You are on page 1of 14

Composites Science and Technology 62 (2002) 13271340

www.elsevier.com/locate/compscitech

Tensile performance improvement of low nanoparticles


lled-polypropylene composites
Chun Lei Wua, Ming Qiu Zhangb,*, Min Zhi Rongb, Klaus Friedrichc
a

Key Laboratory for Polymeric Composite and Functional Materials of Ministry of Education, Zhongshan University,
Guangzhou 510275, PR China
b
Materials Science Institute, Zhongshan University, Guangzhou 510275, PR China
c
Institute for Composite Materials (IVW), University of Kaiserslautern, D-67663 Kaiserslautern, Germany
Received 13 November 2001; received in revised form 19 March 2002; accepted 5 April 2002

Abstract
It was found beforehand that low nanoparticles loaded polymer composites with improved mechanical performance can be prepared by conventional compounding technique in which the nanoparticles are pre-grafted by some polymers using irradiation. To
examine the applicability of the approach, a tougher polypropylene (PP) was compounded with nano-silica by industrial-scale twin
screw extruder and injection molding machine in the present work. The results of tensile tests indicated that the nanoparticles can
simultaneously provide PP with stiening, strengthening and toughening eects at a rather low ller content (typically 0.5% by
volume). The presence of grafting polymers on the nanoparticles improves the tailorability of the composites. Due to the viscoelastic nature of the matrix and the grafting polymers, the tensile performance of the composites lled with untreated and treated
nanoparticles is highly dependent on loading rate. With increasing the crosshead speed for the tensile tests, the dominant failure
mode changed from plastic yielding of the matrix to brittle cleavage.
# 2002 Elsevier Science Ltd. All rights reserved.
Keywords: A. Particle-reinforced composites; B. Mechanical properties; B. Surface treatments; Nanoparticles

1. Introduction
Mineral llers are added to polymers in commercial
production primarily for the reasons of cost reduction
and stiness improvement [1,2]. Although most studies
dealing with the modication of semi-crystalline polymers with inorganic particulates reported embrittling
eects by comparing ultimate elongation and impact
strength of composite materials with those of unlled
resins [35], some researchers showed the enhancement
of toughness in rigid particles lled polypropylene [6,7]
and polyethylene [8,9].
It is worth noting that in the case of micrometer-sized
particulates, high ller content (typically higher than
20% by volume) is generally required to bring the
above-stated positive eects of the llers into play. This
would detrimentally aect some important properties of
the matrix polymers such as processability, appearance,
* Corresponding author. Tel.: +86-20-8403-6576; fax: +86-208403-6564.
E-mail address: ceszmq@zsulink.zsu.edu.cn (M.Q. Zhang).

density and ageing performance. Therefore, a composite


with improved performance and low particle concentration is highly desired. With regard to this, the
newly developed nanocomposites would be competitive
candidates.
The extremely high surface area is one of the most
attractive characteristics of nanoparticles because it
facilitates creating a great amount of interphase in a
composite. Introduction of nanoparticles into a polymer
changes the intermolecular interaction of the matrix
[10]. As estimated by Reynaud et al. [11], an interphase
1 nm thick represents roughly 0.3% of the total volume
of polymer in the case of microparticle lled composites,
whereas it can reach 30% of the total volume in the case
of nanocomposites. That is, the non-negligible contribution made by the interphase provides diverse possibilities of performance tailoring, and is able to
inuence the properties of matrices to a much greater
extent under a rather low nano-ller loading. The crux
of the matter lies in that how to well distribute nanoparticles over a polymer matrix and how to improve
nanoparticles/matrix interaction.

0266-3538/02/$ - see front matter # 2002 Elsevier Science Ltd. All rights reserved.
PII: S0266-3538(02)00079-9

1328

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

From a practical point of view, dispersive mixing in


preparing polymer based particulate composites has
important technical meaning. However, a homogenous
dispersion of nanoparticles in a polymer is very dicult
by using the existing compounding techniques due to
the strong tendency of the ne particles to agglomerate
and the high melt viscosity of the matrix. In many cases,
the so-called nanoparticles lled polymers contain a
number of loosened clusters of nanoparticles, as
demonstrated by the work of Jana and Jain [12] dealing
with untreated nanosilica/polyethersulphone composites.
When the composites are subjected to force, the nanoparticle agglomerates can be split easily and a premature
failure of the materials would thus take place [1215].
To overcome this dilemma and to give full play to
nanoparticles, we used irradiation grafting polymerization method to modify nanoparticles rst, and then
the treated particles were mechanically mixed with a
polymer as usual [16]. Owing to the low molecular
weight nature, the grafting monomers can penetrate
into the agglomerated nanoparticles easily and react
with the activated sites of the nanoparticles inside as
well as outside the agglomerates [17]. As a result, the
following eects can be obtained [18]: (i) Hydrophobicity of the nanoparticles is increased, facilitating
the ller/matrix miscibility; (ii) ller/matrix interaction
is enhanced through the entanglement between the
grafting polymer and the polymer matrix; (iii) the
nanoparticle agglomerates become stronger because
they turn into a nanocomposite microstructure comprising the nanoparticles and the grafted, homopolymerized polymer; (iv) the interfacial characteristics
between the treated nanoparticles and the matrix polymer can be tailored by changing the species of the
grafting monomers and the grafting conditions. In this
context, a uniform dispersion of nanoparticles in the
matrix might no longer be critical.
Mechanical testing of polypropylene lled with nanoSiO2 [18] and nano-CaCO3 [19] demonstrated the feasibility of the above approach. Only a small amount of
modied nanoparticles (typically less than 3% by
volume) can eectively improve modulus, strength,
toughness and thermal deformation temperature of the
matrix polymer. Such an improvement in overall properties of polymers can scarcely be observed in conventional microparticulate composites. It was found that
the deformation habit but not the crystallization characteristics of the matrix polymer remarkably change
with the addition of the treated nanoparticles. To
explain the specic inuence generated by the nanoparticles at low-ller loading regime, a double percolation of stress volumes, characterized by the appearance
of connected shear yielded networks throughout the
composite, was proposed [20].
Considering that the polypropylene used in our previous works [1820] is a brittle type and the composites

were prepared with a lab-scale single screw extruder and


compression molding, the results might not have adequate applicability. Therefore, a commercial polypropylene with higher toughness was compounded with
nano-silica by means of industrial-scale twin screw
extruder and injection molding machine in the current
work. Tensile performance and fractured surfaces are
analyzed as a function of particulate treatment, ller
content and crosshead speed to reveal the structure
property relationships of the composites, and the
mechanical role of the nanoparticles as well.

2. Experimental
2.1. Materials
Isotactic polypropylene (PP) homopolymer T30S1
was supplied by Qilu Petrochemical Industrial Co.,
China. It has a melt ow index of 3.2 g/10 min (2.16 kg
at 230  C). Fumed silica with an average primary particle size of 15 nm and a specic surface area of 374 m2/g
was produced by Shenyang Chemical Engineering Ltd.,
China. Commercial monomers, styrene and ethyl acrylate, were used as grafting monomers without further
purication.
2.2. Irradiation grafting of nano-SiO2
Modication of nano-silica proceeded according to
the following steps. The nanoparticles were pretreated
at 140  C under vacuum for 6 h to eliminate possible
absorbed water on the surface of the particles. Then a
mixture of nanoparticles/monomer (100/20 by weight)
and a certain amount of n-hexane was irradiated by
60
Co g-ray under atmosphere at room temperature.
After exposure to a dose of 4 Mrad, the solvent was
recovered, and the dried residual powder was available
for the subsequent compounding.
2.3. Characterization of the irradiation grafted products
To evaluate the results of grafting and to characterize
the grafted nanoparticles, the grafting polymer and the
homopolymer, which were generated during the irradiation polymerization of the monomers, should be
separated. For this purpose, a certain amount of the
irradiation products were extracted by benzene in a
Soxhlet apparatus for 36 h. In this way the homopolymer was isolated. The residual material was then
dried in vacuum at 80  C until a constant weight was
reached. By using a Shimadzu TA-50 thermogravimetre
(TG) and a Bruker Equinox 55 Fourier transform
infrared spectroscope (FTIR), the weight of the grating
polymer and the chemical structure of the modied
nanoparticles were characterized, respectively. To further

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

separate the grafting polymer from the treated nanoparticles, nano-silica accompanied with the unextractable grafting polymer was immersed in 20% HF
solution for 72 h to remove the inorganic particles. The
molecular weights of the grafting and the homopolymerized polymers were determined by a Waters 991
gel permeation chromatography (GPC), with tetrahydrofuran as the solvent.
To observe the morphologies of the nanoparticles,
untreated SiO2 and grafted SiO2 (without homopolymer) were added into ethanol and toluene to prepare 0.001 g/ml solutions, respectively. With the aid of
sonication for 30 min, the solutions were transferred to
copper gauzes by droppers. After the evaporation of the
solvents, a JEM-100CXZ transmission electron microscopy (TEM) was used to examine the appearance of
the particles.
2.4. Composites preparation and characterization
The nanoparticles were rstly compounded with PP
(1:2 by weight) using an X(S)R-160 two-roll mill at
195  C to produce composite masterbatch. Then, the
masterbatch was mixed with neat PP to dilute the ller
loading to desired values through an SHJN-25 twinscrew extruder at 210230  C. The rotation speed of the
extruder was set to 180 rpm. Finally, the resultant pellets were molded into dog-bone-shaped tensile bars
(ASTM D63897 Type IV specimen) with a CJ150MZ
injection-molding machine at 215  C.
Room temperature tensile testing of the composites
was conducted on a Hounseld-5KN universal testing
machine at crosshead speeds of 10, 30, 50, 100 and 500
mm/min, respectively. Five samples were tested for each
case. The fractured surfaces of the samples were
observed with a Hitachi S-520 scanning electron microscope (SEM) at an accelerating voltage of 20 kV.

3. Results and discussion


3.1. Eect of irradiation grafting polymerization on
nano-SiO2
Since the present work aims to study the eect of
modied nano-silica on the mechanical behavior of PP
composites, variation in the chemical structure of the
particles should be known at the very beginning of the
discussion. FTIR spectra of untreated and treated nanosilica are shown in Fig. 1. To eliminate the inuence of
homopolymers, both polystyrene-grafted nano-SiO2
(SiO2-g-PS) and polyethyl acrylate-grafted nano-SiO2
(SiO2-g-PEA) used for the FTIR examinations were
separated from the homopolymers in advance. In comparison with the spectrum of SiO2 as-received, the
adsorptions at 690, 1460 and 2960 cm 1 appearing in

1329

Fig. 1. FTIR spectra of untreated SiO2 and grafted SiO2.

the spectrum of SiO2-g-PS represent the bending mode


of CH in benzene rings and the stretching modes of
CC and CH, respectively. In addition, the band at
1725 cm 1 in the spectrum of SiO2-g-PEA indicates the
existence of carbonyl groups. These prove that polystyrene (PS) and polyethyl acrylate (PEA) have been
chemically connected to the surface of nano-silica during
the irradiation polymerization processes as expected.
On the basis of above qualitative analysis, the quantitative results of the grafting polymerization on nanosilica are given in Table 1. It is seen that when other
conditions being equal, the percentage grafting and the
grafting eciency of styrene onto the nanoparticles are
much higher than those of ethyl acrylate, while the
monomer conversion of styrene is lower than that of
ethyl acrylate. This reects the dierence in reactive
feature between the monomers. In general, room temperature irradiation grating polymerization onto inorganic particles is controlled by the mechanism of free
radical polymerization. Under the same irradiation
dose, acrylic monomers would generate much more
radicals than styrene, and moreover, ethyl acrylate
radicals have higher activity than styrene radicals [21].
Therefore, the conversion of ethyl acrylate is superior to
that of styrene, leading to higher homopolymer fraction
of PEA than PS. On the other hand, partial surface of
nanosilica might be connected with ethyl acrylate in the
solvent through hydrogen bonding prior to the irradiation processing. When the composite system is exposed
to the irradiation, radicals can be formed on both the
nanoparticles and the ethyl acrylate molecules connected to the particles. The latter would result in either
grafting polymerization or homopolymerization. That
is, the amount of grafted PEA has to be lower as compared with the styrene/silica mixture where no chemical
connection between the monomers and the particles is
established before the irradiation. It should be responsible for the dierence in percentage grafting between
ethyl acrylate/silica and styrene/silica.

1330

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

Table 1
Results of irradiation grafting polymerization onto nano-silica
Monomer

Monomer
conversiona
(%)

Percent
graftingb
(%)

Grafting
eciencyc
(%)

MWd of
grafting
polymer
(104)

MW of
homopolymer
(104)

Styrene
Ethyl acrylate

52.8
93.5

4.64
1.56

43.9
10.5

1.3 (de=1.70)
7.1 (d=2.13)

1.1 (d=1.46)
1.1 (d=1.33)

a
b
c
d
e

Monomer conversion=weight of polymer/weight of monomer.


Percent grafting=weight of grafting polymer/weight of nano-SiO2.
Grafting eciency=weight of grafting polymer/weights of grafting polymer and homopolymer.
MW=weight average molecular weight.
d=Molecular weight polydispersity index.

By further examining the data listed in Table 1, it can


be found that the grafting polymers attached to the
nanoparticles possess higher molecular weights and
broader molecular weight distributions than the homopolymers. Similar phenomena were also reported by
Fukano and Kageyama [22] when they studied radiation
grafting of styrene onto silica-gel. These can be attributed to the characteristics of the grafting reaction.
That is, the reaction was a typical heterogeneous one
because the activated sites were created by irradiation
on the nanoparticles and the chain growth of the grafting polymers had to proceed in solidliquid state. The
mobility of the growing chains was thus worse as compared with a homogeneous reaction in liquid where
both ends of the macromolecular chains can move
freely. As a result, the probability of chain termination
between the radicals became relatively dicult, leading
to higher molecular weight of the grafting polymers. In
addition, because the radicals are not simultaneously
produced on the nanoparticles, the grafting polymers
formed at earlier stage would wrap the surface of the
nanoparticles and reduce the probability of collision
between the monomers and the radicals formed on the
particles at latter stage. This accounts for the higher
molecular weight polydispersity indexes of the grating
polymers.
Morphologies of the nanoparticles before and after
grafting polymerization are illustrated in Fig. 2. A
chain-like branched structure of the agglomerated SiO2
particles [17] can be observed in the solutions. The
smallest perceivable units are approximately 15 nm in
diameter in the case of untreated nanoparticles
[Fig. 2(a)]. When grafting polymers are introduced onto
the particles, the sizes of the agglomerates become larger and the edges are no longer clearly discernible
[Fig. 2(b, c)]. Such a change demonstrates the role of the
grafting polymer, i.e. separating and connecting the
nanoparticles. To estimate the thickness of the polymer
layer adhered to the particles, the data of SiO2-g-PS is
used as an example, i.e. percent grafting=4.64%, density

Fig. 2. TEM microphotos (magnication=105) of (a) SiO2 asreceived, (b) SiO2-g-PS, and (c) SiO2-g-PEA in solvents. In the latter
two specimens, homopolymers were removed in advance.

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

of PS=1.05 g/cm3, specic surface area of the silica=374


m2/g. Supposing a complete coverage on each silica particles, the thickness of the grafting PS should equal to
0.0464/(374 m2/g1.05 g/cm3)0.12 nm. Evidently, this
is a reasonable value as evidenced by the TEM photos
in Fig. 2. Due to the low percent grafting and the thin
grafting polymer layer, there are still many unreacted
hydroxyl groups on the surface of nano-SiO2, which is
responsible for the appearance of the larger agglomerates
of the grafted particles in the solvent.
3.2. Tensile properties of the composites
Fig. 3 shows the results of tensile testing of PP reinforced by nano-silica as a function of ller content,
determined at a moderate crosshead speed of 50 mm/
min. Although both the treated and the untreated
nanoparticles can impart the high stiness of the llers
to the matrix polymer as expected, the composites
incorporated with the modied particles exhibit lower
modulus over the whole range of ller loading of interests [Fig. 3(a)]. Usually the capability of composite
interface to transfer elastic deformation depends to a
great extent upon the interfacial stiness and static
adhesion strength [23,24]. A high interfacial stiness
corresponds to a high composite modulus. Since the
grafting polymers and the homopolymers introduced
onto the nanoparticles form a relatively compliant
interlayer at the particles/matrix interface, the high

1331

stiness of the particles has to be masked under the low


stress level [9] and the composites have to show lower
modulus as compared with the case of untreated SiO2
composites. With a rise in ller content, the increased
amount of grafting polymers further increases the modulus mismatching of the ller and the matrix, and reduces the stiening eect of SiO2, leading to the drop in
Youngs moduli of the composites at high ller regime.
In addition, due to the higher rigidity of PS molecules
than PEA, the interfacial elastic stress is less eciently
transferred in SiO2-g-PEA/PP composites than in SiO2g-PS/PP, especially when the fraction of the modied
nanoparticles is high.
In contrast, an approximately linear composition
dependence of tensile modulus is perceived in the composites lled with SiO2 as-received [Fig. 3(a)]. This can
be interpreted as the absence of a soft interphase and
the appearance of larger agglomerates of the nanoparticles in the matrix as a result of poor ller/matrix
miscibility. Under the same ller volume fraction [25],
the latter eect would provide higher load carrying
ability within small strain range. The above results factually reect the contradiction between the stiening
eect of the rigid particles and the weakening eect of
the soft interlayer.
Fig. 3(b) gives the tensile strengths of the composites
versus SiO2 content. Clearly, both untreated and treated
nanoparticles exhibit the strengthening ability. It is well
known that the tensile strength of a particulate composite

Fig. 3. Tensile properties of PP composites as a function of nano-SiO2 volume fraction: (a) Youngs modulus, (b) tensile strength, (c) elongation to
break, and (d) area under stressstrain curve. Crosshead speed=50 mm/min.

1332

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

Fig. 4. Tensile properties of PP composites as a function of crosshead speed: (a) Youngs modulus, (b) tensile strength, (c) elongation to break, and
(d) area under stressstrain curve.

Fig. 5. Youngs modulus of PP composites as a function of nano-SiO2 volume fraction and crosshead speed.

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

is usually reduced with ller content following a power


law in the case of a poor ller/matrix bonding [26,27].
That means, the strength of the composite cannot be
greater than that of the unlled version because the ller
particles do not bear any fraction of the external load.
This contradicts the results shown in Fig. 3(b). In fact,
when bonding between llers and matrix is strong
enough, the tensile yield strength of a particulate composite can be higher than that of the matrix polymer
[28,29]. Although these models were developed based on
the cases of microsized particulate composites, they are
still valid for the explanation of the composites lled
with nanoparticles [20]. Therefore, the extremely eective improvement of tensile strength of the composites
with grafted nano-SiO2 should result from chain interdiusion and entanglement between the macromolecules of the grafting polymers and the matrix. It is
worth noting that when the amount of the grafted
nanoparticles is increased (e.g. > 0.5 vol%), the content
of compliant PEA chains is raised accordingly in the
interlayer and the interfacial stress transfer eciency
has to be decreased. This accounts for the relatively low
strength of SiO2-g-PEA/PP composites at higher particulate loading.

1333

By comparing Fig. 3(b) with the data of Ref. [18], the


most distinct dierence lies in the results of untreated
SiO2/PP. That is, for the composites prepared by a labscale single screw extruder and compression molding,
the addition of untreated SiO2 lowers the tensile
strength of PP in the lower loading region but then leads
to a slight increase in strength when the particle fraction
reaches 4.68 vol% [18]. With respect to the SiO2/PP
composites of the present work, which were manufactured through twin screw extrusion and injection
molding, a continuous increase in the strength with SiO2
content is detected [Fig. 3(b)]. So far as we know, corotating twin screw extruders are able to provide more
sucient homogenization in comparison with single
screw extruders. The above dierence manifests that the
important role of even distribution of untreated nanoparticles in the composites. For SiO2 as-received, the
more particles are exposed to the matrix polymer, the
more possibly the interaction between the particles and
the matrix can be enhanced. So, an improved homogeneity of the untreated SiO2/PP composites would certainly be benecial to the stress transfer.
This again shows the signicance of grafting modication of the nanoparticles, which reduces the sensitivity

Fig. 6. Tensile strength of PP composites as a function of nano-SiO2 volume fraction and crosshead speed.

1334

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

of composites strength performance to the dispersion


state of the particles. As the grafted nanoparticle
agglomerates turn into a nonocomposite microstructure
consisting of the particles and the grafted, homopolymerized secondary polymer [18], they are brought
into play as an integral when the composites are subjected to the applied force [20]. The situation is completely dierent from the composites with untreated
nanoparticles, in which agglomerated particles have to
be deagglomerated as much as possible to reduce the
probability of premature failure. As a result, a uniform
dispersion of the nanoparticles in the matrix is absolutely necessary for obtaining the reinforcing eect in
case untreated nanoparticles are used.
Failure strain can partially assess the rupture behavior of a composite material. The incorporation of particulate llers usually results in a decrease in this
parameter regardless of the interfacial adhesion [26]. It
is true even in the system exhibiting impressive impact
toughness improvement with the addition of mineral
llers [8]. However, the plots shown in Fig. 3(c)
demonstrate that the values of elongation-to-break of

PP can be signicantly increased by using nano-SiO2,,


implying a failure mechanism dierent from those
involved in conventional composites. Comparatively,
SiO2-g-PEA is able to provide a stable improvement
over the entire ller content range of interests. The
reduction in elongation-to-break of the composite lled
with untreated SiO2 suggests that the llers cause a
reduction in matrix deformation due to an introduction
of mechanical restrains. In contrast, the improvement of
elongation-to-break with the incorporation of the grafted nanoparticles is a result of interfacial viscoelastic
deformation and matrix yielding. Evidently, the grafted
PEA makes more eective contribution.
The area under the tensile stressstrain curve can
more reasonably characterize the toughness potential of
the composites than elongation-to-break under static
tensile loading conditions [30]. As conrmed by
Fig. 3(d), the grafted nanoparticles indeed improve the
ductility of PP at a silica content as low as 0.5 vol%. The
results are somewhat opposite to those observed in Ref.
[18], which reports a deteriorated eect of grafting PEA
on the tensile behavior of the composites. Considering

Fig. 7. Elongation to break of PP composites as a function of nano-SiO2 volume fraction and crosshead speed.

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

that SiO2-g-PEA employed in the present work possesses almost the same percent grafting and grafting
eciency as that in Ref. [18], it can be concluded that
the molecular entanglement between the grafting PEA
and matrix PP was not suciently formed during melt
compounding when the single screw extruder was used.
As a result, localized plastic deformation or matrix
drawing cannot be eciently induced by SiO2-g-PEA as
in the composites prepared by a twin screw extruder.
In the case of untreated nano-SiO2, the areas under
the tensile stress-strain curves of the composites at
higher ller content regime are lower than that of the
neat PP. It implies that the short range interaction at
SiO2/PP interface is not good at inducing plastic deformation of the matrix polymer. Due to the encounter of
the propagating neck with a larger agglomerate, particularly in the case of higher ller content, nal failure of
the composites might be initiated easily [8]. Obviously,
the resultant embrittling eect can be prevented by the
application of grafted nanoparticles.
Since polymer composites maintain the viscoelasticity of
polymers, the dependence of tensile properties on crosshead speed should be known for engineering purposes. As

1335

can be seen from Fig. 4(a), a linear increase in Youngs


modulus with increasing crosshead speed is valid for all
the specimens. The presence of silica leads to the values
of modulus being higher than that for the unlled PP
and being slightly less speed dependent. These observations are exactly as expected.
The values of tensile strength of the materials are
plotted as a function of crosshead speed in Fig. 4(b).
With a rise in crosshead speed, although the strengths
increase in principle, there is a signicant transition in
the slope. Accordingly, the rst derivation of the
strength in Fig. 4(b) with respect to crosshead speed
would yield a peak between 50 and 100 mm/min, which
clearly corresponds to an energy-activated process in
tensile fracture of the materials. Further eorts should
be made to understand whether the Eyrings theory that
considers the eect of an applied stress is to reduce the
height of a potential energy barrier [31,32] is applicable
to the present systems.
Both elongation-to-break and area under the tensile
stress-strain curve have similar dependence on crosshead
speed [Fig. 4(c) and (d)]. With a rise in the speed, a
drastic decrease of the two parameters is followed by a

Fig. 8. Area under stressstrain curve of PP composites as a function of nano-SiO2 volume fraction and crosshead speed.

1336

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

gradual reduction. SiO2-g-PEA/PP is able to keep its


ductility superior to other systems when crosshead
speed is slower than 100 mm/min. As compared with the
plots proles in Fig. 4(b), it can be deduced that dierent failure mechanisms take eects when crosshead
speed is faster or slower than 100 mm/min. Vu-Khanh
and Denault found that the dynamic fracture toughness
of glass-ake/PP composites sharply decreases with
impact speed [33]. It was speculated that due to the low
thermal conductivity of the composite, the relaxation
process in the matrix led to an increase in temperature
at the crack tip. The temperature rise caused a decrease
in the fracture toughness with loading rate. It seems
their analysis can also explain the experimental data
shown in Fig. 4(c) and (d).
To have a more comprehensive understanding of the
interdependence of tensile properties of the composites
on ller content and crosshead speed, three-dimensional
diagrams are drawn in Figs. 58. In the case of a crosshead speed of 10 mm/min, Youngs moduli of SiO2-g-PS/
PP and SiO2-g-PEA/PP composites increase with ller
content and then decrease (Fig. 5). When crosshead
speed is raised, the aforesaid decreasing trend of moduli
is gradually replaced by a slight increase or a plateau.

This is dierent from the performance of SiO2/PP,


which exhibits a continuous increase in the stiness with
silica content at each crosshead speed investigated. Evidently, the viscoelastic nature of the interphase due to
the appearance of grating polymers in SiO2-g-PS/PP
and SiO2-g-PEA/PP composites should be responsible
for the distinct behavior.
The most obvious characteristics of Fig. 6 is that the
strengths measured at 100 and 500 mm/min are much
higher than those obtained at a slower crosshead speed.
As suggested previously, it should be indicative of a
change in failure modes due to the dierent viscoelastic
responses as found in conventional polymers. For SiO2g-PEA/PP composites, the addition of the grafted
nanoparticles used to slightly decrease the strength of
PP at crosshead speeds of 100 and 500 mm/min
[Fig. 6(c)]. It means that the grafting PEA molecules
become less ecient to transfer stress under high strain
rate. In contrast, 0.5 vol% of SiO2-g-PS can still provide
the reinforcing eect at 500 mm/min [Fig. 6(b)].
Figs. 7 and 8 illustrate the elongation-to-break and
the area under tensile stressstrain curve of the composites as a function of silica fraction and crosshead speed.
The untreated SiO2/PP composites exhibit performance

Fig. 9. SEM graphs of tensile fractured surface of: (a) neat PP; (b) and (c) SiO2/PP (SiO2 content=0.86 vol%); (d) SiO2-g-PS/PP (SiO2 content=1.06 vol%); (e) SiO2-g-PEA/PP (SiO2 content=0.82 vol%). Crosshead speed=50 mm/min.

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

dierent from the grafted SiO2 composites especially in the


case of higher particle content. At a silica volume fraction
of about 2.7 vol%, for example, the elongation-to-break
and the area under tensile stress-strain curve of untreated
SiO2/PP are rather small and nearly independent of crosshead speed in comparison with the treated SiO2/PP
composites. This should be interpreted as that splitting
of the large nanoparticle agglomerates accumulated in

1337

the matrix due to the fact that the increased particle


content governs the failure process.
By examining the curves of SiO2-g-PS/PP and SiO2-gPEA/PP determined at dierent crosshead speeds
(Figs. 7 and 8), it can be found that the toughening
eect exerted by the modied particles becomes
remarkable only at moderate speeds. The improved
elongation-to-break and area under tensile stressstrain

Fig. 10. SEM graphs of tensile fractured surface of: (a) and (b) SiO2/PP (SiO2 content=2.74 vol%); (c) and (d) SiO2-g-PS/PP (SiO2 content=2.75
vol%); (e), (f) and (g) SiO2-g-PEA/PP (SiO2 content=2.75 vol%). (g) was taken from the side face of the specimen within the stress-whitened neck
zone. Crosshead speed=50 mm/min.

1338

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

curve represent the improved deformation ability of the


composites in relation to plastic stretching of the matrix
polymer induced by the grafted nanoparticles. In addition, the entanglement between the grafting polymers
and the matrix polymer is also viscoelastic in nature.
3.3. Microscopic observation of fractured surface
To have clear images of the failure patterns of the
composites under tension, SEM fractographs of the
specimens with dierent ller contents tested at dierent
crosshead speeds are discussed hereinafter. Fig. 9 shows
the tensile fractured surfaces of neat PP and the composites with relatively low silica fraction generated at a
crosshead speed of 50 mm/min. The unlled PP has a
relatively smooth fractured surface in association with
terraced markings [Fig. 9(a)], indicating weak resistance
to crack propagation. In the case of SiO2/PP (Vf=0.86
vol%), the fractured surface becomes rougher but the
traces of plastic deformation are still less [Fig. 9(b)].
Many silica agglomerates (41 mm in size) are dispersed
in the matrix without clear signs of stretching of the
surrounding matrix. Occasionally, elongated matrix
polymer can be found around silica agglomerates inside
a large cavity on the composites surface, as highlighted
by an arrow in Fig. 9(c). These demonstrate not only the
insuciently interfacial interaction between the particles
and the matrix, but also the poor toughening capability
of the composites. For SiO2-g-PS/PP and SiO2-g-PEA/
PP composites, the fractured surfaces are full of extensive matrix brils [Fig. 9(d) and (e)]. Therefore, it can be
evidenced that the grafting polymers on the nanoparticles enhance the interfacial interaction and the dissipation of energy through matrix stretching.
When the content of nanosilica approaches around 2.7
vol%, the morphologies of the composites fractured surfaces become somewhat dierent (Fig. 10). A number of
cavities appear on the surface of SiO2/PP composites
[Fig. 10(a)]. In fact, they are produced due to the debonding of the untreated particles, as illustrated by a magnied
view [Fig. 10(b)]. In general, an increased content of
untreated SiO2 would lead to larger agglomerates and
hence greater probability of debonding due to the poor
interfacial adhesion. As there is not enough time for inducing matrix yielding after the extensive particles debonding, the matrix beside the cavities seems to be rather at
[Fig. 10(a)]. This coincides with the reduction of toughness
of SiO2/PP at high SiO2 loading [Fig. 3(c) and (d)].
When SiO2-g-PS is incorporated [Fig. 10(c)], concentric matrix-brillated circles around nanoparticle
agglomerates (as indicated by the upper arrow) and
voids left as a result of agglomerated particles detachment (as indicated by the lower arrow) can be found on
the fractured surface. As suggested in Ref. [18], the
appearance of the brillated matrix circles are probably
the result of a successive debonding of the modied

nanoparticles from the matrix accompanied by an


unconstrained plastic stretching of the interparticulate
matrix ligaments [Fig. 10(d)]. Such a deformation process would certainly consume more energy than that
dominated only by debonding as shown in Fig. 10(a)
and (b). In the case of SiO2-g-PEA/PP, the concentric
brillar circles around nanoparticle agglomerates
emerge next to each other [Fig. 10(e)]. Besides, the
matrix surrounding the agglomerated nanoparticles has
turn into plastically drawn brils [indicated by the
arrow in Fig. 10(f)]. Evidently, the inherent exibility of
PEA has made important contribution. In accordance
with the model describing the double percolation of yielded zones [20], these should result from the superposition
of stress volumes around the agglomerates and the nanoparticles. It explains the cause for that SiO2-g-PEA/PP is
still able to maintain higher static ductility at a relatively
high nanoparticle concentration [Fig. 3(c) and (d)].
Fig. 10(g) exhibits the SEM observation result of the side
surface of the SiO2-g-PEA/PP tensile specimen. Elongated cavities can be seen around partially debonded
nanoparticle agglomerates (indicated by the arrows).
Again, plastic ow of the bridging matrix is clear visible.
On the other hand, Fig. 10(c), (e), (f) and (g) conrm the
estimation that the grafted nanoparticle agglomerates
have turned into a nonocomposite microstructure and

Fig. 11. SEM graphs of tensile fractured surface of: (a) neat PP;
(b) SiO2/PP (SiO2 content=0.86 vol%); (c) SiO2-g-PS/PP (SiO2 content=1.06 vol%); (d) SiO2-g-PEA/PP (SiO2 content=0.82 vol%).
Crosshead speed=10 mm/min.

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

1339

By summarizing the data shown in Fig. 4 and the


fracture morphology of Figs. 11 and 12, it can be
known that the tensile performance of the composites is
close to that of neat matrix polymer when the specimens
are tested under low or high tensile speeds. Neither the
untreated nor the treated particles can take eects under
these circumstances. This is particularly true in the case
of low silica fraction.

4. Conclusions

Fig. 12. SEM graphs of tensile fractured surface of: (a) neat PP;
(b) SiO2/PP (SiO2 content=0.86 vol%); (c) SiO2-g-PS/PP (SiO2 content=1.06 vol%); (d) SiO2-g-PEA/PP (SiO2 content=0.82 vol%).
Crosshead speed=500 mm/min.

are brought into play as an integral. In one sense, the


grafted nanoparticle agglomerates can be taken as the
folded polymer chains conguration because they are
provided with the capability of deforming and releasing
locally concentrated stress instead of simply splitting.
Fig. 11 gives the fractographs of neat PP and the
composites with low SiO2 content tested at a crosshead
speed of 10 mm/min. They all have similar appearances
characterized by ductile failure except some ne microbrils on the surface of SiO2-g-PS/PP and SiO2-g-PEA/
PP. In comparison with the images taken at 50 mm/min
(Fig. 9), no nanoparticles agglomerates are observed
probably because of the shields of the highly elongated
matrix and the low ller content as well.
It is worth noting that the fracture modes of the same
materials can be changed when a higher testing speed is
applied (Fig. 12). The neat PP shows cleavage fracture
feature under the crosshead speed of 500 mm/min
[Fig. 12(a)]. The striation structure resulting from the
joining of dierent fractured planes on the surface of
SiO2/PP composites demonstrates that the particles
have little resistance to the crack propagation
[Fig. 12(b)]. Similarly, the grafted nanoparticles
agglomerates cannot induce eective matrix yielding on a
large scale [Fig. 12(c) and (d)]. The mild proles of the
deformation circles [as indicated by the arrows in
Fig. 12(c) and (d)] suggest a low plastic deformation level.

Based on the above results and discussions, the following statements can be drawn.
(1) The addition of nanoparticles into PP can bring in
both reinforcing and toughening eects at ller content
as low as 0.5 vol%. Such a simultaneous improvement
in modulus, strength and elongation-to-break is hard to
be observed in conventional microsized particulate
composites.
(2) Modication of nanosilica by means of grafting
polymerization helps to provide the composites with
balanced performance. In addition, dierent species of
the grafting monomers result in dierent interfacial
interactions and dierent ultimate properties of the
composites.
(3) With respect to the manufacturing aspect, dispersion homogeneity of the composites lled with
untreated nanoparticles is critical, while it is not necessarily realized in the case of grafted nanoparticles.
(4) As compared with single screw extruder, twin
screw extruder can further decrease the amount of the
nanoparticles needed for the composites performance
enhancement.
(5) The relative increment of the areas under the tensile stress-strain curves of the current composites is
similar to the values reported by Ref. [18], although
there is a signicant dierence in ductility between the
PP used in the two works. This somewhat contradicts
the results of Ren and co-workers, who found that a
tougher PP would gain more remarkable improvement
of fracture toughness with the addition of nanoparticles
[34]. It means that continuous eort should be paid to
understand the role of the matrix toughness in nanoparticles composites.
(6) Owing to the viscoelastic nature of the grafting polymers, the inuence of the modied nanoparticles on the
tensile properties of PP is also a function of loading speed.

Acknowledgements
The authors are grateful to the support of the
Deutsche Forschungsgemeinschaft (DFG FR675/401)
for the cooperation between the German and Chinese
institutes on the topic of nanocomposites. Further thanks

1340

C.L. Wu et al. / Composites Science and Technology 62 (2002) 13271340

are due to the National Natural Science Foundation of


China (Grant: 50133020), the Key Program of the Ministry of Education of China (Grant: 99198), the Team Project of the Natural Science Foundation of Guangdong,
China, the Natural Science Foundation of Guangdong,
China (Grant: 990277), and the Key Program of the
Science and Technology Department of Guangdong,
China (Grant: A10172).

References
[1] Rothon RN. Mineral llers in thermoplastics: ller manufacture
and characterization. Adv Polym Sci 1999;139:67107.
[2] Pukanszky B. Particulate lled polypropylene composites. In:
Karger-Kocsis J, editor. Polypropylene: an A-Z reference.
Kluwer Academic; 1999. p. 57480.
[3] Jancar J, Dibenedetto AT, Dianselmo A. Eect of adhesion on
the fracture toughness of calcium carbonate-lled polypropylene.
Polym Eng Sci 1993;33:55963.
[4] Tjiong SC, Li RKY, Cheung T. Mechanical behavior of CaCO3
particulate-lled b-crystalline phase polypropylene composites.
Polym Eng Sci 1997;37:16677.
[5] Dubnikova IL, Oshmyan VG, Gorenberg AY. Mechanisms of
particulate lled polypropylene nite plastic deformation and
fracture. J Mater Sci 1997;32:161322.
[6] Premphet K, Horanont P. Phase structure of ternary polypropylene/elastomer/ller composites: eects of elastomer polarity. Polymer 2000;41:928390.
[7] Kolarik J, Jancar J. Ternary composites of polypropylene/elastomer/calcium carbonate: eect of functionalized components on
phase structure and mechanical properties. Polymer 1992;33:49617.
[8] Bartczak Z, Argon AS, Cohen RE, Weinberg M. Toughness
mechanism in semi-crystalline polymer blends: II. High-density
polyethylene toughened with calcium carbonate ller particles.
Polymer 1999;40:234765.
[9] Walter R, Friedrich K, Privalko V, Savadori A. On modulus and
fracture toughness of rigid particulate lled high density polyethylene. J Adhesion 1997;64:87109.
[10] Jurga J, Nowicki M, Bula K, Susla B, Rejeibi SS. Eect of heat
treatment on phase behavior and molecular dynamics of minerallled PPS. Mol Cryst Liq Cryst 2000;354:438.
[11] Reynaud E, Gauthier C, Perez J. Nanophases in polymers. Rev
Metallurgie 1999;96(2):16976.
[12] Jana SC, Jain S. Dispersion of nanollers in high performance
polymers using reactive solvents as processing aids. Polymer
2001;42:6897905.
[13] Li JX, Silverstein M, Hiltner A, Baer E. The ductile-to-quasibrittle transition of particulate-lled thermoplastic polyester.
J Appl Polym Sci 1994;52:25567.
[14] Wang Y, Huang JS. Single screw extrusion compounding of particulate lled thermoplastics: state of dispersion and its inuence
on impact properties. J Appl Polym Sci 1996;60:177991.

[15] Sumita M, Tsukumo Y, Miyasaka K, Ishikawa K. Tensile yield


stress of polypropylene composites lled with ultrane particles.
J Mater Sci 1983;18:175864.
[16] Rong MZ, Zhang MQ, Zheng YX, Zeng HM. Chinese patent
application no.: CN99116017, 1999.
[17] Zhang MQ, Rong MZ, Zeng HM, Schmitt S, Wetzel B, Friedrich
K. An atomic force microscopy study on structure and properties
of irradiation grafted silica particles in polypropylene based
nanocomposites. J Appl Polym Sci 2001;80:221827.
[18] Rong MZ, Zhang MQ, Zheng YX, Zeng HM, Walter R, Friedrich K. Structure-property relationships of irradiation grafted
nano-inorganic particle lled polypropylene composites. Polymer
2001;42:16783.
[19] Rong MZ, Zhang MQ, Zheng YX, Zeng HM, Walter R, Friedrich K. Irradiation graft polymerization on nano-inorganic
particles: an eective means to design polymer based nanocomposites. J Mater Sci Lett 2000;19:115961.
[20] Rong MZ, Zhang MQ, Zheng YX, Zeng HM, Friedrich K.
Improvement of tensile properties of nano-SiO2/PP composites in
relation to percolation mechanism. Polymer 2001;42:33014.
[21] Ma RD. Techniques of radiation processing. Sichuan Science and
Technology Press; 1984. p. 3267 [in Chinese].
[22] Fukano K, Kageyama E. Study on radiation-induced polymerization of vinyl monomers adsorbed on inorganic substances I.
Radiation-induced polymerization of styrene adsorbed on several
inorganic substances. J Polym Sci Polym Chem 1975;13:130924.
[23] Kerner EH. Electrical conductivity of composites media. Proc
Phys Soc 1956;B69:8027.
[24] Zhang M, Zeng H, Zhang L, Lin G, Li RKY. Fracture characteristics of discontinuous carbon bre-reinforced PPS and PESC composites. Polym Polym Compos 1993;1:35765.
[25] Ahmed A, Jones FR. A review of particulate reinforcement theories for polymer composites. J Mater Sci 1990;25:493342.
[26] Nielsen LE. Simple theory of stress-strain properties of lled
polymer. J Appl Polym Sci 1966;10:97103.
[27] Nicolais L, Narkis M. Stress-strain behavior of styrene-acrylonitrile/glass bead composites in the glassy region. Polym Eng Sci
1971;11:1949.
[28] Jancar J, Dianselmo A, Dibenedetto AT. The yield strength of
particulate reinforced thermoplastic composites. Polym Eng Sci
1992;32:13949.
[29] Fekete E, Molnar SZ, Kim GM, Michler GH, Pukanszky B.
Aggregation, fracture initiation, and strength of PP/CaCO3
composites. J Macromol Sci Phys 1999;B38:88599.
[30] Friedrich K. Microstructural eciency and fracture toughness of
short ber/thermoplastic matrix composites. Compos Sci Technol
1985;22:4374.
[31] Eyring H. Viscosity, plasticity and diusion as examples of
absolute reaction rates. J Chem Phys 1936;4:28395.
[32] Kinloch AJ, Young RJ. Fracture behaviour of polymers. Applied
Science 1983.
[33] Vu-Khanh T, Denault J. Toughness of reinforced ductile
thermoplastics. J Compos Mater 1992;26:226277.
[34] Ren X, Bai L, Wang G. Reinforcement and toughening of polypropylene composites by nanoparticle CaCO3. Chem World
2000;41(2):837 (in Chinese).

You might also like