Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Polymer Testing 31 (2012) 31–38

Contents lists available at SciVerse ScienceDirect

Polymer Testing
journal homepage: www.elsevier.com/locate/polytest

Material properties

Effect of functionalized graphene on the physical properties of linear low


density polyethylene nanocomposites
Tapas Kuila a, Saswata Bose a, Ananta Kumar Mishra b, Partha Khanra a, Nam Hoon Kim c,
Joong Hee Lee a, b, c, *
a
Department of BIN Fusion Technology, Chonbuk National University, Jeonju, Jeonbuk 561-756, Republic of Korea
b
BIN Fusion Research Team, Department of Polymer & Nano Engineering, Chonbuk National University, Jeonju, Jeonbuk 561-756, Republic of Korea
c
Department of Hydrogen and Fuel Cell Engineering, Chonbuk National University, Jeonju, Jeonbuk 561-756, Republic of Korea

a r t i c l e i n f o a b s t r a c t

Article history: Dodecyl amine-modified graphene (DA-G)/linear low density polyethylene (LLDPE)
Received 6 August 2011 nanocomposites were prepared through solution mixing. Field emission scanning electron
Accepted 19 September 2011 microscopy analysis revealed homogeneous dispersions of graphene layers in the
nanocomposites. X-ray diffraction analysis showed that the average crystallite size of the
Keywords: nanocomposites was increased. However, the % crystallinity was found to decrease due to
Graphene
the formation of a random interface. Dynamic mechanical analysis showed that the storage
Nanocomposites
moduli of the nanocomposites were much higher than that of neat LLDPE. The
Dynamic mechanical analysis (DMA)
Thermomechanical analysis nanocomposites were also more thermally stable than neat LLDPE. Isothermal thermog-
Gas barrier property ravimetry showed that homogeneously distributed graphene could act as a good inhibitor
during thermal degradation of the nanocomposites. Differential scanning calorimetry
showed that the crystallization temperature of the nanocomposites increased with
increasing DA-G content. Thermomechanical analysis showed that the dimensional
stability of the nanocomposites was significantly increased by the addition of the DA-G.
The coefficients of thermal expansion decreased with increasing DA-G content. The
oxygen and nitrogen permeability of the nanocomposites was lower than that of neat
LLDPE.
Ó 2011 Elsevier Ltd. All rights reserved.

1. Introduction carbon nanotubes (CNTs) and carbon nanofibers (CNFs)


have been widely used to prepare polymer nanocomposites
Polymer nanocomposites have attracted considerable [6–10]. Among these, CNTs have been reported to be very
attention due to their wide range of applications such as effective conductive fillers [10,11] in terms of mechanical,
automobiles, aerospace, packaging materials, coatings and electrical and thermal properties. However, the high
pigments and construction engineering [1–5]. Nanofiller production cost of CNT limits their application as nano-
distributed in the polymer matrix can improve mechanical, fillers for the preparation of polymer nanocomposites. The
thermal, gas barrier, flame retardant and electrical prop- discovery of graphene in 2004 has largely circumvented
erties of the nanocomposite materials [1–5]. Carbon-based this problem [12,13].
nanofillers such as carbon black, exfoliated graphite (EG), Graphene, a monolayer of carbon atoms arranged in
honeycomb networks, has extraordinary electrical,
thermal, and mechanical properties arising from its unique
structure [14–16]. Its unique properties make it suitable for
* Corresponding author. Department of BIN Fusion Technology,
Chonbuk National University, Duckjin-dong 1Ga 664-14, Jeonju, Jeonbuk
use as inorganic filler to improve the electrical, thermal and
561-756, Republic of Korea. Tel.: þ82 63 270 2342; fax: þ82 63 270 2341. mechanical properties of polymer composites [10,13–20].
E-mail address: jhl@chonbuk.ac.kr (J.H. Lee). However, pristine graphene is not compatible with organic

0142-9418/$ – see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.polymertesting.2011.09.007
32 T. Kuila et al. / Polymer Testing 31 (2012) 31–38

polymers and does not form homogeneous composites: it 2.3. Preparation of DA-G/LLDPE nanocomposites
requires surface modification prior to nanocomposite
preparation [13]. Graphene has been surface modified by For nanocomposite preparation, a calculated amount of
a range of techniques employing various organo modifying DA-G (0.075, 0.15, 0.45, 0.75, and 1.20 g, respectively) was
agents [10]. Among these reported techniques, the nucle- dispersed in 40 ml xylene by ultrasonication for 30 min. In
ophilic addition of organic molecules to the surface of a 250 ml Wit’s apparatus, 15 g LLDPE was dissolved in
graphene is an effective way to the bulk production of 100 ml xylene at 140  C to obtain a viscous solution. The
surface-modified graphene [13,17]. This work reports the dispersion of DA-G was then added to the polymer solution
use of dodecyl amine (DA) for the surface modification of and stirred for ca. 20 h at 140  C. Finally, the composite
graphite oxide to obtain-DA modified graphene (DA-G). solution was cast onto a glass Petri dish and dried at
Linear low density polyethylene (LLDPE) was selected as reduced pressure inside a vacuum oven at 75  C to remove
the polymer matrix due to its excellent flexibility, the solvent. The samples were then compression molded to
mechanical strength and durability. In addition, it is obtain sheets of DA-G/LLDPE nanocomposites. DA-G/LLDPE
chemically inert to acids and alkali and can easily elongate nanocomposites containing 0.5, 1.0, 3.0, 5.0, and 8.0 wt.% of
while applying a tensile stress. The preparation of DA-G/ DA-G have been designated as LDG-0.5, LDG-1, LDG-3, LDG-
LLDPE nanocomposites by solution mixing techniques (for 5, and LDG-8, respectively.
better dispersion of DA-G in LLDPE) is discussed herein and
the dynamic mechanical analysis, thermomechanical
analysis and gas barrier properties of the resulting nano- 2.4. Characterization
composites are evaluated.
X-ray diffraction (XRD) study of all the samples was
2. Experimental carried out at room temperature on a D/Max 2500V/PC
(Rigaku Corporation, Tokyo, Japan). The crystallite size of
2.1. Materials pure LLDPE and DA-G/LLDPE nanocomposites was carried
out by using the Scherrer formula [22]
Graphite flake was purchased from Sigma Aldrich,
Steinheim, Germany. Sulfuric acid, hydrochloric acid, ethanol Dp ¼ 0:89l=bcos q
and hydrogen peroxide were obtained from Samchun Pure
where b is the half height width (in radian) of the crystal-
Chemical Co. Ltd. (Pyeongtaek-si, Korea). Potassium
line peak, q is the Bragg’s angle and l is the wave length
permanganate (Junsei Chemical Co. Ltd., Tokyo, Japan) was
(0.154 nm for Cu). The degree of crystallinity of the samples
used as oxidizing agent. The organic modifier, dodecyl amine
was calculated from the ratios of integrated intensities of
(DA), and the reducing agent, hydrazine monohydrate, were
the crystalline and amorphous regions. The percentage of
procured from TCI, Tokyo, Japan and used without further
crystallinity, cc was obtained using the following relation-
purification. LLDPE (density, 944 kg/m3 at 23  C) was
ship [23]:
supplied by Hanwha Chemical, Korea. Xylene (Duksan
reagent and chemicals, Youngdeungpo-Gu, Korea) was used Ic
as the solvent to prepare the DA-G/LLDPEcomposite. cc ¼  100
Ia þ Ic
where, Ic and Ia are integrated intensity of the crystalline
2.2. Preparation of dodecyl amine modified graphene (DA-G) and amorphous region.
Fourier transform infrared spectroscopy (FTIR) was
Graphite oxide was prepared by a modified Hummers performed using a Nicolet 6700 spectrometer (Thermo-
method [21]. For surface modification, graphite oxide was scientific, USA) over the wave number range, 4000 -
dispersed in 200 ml of distilled water (2 mg/ml) by 400 cm1. The powdered sample of DA-G was mixed with
ultrasonication for 30 min using a ULH 700S bath-type KBr and pressed into a thin pellet for FTIR study. Pure LLDPE
sonicator. Then, the dispersion of graphite oxide was and its nanocomposites were compression molded to
centrifuged (Beckman Coulter, AllegraÔ X-22R) for prepare a thin film (w 40–50 mm).
15 min at 3000 rpm to remove the unexfoliated graphite Field emission scanning electron microscopy (FESEM)
oxide. The stable brown coloured dispersion was trans- was carried out using a JSM-6701F (JEOL, Japan). Pure
ferred to a 500 ml round bottom flask. 500 mg of dodecyl LLPDE and the nanocomposite samples were coated with
amine (DA) was dissolved in ethanol and the solution was a thin layer of platinum under argon (Ar) atmosphere at
added to the graphite oxide dispersion followed by stir- a vacuum pressure of w 102 Torr prior to being scanned in
ring at room temperature for 24 h. The nucleophilic a high resolution scanning electron microscope. The
substitution occurred through the amine functionality of thickness of the platinum coating was approximately
DA to the epoxy functionality of the graphite oxide. The 20 nm.
reduction of DA functionalized graphite oxide was carried Thermogravimetric analysis (TGA) of neat LLDPE and
out using hydrazine monohydrate. The final product was the nanocomposites (z5.5 mg) was carried out on a Q50
washed with water-ethanol mixture to remove the excess TGA (TA instruments, USA) with a heating rate of 5  C/min
DA adsorbed on the surface of the modified graphene from 60 to 600  C in air. Isothermal TGA was performed
(DA-G). The black powder was dried under vacuum at with the same instrument at constant temperature (300  C)
60  C for 72 h. in the presence of air.
T. Kuila et al. / Polymer Testing 31 (2012) 31–38 33

Differential scanning calorimetry (DSC) analysis of the Table 1


nanocomposites was performed with a TA instruments Q- Crystallite size and crystallinity of pure LLDPE and DA-G/LLDPE nano-
composites determined by XRD analysis.
20 from room temperature to 180  C. First, the sample was
heated from room temperature to þ200  C (5  C/min) and Sample Peak Peak Crystallite Crystallite % crystallinity
held for 10 min at this temperature to eliminate the angle angle size (nm) size (nm) (cc)
(2q1) (2q2) at 2q1 at 2q2
thermal history. Then, the samples were cooled to room
LLDPE 21.37 23.68 13.41 16.31 53.8
temperature at a rate of 5  C/min followed by reheating up
LDG-0.5 21.33 23.58 16.50 20.19 50.5
to þ200  C at the same rate. LDG-1 21.38 23.64 20.12 20.20 50.1
Dynamic mechanical analysis (DMA) of neat LLDPE and LDG-3 21.28 23.64 16.25 16.00 48.1
its nanocomposites were carried out with a Q 800 DMA (TA LDG-5 21.89 24.21 16.25 16.32 47.2
instruments, USA) in tension mode at a frequency and LDG-8 21.55 23.87 13.64 13.48 44.8

heating rate of 1 Hz and 2  C/min, respectively.


In order to measure the dimensional stability of neat
LLDPE and its composites, thermomechanical analysis pure LLDPE. The increase in apparent crystallite size is
(TMA) was carried out with a TMA Q 400 (TA instruments, considered to be due to both an increase in the true mean
USA) at a heating rate of 3  C/min. size and improvements in local lattice order [24]. On the
Oxygen and nitrogen barrier properties were measured contrary, the % crystallinity of the nanocomposites was
on a Toyoseiki (Japan) gas permeability machine at 1 atm gradually decreased with the addition of DA-G. This is
pressure. The sample was cut from the compression mol- probably due to the formation of a random interface. The
ded sheet and used directly for gas barrier property presence of homogeneously distributed DA-G layers in the
analysis. LLDPE matrix inhibited the ordered crystalline structure of
the polymer chains causing a significant decrease in %
3. Results and discussion crystallinity of the DA-G/LLDPE nanocomposites [25].
Fig. 2 shows the FTIR spectra of pure graphite oxide, DA-
Fig. 1 shows the XRD pattern of pure DA-G, pure LLDPE G, LLDPE and DA-G/LLDPE nanocomposite with 3 wt.% DA-
and their nanocomposites with 0.5, 1, 3, 5, and 8 wt.% of DA- G loading. A broad peak at 3437 cm1, assigned to O–H
G. The XRD pattern of pure GO is shown in the inset of Fig. 1. stretching, appeared in the spectra of pure graphite oxide
The 002 reflection peak of graphite oxide appeared at 2q w and DA-G, indicating the incomplete reduction of graphite
11.2 corresponding to the d spacing values of 0.79 nm. oxide to graphene. The peaks at 1721 cm1 and 1066 cm1
Upon surface modification and reduction of graphite oxide, were due to carboxyl and epoxy groups of graphite oxide
the 002 reflection peak fully disappeared due to the and were not observed in the spectrum of DA-G [26,27].
extensive intercalation and exfoliation of graphite oxide by The peak at 1628 cm1 in the spectrum of graphite oxide
dodecyl amine. The spectrum of pure LLDPE showed two indicates the presence of remaining graphitic sp2. The
amorphous peaks at 2q ¼ 21.42 and 23.73 . It was found emergence of peaks at 2950, 2928 and 2857 cm1, corre-
that the XRD pattern of DA-G/LLDPE consisted of the sponding to C–H stretching vibrations of CH3, CH2 and CH
characteristic peaks of pure LLDPE. Table 1 show the crys- groups in DA-G, indicates the successful modification of the
tallite size and % crystallinity of pure LLDPE and DA-G/ graphite oxide. The band at 1644 cm1 was assigned to
LLDPE nanocomposites. The crystallite size of the DA-G/ carbonyl stretching vibration, indicating the formation of
LLDPE nanocomposites was increased compared to the an amide-carbonyl (NHCO) bond between the graphene

Fig. 1. XRD pattern of pure DA-G, LLDPE and their nanocomposites with 0.5,
1, 3, 5, and 8 wt.% of DA-G. Fig. 2. FTIR spectra of (a) graphite oxide, (b) DA-G, (c) LLDPE and (d) LDG-3.
34 T. Kuila et al. / Polymer Testing 31 (2012) 31–38

sheets and DA molecules [28]. The peak at 1560 cm1 other [29]. Their lateral dimension varied between 1–5 mm.
further confirms the formation of amide linkage. LLDPE and The thickness is undeterminable from the FE-SEM images.
its nanocomposites showed bands at 2950, 2928 and In solid powder form, the DA-G exists as aggregates. Fig. 3
2857 cm1, attributable to the C–H stretching vibrations of (c,d) show the surface morphology of the nanocomposites
CH3, CH2 and CH groups. The spectrum of the DA-G/LLDPE with a DA-G content of 3 wt.% (LDG-3). This reveals that the
composite had the peaks at 1470 and 1375 cm1 (defor- graphene sheets were dispersed in the LLDPE matrix. The
mation of CH2 and CH3 groups) shift slightly to 1466 and dispersion of graphene layers in the LLDPE matrix indicates
1370 cm1, possibly due to interactions between the the formation of DA-G/LLDPE nanocomposites. The fracture
modified graphene and LLDPE polymer. surface morphology of neat LLDPE and LDG-3 are shown in
The FE-SEM images of DA-G and DA-G/LLDPE nano- Fig. 3 (e,f). The fracture surface of neat LLDPE was smooth
composites show a flake-like morphology of DA-G (Fig. 3). while the nanocomposites had a comparatively rough
The morphology of DA-G consisted of randomly aggre- surface. The appearance of rougher surface morphology
gated, thin, crumpled sheets closely associated with each and the presence of interfacial interaction between filler

Fig. 3. FE-SEM photomicrographs of (a,b) DA-G, (c,d) surface morphology of LDG-3, (e,f) fracture surface morphology of neat LLDPE LDG-3.
T. Kuila et al. / Polymer Testing 31 (2012) 31–38 35

and polymer is an indication of stronger mechanical Table 2


properties. Therefore, the SEM images of the fracture DMA values of pure LLDPE and DA-G/LLDPE nanocomposites.

surfaces are in good agreement with the observed Sample E0 (MPa) E0 (MPa) E0 (MPa) E0 (MPa) tan da-max
mechanical properties of the nanocomposites. at 75  C at Tg ¼ 28  1  C at 25  C at 50  C at 80  C
The storage moduli (E0 ) of DA-G/LLDPE nanocomposites LLDPE 1821 1046 289 134 0.2389
with different amounts of DA-G were greater than that of LDG-0.5 2167 1310 425 214 0.2271
LDG-1 2272 1387 446 195 0.2356
neat LLDPE (Fig. 4(a)). At 50  C, the modulus of the nano-
LDG-3 2199 1326 447 207 0.2256
composites with 8 wt.% DA-G (LDG-8) was 293 MPa, which LDG-5 2272 1436 496 244 0.2189
is 118% higher than that of neat LLDPE. This is because LDG-8 2528 1582 569 293 0.2134
graphene acts as a reinforcing filler, significantly decreasing
the chain mobility of the LLDPE polymer. The effect of DA-G
content on the tan d of the nanocomposites is shown in
the nanocomposites with DA-G. Similar results were also
Fig. 3(b), and the related values are listed in Table 2. The
observed from plotting loss modulus (E00 ) against temper-
neat LLDPE and the nanocomposites exhibited two relax-
ature. This may be because of the homogeneous distribu-
ations, one at 29  C (b-relaxation) and another in the
tion of graphene in the nanocomposites. Fig. 4(b) also
range of 79–90  C (a-relaxation) [30–32]. The loss factor,
shows that the height of the tan dmax (a-relaxation) peak of
tan d, is the ratio of the loss modulus to the storage
the nanocomposites at 80  C decreases with increasing DA-
modulus and is very sensitive to structural transformations.
G content. This is indicative of the reinforcing effect of DA-G
The tan d peak (b-relaxation) can also be used to identify
in the LLDPE matrix and the decrease in damping behavior
the glass transition temperature (Tg). However, it can be
with increasing DA-G content.
difficult to determine Tg from the tan d peak, as evident
As evident from the DMA results, no significant changes
from Fig. 4(b). It shows a marginal change in both
in Tg was observable with the addition of DA-G to the LLDPE
b-relaxation (Tg) and a-relaxation for both neat LLDPE and matrix. Hence, DSC measurements were carried out to
quantify any changes of melting endotherm and crystalli-
zation temperature in the nanocomposites. The melting
endotherms of the nanocomposites were slightly higher
(w1  C) than that of neat LLDPE (Fig. 5), falling within the
error limit of the instrument. The crystallization tempera-
tures (Tc) of the nanocomposites were comparatively
higher than that of neat LLDPE, ascribable to the nucleation
(seeding) effect of DA-G on the LLDPE matrix [31,33]. Tc
increased with increasing DA-G content, except at 8 wt.%
loading, possibly because the number of nucleation sites
increased with increasing DA-G content. However, the
value of Tc for LDG-8 remained higher than neat LLDPE.
Fig. 6 shows TGA thermograms of neat LLDPE and its
nanocomposites with varying DA-G contents. The onset
degradation temperature of the nanocomposites was
significantly faster than that of neat LLDPE. This is

Fig. 4. Variation of storage modulus (a) and tan d (b) value of the DA-G/ Fig. 5. Cooling and second heating DSC cycle of neat LLDPE and DA-G/LLDPE
LLDPE nanocomposites at different amount of DA-G content. nanocomposites with 0.5, 1, 3, 5 and 8 wt.% of DA-G contents.
36 T. Kuila et al. / Polymer Testing 31 (2012) 31–38

Fig. 6. TGA curves of neat LLDPE and the nanocomposites of DA-G/LLDPE Fig. 7. Isothermal TGA curves of neat LLDPE and the nanocomposites 0.5, 1,
with various amounts of DA-G contents. 3, 5 and 8 wt% DA-G contents.

attributable to the early degradation of dodecyl amine of


functionalized graphene. The final degradation tempera-
ture of the nanocomposites (LDG-3, LDG-5, and LDG-8) was
much higher than that of neat LLDPE. The 50% weight loss
temperatures of LLDPE, LDG-1, LDG-3, LDG-5 and LDG-8
were 407, 408, 413, 442 and 454  C, respectively. Table 3
also lists the degradation temperature for 80% weight loss
as being much higher for the nanocomposites than that for
neat LLDPE. The enhanced thermal stability of the nano-
composites was attributed to the very high aspect ratio of
graphene which prevented the emission of small gaseous
molecules during thermal degradation. The homoge-
neously dispersed graphene also disrupted the oxygen
supply by forming charred layers on the surfaces of the
nanocomposites [6]. This indicates that graphene can act as
a good barrier to prevent the thermal degradation of LLDPE.
This was corroborated by studying the isothermal TGA of
neat LLDPE and the nanocomposites (Fig. 7). The rate
constants for neat LLDPE and the nanocomposites were
calculated considering first order kinetics (W ¼ W0ekt,
where W0 is the initial weight, W is the remaining weight
at time t and K is the first order rate constant). The rate
constant for neat LLDPE was much higher than those of the
nanocomposites (Table 3). The rate of thermal degradation
of the nanocomposites was slower than that of neat LLDPE,
with increasing DA-G content further slowing the degra-
dation (Fig. 7). This can be ascribed to increased char
formation with increased DA-G content.

Table 3
TGA and isothermal TGA data for pure LLDPE and DA-G/LLDPE
nanocomposites.

Sample Onset degradation Temperature Temperature Rate constant


temperature ( C) ( C) at 50% ( C) at 80% (s1)
weight loss weight loss
LLDPE 317 407 443 9.57  104
LDG-1 299 408 443 1.95  104
LDG-3 300 413 448 1.65  104 Fig. 8. (a) Variation of dimensional stability of pure LLDPE and its
LDG-5 300 442 462 6.02  105 composites with the change in temperature and DA-G content, (b) The
LDG-8 302 454 464 4.22  105 coefficient of thermal expansion (CTE) of DA-G/LLDPE nanocomposites in
the range of 40–70  C and 80–110  C.
T. Kuila et al. / Polymer Testing 31 (2012) 31–38 37

The variation of dimensional stability of pure LLDPE and solution and melt mixing [36]. Therefore, the CTE of the
its nanocomposites with the change in DA-G content and nanocomposites decreased regularly with increasing DA-G
temperature are shown in Fig. 8a. The change in dimen- content, with no sharp change observed between 0.5 wt.%
sional stability of pure LLDPE is almost linear at higher and 8 wt.% DA-G loading. It is also interesting to note that
temperatures (70–110  C). On the contrary, the slopes were CTE did not change greatly between the two temperature
much lower for the nanocomposites as compared to that of ranges (40–70  C and 80–110  C).
the pure LLDPE. This is attributed to the increased hardness Oxygen and nitrogen permeability of the LLDPE nano-
proportional to the content of DA-G in the nanocomposites composites are shown in Fig. 9a. The oxygen permeability
[34]. The TMA results showed that the dimensional change was decreased from 36.8 to 19.5 fm/Pa.S with the addition of
of the nanocomposites was significantly decreased by the only 1 wt.% DA-G. At the same amount of DA-G loading,
addition of the DA-G. nitrogen permeability was decreased to 5.7 from 11.9 for pure
Coefficients of thermal expansion (CTE) were calculated LLDPE. This is attributed to the very high aspect ratio of
over two different temperature ranges (40–70  C and 80– graphene which prevents the permeation of gaseous mole-
110  C) from the TMA curves and are shown in Fig. 8b. It was cules through the nanocomposites by creating multiple
determined individually from the linear portions of each layers which forces gaseous molecules to flow through the
thermogram below 70  C and above 80  C. The CTE of neat nanocomposites materials in a “tortuous path”, forming
LLDPE was 575.63  106/ C below 70  C and 977.25  106/ complex barriers to gases, as shown in Fig. 9b. The gas

C above 80  C. A significant change of CTE was observed permeation properties are also largely dependent on the
upon addition of only 0.5 wt.% DA-G. This was attributed to thermal stability of the nanocomposites and the interfacial
the very large surface area and low CTE of graphene [35]. A interaction between the filler and the matrix polymer [37,38].
similar observation was found by Kim et al. for exfoliated The improved gas barrier properties were in good agreement
graphite nanoplatelets/LLDPE nanocomposites prepared by with the morphological characterization and thermal prop-
erties of the nanocomposites. It is interesting to note that no
significant change in permeability (both for O2 and N2) was
observed as DA-G content increased from 1 to 5 wt.%.

4. Conclusions

High performance novel polymer nanocomposites were


prepared by solution mixing of LLDPE with DA-G. FTIR
showed that DA-G was doped onto the LLDPE polymer. FE-
SEM showed the homogeneous dispersion of graphene in
the LLDPE matrix. XRD analysis showed the formation of
DA-G/LLDPE nanocomposites. It was also noted that the
percent crystallinity was decreased in the nanocomposites
due to the formation of random interfaces. The storage
moduli of the nanocomposites were much higher than that
of neat LLDPE. The high surface area and aspect ratio of DA-
G reduced the chain mobility of the nanocomposites. The
nanocomposites were more thermally stable in air than
neat LLDPE. The gas barrier properties of the nano-
composites in terms of oxygen and nitrogen permeability
were also improved compared with the neat LLDPE.
Isothermal TG results of the nanocomposites were in good
agreement with their observed barrier properties. DSC
analysis showed that the crystallization temperature of the
nanocomposites was affected by the presence of DA-G due
to the nucleation effects of DA-G. TMA showed that the
addition of DA-G in the nanocomposites increased its
dimensional stability due to the increased material hard-
ness. The CTE values of the nanocomposites was decreased
significantly with the addition of only 0.5 wt.% of DA-G.

Acknowledgement

This study was supported by the Converging Research


Center Program (2011K000776), the Human Resource
Training Project for Regional Innovation, and the World
Fig. 9. (a) Oxygen and nitrogen permeability of DA-G/LLDPE nano-
Class University (WCU) program (R31-20029) funded by
composites at room temperature and 1 atm pressure, (b) “Tortuous path” the Ministry of Education, Science and Technology (MEST)
model for DA-G/LLDPE nanocomposites. and National Research Foundation (NRF) of Korea.
38 T. Kuila et al. / Polymer Testing 31 (2012) 31–38

References [19] H. Kim, C.W. Macosko, Processing-property relationships of poly-


carbonate/graphene nanocomposites, Polymer 50 (15) (2009)
[1] S.S. Ray, M. Okamoto, Polymer/layered silicate nanocomposites: 3797–3809.
a review from preparation to processing, Prog. Polym. Sci. 28 (11) [20] S. Bose, T. Kuila, M.E. Uddin, N.H. Kim, A.K.T. Lau, J.H. Lee, In-situ
(2003) 1539–1641. synthesis and characterization of electrically conductive polypyrrole/
graphene nanocomposites, Polymer 51 (25) (2010) 5921–5928.
[2] M. Alexandre, Ph. Dubois, Polymer-layered silicate nanocomposites:
preparation, properties and uses of a new class of materials, Mater. [21] W.S. Hummers, R.E. Offeman, Preparation of graphitic oxide, J. Am.
Sci. Eng: R 28 (1–2) (2000) 1–63. Chem. Soc. 80 (1958) 1339.
[3] A. Okada, M. Kawasumi, A. Usuki, Y. Kojima, T. Kurauchi, O. [22] P.P. Kundu, J. Biswas, H. Kim, S. Choe, Influence of film preparation
Kamigaito, Synthesis and properties of nylon-6/clay hybrids. in: procedures on the crystallinity, morphology and mechanical prop-
erties of LLDPE films, Eur. Polym. J. 39 (8) (2003) 1585–1593.
D.W. Schaefer, J.E. Mark (Eds.), Polymer Based Molecular
Composites, MRS Symposium Proceedings, Pittsburgh, vol. 171, [23] S. Bose, N. Pramanik, C.K. Das, A.K. Saxena, Synthesis and effect of
1990, pp. 45–50. polyphosphazenes on the thermal, mechanical and morphological
[4] D.Y. Godovsky, Device applications of polymer-nanocomposites, properties of poly(etherimide)/thermotropic liquid crystalline
Adv. Polym. Sci. 153 (2000) 163–205. polymer blend, Mater. Design 31 (3) (2010) 1148–1155.
[5] E.P. Giannelis, Polymer layered silicate nanocomposites, Adv. Mater. [24] A.M. Hindeleh, D.J. Johnson, Crystallinity and crystallite size
8 (1) (1996) 29–35. measurement in polyamide and polyester fibres, Polymer 19 (1)
(1978) 27–32.
[6] W. Zheng, S.C. Wong, Electrical conductivity and dielectric proper-
ties of PMMA/expanded graphite composites, Compos. Sci. Technol. [25] S. Bose, C.K. Das, A.K. Saxena, A. Ranjan, Compatibilizing effect of
63 (2) (2003) 225–235. functionalized polyphosphazene on the properties of poly(-
[7] K. Kalaitzidou, H. Fukushima, L.T. Drzal, A new compounding phenylene oxide)/vectra a blend system, J. Appl. Polym. Sci. 119 (4)
method for exfoliated graphite-polypropylene nanocomposites with (2011) 1914–1922.
[26] R. Bissessur, P.K.Y. Liu, S.F. Scully, Intercalation of polypyrrole into
enhanced flexural properties and lower percolation threshold,
Compos. Sci. Technol. 67 (10) (2007) 2045–2051. graphite oxide, Synth. Metal 156 (16–17) (2006) 1023–1027.
[8] R.D. Goodridge, M.L. Shofner, R.J.M. Hague, M. McClelland, M.R. [27] H. Wang, Q. Hao, X. Yang, L. Lu, X. Wang, Effect of graphene oxide on
Schlea, R.B. Johnson, C.J. Tuck, Processing of a Polyamide-12/carbon the properties of its composite with polyaniline, ACS Appl. Mater.
nanofibre composite by laser sintering, Polym. Test. 30 (1) (2011) Interfaces 2 (3) (2010) 821–828.
[28] G. Wang, X. Shen, B. Wang, J. Yao, J. Park, Synthesis and character-
94–100.
[9] H. Chen, O. Jacobs, W. Wu, G. Rüdiger, B. Schädel, Effect of dispersion ization of hydrophilic and organophilic graphene nanosheets,
Carbon 47 (5) (2009) 1359–1364.
method on tribological properties of carbon nanotube reinforced
epoxy resin composites, Polym. Test. 26 (3) (2007) 351–360. [29] J. Shen, Y. Hu, C. Li, C. Qin, M. Ye, Synthesis of amphiphilic graphene
[10] T. Kuilla, S. Bhadra, D. Yao, N.H. Kim, S. Bose, J.H. Lee, Recent nanoplatelets, Small 5 (1) (2009) 82–85.
advances in graphene based polymer composites, Prog. Polym. Sci. [30] M. NouriRazavi, J.N. Hay, Time-temperature superposition and
35 (11) (2010) 1350–1375. dynamic mechanical properties of metallocene polyethylene, Iran
Polym. J. 13 (5) (2004) 363–370.
[11] O.K. Park, T. Jeevananda, N.H. Kim, S. Kim, J.H. Lee, Effects of surface
modification on the dispersion and electrical conductivity of carbon [31] R. Pucciariello, V. Villani, G. Giammarino, Thermal behavior of
nanotube/polyaniline composites, Scripta Mater. 60 (7) (2009) 551– nanocomposites based on linear low density poly(ethylene) and
554. carbon nanotubes prepared by high energy ball milling, J. Poym.
[12] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, Y. Zhang, S.V. Res. 18 (5) (2011) 949–956.
[32] E. Kontou, M. Niaounakis, Thermo-mechanical properties of LLDPE/
Dubonos, I.V. Grigorieva, A.A. Firsov, Electric field effect in atomi-
cally thin carbon films, Science 306 (5696) (2004) 666–669. SiO2 nanoocmposites, Polymer 47 (4) (2006) 1267–1280.
[33] C.Y. Li, Polymer single crystal meets nanoparticles, J. Polym. Sci.
[13] S. Stankovich, D.A. Dikin, G.H.B. Dommett, K.M. Kohlhaas, E.J. Zim-
Polym. Phys. B. 47 (24) (2009) 2436–2440.
ney, E.A. Stach, R.D. Piner, S.T. Nguyen, R.S. Ruoff, Graphene - based
[34] K. Agarwal, M. Prasad, R.B. Sharma, D.K. Setua, Studies on micro-
composite materials, Nature 442 (2006) 282–286.
structural and thermophysical properties of polymer nano-
[14] R.D. Dreyer, S. Park, C.W. Bielawski, R.S. Ruoff, The chemistry of
composite based on polyphenylene oxide and ferrimagnetic iron
graphene oxide, Chem. Soc. Rev. 39 (1) (2010) 228–240.
oxide. 30(1) (2011) 155–160.
[15] X. Li, X. Wang, L. Zhang, S. Lee, H. Dai, Chemically derived, ultra-
[35] J.W. Jiang, J.S. Wang, B. Li, Thermal expansion in single-walled
smooth graphene nanoribon semiconductor, Science 319 (5867)
carbon nanotubes and graphene: Nonequilibrium green’s function
(2008) 1229–1231.
approach, Phys. Rev. B. 80 (2009) 205429.
[16] M.J. Allen, V.C. Tung, R.B. Kaner, Honeycomb carbon: a review of
[36] S. Kim, I. Do, L.T. Drazal, Thermal stability and dynamic mechanical
graphene, Chem. Rev. 110 (1) (2010) 132–145.
behavior of exfoliated graphite nanoplatelets-LLDPE nano-
[17] A.B. Bourlinos, D. Gournis, D. Petridis, T. Szabo, A. Szeri, I. Dékány,
composites, Polym. Compos. 31 (5) (2010) 755–761.
Graphite oxide: chemical reduction to graphite and surface modi-
[37] V. Mittal, Gas permeation and mechanical properties of poly-
fication with primary aliphatic amines and amino acids, Langmuir
propylene nanocomposites with thermally-stable imidazolium
19 (15) (2003) 6050–6055.
modified clay, Eur. Polym. J. 43 (9) (2007) 3727–3736.
[18] H. Hu, X. Wang, J. Wanga, L. Wana, F. Liu, H. Zheng, R. Chen, C. Xu,
[38] M.A. Osman, V. Mittal, M. Morbidelli, U.W. Suter, Polyurethane
Preparation and properties of graphene nanosheets-polystyrene
adhesive nanocomposites as gas permeation barrier, Macromole-
nanocomposites via in situ emulsion polymerization, Chem. Phys.
cules 36 (2003) 9851–9858.
Lett. 484 (4–6) (2010) 247–253.

You might also like