Вы находитесь на странице: 1из 8

Composites Science and Technology 77 (2013) 8794

Contents lists available at SciVerse ScienceDirect

Composites Science and Technology


journal homepage: www.elsevier.com/locate/compscitech

Chemical functionalization of graphene oxide toward the tailoring


of the interface in polymer composites
Bin Shen, Wentao Zhai , Mimi Tao, Dingding Lu, Wenge Zheng
Ningbo Key Lab of Polymer Materials, Ningbo Institute of Material Technology and Engineering, Chinese Academy of Sciences, Ningbo, Zhejiang Province 315201, China

a r t i c l e

i n f o

Article history:
Received 28 November 2012
Received in revised form 7 January 2013
Accepted 16 January 2013
Available online 26 January 2013
Keywords:
A. Particle-reinforced composites
B. Interface
B. Mechanical properties

a b s t r a c t
In this work, we demonstrated that the composites with strong interfacial interactions between graphenematrix could achieve excellent mechanical properties even the dispersion of graphene is poor. In
terms of the above reason, an epoxy resin was coupled onto graphene oxide (GO) sheets via the grafting
to method. Since each epoxy chain bears two terminated epoxide groups, it is inevitable that one epoxy
chain connects two GO sheets together, causing the crosslinking of GO layers via the epoxy chain. When
blending these resultant GO (GOepoxy) with polycarbonate (PC), the dispersion was less-than-ideal due
to these crosslinking. However, the residue active sites in the grafted epoxy chains, such as the unreacted
epoxide groups as well as hydroxyl groups, could further react with PC carbonate to form chemical bonds,
leading to strong interfacial interactions between the matrix and GO sheets. Owing to these strong interfacial interactions, the enhancement of the mechanical properties of PC/GOepoxy composites was signicantly higher than that of PC/(GO/epoxy) samples, as well as those shown in other similar works
on thermally reduced graphene oxide (TRG)/PC composites with better dispersion.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
Graphene, a single-atom-thick sheet of hexagonally arrayed
sp2-bonded carbon atoms, has been under the spotlight owning
to its intriguing and unparalleled physical properties [1]. Because
of its novel properties, such as exceptional thermal conductivity
[2]. High Youngs modulus [3], and high electrical conductivity
[4], integration of graphene and its derivations into polymer has
been highlighted [5]. The performance of graphene/polymer nanocomposite depends on the dispersion of graphene in the matrix
and interfacial interactions between the graphene and the polymer
matrix. However, the large surface area of graphene and strong van
der Waals force among them result in severe aggregation in the
composites matrix. To obtain satised performance of the nal
graphene/polymer composites, the issues of the strong interfacial
adhesion between graphenematrix and well dispersion of graphene should be addressed. Thereby, covalent functionalization of
graphene is widely adopted to improve its surface properties.
Currently, covalent functionalization of graphene is always
based on the graphene from previously prepared graphene oxide
(GO) or reduced GO (r-GO), which has multiple oxygenated groups,
such as epoxides and hydroxyls on their basal planes and carboxyls
on the edges [6]. It includes grafting from or grafting on strat Corresponding authors. Tel.: +86 0574 8668 5256; fax: +86 0574 8668 5186.
E-mail addresses: wtzhai@nimte.ac.cn (W. T. Zhai), wgzheng@nimte.ac.cn
(W. G. Zheng).
0266-3538/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.compscitech.2013.01.014

egies. Through the grafting-from method, vinyl polymer molecules such as polystyrene (PS) and poly(methyl methacrylate)
(PMMA) were successfully grafted onto the surface of GO, resulting
in the excellent compatibility and enhanced load transfer between
GO sheets and polymer hosts [79]. On the grafting on approach,
polymer chains can be covalently grafted onto GO sheets by esterication or amidation [1012]; these reactions are conducted between active sites of preformed functionalized polymer chains
and functional groups on GO surface. To the best of our knowledge,
due to the difculty in matching the unique surface chemistry of
GO with the specic functionality on polymer chains, the studies
concerning the grafting-to method are relatively few. More
importantly, if there are many active sites on preformed polymer
and GO sheet, the selectivity of chemical reaction between these
active sites is hardly controlled. For example, the multifunctionalized groups of the preformed polymer chains inevitably act as
covalent cross-linker between graphene sheets to form the crosslinking of graphene; thus, the nal dispersion of these sheets in
the host matrix would be less-than-ideal. Although the concept
that the perfect dispersion of graphene in host matrix is essential
for the good mechanical performances of the matrix is prevalent,
we propose that the composites with strong interfacial interactions
between graphenematrix could also achieve excellent mechanical
properties even the dispersion is poor.
In terms of the above reason, an epoxy resin based on di-glycidyl ether of bisphenol A (DGEBA) was coupled onto GO sheets
via the grafting to method. As shown in Fig. 1, the reaction that

88

B. Shen et al. / Composites Science and Technology 77 (2013) 8794

Fig. 1. Schematic illustration of synthesis of GOepoxy.

takes place is between the carboxylic acid attached to GO surface


and the epoxide ring of epoxy chain. Since each epoxy chain bears
two terminated epoxide groups, it is inevitable that one epoxy
chain connects two GO sheets together, causing the crosslinking
of GO layers via the epoxy chain. When blending these resultant
GO (GOepoxy) with polycarbonate (PC), the dispersion was poor
due to the crosslinking. However, there still have active sites in
the grafted epoxy chains, such as the unreacted epoxide groups
as well as hydroxyl groups. These active sites could further react
with PC carbonate through transesterication and alcoholysis reactions during hot-compression of their composites [1315], leading
to covalent bonds and strong interfacial interactions. Despite the
poor dispersion of GO sheets, the nal PC/GOepoxy composites
still exhibited high mechanical and thermal properties. The strong
interfacial covalent bonding between GO and PC should be responsible for the concurrently improved mechanical and thermal
properties.

2. Experimental
2.1. Materials
The bisphenol-A PC was purchased from GE Plastics Inc. in the
form of pellets and dried at 120 C under vacuum before use. The
epoxy resin (DER 331) was obtained from Dow Chemicals.
Triphenyl phosphine (TPP) was purchased from Aladdin Chemical
Reagents (China). Dichloromethane (CH2Cl2), acetone, dimethyformamide (DMF), tetrahydrofuran (THF) and ethanol were obtained
from Sinoparm Chemical Reagent (China).

2.2. Exfoliation of GO in DMF


Graphite oxide prepared by oxidizing natural graphite based on
a modied Hummers method [16]. Since graphite oxide is hardly
exfoliated in DMF, GO/DMF solution was prepared by a novel
freeze-drying method. Typically, 100 mg GO powder rstly exfoliated in water under sonication. Then, the GO aqueous dispersion
was frozen and placed in a freeze-drying apparatus where the frozen water was removed by sublimation under vacuum, leaving a
uffy GO powder. After that, the uffy GO were sonicated in
DMF for 5 min to obtain GO/DMF dispersion, in which the GO were
well exfoliated and existed as single layer (the details were shown
in the Supporting Information).

2.3. Functionalization of GO with epoxy


Firstly, 10 g DER 331 was dissolved in DMF by vigorous stirring.
Then, the obtained DER 331 solution was added into as-prepared
GO/DMF dispersion and stirred at room temperature for 2 h. Finally, TPP (0.250.5 wt.% to epoxy) was subsequently added to catalyze the reaction [17,18]. The reaction was allowed to proceed
for 24 h at 100 C. After that, the reacted GO was washed with
acetone through a dispersion-ltration-washing method as mentioned in our previous study. During each cycle of vacuumltration, several drops of ltrate were added into water from time
to time, and the absence of cloudiness can be used to indicate the
complete removal of epoxy. The functionalized GOepoxy was
dried in a vacuum oven at 50 C overnight.
2.4. Composites preparation
All PC/GOepoxy or PC/(GO/epoxy) (freeze-dried GO) composites were prepared by solution blending. Typically, GOepoxy or
GO/epoxy was rstly dispersed in THF by ultrasonication (500 W)
for 10 min, and then mixed with PC/THF solution under vigorous
stirring for 6 h. After that, the mixture was precipitated with methanol and dried successively in an air-circulating oven at 80 C for
12 h and a vacuum oven at 100 C for 24 h. Finally, the obtained
composites were compression-molded to 1 mm thick plates using
a hot-press at 250 C under a pressure of 15 MPa for 30 min. For
tensile tests, the obtained plates were compression-molded to
100 lm thick lms.
2.5. Characterizations
Scanning electron microscopy (SEM) observation was performed with a Hitachi S-4800 eld emission SEM at an accelerating
voltage of 4 kV. Fourier transform infrared (FT-IR) measurements
were performed on a Thermo Nicolet 6700 spectrometer. Raman
spectra were excited with a laser of 633 nm and record with Labram spectrometer (Super LabRam II system). Transmission electron
microscopy (TEM) was conducted on a Tecnai G2 F20 transmission
electron microscope with an accelerating voltage of 100 kV. Typical
tapping-mode atomic-force microscopy (AFM) measurements
were performed using Multimode SPM from Digital Instruments
with a Nanoscope V Controller from Veeco. Thermal gravimetric
analyses (TGA) were carried out on Mettler-Toledo TG/DSC 1 thermogravimetric analyzer. The diffraction behavior of the samples
was studied using a Bruker AXS X-ray diffractometer with CuKa

B. Shen et al. / Composites Science and Technology 77 (2013) 8794

89

Fig. 2. (a) Solubility of GO and GOepoxy in CH2Cl2 (down)/H2O(up) mixture (0.5 mg/ml, 2 months standing). (b) FT-IR spectra of GO, GOepoxy, and epoxy. (c) Raman
spectra of nature graphite, GO and GOepoxy with a laser of 633 nm. (d) TGA curves of GO, GOepoxy and pure epoxy with a heating rate of 20 C/min in a nitrogen
atmosphere.

radiation at a generator voltage of 40 kV and a generator current of


40 mA. The linear viscoelastic response of the composites was
measured using dynamic mechanical analysis (DMA SDTA861e,
Mettler-Toledo Instruments). The tensile properties of the lms
were measured on Instron 5567 materials testing system. The
extension rate was 5 mm/min and the gauge length was
50 mm. All samples were cut into strips of 100 mm  10 mm
using a razor blade. In cases, more than ve samples were tested
from which the mean and standard deviation were calculated.
3. Results and discussion
3.1. Chemical functionalization of GO with epoxy
The successful functionalization of GO sheets with epoxy was
rst conrmed by the change of the solubility of GO in organic solvents. Fig. 2a shows the distinct solubility of GO and GOepoxy in
the mixture of CH2Cl2/H2O (1/1, v/v) solvents. As CH2Cl2 and H2O
are immiscible, they are phase-separated with the CH2Cl2 phase
at the bottom of the glass vial. Due to its polar feature, GO is completely dispersed in H2O (the brown part of the left vial). On the
contrary, GOepoxy is completely located and dispersed in CH2Cl2,
and do not undergo sedimentation for 2 months (the black part of
the right vial). The conversion of the hydrophilic GO to the hydrophobic GOepoxy conrms the effectiveness of the functionalization of GO with epoxy. In addition, to exclude the possible
contribution of the physically adsorbed epoxy chains to the dispersibility of GO, the mixture of GO and epoxy resin was prepared and
then dispersed into CH2Cl2. As predicted, GO sheets are completely

precipitated in CH2Cl2, which clarify that the dispersibility is as a


result of the chemical grafting epoxy instead of the physically adsorbed epoxy. Interestingly, there is also a color change from
brown1 GO in water to black GOepoxy in CH2Cl2 (Fig. 2a), which
should be evidence of the solvothermal reduction that occurred during the reuxing of GO in DMF (the details were discussed in the
Supporting Information) [1922].
The evidence for the covalent bonding between GO sheets and
epoxy was further provided by FT-IR spectra. As shown in
Fig. 2b, the absorption peaks of GOepoxy at 1620, 1507, 1232,
1034 and 830 cm1, which are corresponded to absorption of the
benzene ring and ether groups of epoxy, indicates the presence
of the grafted epoxy chains on GO (note that this sample was
exhaustively washed to remove the free epoxy chains adsorbed
on GO sheets prior to the FT-IR tests). Especially, the existence of
the characteristic peak at 1100 cm1, which is related to CAO bond
formed in the reaction between the carboxylic acid and the epoxide ring, provides direct evidence for covalent bonding between GO
sheets and epoxy chains. Moreover, the peak at 1729 cm1 corresponding to m (C@O) of ACOOH on GO shifts to 1720 cm1 due to
the formation of CAO bond. In addition, the peak at 915 cm1,
which corresponds to the epoxide ring in epoxy [23], can still be
seen in the GOepoxy spectrum, indicating the residues of the
epoxide ring after the reaction.
Raman spectroscopy is a convenient and powerful tool to probe
the structure of graphene-based materials. Fig. 2c shows the
1
For interpretation of color in Figs. 2, 4 and 5, the reader is referred to the web
version of this article.

90

B. Shen et al. / Composites Science and Technology 77 (2013) 8794

Raman spectra of natural graphite, GO and GOepoxy. After the


reuxing of GO with epoxy in DMF, the G peak of GOepoxy redshifts from 1588 cm1 to 1582 cm1, getting close to that of pristine graphite (1579 cm1), thus indicating the restoration of the
graphitic sp2 network [24]. The intensity ratio of D and G bands
(I(D/G)), correlating to the disordered and ordered crystal structures of carbon, is inverse to the average size of sp2 domains
[24]. In our case, the grafting reaction is taken place by using carboxyl groups on the surface of GO as the active sites, and thus the
skeletal structure should be greatly preserved after the grafting
[25,26]. However, the I(D/G) of GOepoxy is increased compared
with that of GO. This change is due to the decrease in the average
size of the in-plane sp2 domain upon reduction of GO, and can be
explained if new graphitic domains were created that are smaller
in size to the ones present in GO before reduction, but more
numerous in number [27].
TGA measurement was performed to study the amount of
epoxy chains grafting onto GO sheets, as shown in Fig. 2d. In order
to eliminate the inuence of reduction during the reaction, GO
were reuxed in DMF at the same condition with the functionalization reaction before characterization. In the TGA curve of GO,
there is about 20% weight loss at 600 C, which is ascribed to pyrolysis of the residual oxygen-containing functional groups. The TGA
curve for GOepoxy sample exhibits two major weight loss stages
at 200300 and 300470 C. The former stage is due to the decomposition of the unstable oxygenic groups of the GO component, and
the latter stage above 300 C is mainly attributed to the degradation of the grafted epoxy. Taking into account of the residue at

600 C, the quantity of grafted epoxy in GOepoxy can be calculated to be 15 wt.%.
The microstructure of GOepoxy was investigated by TEM
observation. For the measurements, a droplet of diluted CH2Cl2/
GOepoxy dispersion was deposited onto a copper grid. It reveals
a transparent clean surface with a few thin ripples for the individually exfoliated GO sheet from Fig. 3a. As reported previously, the
ripples are intrinsic properties of thin GO sheets, which is due to
the extra thermodynamic stability of the 2D membranes arising
from microscopic crumpling [28]. In contrast, the TEM image of
GOepoxy is entirely different, which is shown in Fig. 3b. We can
clearly observe that the surface of GO sheets is covered by a thin
coating, which are likely the grafted epoxy. This morphology is
similar with the case of polymer-functionalized graphene or
carbon nanotubes, where a polymer interface layer is created
[2931]. Moreover, since each epoxy chain bears two terminated
epoxide groups, it is inevitable that one epoxy chain connects
two GO sheets together to form the crosslinking of GO layers
(the aggregated structures). To probe the aggregate structure, we
conducted AFM observation on the GOepoxy sample on a freshly
cleaved mica surface. Fig. 3c presents a typical AFM image of a
GOepoxy aggregate, where a lot of GO sheets were observed to
be roughly connected with each other. The average thickness of a
GOepoxy sheet is 2.12 nm according to cross-sectional analysis
(Fig. 3c), which is much higher than that of the individually
exfoliated GO sheets [1,32]. Similarly, the stacking structure of
GOepoxy could be also observed from SEM image (Fig. 3d). Since
GOepoxy solution was in very dilute concentration (the solution

Fig. 3. (a) TEM image of GO sheet. (b) TEM image of GOepoxy. (c) AFM image of GOepoxy sheets with the aggregated structures. (d) SEM image of GOepoxy with the
aggregated structures.

B. Shen et al. / Composites Science and Technology 77 (2013) 8794

was almost colorless and transparent) during the preparation of


AFM and SEM samples, we think that the stacking of different layers of GO during the preparation of TEM and AFM samples was
hard to happen. Actually, GO aggregate almost could not be detected during AFM observation of well-exfoliated GO (freeze-dried
GO) in very dilute concentration. Therefore, such GO aggregate
should be caused by the crosslinking of GO layers.
3.2. Dispersion of GOepoxy into the polymer matrix
For the study of polymer composite properties, GOepoxy were
incorporated into PC using solution-based processing method and
thin-lm samples (100 lm) were prepared using compression
molding. For comparison, GO/epoxy (the quantity of epoxy in
GO/epoxy was 15 wt.%) was also directly mixed with PC matrix
to obtain PC/(GO/epoxy) composites. TEM observation on the
ultrathin sections of the composites was conducted in order to
determine the dispersion. As shown in Fig. 4a and b, GO sheets
were relatively evenly dispersed in the PC/(GO/epoxy) composites
due to the complete exfoliation of GO sheets through the freezedrying method, and no large aggregates, which would result in a
phase-separated structure, could be seen. The existence of small
GO stacks (marked by the yellow circles in Fig. 4b) should arise
from the restacking of GO sheets during the solution-blending process due to the strong van der Waals interactions [33]. For PC/GO
epoxy composites, GO sheets have been functionalized with epoxy
chains and the compatibility between GOepoxy and PC matrix
should be signicantly improved and the dispersion of GOepoxy
should be better than that of PC/(GOepoxy) composites. However,
apparent localization of GO aggregates (Fig. 4e and f) could be seen
in the TEM images of PC/(GOepoxy) composites. We think that
such aggregates should be caused by the signicant crosslinking
of GO layers via the epoxy chains, which made the dispersion more
difcult. In addition to the regions with large aggregates, areas of
small GO stacks were also observed, such as the area marked by
a yellow circle, which might be attributed to the slight crosslinking

91

of GO layers via the epoxy chains. Statistics about the number of


the aggregates revealed that about 10% and 15% of GOepoxy
was poorly dispersed in the composite with 0.5 wt.% and 1.0 wt.%
ller, respectively.
Though the dispersion of GOepoxy was poor due to the crosslinking of GO layers via the epoxy chains, the interfacial interactions between the PC matrix and GOepoxy could be enhanced.
It is because that, during hot-compression of PC/GOepoxy composites as illustrated in Fig. 5a, the residue active sites in the
grafted epoxy chains, such as the unreacted epoxide groups as well
as hydroxyl groups, could react with the PC carbonate to form
chemical bonds, leading to strong interfacial interactions between
the matrix and GOepoxy [1315]. Previous studies have proposed
that transesterication and alcoholysis reactions are the main
model reactions for PC/epoxy subjected to heating at high temperature [14]. Generally, the transesterication takes place between
the carbonate in PC chains and the unreacted epoxide groups in
epoxy. And the alcoholysis reactions occur between the dangling
AOH in epoxy and the carbonate in PC chains at the sites of hydroxyl groups [14]. For PC/(GO/epoxy) composites, it is known that GO
itself already has epoxides and hydroxyl groups, which would react
directly with PC matrix as well. However, during the hot-compression, a high processing temperature of 250 C and long processing
time of 30 min would decompose most of the epoxides and hydroxyl groups on GO [3437]. As a consequence, few chemical bonds
would be formed and hence the interfacial interactions between
the matrix and GO sheets would not be enhanced.
To clearly display that the strong interfaces exist between GO
sheets and the polymer matrix, SEM images were employed to observe the micro-morphology of fracture surface of the composite
with 0.5 wt.% GOepoxy. As shown in Fig. 5b and c, a single GO
sheet was embedded and tightly held to the polymer matrix
(pointed by yellow arrows), and the protruding segment was
thickly coated with a polymer layer, indicating strong polymerGO interactions [38]. In addition, the interface of GO aggregate
with the matrix was also investigated. As shown in Fig. 5d, the

Fig. 4. TEM images of the ultrathin sections of PC/(GO/epoxy) and PC/GOepoxy composites. (a and b) 1.0 wt.% GO/epoxy; (c and e) 0.5 wt.% GOepoxy; and (d and f) 1.0 wt.%
GOepoxy.

92

B. Shen et al. / Composites Science and Technology 77 (2013) 8794

Fig. 5. (a) Schematic illustration of the covalent bonds between GOepoxy and the PC matrix. (bd) SEM images of PC/GOepoxy composite showing fracture surface
topography.

presence of gaps (pointed by pink arrows) between the sheets and


the polymer matrix implied that GO sheets in the aggregate could
not all be wetted by the matrix due to the crosslinking. However,
they were still partially bond to the polymer matrix (pointed by
yellow arrows), though not as tightly as the sheets in Fig. 5b and
c. This phenomenon can be explained in terms of improved interfacial interactions between GO aggregate and the polymer matrix,
possibly due to the formation of chemical bonds.
For strong interfacial interactions, the viscoelasticity of polymer
chains adsorbed on the nanoparticle can be altered, and the length
scale of the connement effects for attractive interactions can be
several hundreds of nanometres [39,40]. The conned polymer
segments exhibit altered relaxation behavior which is directly reected as deviations of the loss modulus of the composites. Therefore, the loss modulus curves of PC/GOepoxy composites and pure
PC were analyzed and these curves were normalized by dividing
the values of the loss moduli by the maximum value of each of
the curves. As shown in Fig. 6, the broadening of the loss modulus
of PC/GOepoxy composites towards higher temperatures was
readily evident. Note that this shift is much more signicant than
that demonstrated for 1 wt.% loading PC/(GO/epoxy) sample and
the higher weight loading (10 wt.%) as-received MWCNT sample
in the literature [39]. It is because that the viscoelastic response
of the PC chains in the PC/GOepoxy sample is altered by both
immobilization of the chains bonded to the GOepoxy surface, as
well as secondary interactions between this adhered polymer layer
and those polymer chains located further from the GOepoxy surface. This shifting of the viscoelastic behavior suggests that the entire population of relaxation modes of the bulk polymer in the PC/
GOepoxy composite changes due to the strong interfacial
interactions.
3.3. Mechanical and thermal properties of the composites
As we know, the mechanical properties of a graphene-based
polymeric nanocomposite greatly depend on the dispersion of

Fig. 6. Normalized loss modulus as a function of temperature for pure PC, PC/GO
epoxy and PC/(GO/epoxy) composites.

graphene sheets in the polymeric matrix as well as interfacial


interactions between graphene sheets and the polymeric matrix.
However, in this work, the dispersion of GOepoxy in the PC matrix was poor, while the interfacial interactions between them
were strong. In order to study the effects of poor dispersion and
strong interfacial interactions on the nal mechanical properties
of the composites, the tensile test for PC/GOepoxy composites
were conducted. The typical stressstrain curves are shown in
Fig. S5 and their mechanical properties are given in Fig. 7a. The
addition of GOepoxy signicantly increases the modulus of the
composites as well as the tensile strength. It is striking to note that,
at the addition of 0.5 wt.% GOepoxy, the Youngs modulus of the
composite increases by 52.3% from 1113.7 to 1695.8 MPa, and
the tensile strength increases by 5.3% from 49.2 to 51.8 MPa, while
the elongation at break only is slightly decreased from 14.8% to
14.3%. For the composite with 1.0 wt.% GOepoxy, the Youngs

B. Shen et al. / Composites Science and Technology 77 (2013) 8794

93

Fig. 7. (a) Tensile properties of various PC/GOepoxy and PC/(GO/epoxy) lms. (b) The reinforced values of PC/TRG composites reported in the literature.

modulus increases by 72.1% from 1113.7 to 1916.8 MPa, and the


tensile strength increases by 20.3% from 49.2 to 59.2 MPa, while
the elongation at break is signicantly decreased from 14.8% to
8.0%, which could be attributed to the increased aggregation of
GOepoxy in the matrix, resulting in a stress concentration near
the GO in the stretching process. However, the reinforced values
measured here are signicantly higher than that of PC/(GO/epoxy)
samples with better dispersion and weak interfacial interactions
(Fig. 7b), as well as those shown in other similar works on thermally reduced graphene oxide (TRG)/PC composites (Fig. 7b)
[41,42]. For example, the enhancement of modulus is 1.5 and 1.6
times higher than that of the PC composites with 0.5 wt.% and
1.0 wt.% TRG [42]. Such superior mechanical properties can certainly be attributed to the strong interfacial adhesion between
GOepoxy and the PC matrix, resulting in effective load transfer
from the matrix to GO sheets. This further suggested that the
composites with strong interfacial interactions between graphenematrix could also achieve excellent mechanical properties
even the dispersion of graphene is poor.
To predict the Youngs modulus of unidirectional or randomly
distributed ller-reinforced composites, the HalpinTsai model is
widely used in the literature [4345]. We also use this model to
predict the Youngs modulus for the PC/GOepoxy composites
(the details were shown in the Supporting Information). As shown
in Fig. 8, it was found that our experimental results was much
higher than the calculated values with randomly distributed GO,
reaching the values with aligned GO. However, GO sheets are

Fig. 8. Experimental Youngs moduli of PC/GOepoxy composites and those


calculated using HalpinTsai models with two extreme cases: the random
orientation and unidirectional dispersion of GO sheets in the PC matrix.

poorly and randomly dispersed in PC/GOepoxy composites, which


could be conrmed by TEM observation (Fig. 4). This phenomenon
was also observed in previous reports [40,46,47]. It is possibly because that the strong interactions between GOepoxy and the
polymer matrix have caused a substantial interphase zone, the adsorbed polymer chains conned in this area have different conformation and viscoelasticity as compared to the bulk polymers
(Fig. 6). Moreover, these interfacial layers also have a signicantly
higher modulus than the bulk polymers. Therefore, the efcient
volume fraction of the ller is much higher than the real volume
fraction of the ller [40,47]. That is why the PC/GOepoxy composites with poor dispersion showed an enhanced reinforcement as
compared to the PC/TRG composites with better dispersion.
Moreover, the thermal stability of these composites was also
mildly increased even with the low loadings of GOepoxy used.
TGA and corresponding differential thermogravimetric analysis
(DTG) results for pure PC and PC/GOepoxy composites are shown
in Fig. 9. The peak temperature (Tp) of the DTG curve represents the
temperature at which the maximum weight loss rate was reached,
as shown in the inserted gure. The Tp of PC/GOepoxy composites
appeared at about 486.7 C and 498.0 C with 0.5 wt.% and 1.0 wt.%
GOepoxy, and was increased by about 9.3 C and 20.6 C, respectively, as compared to pure PC. However, the Tp of PC/TRG composite with 3.0 wt.% TRG was almost not changed compared to pure PC
in the literature [41]. These results suggested that the addition of
GOepoxy make PC more thermally stable, possibly due to the formation of strong interfacial interactions between GOepoxy and
the PC matrix.

Fig. 9. TGA and DTG curves for pure PC and PC/GOepoxy samples at a heating rate
of 20 C/min.

94

B. Shen et al. / Composites Science and Technology 77 (2013) 8794

4. Conclusions
The individually dispersed GO sheets in DMF are obtained by
dispersing freeze-dried GO in DMF under mild sonication. Then,
the epoxy is successfully grafted onto GO sheets via the reaction
between the epoxide groups in epoxy and the carboxy groups on
GO with the assistance of TPP. Since each epoxy chain bears two
terminated epoxide groups, it is inevitable that one epoxy chain
connects two GO sheets together, causing the crosslinking of GO
layers via the epoxy chain, which was conrmed by AFM measurement. Due to these crosslinking, the dispersion of GOepoxy in the
PC matrix was poor. However, the residue active sites in the
grafted epoxy chains could react with PC carbonate to form chemical bonds, leading to strong interfacial interactions between the
matrix and GO sheets. Owing to these strong interfacial interactions, the enhancements of the mechanical properties of PC/GO
epoxy composites was signicantly higher than that of PC/(GO/
epoxy) samples, as well as those shown in other similar works
on TRG/PC composites with better dispersion.
Acknowledgments
The authors are grateful to the National Natural Science Foundation of China (Grants 51003115), and Ningbo Key Lab of Polymer
Materials (Grant No. 2010A22001) for their nancial support of
this study.
Appendix A. Supplementary material
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.compscitech.
2013.01.014.
References
[1] Novoselov KS, Geim AK, Morozov SV, Jiang D, Zhang Y, Dubonos SV, et al. Electric
eld effect in atomically thin carbon lms. Science 2004;306(5696):6669.
[2] Balandin AA, Ghosh S, Bao W, Calizo I, Teweldebrhan D, Miao F, et al. Superior
thermal conductivity of single-layer graphene. Nano Lett 2008;8(3):9027.
[3] Lee C, Wei X, Kysar JW, Hone J. Measurement of the elastic properties and
intrinsic strength of monolayer graphene. Science 2008;321(5887):3858.
[4] Zhang Y, Tan Y-W, Stormer HL, Kim P. Experimental observation of the quantum
Hall effect and Berrys phase in graphene. Nature 2005;438(7065):2014.
[5] Shen B, Zhai W, Chen C, Lu D, Wang J, Zheng W. Melt blending in situ enhances
the interaction between polystyrene and graphene through pp stacking. ACS
Appl Mater Interface 2011;3(8):31039.
[6] Shen B, Lu D, Zhai W, Zheng W. Synthesis of graphene by low-temperature
exfoliation and reduction of graphite oxide under ambient atmosphere. J Mater
Chem C 2013;1(1):503.
[7] Fang M, Wang K, Lu H, Yang Y, Nutt S. Covalent polymer functionalization of
graphene nanosheets and mechanical properties of composites. J Mater Chem
2009;19(38):7098105.
[8] Goncalves G, Marques PAAP, Barros-Timmons A, Bdkin I, Singh MK, Emami N,
et al. Graphene oxide modied with PMMA via ATRP as a reinforcement ller. J
Mater Chem 2010;20(44):992734.
[9] Shen B, Zhai W, Lu D, Wang J, Zheng W. Ultrasonication-assisted direct
functionalization of graphene with macromolecules. RSC Adv 2012;2(11):47139.
[10] Stankovich S, Piner RD, Chen X, Wu N, Nguyen ST, Ruoff RS. Stable aqueous
dispersions of graphitic nanoplatelets via the reduction of exfoliated graphite
oxide in the presence of poly(sodium 4-styrenesulfonate). J Mater Chem
2006;16(2):1558.
[11] Niyogi S, Bekyarova E, Itkis ME, McWilliams JL, Hamon MA, Haddon RC. Solution
properties of graphite and graphene. J Am Chem Soc 2006;128(24):77201.
[12] Quintana M, Montellano A, del Rio Castillo AE, Tendeloo GV, Bittencourt C,
Prato M. Selective organic functionalization of graphene bulk or graphene
edges. Chem Commun 2011;47(33):93302.
[13] Li M-S, Ma C-CM, Lin M-L, Chang F-C. Chemical reactions occurring during the
preparation of polycarbonate-epoxy blends. Polymer 1997;38(19):490313.
[14] Su CC, Woo EM, Chen CY, Wu R-R. N.m.r. and FT i.r. studies on transreactions
and hydroxyl exchanges of bisphenol-A polycarbonate with an epoxy upon
heating. Polymer 1997;38(9):204756.
[15] Su CC, Woo EM. Chemical interactions in blends of bisphenol a polycarbonate
with tetraglycidyl-4,40 -diaminodiphenylmethane epoxy. Macromolecules
1995;28(20):677986.
[16] Hummers WS, Offeman RE. Preparation of graphitic oxide. J Am Chem Soc
1958;80(6):1339.

[17] Sadagopan K, Ratna D, Samui AB. Synthesis and characterization of liquidcrystalline epoxy and its blend with conventional epoxy. J Polym Sci Part A:
Polym Chem 2003;41(21):337583.
[18] Malshe VC, Waghoo G. Chalk resistant epoxy resins. Prog Org Coat
2004;51(3):17280.
[19] Nethravathi C, Rajamathi M. Chemically modied graphene sheets produced
by the solvothermal reduction of colloidal dispersions of graphite oxide.
Carbon 2008;46(14):19948.
[20] Wang H, Robinson JT, Li X, Dai H. Solvothermal reduction of chemically
exfoliated graphene sheets. J Am Chem Soc 2009;131(29):99101.
[21] Zhu Y, Stoller MD, Cai W, Velamakanni A, Piner RD, Chen D, et al. Exfoliation of
graphite oxide in propylene carbonate and thermal reduction of the resulting
graphene oxide platelets. ACS Nano 2010;4(2):122733.
[22] Pham VH, Cuong TV, Hur SH, Oh E, Kim EJ, Shin EW, et al. Chemical
functionalization of graphene sheets by solvothermal reduction of a graphene
oxide suspension in N-methyl-2-pyrrolidone. J Mater Chem 2011;21(10):
33717.
[23] Li M-S, Ma C-CM, Chen J-L, Lin M-L, Chang F-C. Epoxy-polycarbonate blends
catalyzed by a tertiary amine. 1. Mechanism of transesterication and
cyclization. Macromolecules 1996;29(2):499506.
[24] Kudin KN, Ozbas B, Schniepp HC, Prudhomme RK, Aksay IA, Car R. Raman
spectra of graphite oxide and functionalized graphene sheets. Nano Lett
2007;8(1):3641.
[25] Zhang S, Xiong P, Yang X, Wang X. Novel PEG functionalized graphene
nanosheets: enhancement of dispersibility and thermal stability. Nanoscale
2011;3(5):216974.
[26] Tang Z, Kang H, Shen Z, Guo B, Zhang L, Jia D. Grafting of polyester onto graphene
for electrically and thermally conductive composites. Macromolecules
2012;45(8):344451.
[27] Stankovich S, Dikin DA, Piner RD, Kohlhaas KA, Kleinhammes A, Jia Y, et al.
Synthesis of graphene-based nanosheets via chemical reduction of exfoliated
graphite oxide. Carbon 2007;45(7):155865.
[28] Meyer JC, Geim AK, Katsnelson MI, Novoselov KS, Booth TJ, Roth S. The
structure of suspended graphene sheets. Nature 2007;446(7131):603.
[29] Lin Y, Zhou B, Shiral Fernando KA, Liu P, Allard LF, Sun Y-P. Polymeric carbon
nanocomposites from carbon nanotubes functionalized with matrix polymer.
Macromolecules 2003;36(19):7199204.
[30] Xie L, Xu F, Qiu F, Lu H, Yang Y. Single-walled carbon nanotubes functionalized
with high bonding density of polymer layers and enhanced mechanical
properties of composites. Macromolecules 2007;40(9):3296305.
[31] Fang M, Wang K, Lu H, Yang Y, Nutt S. Single-layer graphene nanosheets with
controlled grafting of polymer chains. J Mater Chem 2010;20(10):198292.
[32] McAllister MJ, Li J-L, Adamson DH, Schniepp HC, Abdala AA, Liu J, et al. Single
sheet functionalized graphene by oxidation and thermal expansion of
graphite. Chem Mater 2007;19(18):4396404.
[33] Potts JR, Dreyer DR, Bielawski CW, Ruoff RS. Graphene-based polymer
nanocomposites. Polymer 2011;52(1):525.
[34] Tang H, Ehlert GJ, Lin Y, Sodano HA. Highly efcient synthesis of graphene
nanocomposites. Nano Lett 2011;12(1):8490.
[35] Li W, Tang X-Z, Zhang H-B, Jiang Z-G, Yu Z-Z, Du X-S, et al. Simultaneous surface
functionalization and reduction of graphene oxide with octadecylamine for
electrically conductive polystyrene composites. Carbon 2011;49(14):472430.
[36] Zhang H-B, Wang J-W, Yan Q, Zheng W-G, Chen C, Yu Z-Z. Vacuum-assisted
synthesis of graphene from thermal exfoliation and reduction of graphite
oxide. J Mater Chem 2011;21(14):53927.
[37] Zheng D, Tang G, Zhang H-B, Yu Z-Z, Yavari F, Koratkar N, et al. In situ thermal
reduction of graphene oxide for high electrical conductivity and low
percolation threshold in polyamide 6 nanocomposites. Compos Sci Technol
2012;72(2):2849.
[38] Ramanathan T, Abdala AA, Stankovich S, Dikin DA, Herrera Alonso M, Piner RD,
et al. Functionalized graphene sheets for polymer nanocomposites. Nat Nano
2008;3(6):32731.
[39] Eitan A, Fisher FT, Andrews R, Brinson LC, Schadler LS. Reinforcement
mechanisms in MWCNT-lled polycarbonate. Compos Sci Technol
2006;66(9):116273.
[40] Wan C, Chen B. Reinforcement and interphase of polymer/graphene oxide
nanocomposites. J Mater Chem 2012;22(8):363746.
[41] Potts JR, Murali S, Zhu Y, Zhao X, Ruoff RS. Microwave-exfoliated graphite
oxide/polycarbonate composites. Macromolecules 2011;44(16):648895.
[42] Kim H, Macosko CW. Processing-property relationships of polycarbonate/
graphene composites. Polymer 2009;50(15):3797809.
[43] Kalaitzidou K, Fukushima H, Miyagawa H, Drzal LT. Flexural and tensile moduli
of polypropylene nanocomposites and comparison of experimental data to
HalpinTsai and TandonWeng models. Polym Eng Sci 2007;47(11):1796803.
[44] Schaefer DW, Justice RS. How nano are nanocomposites? Macromolecules
2007;40(24):850117.
[45] Liang J, Huang Y, Zhang L, Wang Y, Ma Y, Guo T, et al. Molecular-level
dispersion of graphene into poly(vinyl alcohol) and effective reinforcement of
their nanocomposites. Adv Funct Mater 2009;19(14):2297302.
[46] Wang Y, Shi Z, Yu J, Chen L, Zhu J, Hu Z. Tailoring the characteristics of graphite
oxide nanosheets for the production of high-performance poly(vinyl alcohol)
composites. Carbon 2012;50(15):552536.
[47] Gao J, Chen F, Wang K, Deng H, Zhang Q, Bai H, et al. A promising alternative to
conventional polyethylene with poly(propylene carbonate) reinforced by
graphene oxide nanosheets. J Mater Chem 2011;21(44):1762730.

Вам также может понравиться