You are on page 1of 8

Industrial Crops and Products 72 (2015) 125–132

Contents lists available at ScienceDirect

Industrial Crops and Products


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

Toughened polyester cellulose nanocomposites: Effects of cellulose


nanocrystals and liquid epoxidized natural rubber on morphology
and mechanical properties
Hanieh Kargarzadeh a , Rasha M. Sheltami a,c , Ishak Ahmad a,∗ ,
Ibrahim Abdullah a , Alain Dufresne b
a
Faculty of Science and Technology, School of Chemical Sciences and Food Technology, Polymer Research Center (PORCE),
Universiti Kebangsan Malaysia (UKM), 43600 Bangi, Selangor, Malaysia
b
The International School of Paper, Print Media and Biomaterials (PAGORA), Grenoble Institute of Technology, BP65,
38402 Saint Martin d’He‘res Cedex, France
c
Chemistry Department, Faculty of Science, University of Benghazi, Benghazi, Libya

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

Article history: Unsaturated polyester resin (UPR) has been modified with a liquid epoxidized natural rubber (LENR). The
Received 17 September 2014 addition of LENR improves the toughness, but it is inevitably accompanied by a significant loss in stiff-
Received in revised form ness, yield strength, and thermal resistance. Incorporation of a rigid material into the LENR–UPR blend,
14 November 2014
such as cellulose nanocrystal (CNC), provided a ternary polyester nanocomposites with desire mechani-
Accepted 22 December 2014
cal properties. In the present work, the effects of different CNC and LENR content on the nanocomposite
Available online 14 January 2015
morphology and mechanical properties were examined. Morphological studies revealed that the size of
the rubber particles is independent of the rubber content due to the chemical interaction of rubber and
Keywords:
Blend
matrix; however, the size of the LENR particles increased with increasing CNC content. Tensile tests indi-
Unsaturated polyester resin cated that the blend’s tensile strength and elastic modulus could be further improved by incorporation
Cellulose nanocrystal of the CNCs. The high crystallinity and surface area of CNCs were the main factors for improvement of
Liquid epoxidized natural rubber tensile properties of the blend. Moreover, the impact energy of the UPR improved when it was modified
Nanocomposite with the LENR, and even further improvement was achieved with the CNC addition to form nanocompos-
Toughness ites. Dynamic mechanical thermal analysis (DMTA) showed that the storage modulus of nanocomposites
with 1.5 wt% rubber content was lower than those with 4.5 wt% LENR; however, both samples showed a
higher storage modulus than the neat UPR and the unfilled blend. The glass transition temperature (Tg ) of
the UPR decreased with the LENR addition, but improved with the addition of CNCs. The nanocomposites
prepared with a higher LENR content displayed better mechanical properties than those with low LENR
content.
© 2014 Elsevier B.V. All rights reserved.

1. Introduction equipment, containers, automobiles, and cultured marble because


of its clarity and excellent chemical and corrosion resistance. The
Unsaturated polyester resins (UPRs) is one of the most popu- major drawback of UPRs is that they are brittle with poor impact
lar thermoset materials used as a matrix for composites, electronic resistance in the cured state. Therefore, toughening UPRs has
been the subject of intense investigation throughout the world
(Kargarzadeh et al., 2014; Thomas et al., 2013).
The most common technique for toughening brittle UPR is incor-
Abbreviations: UPR, unsaturated polyester resin; LENR, liquid epoxidized nat- porating a dispersed elastomeric phase in the resin. In this process, a
ural rubber; CNC, cellulose nanocrystal; MEKPO, methyl ethyl ketone peroxide;
rubber is initially miscible with the resin and a curing agent. When
Tg , glass transition temperature; ENR, epoxidized natural rubber; DMTA, dynamic
mechanical thermal analysis; FESEM, field emission scanning electron microscope; the reaction begins, the rubber generally forms an initial copoly-
TEM, transmission electron microscopy; CTBN, carboxyl-terminated butadiene-co- mer with the resin before phase separation, resulting in the cured
acrylonitrile. thermosetting polymer possessing a dispersed rubbery phase. The
∗ Corresponding author. Tel.: +603 8921 5431/5424; fax: +603 8921 5410.
addition of a rubber modifier improves the fracture toughness, but
E-mail addresses: hanieh.kargar@gmail.com (H. Kargarzadeh),
it is inevitably accompanied by a significant loss in stiffness, yield
gading@ukm.edu.my (I. Ahmad).

http://dx.doi.org/10.1016/j.indcrop.2014.12.052
0926-6690/© 2014 Elsevier B.V. All rights reserved.
126 H. Kargarzadeh et al. / Industrial Crops and Products 72 (2015) 125–132

strength, and thermal resistance. This is not unexpected, as the elas- ing mixture was stirred with a mechanical stirrer and heated at
tic modulus of the toughness modifier is significantly lower than 60 ◦ C until the excess styrene evaporated. The mixture was then
that of the thermosetting matrix (Kargarzadeh et al., 2014). There- cooled to room temperature, the LENR was added to the UPR–CNC
fore, rigid inorganic nanoparticles, such as silica, clay, and carbon and mixed for approximately 2 h before the initiator (1.5% con-
have been used to toughen polyester without adversely affecting centration relative to the resin) was added and stirred for 2 min.
the elastic modulus. Although the polyester nanocomposites show The final mixture was poured into molds and cured at room tem-
better balance properties, the toughening effect of the inorganic perature for 24 h. Different percentages of CNC loading (2, 4 and
nanoparticles was not so prominent compare with the liquid rubber 6 wt%) were prepared. The nanocomposites using the CNC will be
(Afina et al., 2013; Vijayan et al., 2012; Xu et al., 2013). referred to as CNC–LENR–UPR. Two different nanocomposites were
Cellulose nanocrystals (CNCs) are a very important biopolymer prepared, with 1.5 and 4.5 wt% LENR content, chosen on the basis
that have attracted a great deal of interest in the nanocompos- of the mechanical properties (blend with 1.5 wt% LENR indicated
ite field, due to their high modulus, strength, low density, high higher tensile strength and 4.5 wt% LENR performed higher impact
crystallinity, high surface area, unique optical properties, and their energy).
renewability and environmental friendly (Dufresne, 2012; Habibi
et al., 2010). CNCs have been widely used as a reinforcement for a 2.2. Characterization
variety of thermoplastic and thermoset polymeric matrices. How-
ever, reports on the preparation of CNC-reinforced thermosetting 2.2.1. Microscopy
matrices are relatively rare. There have been only a few researchers The fractured surface morphologies of the LENR–UPR blend
who have reported utilizing CNCs in an epoxy resin (Tang and and the CNC–LENR–UPR nanocomposites were investigated using
Weder, 2010; Xu et al., 2013), but to the best of our knowledge, a Zeiss Supra 55VP field emission scanning electron microscope
no report has been found on using an unsaturated polyester as the (FESEM). All samples were sputter-coated with gold prior to obser-
matrix to produce a CNC-reinforced, rubber-modified thermoset. vation to prevent charge build up. In the case of the blend, its rubber
In this study, liquid epoxidized natural rubber (LENR) was used phase was extracted using toluene for 12 h under ambient condi-
to modify the UPR, an orthophthalic polyester resin containing 30% tions.
styrene, and then, the resulting material was combined with CNCs. Transmission electron microscopy (TEM) was conducted using
Both, LENR and CNC have been prepared from natural source and a Philips CM30 microscope to investigate the morphology of
they are good candidates to replace synthetic liquid rubbers and the CNC–LENR–UPR nanocomposite. A thin film (70 nm) of this
inorganic fillers to toughen UPR while keeping desirable mechan- nanocomposite was prepared using cryo-ultramicrotomy (Leica
ical properties. The objective of this research was to investigate EM, FC6) under liquid nitrogen.
the effects of the LENR and CNC content on the morphology and
mechanical properties of the nanocomposites. In particular, the 2.2.2. Tensile test
dispersion of the rubber particles and CNCs, along with extensive Tensile tests of the samples were performed using an Instron
discussion on the matrix interaction and toughening mechanism of Universal Testing Machine, model 5567 according to the ASTM D-
both the blends and the nanocomposites was evaluated. 638-91 standard test methods for examining the tensile properties
of the unsaturated polyester at a 5 mm/min crosshead speed. The
2. Materials and methods specimens were cut from the cured sheets to a 160 × 13 × 3 mm
2.1. Materials and chemicals size, with the shown values taken from an average of 10 measure-
ments.
Raw kenaf (Hibiscus cannabinus) bast fibers, from which CNCs
were prepared, were supplied by KFI Sdn. Bhd. (Malaysia). The UPR 2.2.3. Impact test
was supplied by Revertex Sdn. Bhd. (Malaysia). All other chemi- The impact strength of unnotched specimens was measured
cals used were purchased from SYSTERM, Classic Chemicals Sdn. following the ASTM D4812 impact resistance of plastic test. The
Bhd. (Malaysia). LENR was prepared in our laboratory by photo- tests were run using a RAY-RAN digital universal pendulum impact
sensitized degradation of dried epoxidized natural rubber (ENR) system at room temperature, for an average of 10 samples.
(Abdullah and Zakaria, 1989; Kargarzadeh et al., 2015).
2.2.4. Dynamic mechanical thermal analysis (DMTA)
2.1.2. Cellulose nanocrystal preparation Dynamic mechanical thermal analysis was performed using a
Colloidal suspensions of cellulose nanocrystals (CNC) in water TA Instruments, DMTA instrument model Q 800 according to the
were prepared from sulfuric acid hydrolysis of cellulose extracted ASTM D4065. The purpose of this test was to determine the storage
from kenaf bast fiber as reported in our previous work (Kargarzadeh modulus (E ), tan delta (ı), and glass transition temperature (Tg ) of
et al., 2012). the samples. The peak in the tan delta vs. temperature thermogram
was taken as Tg . Measurements were taken from −100 to 150 ◦ C at a
2.1.3. Preparation of LENR-modified UPR blend 5 ◦ C/min heating rate and a 1 Hz fixed frequency. A single cantilever
To prepare the rubber-modified UPR, LENR was added to the UPR mode was used for the 35 × 13 × 3 mm samples, with values taken
and mixed using an electrical stirrer for approximately 3 h. After from an average of 4 specimens.
that, methyl ethyl ketone peroxide (MEKPO) initiator at a 1.5 wt%
concentration was added and stirred for 2 min. The mixture was 3. Results and discussion
poured into a 3-mm-thick aluminum mold and cured at room tem- 3.1. LENR-modified UPR
perature for 24 h. The rubber content was varied between 1.5 and
6% (w/w) (Kargarzadeh et al., 2015). 3.1.1. Morphological analysis
Fig. 1 shows the typical micrograph of a neat UPR and rubber-
2.1.4. Nanocomposite processing modified polyesters. The neat UPR displayed a smooth and glassy
Freeze-dried CNC were dispersed in styrene by using an ultra- fractured surface with ripples, demonstrative of its poor impact
sonicator for 30 min in an ice bath. The CNC suspension in styrene strength (Ahmad and Hassan, 2010; Kargarzadeh et al., 2015). How-
was mixed with the UPR and sonicated for 30 min. The result- ever, the fractured surface of the LENR–UPR blend [Fig. 1(b–d)]
H. Kargarzadeh et al. / Industrial Crops and Products 72 (2015) 125–132 127

Fig. 1. FESEM micrographs of the neat UPR and LENR–UPR blend: (a) UPR, (b) 1.5 wt%, (c) 4.5 wt%, and (d) the 4.5 wt% blend at higher magnification. White arrows in (c)
showing the rubber particles which remains in the matrix after toluene extraction. White arrows in image (d) show crack pinning effect induced by LENR.

clearly shows two distinct phases, a continues UPR matrix and the neat UPR (40 MPa) increased to 49 MPa after incorporating 1.5 wt%
dispersed rubber particles, that appear as dark spherical phase due LENR, and then it levels off with increased rubber content. The ten-
to the extraction of rubber particles after treatment with toluene sile modulus displayed a similar trend to other rubber–polyester
for 12 h. blends (Ahmad and Hassan, 2010; Ben Saleh et al., 2009; Hisham
In general, the size of the rubber phase increases with increasing et al., 2011), showing a reduction with increasing rubber content.
rubber content (Ben Saleh et al., 2009; Saadati et al., 2005; Thomas The high tensile strength of the 1.5 wt% rubber content
et al., 2010). However, for the LENR–UPR blend, the size of the rub- LENR–UPR blend was due to the enhanced compatibility of the
ber particles was independent of rubber content, resulting from LENR with the UPR matrix via chemical interaction. At this quan-
the high compatibility and miscibility of the LENR due to its chem- tity, the majority of the rubber could interact and cross-link with
ical interactions with the UPR matrix. The likely mechanism and the UPR. In contrast, due to the coagulation of rubber particles at
chemical interaction occurring between the LENR and UPR is shown higher rubber concentration, fewer rubber particles were available
in Fig. 2, illustrating the potential chemical interaction as well as for cross-linking to the UPR, as the majority remained free rubber
hydrogen bonding between the reactive terminal groups (OH or particles within the UPR matrix, thus reducing the tensile strength.
OOH) of the LENR and the carboxyl groups of the UPR, improving A lower tensile modulus of the blend compared to the neat UPR
the compatibility between the rubber phase and polyester. In addi- is attributed to the LENR’s soft segment structure as well as to the
tion, it can be seen that not all of the rubber phase was extracted by intrinsically low modulus of liquid rubbers compared to neat UPR
the toluene which is shown by white arrows in Fig. 1(c), indicating (Seng et al., 2011).
a strong degree of interaction between the LENR and UPR. The impact energy of the UPR increased significantly with the
addition of the rubber. For example, a 4.5 wt% LENR improved
3.1.2. Mechanical properties of the LENR–UPR blend the impact energy by a factor of 156% (Table 1). According to
The effects of LENR content on the tensile and impact properties Ratna (2004), both chemical interaction between the liquid rub-
of the UPR is depicted in Table 1. It is well known that the incor- ber and polyester, and a high degree of rubber particle dispersion
poration of a rubber phase into polyester matrices significantly are required to achieve effective toughening properties. It can be
reduces the tensile strength and modulus of the resultant blend seen from Fig. 1(b and c) that the LENR particles were dispersed
(Ben Saleh et al., 2009; Cherian and Thachil, 2004; Mathew et al., homogenously throughout the matrix. The LENR can also chemi-
2010). However, the LENR–UPR blend shows a different behavior cally interact with the UPR via both the epoxy group and double
for tensile strength. As seen in Table 1, the tensile strength of the band, as suggested in Fig. 2.
The increase in toughness was also due to the amount of elastic
energy stored in the rubber particles while stretching. Therefore,
Table 1
the deformation of the rubber particles in the matrix appeared
The tensile and impact properties of the neat UPR and LENR–UPR blends
(Kargarzadeh et al., 2015). to be responsible for the enhanced stress transfer and consequent
increase in impact resistance. Other toughening mechanisms, such
Sample (all Tensile strength Modulus (MPa) Impact energy
as shear yielding, crack pinning, and crack deflection are also
numbers are (MPa) (kJ/m2 )
“wt%”) responsible for the higher impact resistance of the LENR–UPR
blend. Fig. 1(d) shows crack pinning effect. According to this mech-
UPR 40 ± 3.4 821 ± 50 2.5 ± 0.4
1.5 LENR 49 ± 1.5 624 ± 28 4.0 ± 0.5
anism, the proposed role of the rubber particles is to behave
3.0 LENR 39 ± 0.5 531 ± 12 5.4 ± 0.6 as impenetrable objects that cause the crack to bow out which
4.5 LENR 35 ± 1.5 445 ± 24 6.4 ± 0.6 consumes extra energy. Indirect evidence for the crack pinning
128 H. Kargarzadeh et al. / Industrial Crops and Products 72 (2015) 125–132

Fig. 2. Likely chemical structures of the cured LENR–UPR blend.

mechanism comes from the occurrence “tails” near the particles CNCs dispersed individually or as small aggregates within the UPR
on the fracture surface. The blends with 1.5 wt% and 4.5 wt% LENR, matrix, with the majority aggregated around the rubber particles
with the highest strength and highest impact energy, respectively, within the interphase.
were selected to prepare nanocomposites. The typical fractured surfaces of a LENR–UPR sample and the
CNC–LENR–UPR nanocomposites are shown in Fig. 4. As previously
3.2. Morphology investigation of the nanocomposites mentioned, the fractured surface of LENR modified UPR consists
of two distinct phases: spherical LENR particles (ranging from 1 to
The cryo-ultramicrotomed CNC–LENR–UPR nanocomposite 7 ␮m) dispersed uniformly in a continuous UPR matrix as seen in
thin film with 6 wt% CNC and 4.5 wt% LENR was examined using Fig. 4(a). Fig. 4(b and c) shows the fractured surface micrographs
TEM to study the composite’s morphology at the nanoscale. The of nanocomposites with different CNC loading and 1.5 wt% LENR,
nanocomposite with highest amount of CNC and rubber was cho- while Fig. 4(d and e) illustrates the fractured surfaces of nanocom-
sen to increase the possibility of the presence of both CNC and posite with 4.5 wt% of LENR with 2 wt% and 6 wt% CNC loading,
rubber particles in the nanocomposite thin film. Fig. 3 demon- respectively. By comparing the micrographs it is easy to identify
strates that the nanoscale rod-like CNC shape was preserved the CNC nanoparticles. They appear dotty, corresponding to the
after the nanocomposite preparation. Fig. 3 illustrates that the transverse section of the CNCs. Fig. 4(b–e) shows that the rubber
particles are clearly phase separated and formed as spherical par-
ticles, with the incorporated CNCs. The size of the rubber particles
slightly increased (2–10 ␮m) as the CNC filler loading increased.
However, this phenomenon was more obvious for the nanocom-
posites with 4.5% LENR, potentially explained by the fact that the
increased viscosity hinders phase separation of the rubber from the
UPR matrix, leading to large particle formation during curing (Tang
et al., 2012). Moreover, it can be seen that all of the nanoparticles
are embedded in the matrix, with no debonding or voids observed,
and identifying the CNCs was difficult, indicating they were well
dispersed.

3.3. Tensile properties

The effects of the CNC concentration on the LENR–UPR’s ten-


sile properties were investigated. Fig. 5(a and b) shows the tensile
strength and modulus evolution for the LENR–UPR blend and
CNC–LENR–UPR nanocomposites at different CNC and LENR con-
tent. As discussed previously, the tensile strength of the neat UPR
Fig. 3. TEM image of a 6 wt% CNC and 4.5 wt% LENR, CNC–LENR–UPR increased from 40 to 49 MPa with 1.5 wt% rubber, and decreased
nanocomposite. to 35 MPa with 4.5 wt% rubber. A different behavior was observed
H. Kargarzadeh et al. / Industrial Crops and Products 72 (2015) 125–132 129

Fig. 4. Fractured surface FESEM micrographs of (a) the LENR–UPR blend; (b) 2 wt% CNC–1.5 wt% LENR–UPR; (c) 6 wt% CNC–1.5 wt% LENR–UPR; (d) 2 wt% CNC–4.5 wt%
LENR–UPR; and (e) 6 wt% CNC–4.5 wt% LENR–UPR nanocomposites.

with the CNC addition into the blends with the different rubber con- interface provide bridges between the LENR and UPR phase, result-
centrations. In case of the low rubber content samples, the tensile ing in a better adhesion between the two (Fig. 3); and (3) the rigid
strength reduced with the addition of 2 wt% CNC and then slightly CNC particle effects are more noticeable with an increased elastic
increased with 4 and 6 wt% CNC. The lower tensile strength of the rubber phase. The presence of more rubber reduces the UPR stiff-
nanocomposites with the 1.5 wt% LENR and 2 wt% CNC content may ness, as it is an intrinsically rigid material. Thus, a large quantity of
be due to the dilution effect of CNCs when blended with the matrix rubber introduces a greater softening effect on the matrix, and the
(Mandal and Chakrabarty, 2014). addition of a small amount of rigid particles improved the tensile
The slight increment of tensile strength with 4 and 6 wt% CNC strength significantly. Additionally a CNC 3D network forms, once
was due to the unique behavior of CNC, observed in the TEM image the CNC concentration is increased to 6 wt%; the formation of a 3D
(Fig. 3). The CNCs at the interface provide bridges between the LENR network is an accepted theory for higher mechanical properties of
and UPR phase, resulting in a better adhesion between the two. This nanocomposites.
is because the CNCs can have physical interactions with both the The tensile moduli of the neat UPR and its related nanocom-
LENR’s functional groups (e.g., the reactive end chain group, such posites at different LENR concentrations are depicted in Fig. 5(b).
as the C O, OH, or epoxy groups) (Abdullah, 1994; Abdullah and The results revealed that the tensile moduli of the nanocompos-
Zakaria, 1989) as well as the UPR’s carboxyl groups (Chirayila et al., ites increased with increasing CNC content for both the low and
2014). In addition, it seems that the 3D network of CNC, started to high LENR concentrations. This increase was expected, as the CNC’s
form at this CNC content, thus, tensile strength increased slightly. modulus is extremely high (136–155 GPa) (Eichhorn and Davies,
However, the tensile strength of nanocomposites with 4 and 6 wt% 2006). Moreover, the high surface area provided by the highly crys-
CNC, was lower than of the LENR–UPR blend. talline nanocellulose, increased the surface interaction between
In contrast, the tensile strength of nanocomposite with 4.5 wt% the filler and matrix and led the nanocomposite to display higher
LENR, steadily increased with increasing CNC concentration. The mechanical properties, especially tensile modulus.
increased tensile strength in the 4.5 wt% LENR samples may be As previously mentioned, the reinforcing effect of the CNCs is
ascribed to several reasons: (1) the nanoscale CNC was dispersed more noticeable in the high rubber content samples. For com-
uniformly in the LENR–UPR blend; (2) strong physical interactions parison, the nanocomposites with 1.5 wt% and 4.5 wt% LENR,
occurred, potentially via hydrogen bonding between the CNC to reinforced with 6 wt% CNCs have modulus values around 791 and
both the LENR and UPR, as shown in the TEM image, CNCs at the 778 MPa, corresponding to 27% and 75% improvement from the
130 H. Kargarzadeh et al. / Industrial Crops and Products 72 (2015) 125–132

Fig. 5. Mechanical properties of the neat UPR and the CNC-LENR-UPR nanocomposites relative to CNC content: (a) Tensile strength, (b) tensile modulus, and (c) impact
energy.

non-reinforced LENR–UPR samples, respectively. It can be con- particles act as deflection sites, which can lead to catastrophic fail-
cluded that the effect of the CNC addition overshadowed that of ure of the matrix (Kargarzadeh et al., 2015). However, the impact
the LENR-based modulus reduction. energy at this concentration is still higher than both the neat UPR
and the LENR–UPR blends.
3.4. Impact properties In the case of the CNC reinforced LENR–UPR blend, the tough-
ening mechanism resulted from both the rubber particles and
The impact energies of the LENR–UPR nanocomposites are plot- nanocellulose particles. However, the toughening mechanism from
ted in Fig. 5(c) relative to the neat UPR. The impact energy increases the CNCs is quite different from those of the rubber particles. Obvi-
with increasing CNC filler loading for both types of nanocompos- ously, rubber particle cavitation, shear deformation, crack pinning,
ites. However, for the nanocomposites with the 4.5 wt% LENR, the and crack deflection are the main toughening mechanisms induced
impact energy reduced at the 6 wt% CNC content, as described by the rubber particles. However, the role that CNCs play in the
below. All of the impact energies for the 4.5 wt% rubber content toughening of the nanocomposites cannot be directly observed in
nanocomposite were higher than those with 1.5 wt% LENR. For the FESEM images. According to Xu et al. (2013), the stress concen-
comparison, the maximum impact energy of the 1.5 wt% LENR and tration factor is dependent on particle size. Therefore, the stress
6 wt% CNC nanocomposite was 5.5 kJ/m2 , giving a 37.5% improve- concentration of the CNCs is small and cannot directly induce a
ment over the LENR–UPR sample; while the 4 wt% CNC and the localized shear deformation of the matrix. In addition, the rigid
4.5 wt% LENR nanocomposite had a 7.5 kJ/m2 impact energy, giv- nanoparticles cannot produce a negative hydrostatic stress in the
ing a 17.5% improvement. These impact energy numbers become matrix like the rubber particles, thus, they are not able to effectively
more significant when compared to the neat UPR: ∼96.5% and 200% initiate cavitation of the neighboring matrix. The nanoparticles are
improvement for the maximum impact values of nanocomposites also significantly smaller than the crack-opening displacement,
with 1.5 and 4.5 wt% LENR, respectively. making them unable to induce a plane mechanism such as crack
This significant enhancement was due to the unique properties pinning or deflection (Johnsen et al., 2007). Due to the CNCs’ high
of the cellulose nanocrystals, as their high aspect ratio can resist aspect ratio, the interface area between the CNCs and the matrix is
fractures, resulting in an increased energy dissipation from crack increased, indicating that the energy consumption required for par-
deflection at the filler-matrix interface. In addition, the LENR par- ticle debonding is also increased. Therefore, it is expected that the
ticles alone can bear a load in a triaxial tension and enhance shear energy needed for CNC removal or debonding from the matrix plays
localization by acting as stress concentrators. The parallel effect of an important role for toughening the nanocomposite (Xu et al.,
both the rubber particles and the CNCs are the primary reason for 2013).
improved impact energy in the nanocomposites. Moreover, the hydrophilic nature of the CNCs and the formation
The lower impact energy at 6 wt% CNC loading with the 4.5 wt% of a percolation network resulting from strong hydrogen bonding
LENR was likely due to the agglomeration of CNCs combined with between CNCs, especially at higher loading, is another potential
the formation of larger rubber particles. In fact, the large rubber reason for the higher impact properties. The CNC network behaves
H. Kargarzadeh et al. / Industrial Crops and Products 72 (2015) 125–132 131

Fig. 6. Dynamic mechanical properties of the neat UPR, LENR-UPR blend and CNC–LENR–UPR nanocomposites reinforced with 6 wt% CNC; (a) Storage modulus, (b) a close
up of a region in (a) and (c) tan ı. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

as an impact modifier, absorbing energy to prevent crack prop- moduli were higher than the neat UPR (4560 MPa), indicating that
agation through the matrix, similar to the rubber particles. The the modified nanocomposites are stiffer (Fig. 6a).
unique behavior of the CNCs, with localized agglomeration at the The storage moduli in the rubbery region (150 ◦ C) improved
LENR–UPR interphase (Fig. 3), facilitates an efficient energy trans- with increasing CNC content. For instance, the storage mod-
fer process to toughen the nanocomposite. The network of CNCs ulus of a blend with 6 wt% CNC increased by 88% and 183%
also act as stress concentrators during fracture, inducing a localized from the LENR–UPR blends alone, for the low (1.5 wt%) and
plastic deformation of the surrounding matrix that result in stress- high (4.5 wt%) rubber content nanocomposites, respectively. This
whitened zones on the fracture surface of the nanocomposite. This immense increase can be explained by the fact that the rubber
stress-whitening effect was related to a local plastic deformation particles become extremely soft at temperatures above their Tg
at the crack edge, which can further explain the increased fracture (150 ◦ C), and the CNCs’ rigid reinforcement effect became notice-
toughness of the nanocomposites compared to the LENR–UPR alone able, restricting the plastic. This phenomenon is more obvious
(Fig. 4). for the 4.5% LENR nanocomposites, especially at higher CNC con-
Table 1 and Fig. 5 show that choosing a careful match between tent. Moreover, the formation of a CNC rigid network, with the
concentrations of the two additives (LENR and CNC) can provide a CNCs connected by hydrogen bonding, could be another reason for
higher toughness than either the LENR–UPR alone or the neat UPR. the nanocomposites’ excellent mechanical properties (Azizi Samir
Therefore, using a ternary mixture of the LENR–UPR combined with et al., 2005; Goffin et al., 2011; Lu et al., 2008; Raquez et al., 2012).
the CNC represents a good strategy for achieving balance between Fig. 6(c) delineates the variation of tan ı as a function of temper-
the tensile modulus and toughness. ature for the neat UPR, the LENR–UPR blends and their 6 wt% CNC
nanocomposites. The tan ı peak, associated with Tg , was obtained
at 94 ◦ C for the neat UPR. As shown in Fig. 6, the incorporation of
1.5 and 4.5 wt% LENR into the UPR matrix reduced the neat UPR’s
3.5. Dynamic mechanical thermal analysis (DMTA)
Tg from 94 ◦ C to 90 and 89 ◦ C, respectively, due to the low modulus
The dynamic mechanical properties of both the neat UPR and nature of elastomers (Kargarzadeh et al., 2015). Additionally, the
the nanocomposites are shown in Fig. 6. Fig. 6(a) shows that modified UPR matrixes had reduced tan ı values, resulting from
the neat UPR has a storage modulus of ∼4560 MPa at −100 ◦ C. enhanced damping properties in the matrix induced by the LENR
At this temperature, the polymer is in the glassy state and the particles.
modulus remained approximately constant with temperature. At Incorporation of the CNCs into the LENR–UPR blend slightly
around 50 ◦ C, the modulus began dropping sharply, corresponding increased the Tg values. Based on the Fig. 6, the Tg of the 1.5 wt%
to the relaxation of the polyester. In contrast, both the LENR–UPR LENR and 6 wt% CNC nanocomposite increased from 90 to 92 ◦ C,
blends and the nanocomposites displayed a two-step drop in the while that of the 4.5 wt% LENR and 6 wt% CNC nanocomposite
modulus (Fig. 6(b)). The low temperature modulus drop, which shifted from 89 to 92 ◦ C. These Tg values fall between the Tg range
increased with increasing of rubber content, is associated with of the blend and neat UPR. The increased Tg may be attributed
the glass transition temperature (Tg ) of the LENR (Kargarzadeh to increased adhesion between the polymer matrix and nanocel-
et al., 2015). The results revealed that the UPR’s storage modulus lulose particles, or it could result from the CNC’s nanometer size
increased to 4819 MPa and 4619 MPa after modification with 1.5 and homogenous dispersion through the matrix and rubber inter-
and 4.5 wt% LENR, due to the high density of crosslinking within the phase. In fact, the CNCs can have both a strong intramolecular
LENR–UPR blend. However, the blend with the 4.5 wt% LENR dis- interaction and a physical interaction with the matrix and LENR
played a lower value than that with the 1.5 wt%. A similar trend was particles, which can affect the polymer segment motion, reduc-
reported for a blend of epoxy and carboxyl-terminated butadiene- ing polymer chain mobility, a typical effect for CNC inclusion in
co-acrylonitrile copolymer (CTBN) (Thomas et al., 2004). a polymer system. The CNCs’ reinforcing effect is a direct result
Upon addition of the CNCs, the storage moduli increased with of the interaction between the filler and matrix, as well as the
increasing filler content for both the low and high rubber con- rigidity of the CNCs’ web-like structure (Lu et al., 2008). It appears
tent nanocomposites, becoming higher than the LENR–UPR blends that the CNC–LENR–UPR nanocomposites primary reinforcement
alone (Fig. 6a). This phenomenon showed the CNCs’ superior occurred from this CNC web-like structure of CNCs in addition to the
reinforcing effect in the LENR–UPR matrix at the glassy region CNCs’ high crystallinity. In comparison to inorganic nanoparticles,
(−100 ◦ C). The results indicate that the well-dispersed CNCs stiffen a reduction in Tg has been reported for nanocomposites prepared
the LENR–UPR matrix, especially for the blend with a higher rubber with CTBN, nanoclay, and epoxy (Chonkaew et al., 2013) as well as
content. It is worth noting that all of the nanocomposites’ storage CTBN, nanosilica, and epoxy (Tang et al., 2012).
132 H. Kargarzadeh et al. / Industrial Crops and Products 72 (2015) 125–132

Additionally, the magnitude of the tan ı peak increased slightly Chirayila, C.J., Joya, J., Mathewa, L., Koetzb, J., Thomas, S., 2014. Nanofibril
with the incorporation of CNCs into the LENR–UPR blends. This reinforced unsaturated polyester nanocomposites: morphology, mechanical
and barrier properties viscoelastic behaviorand polymer chain confinement.
incremental increase in peak height could result from the same Ind. Crops Prod. 56, 246–254.
polymer chain restriction occurring from the highly rigid web-like Chonkaew, W., Sombatsompop, N., Brostow, W., 2013. High impact strength and
CNC structure running through it, thus reducing the damping effect low wear of epoxy modified by a combination of liquid carboxyl terminated
poly (butadiene-co-acrylonitrile) rubber and organoclay. Eur. Polym. J. 49,
provided by the rubber particles, increasing the tan ı peak intensity. 1461–1470.
Dufresne, A., 2012. Nanocellulose: potential reinforcement in composites. In: John,
4. Conclusions M.J., Thomas, S. (Eds.), Natural Polymers: Volume 2: Nanocomposites. Royal
Society of Chemistry, Cambridge, pp. 1–32.
Eichhorn, S., Davies, G., 2006. Modelling the crystalline deformation of native and
An unsaturated polyester resin was toughened by incorporating regenerated cellulose. Cellulose 13, 291–307.
LENR. Tensile tests showed a significant reduction in strength and Goffin, A.L., Raquez, J.M., Duquesne, E., Siqueira, G., Habibi, Y., Dufresne, A., Dubois,
P., 2011. From interfacial ring-opening polymerization to melt processing of
modulus with increased rubber content, with the exception of an cellulose nanowhisker-filled polylactide-based nanocomposite.
increase observed for the tensile strength of 1.5 wt% LENR blend. Biomacromolecules 12, 2456–2465.
The size of LENR particles was independent of the LENR content Habibi, Y., Lucia, L., Rojas, O.J., 2010. Cellulose nanocrystals: chemistry
self-assembly, and applications. Chem. Rev. 110, 3479–3500.
due to the chemical interaction of rubber and matrix. Impact energy
Hisham, S.F., Ahmad, I., Daik, R., Ramli, A., 2011. Blends of LNR with unsaturated
increase by the factor of 156% with 4.5 wt% LENR. The CNC incor- polyester resin from recycled PET: comparison of mechanical properties and
poration significantly improved the mechanical properties of the morphological analysis with the optimum blend by commercial resin. Sains
LENR–UPR blend. CNC-based nanocomposites were prepared with Malaysiana 40, 729–735.
Johnsen, B., Kinloch, A., Mohammed, R., Taylor, A., Sprenger, S., 2007. Toughening
two different rubber concentrations (1.5 and 4.5 wt%). A signifi- mechanisms of nanoparticle-modified epoxy polymers. Polymer 48,
cant enhancement was observed in tensile modulus, when the CNCs 530– 541.
were added, due to their high crystalinity and stiffness. Impact tests Kargarzadeh, H., Ahmad, I., Abdullah, I., 2014. Liquid rubbers as toughening agents.
In: Thomas, S., Sinturel, C., Thomas, R. (Eds.), Micro- and Nano-structured
revealed that the UPR’s impact resistance increased with the LENR Epoxy/Rubber Blends. Wiley-VCH Verlag GmbH & Co. KGaA, Germany,
addition, and it improved further with the CNC addition, due to the pp. 31–52.
strong chemical interaction between the LENR and UPR, as well as Kargarzadeh, H., Ahmad, I., Abdullah, I., Dufresne, A., Zainudin, S.Y., Sheltami, R.M.,
2012. Effects of hydrolysis conditions on the morphology, crystallinity, and
the formation of a 3D CNC network that inhibits crack growth and thermal stability of cellulose nanocrystals extracted from kenaf bast fibers.
acts as a stress concentrator. The DMTA showed that the storage Cellulose 19, 855–866.
modulus of the LENR–UPR blends was higher than the neat UPR due Kargarzadeh, H., Ahmad, I., Abdullah, I., Thomas, R., Dufresne, A., Thomas, S.,
Hassan, A., 2015. Functionalized liquid natural rubber and liquid epoxidized
to a high crosslinking density, and the storage modulus increased natural rubber: a promising green toughening agent for polyester. J. Appl.
with CNC addition. The Tg shifted to a lower temperature for the Polym. Sci., http://dx.doi.org/10.1002/app.41292.
LENR–UPR blends; however, it increased with the incorporation of Lu, J., Wang, T., Drzal, L.T., 2008. Preparation and properties of microfibrillated
cellulose polyvinyl alcohol composite materials. Compos. Part A: Appl. Sci.
the CNCs. A better balance of material properties was observed for
Manuf. 39, 738–746.
the nanocomposites, i.e., a higher modulus, Tg , and impact energy. Mandal, A., Chakrabarty, D., 2014. Studies on the mechanical, thermal,
Moreover, the effects of the CNCs were enhanced in the nanocom- morphological and barrier properties of nanocomposites based on poly(vinyl
posites having higher LENR content. Finally, as the LENR and CNCs alcohol) and nanocellulose from sugarcane bagasse. J. Ind. Eng. Chem. 20,
462–473.
are green, natural materials, they are good candidates to replace Mathew, V.S., Sinturel, C., George, S.C., Thomas, S., 2010. Epoxy resin/liquid natural
synthetic liquid rubbers and inorganic fillers to toughen UPR while rubber system: secondary phase separation and its impact on mechanical
keeping desirable mechanical properties. properties. J. Mater. Sci. 45, 1769–1781.
Raquez, J.M., Murena, Y., Goffin, A.L., Habibi, Y., Ruelle, B., Debuyl, F., Dubois, P.,
2012. Surface modification of cellulose nanowhiskers and their use as
Acknowledgments nanoreinforcers into polylactide: a sustainably-integrated approach. Compos.
Sci. Technol. 72, 544–549.
Ratna, D., 2004. Rubber toughened epoxy. Macromol. Res. 12, 11–21.
The authors acknowledge financial support from the Saadati, P., Baharvand, H., Rahimi, A., Morshedian, J., 2005. Effect of modified liquid
ministry of Higher Education (MOHE) under FRGS grant rubber on increasing toughness of epoxy resins. Iran. Polym. J. 14,
FRGS/2/2013/SG06/UKM/02/2 and the Ministry of Science, 637–646.
Seng, Y.L., Ahmad, S.H., Rasid, R., Noum, S.Y., Hock, Y.C., Tarawneh, M.A., 2011.
Technology, and Innovation (MOSTI) under the ScienceFund grant Effect of liquid natural rubber (LNR) on the mechanical properties of LNR
03-01-02-SF1106. toughened epoxy composite. Sains Malaysiana 40, 679–683.
Tang, L., Weder, C., 2010. Cellulose whisker/epoxy resin nanocomposites. ACS Appl.
Mater. Interfaces 2, 1073–1080.
References Tang, L.C., Zhang, H., Sprenger, S., Ye, L., Zhang, Z., 2012. Fracture mechanisms of
epoxy-based ternary composites filled with rigid–soft particles. Compos. Sci.
Abdullah, I., 1994. Liquid natural rubber: preparation and application. In: Ghiggino, Technol. 72, 558–565.
K. (Ed.), Progress in Pacific Polymer Science 3. Springer, Berlin, Heidelberg, pp. Thomas, R., Abraham, J., Thomas, P.S., Thomas, S., 2004. Influence of
351–365. carboxyl-terminated (butadiene-co-acrylonitrile) loading on the mechanical
Abdullah, I., Zakaria, Z., 1989. Pendepolimeran fotokimia getah asli. Sains and thermal properties of cured epoxy blends. J. Poly. Sci. 42,
Malaysiana 18, 99–109. 2531–2544.
Afina, M.R.A., Najmi, B.N., Shakirah, S.S., Surip, S.N., 2013. Mechanical and thermal Thomas, R., Ahmad, I., Ahmad, S.H., Koshy, S., 2013. Blend and IPNs of natural
properties of rubber toughened carbon black-filled polyester composite. Adv. rubber with thermosetting polymers. In: Thomas, S., Chan, C.H., Pothen, L.A.,
Mater. Res. 812, 163–168. Rajisha, K.R., Maria, H.J. (Eds.), Natural rubber materials: Volume 1: Blend and
Ahmad, I., Hassan, F.M., 2010. Preparation of unsaturated polyester liquid natural IPNs. RCS Publishing, Cambridge, UK, pp. 336–348.
rubber reinforced by montmorillonite. J. Reinf. Plast. Compos. 29, 2834–2841. Thomas, R., Boudenne, A., Ibos, L., Candau, Y., Thomas, S., 2010. Thermophysical
Azizi Samir, M.A.S., Alloin, F., Dufresne, A., 2005. Review of recent research into properties of CTBN and HTPB liquid rubber modified epoxy blends. J. Appl.
cellulose whiskers, their properties and their application in nanocomposite Poly. Sci. 116, 3232–3241.
field. Biomacromolecules 6, 612–626. Vijayan, P.P., Puglia, D., Jyotishkumar, P., Kenny, J.M., Thomas, S., 2012. Effect of
Ben Saleh, A.B., Mohd Ishak, Z.A., Hashim, A.S., Kamil, W.A., 2009. Compatibility, nanoclay and carboxyl-terminated (butadiene-co-acrylonitrile) (CTBN) rubber
mechanical, thermal, and morphological properties of epoxy resin modified on the reaction induced phase separation and cure kinetics of an epoxy/cyclic
with carbonyl-terminated butadiene acrylonitrile copolymer liquid rubber. J. anhydride system. J. Mater. Sci. 47, 5241–5253.
Phys. Sci. 20, 1–12. Xu, S.A., Wang, G.T., Mai, Y.W., 2013. Effect of hybridization of liquid rubber and
Cherian, A.B., Thachil, E.T., 2004. Modification of isophthalic unsaturated polyester nanosilica particles on the morphology, mechanical properties, and fracture
resin using elastomers bearing reactive functional groups. Prog. Rubber Plast. toughness of epoxy composites. J. Mater. Sci. 48, 3546–3556.
Recycl. Technol. 20, 247–266.

You might also like