloric acid and (SEM, Leica Cambridge S360, Cambridge, England, UK).
barium chloride were purchased from Merck (­Darmstadt, The dispersed GNS in acetone were dropped onto a glass
Germany). Epoxy resin (M506) based on bisphenol-F and lam, allowed to dry and then coated with gold for SEM
the hardener (HA11) polyamine were supplied by Mavad
Mokarrar Company (Tehran, Iran).

Graphene agitation by ultrasonic in


acetone
2.2 Prepartion of GNS
Epoxy resin with graphene agitation by
The GNS were prepared by the thermal expansion and ultrasonic in acetone
reduction method [20]. Sulfuric acid (114 ml, 98 wt.%)
and nitric acid (67 ml, 67 wt.%) with 6 g of graphite were
Removal of acetone
stirred and mixed in ice-water bath. After 15 min, potas-
sium chlorate (63 g) was added gradually to the mixture
and the mixture was blended for 96 h. The mixture was Hardner addition and degassing
then washed with hydrochloric acid solution (0.1 mol/l),
and this continued until all sulfate ions disappeared and
nothing remained to react with the barium chloride solu- Injection of nanocomposites

tion. Afterward, this graphite oxide was dried at 70°C.


Then 3 g of this graphite oxide was held for 30  s under Figure 1 Schematic of the processing procedure.

Brought to you by | New York University Bobst Library Technical Services


Authenticated
Download Date | 7/1/15 1:07 PM
R. Eqra et al.: Polymer nanocomposites      259

imaging. The tensile and bending properties of the nano- expansion step and the GNS were successfully formed
composite samples were measured under a 100-kN load with disordered planes.
cell with a cross-head speed of 1 mm/min based on the The Raman spectroscopy method is a well-estab-
ASTM D638-03 standard. Three-point flexural test was per- lished technique for characterizing GNS. The G band at
formed on samples 10 mm wide, 100 mm long and 4 mm nearly 1590  cm-1 relates to the zone center phonons of
thickness with a support span of 60  mm at a constant E2g symmetry at the Γ point of the first Brillouin zone of
cross-head speed of 1 mm/min according to the ASTM D790 graphite and includes bond stretching of sp2 carbon pairs
standard using a mechanical tester [Hegwald Peschke (HP) in the graphite plane [14, 21]. The D band at approximately
inspect 50, Germany]. The Fourier transform infrared (FTIR) 1350  cm-1 is attributed to the K-point phonon of the A1g
spectra of GNS and nanocomposite samples were obtained symmetry, which is associated with disordered sp3 hybrid-
using a Tensor 27 FTIR (Bruker, Ettlingen, Germany) ized carbon, present as impurities and defects in the gra-
between 500 and 4000 cm-1. The GNS or the nanocomposite phene structure, and its intensity is dependent on the
sample powder mixture with the KBr powders was molded presence of six-fold aromatic rings. Thus, the ratio of the
into discs for this characterization. Raman spectroscopy intensities of the D and G bands (ID/IG) is usually indica-
characterization was done using a SENTERRA-2009 system tive of the relative disorder present in graphitic structures
(Bruker) with a 785-nm Ar laser between 200 and 3500 cm-1. [22–24]. Figure 3 shows the Raman spectra of graphite,
graphite oxide and synthesized GNS. For graphite, the
strong G band at 1580 cm-1 is reflected to the highest equi-

3 Results and discussion librium stretching of sp2 hybridized carbon atom in its
planes and the weak D band at 1310 cm-1 is due to the edge
effect and inherent defects [21, 23]. In contrast, the ID/IG
3.1 Characterization of GNS ratio of 0.35 of graphite confirmed its relatively low defect
structure. Meanwhile, oxidation of graphite substantially
The XRD results of graphite, graphite oxide and gra- changed its structure such that the G band was broadened
phene are shown in Figure 2. Graphite shows a sharp and the D band showed a higher relative intensity, which
peak at 2θ = 26.5°, which is related to the diffraction of the was attributed to the increase in impurities and defects
repeated plane of graphite with an interlayer distance of in the graphite structure. Thus the ID/IG ratio rose to 1.75.
3.35 Å according to Bragg’s law (nλ = 2dsinθ). A relatively Moreover, the intensity of the D and G bands decreased
low peak at 2θ = 12.6° confirmed that the graphite planes and increased, respectively, during the transformation of
intercalated and graphite oxide formed with an interlayer graphite oxide to graphene by the heat reduction process
spacing of 7.8 Å at the oxidation step. GNS did not show due to graphite “self-healing” [21, 25], which caused the
any peak, indicating that the graphite oxide was simul- reduction in the ID/IG ratio to 1.3. Unlike in other research
taneously exfoliated and reduced during the thermal reports [21, 25], the shift of the G band was not observed
in thermally reduced graphene, which can be attributed
to the higher disorder of the graphite used in this research
25,000

Graphite
20,000
Graphite oxide
Purchased graphite
Synthesized graphene ID/IG=0.35
Intensity (a.u.)

15,000
Intensity (a.u.)

Synthesized graphite oxide


10,000 ID/IG=1.75

5000
Synthesized graphene
ID/IG=1.3

0
0 10 20 30 40 500 1000 1500 2000 2500
2θ Wavenumber (cm-1)

Figure 2 XRD patterns of the graphite, graphite oxide and synthe- Figure 3 Raman spectra of the graphite, graphite oxide and synthe-
sized graphene. sized graphene.

Brought to you by | New York University Bobst Library Technical Services


Authenticated
Download Date | 7/1/15 1:07 PM
260      R. Eqra et al.: Polymer nanocomposites

(ID/IG = 0.35) compared to the values of 0.05 and 0.2 used in


other research studies.
Figure 4 shows a direct comparison between FTIR
spectra of the graphite, graphite oxide and synthesized
GNS. The peaks at 1050, 1440 and 1625 cm-1 and the strong
broadened peak at about 3430  cm-1 are attributed to the
C-O stretching vibrations of the epoxide groups, the C-OH
stretching vibrations of the carboxyl groups, the C = O
stretching vibrations of the carbonyl groups and the O-H
stretching vibrations of the hydroxyl groups, respectively
[12, 25, 26]. As seen in the figure, the peaks did not change
in the synthesized process of graphene. Furthermore, the
intensity of the peaks increased in the oxidation of graph-
ite, which was assigned to the oxygen functionalities in the
graphite oxide. After the formation of reduced graphene,
the bands around 1050, 1440, 1625 and 3430 cm-1 were still Figure 5 SEM image of the synthesized graphene.
present on the graphene surface with lower intensities [22,
27]. A scanning electron microscope image of the synthe-
sized GNS is illustrated in Figure 5. Wrinkled disordered 14, 18, 28]. By further increase of GNS content (zone II),
sheets of the synthesized graphene were observed in this the strength dropped sharply by crack initiation in the
figure and were confirmed by the XRD results. agglomerated regions of GNS. Meanwhile, the strain to
fracture of nanocomposite decreased slightly because the
crack propagation was blocked or deflected by the GNS.
3.2 Tensile characteristics In zone III (0.3–0.5 wt.%), the nanocomposite strength
gently decreased, but the strain to fracture of nanocom-
The uniaxial tension test results are illustrated in Figure 6. posites decreased, which could be attributed to easy
The strength and strain to fracture of pure epoxy and crack propagation through a closed and large number
nanocomposite samples are shown in Figure 7. In zone I, of agglomerated regions. The strength and strain to
with a GNS content of 0.1 wt.%, the strength and strain fracture of nanocomposites are almost flattened in zone
of composite samples increased to 13% and 3.1%, respec- IV because the GNS were nearly saturated in the epoxy
tively, compared to the pure epoxy samples, due to a good matrix. The above results show that the GNS act as a rein-
dispersion and excellent intrinsic strength (130 GPa) of forcement in zone I but as a defect in zone II and zone III.
the GNS that were applied to reinforce the composites. In contrast, the roughness or total fracture surface
The same results were reported by other researchers [12, increased with increasing GNS content up to 0.5 wt.% and

Graphite
90
Graphite oxide
80
Synthesized graphene Strain rate: 0.002 S-1
70
Absorbance (a.u.)

60
Stress (MPa)

50
Pure epoxy
40
0.1 wt.% graphene
30
0.3 wt.% graphene
20
0.5 wt.% graphene
10 1 wt.% graphene
0
500 1500 2500 3500 4500 0 0.02 0.04 0.06 0.08 0.1 0.12
Wavenumber (cm-1) Strain

Figure 4 FTIR spectra of the graphite, graphite oxide and synthe- Figure 6 Stress-strain curves of the nanocomposite samples and
sized graphene. pure epoxy.

Brought to you by | New York University Bobst Library Technical Services


Authenticated
Download Date | 7/1/15 1:07 PM
R. Eqra et al.: Polymer nanocomposites      261

90 12 5
11.5
4.5
85
I II III IV 11
Tensile strength (MPa)

Strength
Strain 10.5 4
80

Toughness (J/cm3)
10

Strain (%)
3.5
75 9.5
9 3
70
8.5
2.5
65 8
7.5 2
60 7
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.5
Graphene (wt.%) 1
0 0.5 1 1.5
Figure 7 Effect of the graphene content on the tensile strength and Graphene (wt.%)
strain to fracture of the nanocomposite samples and pure epoxy.
Figure 9 Toughness of the nanocomposite samples.

was approximately fixed at higher GNS content (Figure 8). in Figure 10. This result was pointed out by other research-
As shown in Figure 9, however, the total energy required to ers, too [12, 14, 28].
produce this surface (toughness or total area under stress
and strain curve) decreased after the addition of 0.1 wt.%
of GNS content. This contradiction confirmed that the pro- 3.3 Flexural characteristics
duction of new fracture surface required less energy owing
to the agglomeration of GNS. Also, the dramatic decline in Flexural test results are illustrated in Figures 11, 12 and
toughness in the range of 0.3–0.5 wt.% of GNS content con- 13. As shown in Figures 12 and 13, the flexural strength
firmed that the toughness did not considerably decrease and flexural strain enhanced and reached 3.3% and 29%,
while the GNS were acting as a reinforcement, even after respectively, for the 0.1 wt.% nanocomposite sample com-
the initiation of cracks due to the agglomeration of GNS pared to pure epoxy, and decreased with increasing GNS
(zone II in Figure 7). However, GNS noticeably enhanced content. Similar information has been reported by other
the elastic modulus of the nanocomposites owing to their investigators [25, 29]. Like the tension results, the flexural
excellent elastic modulus (1000 GPa), which is illustrated strength and flexural strain increased up to 0.1 wt.% GNS

A B C

Figure 8 SEM images of the epoxy-graphene nanocomposite samples. (A) 0.1 wt.%, (B) 0.5 wt.% and (C) 1 wt.%.

Brought to you by | New York University Bobst Library Technical Services


Authenticated
Download Date | 7/1/15 1:07 PM
262      R. Eqra et al.: Polymer nanocomposites

3.5

3
Young’s modulus (GPa)

2.5

2
Strain rate: 0.002 S-1
1.5

0.5

0
0 0.1 0.2 0.3 0.4 0.5 0.6 Figure 12 Flexural strength of the nanocomposite samples and
Graphene (wt.%) pure epoxy.

Figure 10 Effect of the graphene content on the Young’s modulus of


the nanocomposite samples.
the tensile strength remained greater up to 1 wt.%. It is
attributed to the better tensile properties of graphene
relative to its flexural properties. The agglomeration char-
due to the reinforcement effect of GNS and then decreased
acteristic of graphene, which acts as a defect in epoxy
as the GNS content increased owing to its agglomeration.
matrix, is more destructive in the reduction of flexural
The flexural modulus of nanocomposite samples is shown
strength of nanocomposite samples. In contrast, tensile
in Figure 14 as a function of the GNS content. The flexural
strain to fracture was reduced below the pure epoxy at
modulus increased slightly as the GNS content increased,
0.5 wt.% graphene content owing to the large number
and 7% enhancement with respect to pure epoxy was
attained for 1 wt.% GNS. The flexural strength and the
flexural modulus did not dramatically increase by adding
GNS fillers, but flexural strain was remarkably enhanced
for 0.1 wt.% GNS, which indicates a good interface
bonding of GNS with epoxy, and the growth of crack in
nanocomposites was delayed. In contrast, compared with
tensile strength and tensile modulus, the flexural strength
and flexural modulus of the nanocomposite samples did
not increase significantly, which points to better tensile
properties than the flexural properties of GNS.
The general results of tensile and flexural tests are
illustrated in Table 1. As shown in the table, the flexural
strength of nanocomposite samples containing 0.5 wt.% Figure 13 Flexural strain of the nanocomposite samples and pure
graphene was less than that of the pure epoxy. Meanwhile, epoxy.

Figure 11 Flexural stress-strain curves of the nanocomposite Figure 14 Effect of the graphene content on the flexural modulus of
samples and pure epoxy. the nanocomposite samples.

Brought to you by | New York University Bobst Library Technical Services


Authenticated
Download Date | 7/1/15 1:07 PM
R. Eqra et al.: Polymer nanocomposites      263

PIOE of closely agglomerated regions of graphene that caused

7.5
5
7.3
0.59
0
easier crack propagation through nanocomposite samples
than the pure epoxy.

3107 
3034 
3101 
2907 
2890 
Flexural 
modulus (MPa)

3.4 FTIR spectroscopy results

Figure 15 shows the FTIR spectrum of nanocompos-


ite samples and pure epoxy. The broad band around
-25.6 
1.48 
21.3 
29.7 
0 
PIOE 

3300  cm-1 is attributed to the stretching vibration of the


hydroxyl groups (O-H). The peaks at 2925 and 2855 cm-1 are
4.54 
6.19 
7.4 
7.91 
6.1 
Flexural 
strain (%)

related to the stretching vibration group (C-H) in the CH2


and CH3 bands, respectively. The peaks at 1726, 1226 and
1103 cm-1 are in the range of the stretching vibration bands
of the carbonyl (C = O), carboxyl (C-OH) and epoxide (C-O)
-2.57 
-1.25 
0 
3.3 
0 
PIOE 

groups, respectively.

3.5 Raman spectroscopy results


94.5 
95.79 
97 
100.2 
97 
Flexural strength 
(MPa)

The Raman spectroscopy of nanocomposite samples is


shown in Figure 16. The D band remained constant and
its intensity increased with increasing GNS content. The
0.1, 0.3 and 1 wt.% nanocomposite samples exhibited the
– 
158 
– 
32.9 
0 
PIOE 

G-band position at 1610, 1600 and 1580 cm-1, respectively,


which showed the G-band position moving to the left as
– 
3690 
– 
1900 
1430 
Tensile modulus 
(MPa)

the GNS content increased. The above results indicated


that, with the increase in GNS content of the nanocompos-
ite samples, the amount of GNS agglomeration increased
[12, 22]. The same results were obtained in the tension
characterization section.
-5.1 
-7 
2 
3.1 
0 
PIOE 

3.6 M
 orphology and toughening
mechanisms
9.3 
9.1 
10 
10.1 
9.8 
Tensile strain 
to fracture (%)

Figure 17 shows the morphology of the fracture surface


of epoxy samples with the addition of GNS. In contrast,
Table 1 Summary of the tensile and flexural test results.

2.2 
4.3 
7.3 
13 
0 
PIOE 

71 
72.5 
74.6 
78.6 
69.5 
Tensile strength 
(MPa)

PIOE, Percent increase over pure epoxy.


1 wt.% graphene  
0.5 wt.% graphene 
0.3 wt.% graphene 
0.1 wt.% graphene 

Pure epoxy
Sample

Figure 15 FTIR spectra of the nanocomposite samples and pure


epoxy.

Brought to you by | New York University Bobst Library Technical Services


Authenticated
Download Date | 7/1/15 1:07 PM
264      R. Eqra et al.: Polymer nanocomposites

1600
0.1 wt.% graphene
but GNS in the nanocomposite samples act as an obstacle
1400 0.3 wt.% graphene to prevent the crack growth and straight extent of the frac-
ture surface. This caused a crack growth path deflection,
Raman intensity (a.u.)

1200 1 wt.% graphene


which improved the toughness of the nanocomposites
1000
(Figure 18B and C). Another toughening mechanism is the
800
GNS pullout, which is shown in Figure 17C. It should be
600
mentioned that tensile stress is transferred to the GNS and
400 stretches it in the tension direction (Figure 18C). Finally,
200 the enhancement of the mechanical properties of the
0 nanocomposite samples at 0.1 wt.% GNS content can be
1000 1200 1400 1600 1800 2000
attributed to the homogeneous dispersion of graphene,
Wavenumber (cm-1)
which was confirmed by the uniform fracture surface
Figure 16 Raman spectra of the nanocomposite samples. shown in Figure 17B, and the formation of cracks in the
agglomeration area of GNS at higher graphene contents,
which has destructive effects on the mechanical proper-
Figure 18 shows the morphology of epoxy and its nano- ties, is shown in Figure 18D.
composites parallel to the tension axes. As seen in the Figure 19 illustrates the flexural fracture surface of
figure, the smooth and mirror fracture surface of the the nanocomposite samples and pure epoxy. Similar to
tension test of the pure epoxy sample is converted to the tension results, the mirror fracture surface of pure
rough surface by adding GNS (Figure 17A,B). Acceptable epoxy changed to rough surface for the nanocomposite
interface bonding and agglomeration of GNS in the epoxy samples owing to the reinforcement of graphene and its
matrix are shown in Figure 17C and D, respectively. Fur- crack path deflection toughening mechanism (Figure
thermore, the smooth surface of the pure epoxy sample 19A–C). The agglomeration produced by the increased
caused crack growth and the extent of the failure of the amount of GNS in the nanocomposite samples is shown
surface to spread easily without any barrier (Figure 18A), in Figure 19D.

A B

C D

Figure 17 SEM images. (A) The mirror fracture surface of pure epoxy. (B) The fracture surface of the 0.1 wt.% nanocomposite sample.
(C) The fracture surface parallel to tension axes of the 0.1 wt.% nanocomposite sample. (D) The fracture surface of the 1 wt.% nanocomposite
sample.

Brought to you by | New York University Bobst Library Technical Services


Authenticated
Download Date | 7/1/15 1:07 PM
R. Eqra et al.: Polymer nanocomposites      265

A B

Graphene nanosheets

C D

Stretched graphene
nanosheets Crack

Figure 18 SEM images of the fracture surface parallel to tension axes. (A) Pure epoxy. (B) 0.1 wt.% nanocomposite sample. (C) 0.3 wt.%
nanocomposite sample. (D) 1 wt.% nanocomposite sample.

4 Conclusions epoxy samples increased by 13% and 3.1%, respectively.


By further increase of GNS content, the nanocomposite
With loading GNS of up to 0.1 wt.%, the tensile strength strength decreased significantly owing to the crack initia-
and strain of composite samples compared to the pure tion in the agglomerated regions of GNS. The toughness

A B

C D

Agglomeration

Figure 19 SEM images of the flexural fracture surface. (A) The mirror surface of pure epoxy. (B, C) 0.1 wt.% nanocomposite sample.
(D) 0.5 wt.% nanocomposite sample.

Brought to you by | New York University Bobst Library Technical Services


Authenticated
Download Date | 7/1/15 1:07 PM
266      R. Eqra et al.: Polymer nanocomposites

of the nanocomposite samples increased, which was [4] Tkalya EE, Ghislandi M, De With G, Koning CE. Curr. Opin. Col-
the result of the GNS content and was not considerably loid Interface Sci. 2012, 17, 225–232.
[5] Navoselov KS, Geim AK, Morozov SV, Jiang D, Zhang Y, Dubonos
reduced until the GNS acted as a reinforcement even
SV, Grigorieva IV, Firsov AA. Science 2004, 306, 666–9.
after the initiation of cracks due to the agglomeration of [6] Gong JR. Graphene — Synthesis, Characterization, Properties
GNS. GNS noticeably enhanced the elastic modulus of the and Applications. InTech: Rijeka, 2011.
nanocomposite samples owing to their excellent elastic [7] Wajid AS, Das S, Irin F, Tanvir Ahmed HS, Shelburne JL, Parviz
modulus. D, Fullerton RJ, Jankowski AF, Hedden RC, Green MJ. Carbon
2012, 50, 526–534.
The flexural strength and flexural strain were
[8] Kim H, Abdala AA, Macosko CW. Macromolecules 2010, 43,
enhanced and reached 3.3% and 29.7%, respectively, for 6515–6530.
the 0.1 wt.% nanocomposite sample compared to pure [9] Stankovich S, Dikin DA, Dommett GH, Kohlhaas KM, Zimney EJ,
epoxy, and decreased with increasing GNS content. The Stach EA, Piner RD, Nguyen ST, Ruoff RS. Nat. Lett. 2006, 442,
flexural modulus increased slightly as the GNS content 282–286.
increased, and a 7% enhancement with respect to pure [10] Potts JR, Dreyer DR, Bielawski CW, Ruoff RS. Polymer 2011, 52,
5–25.
epoxy was attained for the 1  wt.% GNS. The flexural
[11] Rafiee MA, Rafiee J, Wang Z, Song H, Yu ZZ, Koratkar N. ACS
strength and flexural modulus did not dramatically Nano 2009, 3, 3884–3890.
increase by adding GNS fillers, but flexural strain was [12] Shen XJ, Liu Y, Xiao HM, Feng QP, Yu ZZ, Fu SY. Compos. Sci.
remarkably enhanced, which confirmed the good inter- Technol. 2012, 72, 1581–1587.
face bonding of GNS with epoxy. [13] Zaman I, Lip TM, Le QH, Ma J. 18th International Conference on
Composite Materials, South Korea, August 21–26, 2011.
Compared with tensile strength and tensile modulus,
[14] Chatterjee S, Wang JW, Kuo WS, Tai NH, Salzmann C, Li WL,
the flexural strength and flexural modulus of the nano-
­Hollertz R, Nüesch FA, Chu BT. Chem. Phys. Lett. 2012, 531, 6–10.
composite samples did not increase dramatically, which [15] Tien DH, Park J, Han SA, Ahmad M, Seo Y. J. Kor. Phys. Soc.
points to the better tensile properties rather than the 2011, 59, 2760–2746.
flexural properties of GNS. The tensile strain to fracture [16] Tien DH, Park J, Han SA, Hong S, Seo Y. 18th International
and the flexural strength of the nanocomposite samples ­Conference on Composite Materials, Korea, 2011.
[17] Liang J, Wang Y, Huang Y, Ma Y, Liu Z, Cai J, Zhang C, Gao H,
were less than those of pure epoxy at 0.5 wt.% graphene
Chen Y. Carbon 2009, 47, 922 –925.
content, but the tensile strength of the nanocomposite [18] Wang X, Xing W, Zhang P, Song L, Yang H, Hu Y. Compos. Sci.
samples remained greater than that of pure epoxy, up to Technol. 2012, 72, 737–743.
1 wt.% GNS. [19] Yang SY, Lin WN, Huang YL, Tien HW, Wang JY, Ma CC, Li SM,
The major toughening mechanism of graphene- Wang YS. Carbon 2011, 49, 793 –803.
[20] Guo P, Song H, Chen X, Ma L, Wang G, Wang F. Anal. Chim. Acta
epoxy nanocomposites is crack path deflection by
2011, 688, 146–155.
GNS, which caused the mirror fracture surface of pure
[21] Teng CC, Ma CC, Lu CH, Yang SY, Lee SH, Hsiao MC, Yen MY,
epoxy to convert to the rough fracture surface of epoxy Chiou KC, Lee TM. Carbon 2011, 49, 5107–5116.
containing GNS, although the graphene pullout was [22] Villar-Rodil S, Paredes JI, Martinez-Alonso A, Tascón JM.
also seen in the fracture surface of graphene-epoxy J. Mater. Chem. 2009, 19, 3591–3593.
nanocomposites. [23] Singh V, Joung D, Zhai L, Das S, Khondaker SI, Seal S. Prog.
Mater. Sci. 2011, 56, 1178–1271.
[24] Din Khan SU, Arora M, Wahab MA, Saini PJ. Polymers 2014,
2014, 1–7.
[25] Naebe M, Wang J, Amini A, Khayyam H, Hameed N, Li L, Chen Y,
References Fox B. Sci. Rep. 2014, 4, 1–7; doi: 10.1038/srep04375.
[26] Yan Zhao Y, Liu W, Zheng H. 18th International Conference on
[1] Galpaya D, Wang M, Liu M, Motta N, Waclawik E, Yan C. Composite Materials, Korea, August 21–26, 2011.
­Graphene 2012, 1, 30–49. [27] Zhu X, Liu Q, Zhu X, Li C, Xu M, Liang Y. Int. J. Electrochem. Sci.
[2] Du JH, Cheng HM. Macromol. Chem. Phys. 2012, 213, 2012, 7, 5172–5184.
1060–1077. [28] Ovid’ko IA. Rev. Adv. Mater. Sci. 2013, 34, 19–25.
[3] De Bellis G, Tamburrano A, Dinescu A, Santarelli ML, Sarto MS. [29] Lee SY, Chong MH, Park M, Kim HY, Park SJ. Carbon Lett. 2014,
Carbon 2011, 49, 4291–4300. 15, 67–70.

Brought to you by | New York University Bobst Library Technical Services


Authenticated
Download Date | 7/1/15 1:07 PM

You might also like