Vous êtes sur la page 1sur 24

Applied Catalysis A: General 475 (2014) 386409

Contents lists available at ScienceDirect

Applied Catalysis A: General


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

Review

A review of clay-supported ZieglerNatta catalysts for production of


polyolen/clay nanocomposites through in situ polymerization
Sadegh Abedi, Majid Abdouss
Chemistry Department, Amirkabir University of Technology, Hafez Ave, Tehran, Iran

a r t i c l e

i n f o

Article history:
Received 7 November 2013
Received in revised form 5 January 2014
Accepted 10 January 2014
Available online 21 January 2014
Keywords:
Clay
ZieglerNatta catalyst
Nanocomposites
Polyolen

a b s t r a c t
Polyolen/clay nanocomposites are mainly prepared by melt mixing, solution, and in situ polymerization
approaches. Among them, the last one is more attractive because the dispersion obtained through in situ
polymerization is the most efcient, particularly, in the fully exfoliated polyolen/clay nanocomposite
formation. The clay-supported coordination catalysts, i.e. ZieglerNatta, metallocene, and late transition
metal catalysts, were applied to olens polymerization in order to produce the polyolen/clay nanocomposites. This review presents the studies reported on the use of the clay-supported ZieglerNatta catalysts
for polymerizing the olens, which provided that detailed examples reported in the open literature. The
emphasis is placed on the production of the polyolen/clay nanocomposites via in situ polymerization.
2014 Elsevier B.V. All rights reserved.

Contents
1.
2.

3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Generalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
Structure and chemistry of layered silicates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Types of nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.1.
Phase separated microcomposite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.2.
Intercalated nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.3.
Exfoliated nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
Preparation of nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.1.
Exfoliation-adsorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.2.
Melt intercalation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.3.
Template synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.4.
In situ intercalative polymerization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
In situ polymerization of olens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.1.
Metallocene catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.
Late transition metal catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.
ZieglerNatta catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.1.
ZieglerNatta catalysts for production of PE/clay nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.2.
ZieglerNatta catalysts for production of PP/clay nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.3.
Morphology of nanocomposites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

386
387
387
387
388
388
388
388
388
388
389
390
390
390
391
392
392
398
406
407
407

1. Introduction

Corresponding author. Tel.: +98 21 64542765; fax: +98 21 64542762.


E-mail address: phdabdouss44@aut.ac.ir (M. Abdouss).
0926-860X/$ see front matter 2014 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apcata.2014.01.028

Polyolens are the most important commercial thermoplastics


which have the largest tonnage in the polymeric industries, more
than 60% of the total worldwide polymer production, because of
the abundance of the monomers and being easily molded by many

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

different processes. In addition, these polymers enjoy the valuable


physical and mechanical properties which lead to a variety of applications in different industries. However, such products suffer from
some drawbacks including mechanical properties. These properties
can outstandingly be further enhanced by adding the llers and
reinforcements. Many different types of materials including calcium carbonate, talc, carbon black, silica, etc. have been introduced
into the polymers to provide a synergistic improvement for processibility and nal product properties [14].
In the last few years, a great attention has been paid to the
preparation of polyolenic nanocomposites using nanoparticles
due to the dramatic modication in the physical and mechanical properties of polymers. Such properties can be achieved by
adding just a small fraction of nanoparticles including CaCO3
[510], clay [1115], talc [1618], carbon nanotube [1923], carbon black [2428], silica [2935], etc. to the polymer matrix. Totally,
the nanocomposite technology offers superior mechanical, thermal
and barrier properties which are not achievable in the conventional
polymer composites [4,3639].
The synthesis of the polymer/clay nanocomposites can be performed by three methods namely, melt processing, solution and
in situ polymerization [4,3641]. The melt processing of the polyolens with nanollers is often led to insufcient ller dispersion.
It leads to the aggregation and intercalation which decrease the
mechanical properties, especially at high ller contents. The solution method suffers from some drawbacks including the low
solubility of polyolen in the low boiling organic solvents. Another
problem of both above methods is the hydrophilic nature of most
inorganic llers, the hydrophobic nature of the polyolens including polyethylene (PE) and polypropylene (PP). The diverseness
results in the weak interfacial adhesion between the ller and the
polyolen matrix and also the low mechanical properties. Therefore, the llers must be modied by the surface active agents. These
disadvantages can be solved by in situ polymerization [42,43].
Totally, the in situ polymerization is more effective method in
the nanocomposite formation than the melt processing method
according to the comparison of the silicate dispersion data [43,44].
In addition, an in situ polymerization is practiced by introducing
an olen polymerization catalyst into the interlayer galleries of
the silicate layer of the clay where in situ olen monomer polymerization will generate sufcient heat. This is combined with the
physical expansion effect in the continuously growing polyolen
macromolecules to effectively promote the delamination into exfoliation state and homogeneous dispersion of the silicate layers in
the polyolen matrixes.
The in situ polymerization technique is generally believed to
be the way to access to the polyolen/clay nanocomposites due to
the successful bypassing the rigorous thermodynamic requirement
associated with the polymer intercalation process. In addition to an
easy accomplishing a nano-scale dispersion, an in situ polymerization also allows the versatile molecular design of the polyolen
matrix. This paves a route to not only gaining different polyolen
nanocomposites with expanded property range by exible tuning
the matrix composition and structure, but also the rational designs
of the interface between the nanoparticles and the polyolen
matrix. For example, by implanting polar functional groups into
the chain of the polyolens can lead to strong interfacial interaction with the nano-particles in order to enhance their nano-effects
on polyolen properties [45,46]
Metallocenes and ZieglerNatta catalysts supported on the
clay have been widely employed for the in situ polymerization
of ethylene [4754]. The clay-supported metallocenes have been
comprehensively considered in reference [55]. In addition, the claysupported late transition metal catalysts have been also reported in
some papers used for the preparation of polyolen/clay nanocomposites via in situ polymerization [5661].

387

Fig. 1. Structure of 2:1 phyllosilicates [4].

The preparation of the polyolen nanocomposites, using a


typical ZieglerNatta catalyst by in situ polymerization, has the
advantage of using industrial catalysts. It is the most cost-effective
one among the coordination catalysts.
The aim of this article is to provide a comprehensive review
of the clay-supported ZieglerNatta catalysts used to obtain
polyolen-layered silicate nanocomposites. In addition, the catalyst performance in the olen polymerization and also the
properties of the obtained nanocomposites has been considered.
2. Generalities
2.1. Structure and chemistry of layered silicates
The usual layered silicates used in nanocomposites, belong to
the structural category, are known as the 2:1 layered or phyllosilicates (Fig. 1). Their crystal lattice consists of layers made up of two
tetrahedrally coordinated silicon atoms fused to an edge-shared
octahedral sheet of either alumina or magnesia. The layer thickness
is around 1 nm and the lateral dimensions of these layers may vary
from 30 nm to several microns and even larger depending on the
particular silicate. Stacking of the layers leads to a regular Van der
Waals gap between them called the interlayer or gallery. Isomorphic substitution within the layers (for example, Al3+ replaced by
Mg2+ or Fe2+ , or Mg2+ replaced by Li+ ) generates negative charges
which are counterbalanced by alkali or alkaline earth cations situated inside the layers.
Montmorillonite (MMT), hectorite and saponite are the most
commonly used layered silicates. Their chemical structures are
shown in Table 1.
2.2. Types of nanocomposites
Layered silicates have a general thickness in the order of 1 nm
and very high aspect ratios (typically, 101000). A few weight percentages of layered silicates, when fully dispersed in the polymer
Table 1
Chemistry formula and characteristic parameter of phyllosilicatesa [4].
2:1 Phyllosilicates

Chemical formula

Particle length (nm)

Montmorillonite
Hectorite
Saponite

Mx (Al4x Mgx )Si8 O20 (OH)4


Mx (Mg6x Lix )Si8 O20 (OH)4
Mx Mg6 (Si8x Alx )Si8 O20 (OH)4

100150
200300
5060

a
M, monovalent cation; x, degree of isomorphous substitution (between 0.5 and
1.3).

388

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

Fig. 2. Types of polymer/silicate nanocomposites: (a) phase separated microcomposite (b) intercalated nanocomposite (c) exfoliated nanocomposite [4].

matrix, can create much higher surface area for the polymer/ller
interaction compared to the conventional composites.
There are thermodynamically three main types of polymer/layers clay of nanocomposites depending on the strength of
the interfacial interactions between the polymer matrix and layered silicate and also the method of nanocomposite preparation.
Different types of nanocomposites are shown in Fig. 2.
2.2.1. Phase separated microcomposite
A phase separated composite (Fig. 2a) is obtained while the polymer is unable to intercalate the silicate sheets. In this case, the
mechanical properties remain in the same range like the traditional
microcomposites.
2.2.2. Intercalated nanocomposites
Insertion of a polymer matrix into the layered silicate structure
occurs in a crystallographically regular fashion, regardless of the
clay to polymer ratio. The intercalated nanocomposites are normally interlayered by a few molecular layers of the polymer. The
properties of the composites typically resemble those of ceramic
materials (Fig. 2b).
2.2.3. Exfoliated nanocomposites
An exfoliated or delaminated structure is obtained when the
silicate layers are completely and uniformly dispersed in a continuous polymer matrix (Fig. 2c). In an exfoliated nanocomposite, the
individual clay layers are separated in a continuous polymer matrix
by an average distance that depends on the clay loading. Generally,
the clay content of an exfoliated nanocomposite is much lower than
that of an intercalated nanocomposite.
In addition, the silicate layers may collapse on each other.
Conceptually, this is the same as intercalated nanocomposites.
However, the silicate layers are sometimes occulated due to the
hydroxylated edgeedge interaction of the silicate layers.
The nanocomposite structure is usually evaluated through the
transmission electron microscopy (TEM) and X-ray diffraction
(XRD) techniques. In the clay-based nanocomposites, the repetitive multilayer structure is well preserved, allowing the interlayer
spacing to be determined. The intercalation of the polymer chains
usually increases the interlayer spacing, in comparison with the

spacing of the organoclay used (Fig. 3), leading to shift the diffraction peak onto lower angle values.
As far as the exfoliated structure is concerned, no more diffraction peaks are visible in the XRD diffractograms, either because
of very large spacing between the layers (i.e. exceeding 8 nm in
the case of ordered exfoliated structure) or because the nanocomposite does not present any ordering anymore. In the latter case,
transmission electronic spectroscopy is used to characterize the
nanocomposite morphology (Fig. 3). TEM shows not only the exfoliation structure of the clay layers, but also the order of the dispersion
of the clay layers in the polymer matrix.
2.3. Preparation of nanocomposites
The synthesis of polymer/clay nanocomposites can be performed in four methods which are discussed as follows
[4,40,41,37,62].
2.3.1. Exfoliation-adsorption
The layered silicate is exfoliated into single layers using a
solvent in which the polymer (or a prepolymer in case of insoluble polymers such as polyimide) is soluble. This technique has
been widely used with water-soluble polymers to produce intercalated nanocomposites based on poly(vinyl alcohol), poly(ethylene
oxide), poly(vinylpyrrolidone) or poly(acrylic acid). Fig. 4 illustrates
that the stacked (intercalated) and isolated (exfoliated) silicate layers can be observed in the TEM micrograph of the nitrile copolymer
lled with montmorillonite by the exfoliation-adsorption method
[37].
It is well known that such layered silicates, owing to the weak
forces that stack the layers together, can be easily dispersed in an
adequate solvent. The polymer then adsorbs onto the delaminated
sheets and when the solvent is evaporated (or the mixture precipitated), the sheets reassemble, sandwiching the polymer to form, in
the best case, an ordered multilayer structure.
2.3.2. Melt intercalation
The layered silicate is mixed with the polymer matrix in the
molten state. Under this condition, if the layer surfaces are sufciently compatible with the chosen polymer, the polymer can crawl

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

389

Fig. 3. XRD patterns and TEM images of three different types of nanocomposites [4].

along the interlayer space and form either an intercalated or an


exfoliated nanocomposite. In this technique, no solvent is required.
The melt processing of the polyolens with nanoparticles often
leads to an insufcient nanoparticles dispersion, which results in
aggregation and intercalation, especially at the high contents of
the nanoparticles which in turn reduce the mechanical properties
of the nanocomposites.
Currently, the melting process is commercially the dominative
method for producing the polyolens/clay nanocomposites. However, it suffers from some disadvantages including the insufcient
compatibility of the clay and polyolens. In addition, the organoclays modied with alkyl ammonium are usually decomposed at
temperature higher than 140 C; whereas, the processing temperature is commonly within the range of 190220 C.

Fig. 4. TEM micrograph of the nitrile copolymer lled with montmorillonite by the
exfoliation-adsorption method [37].

2.3.3. Template synthesis


This technique, in which the silicates are formed in situ in an
aqueous solution containing the polymer and the silicate building
blocks, has been widely used for the synthesis of double-layer
hydroxide-based nanocomposites. In this technique, based on

390

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

as trichlorobenzene and toluene (Fig. 6). The results showed that


only a part of the polyethylene could be extracted; whereas, the
homopolyethylene was totally extracted.
3. In situ polymerization of olens
The different types of the catalyst for producing the polyolen/clay nanocomposites during the polymerization are briey
discussed below. The ZieglerNatta catalysts will be considered in
more details due to their more practical uses in the commercial
polyolenic processes.
3.1. Metallocene catalysts

Fig. 5. TEM micrograph of Na-montmorillonite exfoliated in HDPE after in situ intercalative polymerization of ethylene [37].

self-assembled forces, the polymer aids the nucleation and growth


in the inorganic host crystals and gets trapped into the layers as
they grow [37].
2.3.4. In situ intercalative polymerization
In this technique, the layered silicate is swollen with the liquid
monomer (or a monomer solution) so as the polymer formation can
occur in the intercalated sheets. Polymerization can be initiated by
heat, radiation, diffusion of a suitable initiator or an organic initiator
or catalyst xed through cationic exchange inside the interlayer
before the swelling step by the monomer.
Overall, the disadvantages of the solution and melting processes
can be solved by in situ polymerization [42,43]. In situ polymerization is more effective in nanocomposite formation than the melt
processing method according to the comparison of the silicate dispersions. In fact, the in situ polymerization allows the intercalation
of highly reactive and catalytically active organometallic species
to several layers of alumino-silicates. These intercalated species
can react with the monomer, leading to the polymerization inside
the lamellar porosity [4246]. The examination of the TEM picture
reveals that, besides some stacked multilayered silicates, one can
observe well exfoliated and dispersed nanosized sheets (Fig. 5).
Shin et al. [47] showed that the PE produced through in situ
polymerization was chemically linked to the silicate layers. In
this case, the interaction between polymer-clay could be high.
Consequently, it would be conducted to more enhancements in
the nanocomposite properties. Extraction of PE chains from the
PE/clay nanocomposite was carried out using hot solvent such

Metallocene/methylaluminoxane (MAO) catalysts exhibited


high activity for the production of the designed polyolens and
engineering plastics. The characteristics of the polymers, such as
thermal resistance, hardness, impact strength and transparency,
can be precisely controlled through the metallocene structure
[39,6365].
Metallocene catalysts, supported on the clay, have been used
to polymerize the olens in order to produce polyolen/clay
nanocomposites [6673]. Several reviews have been published
describing the methods used to support metallocenes on silica and
other inorganic and organic carriers. MAO does not only remove
the acidic protons seriously affecting the anchorage of the catalysts
and deactivating active sites on the internal surfaces of montmorillonite, but possibly plays other important role [74].
The clay, metallocene catalyst and MAO were added sequentially through three different routes [55,7578]:
1. The clay is directly mixed with olen monomer, catalyst and
cocatalyst to begin the olen polymerization.
2. The clay is rst treated with the MAO cocatalyst and then mixed
with the metallocene catalyst. This mixture is used in the olen
polymerization.
3. The catalyst is initially supported on the clay and then activated
by a cocatalyst in the polymerization reactor.
Among them, it was proven that route 2 is the most efcient approach for producing polymer/clay nanocomposites using
metallocene-type catalysts. These catalyst-supporting methods are
illustrated in Fig. 7.
Route 3 showed the least polymerization performance. Such a
low catalytic activity of the clay-supported metallocene may be due
to the reason of the extensive decomposition of the prototypicallysensitive catalyst on the strongly Bronsted-aciditic clay surface
[57].
Polymerization activity was much enhanced after the passivity
of the clay with an alkyl aluminum cocatalyst. Thus, when the
clay, pretreated with methylaluminoxane, was used to support
Cp2 ZrMe2 , the resulting catalyst was very active towards the
ethylene polymerization. However, the polymerization still took
place mostly in the solution phase. The clay could be recovered

Fig. 6. Synthesis approach using bifunctional organic modier to produce polyethylene chemically linked delicate layers prepared by in situ polymerization [47].

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

391

Table 2
Summary of results for screening ethylene polymerization [80].

Fig. 7. Different metallocene supporting approaches [55].

and separated from the polyethylene product at the bottom of the


reactor. Leaching MAO out of the clay surface via ligand metathesis
presumably allows the formation of highly active metallocenium
cations in the solution phase [57].
It becomes a highly active ethylene polymerization catalyst
when anchored on hydroxyl-terminated oxides such as silica or
alumina while the Zr(CH2 C6 H5 )4 catalyst itself is inactive towards
-olens. In the case of the alumina-supported catalyst, Lewis
acid sites on the surface may activate the catalyst by abstracting a benzyl ligand from the anchored active site. When a bright
yellow solution of Zr(CH2 C6 H5 )4 was added to a slurry of montmorillonite in toluene, an immediate color change occurred. Both
solution and montmorillonite turned to dark green, and the solid
phase was inactive for the polymerization of ethylene. Changing
the color suggests that the organozirconium complex decomposed
in contact with the clay. However, an active catalyst was obtained
when the montmorillonite was pretreated with either (CH3 )3 SiCl
(TMS) or triisobutylaluminum (TIBA). Both of the pretreatments
remove water and cap the surface hydroxyl groups. Addition of
Zr(CH2 C6 H5 )4 to TMS-capped montmorillonite gave a magenta
catalyst with appreciable activity (120 kgPE/molZr h) towards the
ethylene polymerization [79,80].
Maneshi et al. [81] showed that the type of the clay had a deep
effect on the catalyst preparation and consequently, its performance in the ethylene polymerization. Among the clays tested in
the polymerization experiments, only Cloisite 93A and Na+ MMT
were active in the polymerization. The other clay samples were
inactive or could not be recovered from the washing steps due to the
formation of stable colloids in the solvent. A summary of supporting
and polymerization results is shown in Table 2.
Cloisites-10A and 15A formed stable colloids in toluene and
could not be effectively washed with toluene (solvent). Therefore,
they were not tested any further. Cloisite-20A formed a colloid
with very long decantation time and was not used in the ethylene
polymerization. Cloisite-25A also formed a colloid, but a shorter
decantation time (8 h) was required to achieve good separation
from the supernatant; however, no activity was observed during
the screening polymerizations. Cloisite-30B did not form a colloid
in toluene and the particles sedimented in about 2 h during the

Clay sample

Sedimentation rate

Catalyst yield (g/h)

Na+ MMT
Cloisite-10A
Cloisite-15A
Cloisite-20A
Cloisite-25A
Cloisite-30B
Cloisite-93A

Fast (<2 h)
Very slow
Very slow
Slow (>24 h)
Moderate (8 h)
Fast ( 2 h)
Slow (<24 h)

2.16

Very small
3.21

washing steps; however, no noticeable activity was observed in the


ethylene polymerization. Na+ MMT sedimented quickly in toluene
(<2 h), but had a limited polymerization activity. Cloisite-93A sediment occurred slowly (24 h), but the supported metallocene had
the highest activity among all of the Cloisites and Na+ MMT investigated.
Maneshi et al. [82] hypothesized that the Cloisites, formed stable colloids in toluene, such as 10A and 15A, had a larger loading of
organic modication, with respect to the cation exchange capacity
of the original Na+ MMT. Due to the high compatibility between the
organic modication and toluene, sedimentation was extremely
slow and they could not be washed with fresh solvent to remove
the unsupported catalyst molecules [82].
Although polyolen/clay nanocomposites were successfully
produced by the use of the metallocene catalysts, these catalysts
suffer from certain drawbacks. The most crucial restriction is that
an excessive amount of MAO as cocatalyst is needed for the activation of the metallocene catalysts [8386].
3.2. Late transition metal catalysts
In 1995, Brookhart and co-workers turned a new page in the
research on the transition metal complex catalysts when they
introduced the novel nickel and palladium complexes as the
polymerization catalyst precursors. These complexes, which were
highly active in the ethylene polymerization and produced high
molar mass polymers, initiated a period of intensive research on
the late transition metal catalysts. The aim of such researches was
the same as single site metallocene studies: to understand the relation between catalyst structure and polymer microstructure and to
also improve the catalyst performance. Then, it was discovered that
iron and cobalt complexes were highly active toward the polymerization of ethylene [8789].
There are some reports on the in situ polymerization of olens
with the clay-supported late transition metal catalysts towards the
preparation of the polyolen/clay nanocomposites [5661,90,91].
Heinemann et al. [56] reported the in situ polymerization of
ethylene to produce the PE nanocomposites with N,N -bis(2,6diisopropylphenyl-1,4-diaza-2,3-dimethyl-1,3-butadienenickel
and {[ArN = C(Me) C(Me) = NAr]Pd(CH3 )(NC CH3 )} + BAr4 . They
showed that the catalyst performance and polymerization rate signicantly depend on the types and contents of layered silicates and
silicate modiers, comonomer, and late transition metal catalysts.
These factors inuence the properties of the produced nanocomposites [56]. Bergman et al. [92] intercalated a Brookhart type
palladium catalyst {[2,6-Pr2 C6 H3 N C(Me)-C(Me) NC6 H3 Pr3 -2,6]
Pd(CH2 )3 CO2 Me}(B[C6 H3 (CF3 )2 ]4 ) into synthetic uorohectorite.
Ray et al. [44] chose 2,6-bis[1-(2, 6-diisopropylphenylimino)ethyl]
pyridine iron(II) dichloride as a catalyst for the synthesis of the
PE/clay nanocomposites. Mignoni and co-workers [60] showed that
the nanocomposites of exfoliated montmorillonite in polyethylene
can be obtained through using the combination of 1,4-bis(2,6diisopropylphenyl) acenaphthene diimine dichloro nickel(II),
montmorillonite, and methylaluminoxane or trimethylaluminum

392

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

clay may occur. Fig. 10 shows the TEM analysis of a 10 wt% clay-PE
nanocomposite before and after the annealing at 170 C in 30 min,
including little evidence for reaggregation. In fact, the high molecular weight of PE (ca. 500 kg/mol) produced by the clay-supported
catalysts and the high viscosity of the nanocomposites stabilize the
clay dispersion in the polymer matrix [57].
3.3. ZieglerNatta catalysts
Fig. 8. Proposed mechanism of clay activation for nickel catalyst [57].

(TMA) to polymerize ethylene. Fig. 8 shows the mechanism of the


catalyst supporting on the clay.
The TEM studies have shown that the late transition metal
catalysts could acceptably exfoliate the silicate layers during the
polymerization by the intercalation of PE between the clay layers
(Fig. 9). However, it is possible that a partial exfoliation of the clay
layers occurs and some stacks are obtained which contain a few
layers. In addition, a reasonable dispersion of the clay layers in the
polymer matrix was reported. The effect of highly dispersed clay on
the mechanical properties of polyolens is one of the motivations
for producing such nanocomposites [44,57].
The stability of the clay exfoliation in the absence of compatibilizers is a potential issue during processing or other thermal
treatments, when the separation phase and reaggregation of the

The literature studies show that producing the polyolen/clay


nanocomposites by ZieglerNatta catalysts is an interest in many
academic and industrial research centers. In this section, the
clay-supported ZieglerNatta catalysts for the preparation of polyolen/clay nanocomposites through in situ polymerization will be
discussed.
3.3.1. ZieglerNatta catalysts for production of PE/clay
nanocomposites
The clay-supported ZieglerNatta catalysts were successfully
applied to the production of PE/clay nanocomposites through
in situ polymerization. The literature studies illustrate that
two types of the catalysts have been used including the clayunisupported and clay-bisupported ZieglerNatta catalysts for the
olen polymerization which will be discussed in the following.

Fig. 9. TEM micrographs at different magnications [88].

Fig. 10. TEM images of a PE-clay nanocomposite (10 wt% clay) (a) as prepared, using clay-supported (b) after annealing at 170 C for 30 min [57].

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409


Table 3
Ethylene concentration and activity at different temperaturesa [92].

Ethylene concentration (mol/L)


Activity (gPE/gTi h)

Table 4
Inuence of different cocatalysts on the activity of activated clay [93].

Polymerization temperature ( C)

Measurement

393

30

40

50

60

0.1405
1311

0.1295
1826

0.1190
2090

0.1080
1331

Polymerization conditions: Al/Ti: 15 molar ratio, Al (i-Bu)3 , pressure: 1 atm., solvent: heptane.

3.3.1.1. Clay-unisupported ZieglerNatta catalysts. In 2000, Rong


et al. [93] produced a novel organic/inorganic nanocomposite
of polyethylene through the in situ coordinated polymerization
approach. They used ZieglerNatta catalysts supported on the surface of a nanoscale crystal ber (palygorskite) and subsequently,
ethylene was supplied to the activated surface and the polymerization was initiated onto the surface of the bers. As the
polymerization proceeded, the surface of the ber was gradually
covered by the forming polymer, until the ber was fully encapsulated by polyethylene. The roles of the palygorskite were twofold:
- It was a support for the catalyst for initiating the polymerization.
- Reinforcement as a nanoparticle after the polymerization was
stopped.
Three possible reactions between TiCl4 and palygorskite may
take place in the process of the palygorskite activation when supporting TiCl4 :
1. TiCl4 reacts with water of palygorskite and forms titanium oxides
or their derivatives.
2. TiCl4 reacts with SiOH located on the surface of the clay bers
and is supported on it.
3. TiCl4 coordinates with the magnesium vacancies in the clay
structure and is supported on it.
Because the amount of SiOH in the palygorskite is low, the
reaction between SiOH and TiCl4 is negligible.
If the calcining temperature is low, there is relatively a large
amount of water present in palygorskite. The water reacts preferably with TiCl4 and many noncatalytic titanium species are formed
when TiCl4 is introduced. In this case, the activity of the activated
palygorskite is low. However, if palygorskite was treated at the elevated temperatures, most of the water content would be driven out
and the reaction between TiCl4 and magnesium would become the
main reaction. In this case, the highest activity was achieved [93].
Rong et al. also showed that the polymerization temperature
and the amount of activator had a deep effect on the catalyst activity
and the molecular weight of the nanocomposites [93].
According to Table 3, the polymerization temperature has a
great inuence on the polymerization. Limited to the boiling point
of the solvent, the polymerization temperature was controlled not
to be higher than 60 C under the normal pressure. The rate of
polymerization initiated by ZieglerNatta catalysts was controlled
through two parameters: a higher polymerization temperature
results in a higher activity, but a lower ethylene concentration. For
this reason, the overall activity exhibits an increase.
The cocatalyst plays an important role in the olen polymerization. The effect of the Al/Ti on the catalyst activity is listed in
Table 4. It is seen that the activity displays a maximum as the Al/Ti
ratio increases. This is normal for the ZieglerNatta type catalysts
and the decrease in the activity at high Al/Ti ratios to be explained
by the overall reduction of the active sites [94,95].
As shown in Table 4, besides the Al/Ti ratio, the type of the cocatalyst also affects the activity of the polymerization. The order of the

Measurement

Activity (gPE/gTi h)

Cocatalyst
AlMe3

AlEt3

Al(i-Bu)3

366

847

1826

Polymerization conditions: Temp.: 40 C, Al/Ti: 30, pressure: 1 atm, solvent: heptane.

activity is Al(i-Bu)3 > AlEt3 > AlMe3 . The reason may lie in the low
reduction capability of Al(i-Bu)3 [93].
The activity of the catalysts supported on palygorskite was very
low compared to the commercial ZieglerNatta catalyst used for
the ethylene polymerization. In fact, the activity of such catalysts is
almost similar to the rst and second generations of ZieglerNatta
catalysts which have low activity [3,65,96,2,97].
Jin et al. [41] illustrated that the xation of the catalyst on
sodium montmorillonite (MMT-Na) and organophilically modied
montmorillonite (Cloisite 30B, MMT-OH) can be used for producing the in situ polymerized nanocomposites. The catalysts showed
a low activity and a fast decrease in the polymerization rate during the polymerization which resulted from the deactivation of
titanium catalyst by the dangling hydroxyl group existing in the
MMT-OH and MMT-Na. Meanwhile, the activity of MMT-Na supported catalyst indicated being higher than that of the MMT-OH
supported catalyst, 60 kgPE/molTi h and 19.9 kgPE/molTi h in MMTNa and MMT-OH, respectively. The reason is that MMT-OH contains
higher amounts of hydroxyl groups than MMT-Na.
Ramazani and co-workers [98] calcinated the clay and then
used it as a support for the preparation of the clay-supported
ZieglerNatta catalyst. In fact, the existing water in the clay can
react with the catalyst and can thus disturb supporting of the catalyst on the clay. Therefore, water should be removed from the clay
by suitable heat treatment. Smectite clay has three types of water
or hydroxide groups in or on its crystallites structure. Physically
absorbed water is held by less strong interaction and is found on the
surface of the clay on defect sites or on sites of broken bond of silicate structure and eliminated at 8090 C. Bound water associated
with geometric structure around a cation, found between the sheet
layers of smectites, is the second type of hydroxide groups. This
water normally leaves the clay structure at 100200 C. Crystalline
water which is the last type of hydroxide source, is found within
the sheet layers as OH unites. To remove this water which is more
rmly bound to the structure, temperature of 500 C or more is necessary. To make sure about removing all kind of hydroxide groups
from the clay, samples of the clay were calcinated at 600 C for about
6 h and subsequently, stored under inert gas. Fig. 11 shows the FTIR spectra of MMT samples. The OH absorption peaks for different

Fig. 11. FT-IR spectra of MMT samples calcinated at 200 and 600 C [98].

394

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

Table 5
The effect of Al/Ti molar ratio on polymerization activity [98].
Al/Ti

Activitya (gPE/mmolTi h)

Activityb (gPE/mmolTi h)

97
130
178
190
285

62
65
140
125
120

70
73
193
149
141

a
b

TEA as activator.
TIBA as activator.

types of bound water appear at different wavelengths. The peaks


between 3600 and 3700 cm1 stand for the surface water. The three
peaks between 3413 and 3524 cm1 are the reection of crystalline
water. The peak at 1649 cm1 stands for the surface adsorbed water
and the peak at 1626 cm1 is due to coordinated water. Fig. 11 also
shows that increasing the clay heat treatment temperature from
200 C to 600 C changes the relative size of hydroxyl peaks [98].
It was shown that the Al/Ti molar ratio had strong inuence on
the polymerization activity. In addition, it seems that tri-isobutyl
aluminum (TIBA) as a cocatalyst, was more effective on the catalyst
activity than triethylaluminum (Table 5).
It can be seen that as the Al/Ti molar ratio increased up to
178, the activity increased and further increase in this molar ratio
resulted in decreasing the activity. This behavior can be normally
observed in the ZieglerNatta catalysts. Such a decrease in the catalyst activity at high Al/Ti weight ratios can be attributed to the
overall reduction of the active sites [93]. The results also show that
TIBA was more effective activator than TEA in the ethylene polymerization, which may be attributed to the low reduction capability
of TIBA [98]. Although the activity of the catalyst was enhanced
using TIBA as an activator, it was low compared to the commercial
ZieglerNatta catalysts.
3.3.1.2. Clay-bisupported ZieglerNatta catalysts. It has been shown
that the clay-unisupported ZieglerNatta catalysts did not have
suitable activity for the ethylene polymerization; while the high
activity is essential to use a catalyst in olens polymerization, particularly, in the industrial processes.
Some
researchers
synthesized
the
clay-bisupported
ZieglerNatta catalysts for improving the catalyst performance
in olen polymerization. For this purpose, the clay was mainly
treated with magnesium compounds which were well known as
efcient supports for this type of the catalysts. Three different
approaches were applied to produce the clay bisupports.
1. Mixing the clay and support powder like the clay and MgCl2 . In
this method, the clay and solid support are mixed and ball-milled
and following, the resulting bisupport is used for the catalyst
preparation.
2. The adduct MgCl2 as the second support is mixed with the clay.
The obtained solid is used as a bisupport.
3. A homogenous compound is xed on the clay surface and then,
the resulting bisupport is applied for the Ti and V xation.
Yang et al. [99] prepared the PE/MMT nanocomposite using a
MMT/MgCl2 /TiCl4 catalyst. MgCl2 was rst deposited in the layer
surfaces of MMT, then MMT/MgCl2 /TiCl4 was prepared by loading
TiCl4 similar to a typical highly active ZieglerNatta catalyst. During the ethylene polymerization process, the layers of MMT could
be exfoliated by polymerization force and the polyethylene/MMT
nanocomposite is produced in situ.
It is well known that MgCl2 dissolves in alcohols and
forms MgCl2 nROH complexes in the homogeneous solution
[95,100104]. Montmorillonite can be swollen to alcohols. When
MMT is immersed in MgCl2 /alcohol solution, MMT can be swelled.

Fig. 12. Scheme of intercalated catalyst [99].

Thus, it allows the diffusion of MgCl2 -nROH complexes into the


spacing between the MMT layers. After the removal of alcohol,
MgCl2 may well deposit in/between the surfaces of the layered
MMT as microcrystallite.
Overall, the key factor in the preparation of the polyolen/layered MMT nanocomposites by in situ polymerization is
that the catalyst species can be intercalated into the MMT layers efciently without sacricing the high catalytic activity for the
catalyst. The possible reactions are schematically shown in Fig. 12.
In typical highly active ZieglerNatta catalysts, MgCl2 is an
essential component to promote the activity and polymerization
behavior of the catalyst. Additionally, it may form a single molecular layer along the inner surfaces of MMT layers, avoiding the
formation of non/low active species (SiOTi) which results from
the reaction of OH on the MMT layer surface with TiCl4 . Increasing
the MMT interlayer spacing, it preliminarily makes TiCl4 easier to
intercalate into MMT layers [101,105].
The polymerization of ethylene by the MMT/MgCl2 /TiCl4 catalyst showed that the catalyst had fairly good activity up to
748 kgPE/molTi h [99].
It is possible to obtain the PE/MMT nanocomposites through
different clay content by changing the polymerization conditions
such as polymerization temperature, ethylene pressure, and Al/Ti
molar ratio [106]. However, the molecular weight of the nal product may be altered as well, which is undesirable if one wishes to
examine the inuence of MMT content on the physical properties of the composites. Table 6 shows that the molecular weight
of the nanocomposites did not alter signicantly by changing the
polymerization time. In fact, it remained around 1 106 g/mol, on
average, in the case of lack of a chain transfer agent. The MWD of
the nanocomposites decreased while increasing the polymerization time, reducing from 7.8 after 45 min of polymerization to 3.9
after 180 min of polymerization. Therefore, as shown in Table 6,
changing the polymerization time is a simple and efcient approach
to obtain the PE/MMT nanocomposites of different MMT contents
that have comparable molecular weights [99].
The xylene-extraction tests using a soxhlet extractor showed
that over 50 wt% of the nanocomposites was not extractable, indicating the strong interaction between the MMT layers and PE
molecular chains [99].
Cui et al. [107] reported that the complete exfoliation of the clay
layers during the vanadium-based ZieglerNatta polymerization of
ethylene has been successfully carried out using clay and MgCl2
hybrid supports. MgCl2 offers catalyst loading sites, and the vanadium catalyst is avoided from direct anchoring on the surface of the
clay. So, the interactions of the clay/MgCl2 /VOCl3 catalyst caused
high activity for the ethylene polymerization [100]. It is proven that
Table 6
Results of polymerization [99].
Sample

Time
(min)

Activity
(kgPE/molTi h)

Clay content
(wt%)

Mw 106
(g/mol)

MWD

PE1
PE2
PE3
PE4
PE5
PE6

180
150
120
90
60
45

280
291
320
472
641
748

0.08
0.13
0.16
0.22
0.56
0.67

1.1
1.1
1.2
1.0
0.9
0.8

3.9
3.9
3.8
4.3
5.2
7.8

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409


Table 7
Ethylene polymerization with intercalation catalyst V-MMT [107].
Initial clay
loading (g)
0
0.15
0.10
0.10

Polymerization
time (h)
2
12
24
12

Activity
(kgPE/molV h
atm)
2.60
50.7
28.0
36.2

Clay mass
fraction (wt%)
0
1.61
2.43
3.76

395

Table 8
Clay-BOM supports, catalysts and their Ti content [46].
Extracted in
decalin (%)
82.7
73.4
40.4
23.4

the vanadium-based supported catalysts have relatively high initial


activity for the ethylene polymerization [101].
The MMT/MgCl2 nEtOH/VOCl3 /TEA catalyst system acted as an
effective catalyst for the ethylene polymerization, as indicated
in Table 7. Polyethylene with different clay loading was synthesized by varying the polymerization time and the amount
of the catalysts. The V-MMT catalyst showed fairly good activity up to 76 kgPE/molV h atm compared with the homogeneous
vanadium catalyst, behaving as the high active MgCl2 -supported
ZieglerNatta catalysts.
Extraction of polyethylene chains from the hybrid PE/clay composites was carried out using reux solvent (decalin) in a soxhlet
extractor in 24 h. The results showed that only part of PE could be
extracted. In addition, the extracted amount of PE decreased via
increasing the clay loading in the PE matrix. When the clay is dispersed in the PE matrix, the clay layers can stiffen the PE chains and
impede the movements of PE chains, especially when the particular
network is formed [99].
The interaction between the silicate layers of the clay and the
polymer matrix was studied by FT-IR characterization. The representative FT-IR spectra of the pristine Kunipia F, pure PE and a series
of decalin-extracted PE/clay nanocomposites are shown in Fig. 13.
In the IR spectrum of Kunipia F, the characteristic absorbance
bands occur in the following assignment: OH stretching at
3632 cm1 , coordination and absorbed water peak at 3430 and
1640 cm1 , Si O stretching at 1040 cm1 and Al O stretching at
550 cm1 [85]. The characteristic vibration bands of PE are C H
stretching at 2855 and 2950 cm1 and 1470 cm1 (C H ). In the IR
spectrum of decalin-extracted PE/clay nanocomposites, the presence of both characteristic group frequencies of PE and MMT can
be found. These results show that only a part of PE can be extracted
from the nanocomposites, and the rest of the PE chains stay immobilized inside and/or on the layered silicates. This conrms that
strong interactions exist between the nanometric silicate layers and
the PE segments.

Support and Catalyst No.

Clay (g):BOM (g)


Ti (wt%)

15:0
1.01

15:2
1.12

15:4
1.23

15:8
2.05

15:12
3.01

As discussed, the clay-MgCl2 -bisupported catalysts showed


high activity through the ethylene polymerization. However, since
MgCl2 is a solid powder, it cannot effectively diffuse in the clay
layers and forms a homogenous bisupport. In addition, it is possible there was free or unreacted MgCl2 in the bisupport compound
which cannot produce the given nanocomposites. In other words,
the free part of MgCl2 may produce the pure PE which does not
have any interactions with the nanoparticles.
Abedi et al. used the clay/BOM (butyl octyl magnesium)bisupported ZieglerNatta catalysts for the polymerization of
ethylene towards the preparation of the PE/clay nanocomposites
via in situ polymerization [46,108,109]. In these catalysts, BOM
is a homogenous (liquid) compound which can easily diffuse in
the silicate layers of the clay and is homogenously dispersed.
This dispersion results in being produced a homogenous clay-BOM
bisupport.
In general, ZieglerNatta catalysts supported on the clay for the
preparation of PE/clay nanocomposites suffer from some deciencies including low yield and undesirable morphology which are the
basic main restriction for the production of PE/clay nanocomposites
by in situ polymerization, especially in the industrial perspective
[42,43]. The clay/BOM-bisupported catalysts showed a reasonable
activity in the ethylene polymerization. For this purpose, the clay
was reacted with different amounts of BOM in order to modify the
clay towards achieving suitable supports for the preparation of the
clay-based ZieglerNatta catalysts for the ethylene polymerization
(Table 8). The table shows that the content of Ti loaded on the
catalysts rose by increasing the amount of BOM treated with the
clay.
Table 9 shows the BET surface area of the supports (No. 1 and
2) and catalysts (No. 1 and 2). According to the table, the surface
area of the clay was enhanced under the reaction of the clay with
BOM and also during the catalyst preparation. Overall, an increase
in the surface can lead to the improvement in the behavior of the
ZieglerNatta type catalysts in the olens polymerization [96]. In
addition, the pore volume of the supports and catalysts were relatively low compared to the common ZieglerNatta catalysts [95].
Fig. 14 illustrates the FT-IR spectra of the clay, clay-BOM support and catalysts No. 1 and 2. As observed in Fig. 14a, the clay
showed the absorption bands at 3632 and 1046 cm1 which can
be contributed to N C and N H groups of the alkyl ammonium
used for the clay modication. In addition, the alkyl group showed
the absorption bands at 2924 and 2851 cm1 corresponding to
C H. Compared to Fig. 14b, the intensity of the absorption bands
remarkably decreased, especially at 3632 cm1 . This indicates that
a large portion of alkyl ammonium in the clay was removed by
BOM. According to Fig. 14c, the untreated clay also lost its alkyl

Table 9
BET surface area analysis [46].

Fig. 13. FT-IR spectra of pristine Kunipia F (a) pure PE (b) decalin-extracted PE/clay
nanocomposites with different clay loadings (c) PE/clay 1.6 wt% (d) PE/clay 2.4 wt%
(e) PE/clay 3.1 wt% [107].

Sample

Surface area
(m2 /g)

Pore volume
(mL/g)

Average pore
diameter ()

Support No. 1
Support No. 2
Catalyst No. 1
Catalyst No. 2

8.7
11.9
19.6
25.1

0.041
0.076
0.091
0.096

93
129.2
76.4
92.6

396

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

Fig. 14. FT-IR spectra of (a) clay (b) clay/BOM (c) clay-supported catalyst (d)
clay/BOM-bisupported catalyst [46].
Table 10
Yield of produced catalystsa [46].
Catalyst No.

Yield (gPE/molTi)

1
2
3
4
5

95.2
1217.4
2346.1
2209.3
2206.3

Fig. 15. Polymerization rate of prepared catalysts [46].


Table 12
Effect of hydrogen on catalyst yielda [46].
H2 (bar)

Polymerization conditions: PC2 = 8 bar, Time = 2 h, PH2 = 0, TEA/Ti = 20, Temperature = 70 C.


Yield (kgPE/molTi)

ammonium during the titanation (catalyst preparation). In addition, catalyst No. 2 contained some alkyl ammonium (Fig. 14d).
According to Table 10, the yield of the catalysts increased while
increasing the clay/BOM ratio, rising from 65.20 kgPE/g Cat in catalyst No. 1 to 2206.3 kgPE/g Cat in catalyst No. 5. This effect can be
related to the increase of the amount of BOM and consequently, Ti
loaded on the clay-BOM support in which the clay-BOM component acts as a suitable support in the preparation of the catalyst.
Therefore, BOM offers higher active sites. In addition, since the
presence of nitrogen groups on the clay surface acts as poison for
the ZieglerNatta type catalysts, removal of such groups through
the clay treatment by using BOM reduces their adverse effects on
the catalyst performance. The results indicate that the clay-BOM
supported catalysts enjoyed an acceptable yield in the ethylene
polymerization (Fig. 15). According to the gure, the polymerization rate of the catalysts enhanced by increasing their yield. Similar
to the yield of catalyst No. 1, this catalyst showed the least polymerization rate; conversely, catalyst No. 5 had the highest yield.
Although catalysts No. 1 and 2 showed low yields, they displayed
smooth polymerization rate; whereas, catalysts No. 3, 4, and 5
showed a rapid drop in their rate during the polymerization. However, the polymerization rate reached almost a stable state after
about 1 h of polymerization.
According to Table 11, the maximum yield of the catalyst was
observed at the TEA/Ti molar ratio of 2030. This means that the
TEA/Ti ratio of 2030 was required for obtaining the maximum activation of the catalyst, but higher levels of the cocatalyst had an
adverse effect on the catalyst. The reduction of the catalyst activity
and its rate in the presence of the high cocatalyst concentration can
be attributed to the overall reduction of the active sites [93].

1217.4

739.1

521.7

347.8

4
130.43

Polymerization conditions: Catalyst No. = 2, temperature ( C) = 70, time = 2 h,


PC2 = 8 bar, TEA/Ti = 20.

As Table 11 shows, the yield of the catalyst passed a maximum


at 6070 C and then decreased with increasing temperature. This
result might be ascribed to an irreversible destruction of active
sites at higher temperature under the polymerization conditions
[95,110].
According to Table 11, the catalyst yield was roughly proportional to the monomer pressure, rising from 347.8 kgPE/mol Ti at
2 bar to 1434.8 kgPE/mol Ti at 10 bar. This can be contributed to
the increase of the monomer concentration in the polymerization
system at higher pressures.
Hydrogen is used as a chain transfer agent for controlling molecular weight in the ZieglerNatta type polymerization [111113].
Table 12 shows the effect of hydrogen on the yield of catalyst No. 2.
According to the table, the yield of the catalyst showed a decrease
in the presence of hydrogen, reducing from 1217.4 kgPE/mol Ti
in the absence of hydrogen to 130.4 kgPE/molTi in the presence
of 4 bar of hydrogen. The main reason for such a decline in the
catalyst yield was the reduction of the partial pressure of the
monomer in the polymerization system by increasing the amount
of hydrogen. Furthermore, the slow addition of the monomer to
the catalysthydrogen bond formed in the chain transfer stage to
hydrogen causes an adverse effect on the catalyst yield [114].
3.3.1.3. Properties of PE/clay nanocomposites.
3.3.1.3.1. Physical properties. Some articles reported that the
values of the melting temperature (Tm ), melting enthalpy (Hm ),
crystallization temperature (Tc ) and crystallinity degree (Xc ) of the

Table 11
Effect of polymerization conditions on catalyst yield [46].
Al:Tia

Yield (kgPE/molTi)

Temperature ( C)b

Yield (kgPE/molTi)

P (bar)c

Yield (kgPE/molTi)

10
20
30
40
50

826.1
1217.4
1173.9
1043.5
869.6

50
60
70
80
90

1043.5
1730.4
1217.4
956.5
782.6

2
4
6
8
10

347.8
652.2
869.0
1217.4
1434.8

a
b
c

Polymerization conditions: Catalyst No. = 2, PC2 = 8 bar, time = 2 h, PH2 = 0, temperature = 70 C.


Polymerization conditions: Catalyst No. = 2, PC2 = 8 bar, time = 2 h, PH2 = 0, TEA/Ti = 20.
Polymerization conditions: Catalyst No. = 2, temperature ( C) = 70, time = 2 h, PH2 = 0, TEA/Ti = 20.

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

397

Table 13
Physical properties of nanocomposites [108].
Sample

Clay (wt%)

Tm ( C)

Tc ( C)

Hm (cal/g)

Xc (%)

Density (g/cm3 )

Bulk density (g/cm3 )

PE1
PE2
PE3
PE4
PE5

0.24
1.22
2.76
3.45
4.47

135.3
136.2
133.9
135.1
135.5

117.5
118.3
114.5
118.1
118.5

33.6
34.1
34.2
34.73
36.2

51.8
52.6
52.7
53.6
55.9

0.928
0.933
0.939
0.948
0.959

0.19
0.18
0.22
0.21
0.19

polyolen/clay nanocomposites may be different from those of the


pure polyolens. Overall, the Tm and Tc values of such nanocomposites are often close to the pure polyolens; whereas, the values
of the Hm , and Xc are under the inuence of the clay contents
[83,115].
According to Table 13, Tm and Tc of the PE/clay nanocomposites
were about 135 C and 117 C which were close to the pure PE. In
addition, the crystallinity degree of the nanocomposites was in the
range of 5156% which was lower than the pure PEs. The obtained
results were consistent with the reported data in the literature
[108].
Nanoscale clay layers affect the crystallization in two opposite
ways. On the one hand, the interaction between the clay layers and PE chains decrease the number of crystalline PE chains
by preventing the cell growth mechanism during the crystallization. On the other hand, the nucleation of the clay results in more
perfect crystallization structure. The nucleation of the clay can relatively increase the number of the crystalline. For this reason, the
crystallinity of the nanocomposites slightly enhances at the high
amounts of the clay [4,116].
The density of the nanocomposites generally enhances whilst
increasing the clay content [49]. The density of the nanocomposites increased from 0.928 g/cm3 in PE1 to 0.959 g/cm3 in PE5 as the
clay content increased. It is clear that the density of the nanocomposites was less than the pure PEs in spite of using no comonomer
in the polymerization. This can be contributed to a decrease in the
crystallinity of the nanocomposites compared to the neat polyethylene.
The bulk density of the polymer/clay nanocomposite powders
is around 0.180.22 g/cm3 which is lower than that of the current
commercial products. The ag-like morphology of the powder is
responsible for the low bulk density of the nanocomposites [108].
3.3.1.3.2. Mechanical properties. The mechanical properties of
the PE/palygorskite nanocomposites are shown in Table 13. According to the table, the tensile modulus increased and the elongations
at break decreased systematically by increasing the content of
palygorskite in the nanocomposite produced through in situ polymerization [105].
Palygorskite is an inorganic ber with much higher modulus
than PE. Mixed with PE, the inorganic ber will undoubtedly stiffen
PE. At the same time, the inorganic ber will also impede the movements of the polymer chains, especially when the particular net
work (gel) is formed. The chains cannot easily orient themselves

Fig. 16. Model structure of organic/inorganic network [105].

to comply with the stress. Consequently, the elongation at break


decreases as the content of palygorskite increases. For the same
reason, the tensile strength becomes rather undened than palygorskite being introduced. The segments undergo an orientation
to a dened extent before they break when the neat PE is subjected to the tensile stress. However, when such an orientation is
inhibited, a break can occur at any extent of the orientation. The
tensile strength is therefore not well-dened. The better denition
is the yield strength, since it occurs when the lamellae start to slide,
which is not impeded by palygorskite. Meanwhile, no sample broke
in the impact test. This is probably attributed to the strong entanglement resulting from the chain brush, macromolecular comb or
structure (Fig. 16).
The properties of the PE/palygorskite composites produced
via melt blending are also shown in Table 14. Since the modulus is dependent on the chemical composition and not on the
microstructure, the moduli are in the same order as those of
the PE/palygorskite produced through in situ polymerization.
However, both the tensile strength and elongation at break are
much lower than those of the nanocomposite produced by in situ

Table 14
Mechanical properties of PE/palygorskite composites [105].
Samples

Palygorskite content (%)

Tensile strength (MPa)

Elongation at break (%)

Tensile modulus

Yield strength (MPa)

Nanocomposites produced by in situ


polymerization

0.5
3.1
5.5
9.9
18.7
24.5
3.0
8.0
16.0
25.0

28.9
40.9
36.9
31.8
33.7
27.2
14.0
23.4
20.2
19.8

396.8
266.9
205.8
214.2
180.5
80.8
5.2
33.8
5.1
5.1

355.1
448.6
545.9
850.0
878.5
1167.6
572.7
400.6
850.0
816.0

20.4
21.9
22.4
25.6
26.4
26.3

Nanocomposites produced by
melting process

398

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

Fig. 17. Thermal stability of PE/clay nanocomposites [108].

polymerization. This can easily be explained by the poor bonding


between PE and palygorskite [105].
Abedi reported that the impact strength of the PE/clay nanocomposites was rst enhanced an increase in the clay contents
and then, there was a decline in the impact strength of the
nanocomposites containing a high amount of clay [108]. It was
comparable to that of the ultra-high molecular weight polyethylene (Mw > 2 106 g/mol), while the nanocomposites had much
lower molecular weight [117].
3.3.1.3.3. Thermal properties. The thermal stability of the
PE/clay nanocomposites, investigated by TGA analysis, is shown in
Fig. 17 [108]. According to the gure, the thermal stability of the
nanocomposites are slightly enhanced and then reduced through
increasing the clay loadings. The thermal decomposition temperatures for PE1 (0.24%), PE2 (1.22%), PE3 (2.76%), PE4 (3.45%) and
PE5 (4.47%) have been about 471, 478, 477, 476 and 469 C, respectively; whereas, the thermal decomposition temperature for the
pure PE is around 415 C. It can be concluded that the clay layers
play an important role in the improvement of the thermal stability in the nanocomposites. In fact, the good barrier action of the
clay layers increases the thermal stability of the nanocomposites.
However, the thermal decomposition temperature of the nanocomposites shifts to lower temperatures by increasing the clay contents
because the clay can also catalyze the degradation of the polymer
matrix. This action will reduce the thermal stability of the PE/clay
nanocomposites at high clay contents in the PE matrix [71,116].
Table 15 shows the OIT data of the PE/Clay nanocomposites.
The results indicate that the increase of the clay loadings slightly
improves the OIT property of the polyethylene because of the barrier role of the clay. Overall, the effect of the clay on the OIT was
low compared to that of the traditional antioxidants [108].

Fig. 18. WAXD patterns of Na+ MMT, organo-MMT and PP/MMT nanocomposites
[118].

3.3.2.1. Clay-unisupported ZieglerNatta catalysts. In 2001, Ma et al.


[118] reported that the PP/clay nanocomposites were synthesized
by intercalative polymerization. The idea of intercalative synthesis of the PP/clay nanocomposites was from the concept of
the polymerization-lling synthesis (PFS) of polyolen composites proposed by Dyachkovskii. In the PFS concept, the polyolen
composites were obtained through polymerizing the corresponding monomer on the surface of the inorganic ller (e.g. kaolin),
which was activated by the ZieglerNatta catalyst [119,120].
Ma used the layer structure silicate of Na+ MMT as the ller
instead of the kaolin [118]. After being modied by the intercalative agent, i.e. hexadecyl-octadecyl trimethylammonium chloride,
the modied clay was grounded in MgCl2 . Then, the resulting mixture was treated with the internal donor and TiCl4 to produce
the activated clay/MgCl2 -supported ZieglerNatta catalyst for the
propylene polymerization in order to prepare the PP/clay nanocomposites through in situ polymerization. The PP/MMT1 (2.5 wt%),
PP/MMT2 (4.6 wt%), PP/MMT3 (8.1 wt%) and PP/MMT4 (10.4 wt%)
samples were obtained by controlling the polymerization time.
Meanwhile, there was no data on the catalyst yield in the report
[118].

3.3.2. ZieglerNatta catalysts for production of PP/clay


nanocomposites
The synthesis of the PP/clay nanocomposites is going to become
an active eld in industrial and academic research. Similar to the
ZiglerNatta type catalysts, applied for the preparation of PE/clay
nanocomposites, two types of the catalysts were used for the
producing the PP/clay nanocomposites that are considered in the
following.

Table 15
OIT values of the produced PE/clay nanocomposites [108].
Samples

Induction time (s)

PE1
PE2
PE3
PE4
PE5

24.6
31.2
33.6
34.6
35.4

Fig. 19. TEM images of PP/MMT nanocomposite [118].

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

399

Table 16
Results of XPS analysis of catalyst samples [122].
Samples

MgCl2 /TiCl4 -EB


MMT/TiCl4 -EB
MMT/MgCl2 /TiCl4 -EB

Fig. 20. X-ray diffraction patterns of (a) MMT (b) MMT/MgCl2 (c) MMT/MgCl2 /TiCl4 EB [122].

Fig. 18 shows the WAXD patterns of the Na+ MMT, organo-MMT


and PP/MMT nanocomposites. The Na+ MMT and organo-MMT display their characteristic peaks at 2 = 7.12 and 4.54 , respectively,
corresponding to (0 0 1) diffraction of the layer structure of the
Na+ MMT and organo-MMT, respectively. However, the characteristic peak disappears in the WAXD pattern of PP/MMT2 with the clay
content of 4.6 wt%, as shown in Fig. 18. Consequently, a nanocomposite is commonly considered to be an exfoliated nanocomposite
[118,121].
The nanostructure of the PP/Na+ MMT nanocomposite was studied by the TEM observation (Fig. 19). The dark lines present an
individual clay layer, whereas the bright area represents the PP
matrix. It can be seen that the PP chains have been intercalated
into the clay layers. The clay layers are packed along the main axis
direction with a low angle. The TEM image also shows the packing
of the clay layers containing about two to seven individual layers.
The gallery distance is about 410 nm. The width of the stacking
clay layers is about 470 nm. It can be concluded that the stacked
clay layers were exfoliated into nanometer-size layers and uniformly dispersed in the PP matrix during the polymerization. In
fact, the layered structure of the clay was destroyed by growing PP
macromolecules chains [118].
3.3.2.2. Clay-bisupported ZieglerNatta catalysts. Zhao et al. [121]
suggested the MMT/MgCl2 /TiCl4 /ethylbenzoate (EB) catalyst for
the polymerization of propylene to produce the PP/clay nanocomposites. To prepare the catalyst, MMT was rst immersed in a
homogeneous MgCl2 /EtOH heptane solution, so called adduct,
where the complex of MgCl2 nEtOH was able to diffuse into the
swelled interlayers of MMT by chemical or physical absorption.
Then, MgCl2 was reformed after removing EtOH from the system. By
loading TiCl4 and the internal donor EB, the MMT/MgCl2 /TiCl4 -EB
catalyst with the intercalated structure was formed [122].
The intercalation of the catalyst components into the MMT layers was essential for obtaining composites with the nano-scale
MMT layers dispersed in the polymer matrix by using the in situ
polymerization approach. The WAXD diffraction patterns clearly

Binding energy Eb in eV
Mg2s

Cl2p

Ti2p3/2

51.2

51.1

199.0
198.6
199.5

458.2
459.2
458.1

showed a continuous layer spacing enlargement of MMT when the


catalyst components were successively intercalated into the layers
of MMT (Fig. 20).
As shown in Fig. 20, the pristine MMT had a strong 0 0 1 peak at
2 = 7.64 , corresponding to an interlayer spacing of 1.15 nm, while
with intercalation of MgCl2 , the 0 0 1 peak shifted to 2 = 6.34 corresponding to an interlayer spacing of 1.39 nm, and nally with
TiCl4 -EB intercalation, the peak shifted to 2 = 5.18 corresponding to an interlayer spacing of 1.71 nm. The continuous enlarged
interlayer spacing of MMT successive intercalation of the catalyst
components is schematically shown in Fig. 21 [122].
It is worthwhile to understand whether TiCl4 was supported
on MgCl2 microcrystalites or on the layer surface of MMT. The
latter has been conrmed by the very low activity of the catalysts with regard to the olen polymerization [105]. To clarify
this suspicion, the binding energies (Eb ) of Ti2p3/2 of the catalysts measured by X-ray photoelectron spectroscopy (XPS) were
investigated. As tabulated in Table 16, the Eb values of Ti2p3/2 in
the MMT/MgCl2 /TiCl4 -EB catalyst were almost the same as in the
MgCl2 /TiCl4 -EB catalyst, but lower than that in the MMT/TiCl4 EB catalyst. It indicates that the chemical environment of the Ti
species in the MMT/MgCl2 /TiCl4 -EB catalyst is the same for the
MgCl2 /TiCl4 -EB catalyst, but different from that in the MMT/TiCl4 EB catalyst. From this observation, it can be concluded that the Ti
species in the MMT/MgCl2 /TiCl4 -EB catalyst is mainly supported on
MgCl2 , probably through a Cl bridge, instead of the layer surface of
MMT [123].
Table 17 shows the results of the propylene polymerization
by the above catalysts. As shown in Table 17, the activity of
the MMT/MgCl2 /TiCl4 -EB catalyst was comparable to that of the
general type ZieglerNatta catalyst, MgCl2 /TiCl4 -EB, further conrmation that TiCl4 was mainly supported on MgCl2 . The activity
of the MMT/TiCl4 -EB catalyst in the propylene polymerization was
very low. This means that the direct titanation of the MMT led to
an inefcient catalyst. Nevertheless, the isotactic index (I.I.) of the
nanocomposites was lower than that of general PP, probably due to
the complicated chemical or physical properties of the MMT layer
surface. Meanwhile, the PP/clay nanocomposites were obtained
through varying the polymerization conditions [122].
Du et al. [124] reported that imidazolium-modied MMTs
(IOHMMT), made by ion exchange of Na+ MMT with 3-hexadecyl1-(2-hydroxy-ethyl)-3H-imidazol-1ium bromide (IOH), were used
for the preparation of the clay-supported ZieglerNatta catalysts
in order to polymerize propylene towards the production of the
PP/clay nanocomposites. For this purpose, the adduct MgCl2 nROH
is prepared by dissolving MgCl2 in 2-ethylhexan-1-ol at 130 C
for 2 h, followed by the addition of phthalic anhydride. The
IOHMMT/toluene mixture was added to the MgCl2 nROH solution.

Fig. 21. Scheme of successive intercalation of catalyst components [122].

400

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

Table 17
The polymerization results [122].
Catalyst

Al/Ti (molar ratio)

Ti/DDSa (molar ratio)

Activity (kg/mol Ti h)

I.I. (wt%)

Mw (104 g/mol)

MMT (wt%)

MgCl2 /TiCl4 -EB

40
20
30
40
40

4
4
4
4
4

170
240
310
110
trace

97
87
85
89

14.6
13.9
15.3
15.7

1.1
0.8
2.4

MMT/MgCl2 /TiCl4 -EB


MMT/TiCl4 -EB
a

Dimethyldimethoxysilane.

Table 18
The results of the propylene polymerization using the MMT/MgCl2 /TiCl4 catalyst [124].
MgCl2 /surfactant (molar ratio)

Al/Ti (molar ratio)

Activity (kg PP/mol Ti h)

Clay (wt%)

I.I. (wt%)

Tm ( C)

Hm (J/g)

Mw (104 )

MWD

4
4
4
2

50
80
100
100

77.9
53.0
31.1
19.1

2.0
3.0
5.0
19.0

94
93
92
94

161
161
159
159

69.8
69.1
72.1
75.4

61.1
42.5
43.0
58.5

5.9
6.5
7.5
6.0

The resulting MMT-MgCl2 nROH was activated by treating with


TiCl4 . In this way, the MMT/MgCl2 /TiCl4 catalyst was used in the
propylene polymerization. The polymerizations were initiated by
the above catalysts activated by Et3 Al. A summary of conditions
and results for in situ polymerization and typical physical properties of the PP/IOHMMT nanocomposites is listed in Table 18. The
results showed that the activity of the catalysts could be adjusted
by changing polymerization parameters as well as the molar ratio
of MgCl2 to the surfactant [124].
The isotacticity index and differential scanning calorimetry
(DSC) results indicate that highly isotactic PP can be produced by
using the above procedures. The I.I. values of PP in the nanocomposites are as high as 94%, while the melting points of the PP composites
are as high as 161 C. The molecular weight of PP in these nanocomposites varies typically from 420 to 610 kg/mol. The polydispersity
index (PDI) of PP was in the range of 68 [124].
The verication of the existence of the new complex formed by
IOH and MgCl2 nROH has become the key point in order to prove
that the ZieglerNatta catalyst can be enriched inside the galleries
of MMT. FT-IR, as an important method in the characterization of
the complex formation, was used to characterize the interaction
between IOH and MgCl2 nROH. In order to avoid the disturbance of
MMT absorption peaks on the analysis of the complex formation in
FT-IR spectra, IOH is mixed with MgCl2 nROH instead of IOHMMT
to form the complex-I. The FT-IR spectra of IOH and complex-I are
shown in Fig. 22. It can be seen in Fig. 22(b) that the C H absorption band of aromatic rings at 800900 cm1 became signicantly
weak, and at the same time, a signicant shift of the C O H absorption band from 1066 to 1031 cm1 was observed. These results
indicate that a new complex was formed and MgCl2 nROH could
be bonded with IOH in the galleries of MMT through the hydrogen bond between the hydroxyl groups of the alkylimidazolium
surfactant and MgCl2 nROH [124].
XPS was used to reveal the chemical environment of Ti in
the nal compound catalyst. The results showed that the chemical environment of Ti (Ti2p1/2 465.1 eV; Ti2p3/2 458.9 eV) in the
catalyst was essentially the same as the commercial TiCl4 /MgCl2
catalyst (Ti2p1/2 465.0 eV, Ti2p3/2 459.2 eV), indicating that the
majority of bonding occurred between TiCl4 and MgCl2 . Both FT-IR
and XPS results indicate that a controlled amount of ZieglerNatta
catalyst could be formed in the galleries of MMT [124].
As shown in Fig. 23, TEM images illustrate an exfoliation of
clay into very small stacks, including single clay plates. The particle size distribution was fairly uniform. Micron-sized clays were
scarcely observed in the PP matrix. When the IOHMMT content
was increased to 19.0 wt%, as can be seen in Fig. 23(d), only a small
fraction of the MMT was stacked. This indicates that propylene

Fig. 22. FT-IR spectrum of (a) the surfactant and (b) MgCl2 nROH compound/surfactant (molar ratio = 1:1, n = 2.8) [124].

polymerization predominantly occurred inside the galleries of the


silicate layers of MMT and contributed to the exfoliation of the clay
layers [124].
KeFang [125] used the MMT/MgCl2 /TiCl4 catalyst for the copolymerization of propylene and 5-hexenyl-9-BBN. The synthetic
route to the PP/MMT nanocomposites is shown in Fig. 24.
The reaction conditions and properties of the nanocomposites
are summarized in Table 19. The intercalated MMT/MgCl2 /TiCl4
catalyst was quite effective in the copolymerization of propylene
and 5-hexenyl-9-BBN [125].
Fig. 25 compares 1 H NMR spectra of the two nanocomposite samples. The chemical shift around 3.6 ppm assignable to the
CH2 OH group was obvious in the 1 H NMR spectrum of PPOH/OMMT (Run 2) nanocomposites. This chemical shift did not
appear in the neat PP/OMMT (Run 1) spectrum. These results
Table 19
Conditions and results of propylene homo-and co-polymerization with 5-hexenyl9-BBN [125].
Run

Comonomer
(mmol)

Activity
(kgPP/molTi h)

Clay
(wt.%)

()

Tm ( C)

Hm
(J/g)

1
2

0
18.0

7.8
9.5

1.30
1.34

3.5
3.4

159.5
154.7

75.0
54.2

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

401

Fig. 23. TEM images of PP/MMT nanocomposites with different amounts of MMT (a) 2.0 wt% (b) 3.0 wt% (c) 5.0 wt% (d) 19.0 wt% [124].

proved that hydroxyl groups were introduced into the PP matrix.


The copolymer composition was calculated by the comparison
of the peak intensity of the chemical shift around 3.6 ppm and
0.52 ppm. Thus, the hydroxyl group content in the PP/MMT
nanocomposite sample was found to be 0.15 mol% [125].
Baniasadi and co-workers investigated the production of the
PP/clay nanocomposites using the clay/Mg(OEt)2 /TiCl4 bisupported
catalyst and compared them to the PP/clay nanocomposites produced by the melt blending method with the same clay content
[115]. Mg(OEt)2 , as the second support which is converted to MgCl2

during the titanation with TiCl4 , was used to avoid the direct contact of the active sites with the clay [107]. The activity of the above
catalyst in the propylene polymerization was not reported.
Marques et al. [126] reported the use of the ZieglerNatta catalyst based on the clay/MgCl2 -bisupported catalyst to prepare the
PP/clay nanocomposites. Three catalysts, containing catalyst without clay (Catalyst-1), catalyst with the untreated clay (Catalyst-2)
and catalyst with the clay treated by triethylaluminum (Catalyst-3)
were synthesized. Table 20 shows the results obtained through the
polymerization carried out with the prepared catalysts, as well as

Fig. 24. Preparation of PP/MMT nanocomposites via in situ intercalative copolymerization of propylene and 5-hexenyl-9-BBN [125].

402

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

Fig. 25. 1 H NMR spectra of neat PP/MMT (Run 1) and PP/MMT (Run 2) nanocomposite samples [125].

the characterization of catalysts in terms of titanium content and


the polymer isotacticity index [126].
Catalyst-2 presented the increased amount of titanium in comparison with that of the reference catalyst (Catalyst-1), while
Catalyst-3 has decreased the titanium content. The increased value
was probably due to the xation of high amounts of Ti in the
hydroxyl groups of the untreated clay, although it is well known
that the obtained SiOTiCl3 sites in the clay platelets surface will
not act as an active polymerization catalyst after introduction of
TEA in the reaction medium as a cocatalyst [127]. Actually, the catalyst activity was low despite of its higher Ti content. On the other
hand, the alkyl aluminum pretreatment performed on the clay has
decreased the Ti content in the catalyst, meaning that the clay was
inert for the xation of TiCl4 . In fact, an impregnation of the clay
has been directly performed with TiCl4 and the amount of Ti xed
in this material was less than 0.5 wt%. Therefore, there is an actual
active support for TiCl4 in the bisupported ZieglerNatta catalyst
[126].
Table 20 also shows that the addition of the commercial clay led
to a great decrease in activity for Catalysts-2 and 3 compared with
Catalyst-1. This may be due to the connement of the active sites
in the interlayer of the clay at the beginning of the polymerization.
Meanwhile, the stereospecicity without the clay was higher than
that of the clay-supported catalyst [126].

Table 20
Results of polymerization yield and the characterization analysis of catalysts and
the obtained polypropylenes [126].
Catalyst

Yield (gPP/gCat)

I.I. (%)

Catalyst-1 (2.4 wt% of Ti)


Catalyst-2 (5.5 wt% of Ti)
Catalyst-3 (1.4 wt% of Ti)

100140
1418
1113

9398
8287
7886

Fig. 26. Methine region of the 13 C NMR spectra of polypropylene (a) isotactic and
(b) atactic fractions of the composite (9.5 wt% ller) [128].

Dias et al. synthesized the clay/MgCl2 /BuOH/EB/TiCl4 catalyst for the propylene polymerization in which two types of the
clay (sodium MMT (Gelmax 400) and organically modied MMT
(Cloisite-30B)) were used [128].
The polymerization activity had a large variation depending on
the catalyst and reaction conditions. Although low activity was
observed in some cases, the signicant activity was obtained using
the sodium clay as a support without internal donor (e.g. 6.4 kg PP/g
Ti) [127].
The supported catalysts prepared with sodium and organophilic
clays showed different colors, put in contact with air, indicating
that different active species were formed depending on the type
of clay used. However, the catalysts were maintained under an
inert atmosphere to avoid the deactivation by moisture and oxygen
[128].
The catalysts which excluded low activity produced polymers
with a high content of the ller (clay). For instance, the polymerization using the catalyst prepared with the organophilic clay
without internal donor (MMT/MgCl2 /TiCl4 /TEA) and reaction times
of 60 min resulted in small yield, generating polypropylene composites with 76.3 wt%. In contrast, the polymerizations carried out
with the catalyst prepared with the similar clay containing EB as the
internal donor (MMT/MgCl2 /EB/TiCl4 /TEA) showed better yields
and activities (e.g. 669 g PP/g Ti h) [128].
In some cases, the produced PP did not show the characteristic
peak attributed to clay d-spacing, suggesting a high degree of clay
exfoliation. Nevertheless, the isotacticity index evaluated by the
heptane soluble fraction did not surpass 74.5% [128].
Reactions performed with the catalysts synthesized with
sodium clay, EB as internal donor and TEA as cocatalyst
(MMT/MgCl2 /EB/TiCl4 /TEA) showed extremely low yields. These

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

Fig. 27. Dynamic storage modulus (E ) of PP and PP/clay nanocomposites [118].

403

Fig. 29. Thermogravimetric analysis of pure PP and PP/clay nanocomposites [118].

results were attributed to the low Ti impregnation in the catalysts.


This fact was observed particularly for the catalysts made from
sodium clay and EB [128].
Fig. 26 shows the methyl region (2022 ppm) of the 13 C NMR
spectra of heptane soluble and insoluble fractions of polypropylene. These spectra demonstrated that the insoluble fraction was a
very isotactic PP and the soluble fraction was completely the atactic
polymer [128].
The pentad distribution for these fractions showed that the
polypropylene isotactic chains contained signicant amounts of
stereo errors. The mmmm pentad content was ranging from 73
to 88% which was lower than that expected for the isotactic fraction obtained through the third generation ZieglerNatta catalyst
[95,129]. As expected, the content of isotactic mmmm pentad
increases to 88% using the internal donor while it is 73% when the
donor is not used [128].
3.3.2.3. Properties of PP/clay nanocomposites.
3.3.2.3.1. Physical properties. The physical properties of the
PP/clay nanocomposites may be inuenced by the presence of
the clay in the nanocomposites. The value of Tm of the PP/clay
nanocomposites is nearly close to the pure PP; whereas, the value
of Xc of such nanocomposites may change. As shown in Table 21,
the value of Tm of the PP/MMT nanocompoites was around 161 C
and that of their Xc was about 34% which slightly less than the pure
PP [124].

Fig. 30. Dependency of derivation weight loss on temperature of PP and PP/clay


nanocomposites [118].

3.3.2.3.2. Mechanical properties. The tensile test results


showed that the yield strength and tensile modulus of in situ
prepared samples were considerably improved by the clay loading.
Another conclusion from the tensile test is that an increase in the
yield strength and tensile modulus of the in situ polymerized sample relative to the melt blended sample caused better dispersion of
the clay in the polymer matrix. Consequently, a stronger interfacial
adhesion was created between the polymer and the exfoliated clay
layers [115].
The storage modulus (E ) versus temperature (T) spectra of the
PP and PP/clay nanocomposites are shown in Fig. 27. The nanocomposites had higher E values than those of the pure PP over the
whole temperature range. To clarify the effect of the clay on the
E , the storage modulus of PP/MMT relative to that of the pure PP
(E PP/MMT /E PP ) was plotted in Fig. 28 [118]. Whit increasing the clay
content, the PP/clay nanocomposites show a substantial increase in
E , especially at temperatures higher than the glass transition temperature (Tg ). The E is almost three times bigger than the pure PP
one by introducing about 8 wt% of the clay. (Table 22)
Table 21
Physical properties of the PP/MMT
TiCl4 /MgCl2 /IOHMMT catalyst [124].

Fig. 28. Relative dynamic storage modulus (E co /E PP ) as a function of temperature
[118].

nanocomposites

produced

by

Clay content (%)

Tm ( C)

H (J/g)

Xc (%)

2
3
5

161
161
159

69.8
69.1
72.1

33.7
33.4
34.8

404

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

Fig. 31. SEM micrographs of Cloisite 30B (a) 500 and (b) 1000; and catalyst prepared with the clay (c) 500 and (d) 1000 [134].

Fig. 32. SEM micrographs of Gelmax 400 (a) 500 and (b) 1000; and catalyst prepared with the clay (c) 500 and (d) 1000 [134].

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

Fig. 33. SEM micrographs of PP particles obtained with catalyst supported on sodium MMT (a) 70 (b) 150 (c) 500 and (d) 1000 [134].

Fig. 34. SEM micrographs of PP particles obtained with catalyst supported on organically modied MMT (a) 70 (b) 150 (c) 500 and (d) 1000 [134].

405

406

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

Fig. 35. SEM micrograph of (a) clay (b) catalyst [135].

3.3.2.3.3. Thermal properties. According to Table 22, the HDT


and Tg values of the nanocomposites are increased with increasing
the clay contents. By introducing about 8 wt% of the clay, the HDT
of the nanocomposite increased by 40 C over that of the pure PP
[118].
A comparison of the thermal decomposition temperature of the
PP/clay nanocomposites with that of the pure PP shows that the
thermal stability of the nanocomposites is much higher than that of

the pure PP (Fig. 29). The onset of the thermal decomposition temperature of the nanocomposite was shifted to higher temperatures
by increasing the clay contents (Fig. 30). [118].
Fig. 31 shows the dependency of derivation weight loss on the
temperature of PP and PP/clay nanocomposites. It shows that, by
introducing about 10 wt% of the clay in the nanocomposite, the
temperature of maximum derivative weight loss increased by 44 C
over that of the pure PP.

Table 22
Effect of the clay content on E , Tg and HDT of PP/clay nanocomposites [118].

3.3.3. Morphology of nanocomposites


The control of the morphology of the polyolens powder is one
of the challenging factors in the olens polymerization due to their
morphology inuences on the polymerization process. It is well
known that the polymer particles tend to replicate the shape of
the ZieglerNatta catalysts, produced during the polymerization,
so called replication phenomenon. In other words, the catalyst
particles act as a template for the growth of the polymer particles
[130132]. In the polyolen production processes, the spherical

E (GPa)

Samples

PP
PP/MMT (2.5 wt%)
PP/MMT (4.6 wt%)
PP/MMT (8.1 wt%)

Tg ( C)

HDT ( C)

6.2
12.1
9.0
8.0

110
138
144
151

40 C

20 C

80 C

120 C

1.42
2.08
2.22
2.43

0.38
0.76
0.82
0.98

0.21
0.36
0.46
0.54

0.12
0.24
0.28
0.36

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

shape is the favorite morphology for the polymerization processes.


This morphology can be easily obtained through the third and
fourth generations of the ZieglerNatta catalyst. In addition, the
polymerization conditions have deeply effects on the morphology
of the polymers [96,129,133].
The literature survey shows that the most of the claysupported catalysts have displayed an irregular morphology
[108,109,134,135]. The SEM micrographs of the clays and typical
catalysts prepared from them are shown in Figs. 3133. Particles of
the neat Cloisite 30B (Fig. 31a and b) appear to be compact particles
with a large distribution of sizes and average diameter about 15 m.
Particles of the sodium clay GELMAX 400 presents larger particles
(average around 20 m) (Fig. 32a and b) with more regular shape
and less ne particles than the organophilic clay [134].
Fig. 31c and d clearly shows the marked changes in the shape
of the particles when comparing the neat organophilic clay and
the treated clays (catalysts). This result is supported by the micrograph of the catalyst in which particles present completely irregular
contours, different from the neat clay, indicating marked changes
in their structure and probable partial platelet exfoliation or the
beginning of this process [134].
In the catalyst, prepared from the sodium clay (Fig. 32c and d),
more regular shapes and larger particles are observed. The micrograph also shows less ne content than the one observed for the
catalyst. The results indicate that more signicant changes in the
particle morphology took place after the treatment with the metal
chlorides for the organophilic clay [134].
The micrographs of polymer particles obtained from propylene
polymerization with the Cloisite 30B-based catalyst (Fig. 33) show
that these particles have shapes similar to those of the catalyst
particles, suggesting the replica phenomenon. In this case, polymer particles with dimensions around 400 m and the formation
of microbrils (Fig. 33d) with diameter smaller than 10 m are seen
[134].
The micrographs in Fig. 34 show that the polymers produced
by the sodium clay-supported catalyst seem to have replicated the
shape of the catalyst particles. The gures also showed a certain
degree of porosity in the polymer particles. This can be attributed
to the catalyst particle fragmentation because of the pressure
exercised by the growing polymer chain during polymerization.
Although the polymer particles replicate the catalyst particles, this
replica takes place even by the catalyst particle fragmentation and
polymer particle with a dimension over 600 m [134].
The SEM studies on the clay show that the surface of the clay is
ag-like morphology [108,135]. Similarly, the catalyst displays also
the irregular and ag-like shape (Fig. 35).

4. Conclusions
The synthesis of the polymer nanocomposites is an integral
aspect of the polymer nanotechnology. By inserting the nanometric
inorganic compounds, the properties of polymers improve, hence
having a lot of applications depending on the inorganic materials present in the polymers. The improvements obtained from the
polyolens/clay nanocomposite structure can make this commercial thermoplastic polymer more suitable for many applications,
albeit in a low-prole manner and slower than had been anticipated. This pace is expected to speed up dramatically, as indicated
by the enthusiasm of researchers and industry around the world.
Among the melt blending, solution and in situ polymerization
approaches for the preparation of the PE/clay nanocomposites, the
latest is more popular; however, it suffers from some deciencies.
The most critical restriction is that the clay-supported catalysts
have not shown acceptable performance in the olens polymerization.

407

Although the catalysts supported on the clay did not show


high activity in the polymerization, the catalysts supported on the
clay treated with a second support such as MgCl2 and BOM had
signicantly higher activity. In other words, the clay-bisupported
ZieglerNatta catalysts displayed a reasonable efciency in the
olen polymerization. In addition, the type of clay showed a
remarkable effect on the catalyst and polymerization performance.
XPS study shows that the Ti species in the bisupported catalysts are
mainly supported on the second support.
The polymerization conditions and the type of the activator
have a deep effect on the catalyst yield and its kinetic behavior
in the olen polymerization. The clay content of the polymer can
be controlled by changing the polymerization conditions.
The XRD and TEM analyses indicated the occurrence of exfoliation in the silicate layers and their well dispersion in the polymer
matrices by polymerizing of the olen monomer into the clay layers. This structure led to an improvement in the thermal stability
and mechanical properties of the polyolens/clay nanocomposites
compared to the pure polyolens.
The SEM images illustrate that the clay and catalyst have
irregular and ag-like shape. The morphology of the nanocomposite particles has been very similar to those of the catalyst particles
due to the replica phenomenon.
It was found that a part of the polymer could be extracted from
the polymer/nanocomposite produced through in situ polymerization. This indicates that there is a strong interaction between
the clay and polymer resulting from the formation of the chemical
bonds. This explains the desired mechanical properties obtained
from the polyolens/clay nanocomposites produced through in situ
polymerization.
At present, there are two main challenging matters for the
production of the polyolen/clay nanocomposites through in situ
polymerization by ZieglerNatta type catalysts. The rst matter
is the yield of the catalysts. Although the bisupported catalysts
show reasonable activity, the amount of the residue Ti in the
nanocomposites seems still high. Therefore, the catalysts should
be synthesized that show high activity based on the used Ti and
also produce the nanocomposites having reasonable clay contents
(e.g. 12 wt%). The second matter is the poor morphology and low
bulk density of the polymer powders.

References
[1] J. Peacock, Handbook of Polyethylene: Structures, Properties and Applications, rst ed., Marcel Dekker, New York, 2000.
[2] G. Piatti, Advances in Composite Materials, rst ed., Applied Science Publisher,
London, 1978.
[3] B. Sedlacek, Polymer Composites, rst ed., Walter de Gruyter, Berlin, 1986.
[4] S.S. Ray, M. Okamoto, Prog. Polym. Sci. 28 (2003) 15391641.
[5] M.A. Al-Eshaikh, M.I. Qureshi, Int. J. Phys. Sci. 8 (2013) 167174.
[6] M. Avella, M.E. Errico, G. Gentile, Macromol. Symp. 234 (2006) 170175.
[7] C.M. Chan, J. Wu, J.X. Li, Y.K. Cheung, Polymer 43 (2002) 29812992.
[8] C.B. Patil, U.R. Kapatil, D.G. Hundiwale, P.P. Muhalikar, J. Mater. Sci. 44 (2009)
31183124.
[9] G. Kozlov, Z. Aphashagova, A. Malamatov, G. Zaikov, Chem. Chem. Technol. 6
(2012) 113118.
[10] B. Andricic, T. Kovacic, S. Perinovic, A. Grigic, Macromol. Symp. 263 (2008)
96101.
[11] J.J. Park, Trans. Elec. Elec. Mater. 14 (2013) 4346.
[12] J. Longun, J.O. Iroh, J. Appl. Polym. Sci. 128 (2013) 14251435.
[13] C. Palumbo, A. Baldini, F. Cavani, P. Sena, M. Benincasa, M. Ferretti, D. Zaffe,
Micron 47 (2013) 158.
[14] M.L.Q.A. Kaneko, M.I. Felisberti, M.C. Goncalves, I.V.P. Yoshida, Polym. Comp.
34 (2013) 194203.
[15] M.H. Abdolrasouli, E. Behzadfar, H. Nazockdast, F. Sharif, J. Appl. Polym. Sci.
125 (2012) 435444.
[16] K.S.K.R. Patnaik, K.S. Devi, V.K. Kumar, Int. J. Chem. Eng. Appl. 1 (2010)
346353.
[17] S. Bocchini, A. Frache, Exp. Polym. Lett. 7 (2013) 431442.
[18] S. Sakthivel, B. Pitchuman, Particle Sci. Tech. Int. J. 29 (2011) 441449.
[19] I. Ghasemi, A.T. Farsheh, Z. Masooni, J. Vinyl Add. Technol. 18 (2012)
161167.

408

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409

[20] C. Liu, U.F. Luo, Z.X. Jia, B.C. Zhong, S.Q. Li, B.C. Guo, D.M. Jia, Int. J. Polym.
Mater. (62) (2013) 128132.
[21] M. Salami-Kalajahi, V. Haddadi-Asl, F. Behboodi-Sadabad, S. Rahimi-Razin, H.
Roghani-Mamaqani, Polym. Comp. 33 (2012) 215224.
[22] V. Sister, E. Ivannikova, S. Lomakin, A. Yamchuk, Chem. Petro. Eng. 47 (2011)
741748.
[23] M. Rahmat, P. Hubert, J. Nano Res. 1819 (2012) 117128.
[24] M. Knite, V. Tateris, A. Kiploka, J. Kaupuzs, Sens. Actuators A Phys. 110 (2004)
142149.
[25] X. Ye, P.J. Yao, M.M. Xi, L.Q. Zhang, Plast. Rub. Comp. 42 (2013) 5058.
[26] T.V. Kosmidou, A.S. Vatalis, C.G. Delides, E. Logakis, P. Pissis, G.C. Papanicolaou,
Exp. Polym. Lett. 2 (2008) 364372.
[27] X. Wang, J. Zhao, M. Chen, L. Ma, X. Zhao, Z.M. Dang, Z. Wang, J. Phys. Chem.
117 (2013) 14671474.
[28] M. Kanapitsas, H. Zois, A.Z. Stimoniaris, C.G. Delides, Nanotechnology 1 (2012)
582585.
[29] J. Cai, S. Liu, J. Feng, S. Kimura, M. Wada, S. Kusa, L. Zhang, Angew. Chem. Int.
Ed. 51 (2012) 20762079.
[30] W.E. Zyl, M. Garcia, B.A.G. Schrauwen, B.J. Kooi, J.T.M. Hosson, H. Verweij,
Macromol. Mater. Eng. 287 (2002) 106110.
[31] M. Sadeghi, G. Khanbabaei, A.H. Saedi-Dehaghani, M. Sadeghi, M.A. Aravand,
M. Akbarzadeh, S. Khatti, J. Membr. Sci. 332 (2008) 423428.
[32] A. Zhu, A. Cai, J. Zhang, H. Jia, J. Wang, J. Appl. Polym. Sci. 108 (2008)
21892196.
[33] S. Chuayjuljit, N. Sukkasem, Adv. Mater. Sci. 488489 (2012) 633637.
[34] D.M. Qi, Y.Z. Bao, Z.X. Weng, Z.M. Huang, Polymer 47 (2006) 46224629.
[35] M.J. Tommalieha, A.M. Zihlifa, G. Ragosta, J. Exp. Nanosci. 6 (2011)
652664.
[36] S.S. Ray, M. Okamoto, Macromol. Rapid Commun. 24 (2003) 815840.
[37] M. Alexandre, P. Dubois, Mater. Sci. Eng. 28 (2000) 163.
[38] D. Bikiaris, Materials 3 (2010) 28842946.
[39] W. Kaminsky, A. Funck, Macromol. Symp. 260 (2007) 18.
[40] M.W. Weirner, H. Chen, E.P. Giannelis, D.Y. Sogah, J. Am. Chem. Soc. 121 (1999)
16151616.
[41] Y.H. Jin, H.J. Park, S.S. Im, S.Y. Kwak, S. Kwak, Macromol. Rapid. Commun. 23
(2002) 135140.
[42] M. Oliveira, M.F. Marques, Chem. Chem. Technol. 5 (2011) 201207.
[43] Y. Huang, K. Yang, J.Y. Dong, Macromol. Rapid. Commun. 27 (2006)
12781283.
[44] S.S. Ray, G. Galgali, A. Lele, S. Sivaram, J. Polym. Sci. Part A: Polym. Chem. 43
(2005) 304318.
[45] Q.Y. Wei, D.J. Yong, Chin. Sci. Bull. 53 (2009) 3843.
[46] S. Abedi, M. Abdouss, J. Appl. Polym. Sci. 128 (2013) 18791884.
[47] S.Y.A. Shin, L.C. Simon, J.B.P. Soares, G. Scholz, Polymer 44 (2003)
53175321.
[48] D.H. Lee, H.S. Kim, K.B. Yoon, K.E. Min, K.H. Seo, S.K. Noh, Sci. Technol. Adv.
Mater. 6 (2005) 457462.
[49] J.M. Hwu, G.J. Jiang, J. Appl. Polym. Sci. 95 (2005) 12281236.
[50] A. Maneshi, J.B.P. Soares, L.C. Simon, Macromol. Chem. Phys. 212 (2011)
20172028.
[51] F.C. Fim, J.M. Guterres, N.R.S. Basso, G.B. Galland, J. Polym. Sci. Part A: Polym.
Chem. 48 (2010) 692698.
[52] G. Leone, F. Bertini, M. Canetti, L. Boggioni, P. Stagnaro, I. Tritto, J. Polym. Sci.
Part A: Polym. Chem. 46 (2008) 53045390.
[53] T.S. Halbach, R. Mulhaupt, Polymer 49 (2008) 867876.
[54] Z.V.P. Murthy, P.A. Parikh, Quim. Nova 34 (2011) 11571162.
[55] V. Mittal, In-situ Synthesis of Polymer Nanocomposites, Wiley-VCH Verlag &
Co. KgaA, Weinheim, 2012.
[56] J. Heinemann, P. Reichert, R. Thomann, R. Mulhaupt, Macromol. Rapid Commun. 20 (1999) 423430.
[57] S.L. Scott, B.C. Peoples, C. Yung, R.S. Rojas, V. Khanna, H. Sano, T. Suzuki, F.
Shimizu, Chem. Commun. (2008) 41864188.
[58] A. He, L. Wang, W. Yao, B. Huang, D. Wang, C.C. Han, Polym. Degrad. Stab. 95
(2010) 651656.
[59] V. Monteil, J. Stumbaum, R. Themann, S. Mecking, Macromolecules 39 (2006)
20562062.
[60] M.L. Mignoni, J.V.M. Silva, M.O. Souza, R.S. Mauler, R.F. Souza, K.B. Gusmao, J.
Appl. Polym. Sci. 122 (2011) 21592165.
[61] S. Ray, G. Galgali, A. Lele, S. Sivaram, J. Polym. Sci. Part A: Polym. Chem. 43
(2005) 304318.
[62] C. Oriakhi, Chem. Br. 34 (1998) 5962.
[63] J.C.W. Chien, D. He, J. Polym. Sci. Polym. Chem. Ed. 29 (1991) 16031607.
[64] K. Carrado, A.L. Xu, Chem. Mater. 101 (1998) 14401445.
[65] S. Abedi, V. Haddadi-Asl, Olens Coordination Polymerization, NPC Press,
Tehran, 2007.
[66] L. Cui, H.Y. Cho, J.W. Shin, N.H. Tarte, S.I. Woo, Macromol. Symp. 260 (2007)
4957.
[67] C. Liu, T. Tang, D. Wang, B. Huang, J. Polym. Sci. Part A: Polym. Chem. 41 (2003)
21872196.
[68] Y.J. Huang, Y.W. Qin, J.Y. Dong, X. Zhao, X. Hu, J. Appl. Polym. Sci. 123 (2012)
31063116.
[69] M. Alexandre, E. Martin, P. Dubois, M. Garcia-Marti, R. Jerome, Macromol.
Rapid Commun. 21 (2000) 931936.
[70] R. Xalter, T.S. Halbach, R. Mulhaulp, Macromol. Symp. 236 (2006) 145151.
[71] J. Wang, Z. Lui, C. Cuo, Y. Chen, D. Wang, Macromol. Rapid Commun. 22 (2001)
14221426.

[72] J. Quian, C.Y. Cuo, Open Macromol. J. 4 (2010) 114.


[73] A. Carrero, R. Grieken, I. Suarez, B. Paredes, J. Appl. Polym. Sci. 126 (2012)
987997.
[74] C. Liu, T. Tang, Z. Zhao, B. Huang, J. Polym. Sci. Part A: Polym. Chem. 40 (2003)
18921898.
[75] S.W. Kuo, W.J. Huang, S.B. Huang, H.C. Kao, F.C. Chang, Polymer 44 (2003)
77097719.
[76] W. Kaminsky, H. Winkelbach, Top. Catal. 7 (1999) 6167.
[77] D.A. Estenoz, M.G. Chiovetta, J. Appl. Polym. Sci. 81 (2001) 285311.
[78] G. Fink, B. Steinmetz, J. Zechlin, C. Przybyla, B. Tesche, Chem. Rev. 100 (2000)
13771390.
[79] D.G.H. Ballard, E. Jones, R.J. Wyatt, R.T. Murray, P.A. Robinson, Polymer 15
(1974) 169174.
[80] E.G. Howard, B.L. Glazar, J.W. Collette, Ind. Eng. Chem. Prod. Res. Dev. 20 (1981)
421428.
[81] A. Maneshi, J.B.P. Soares, L.C. Simon, Macromol. Chem. Phys. 212 (2011)
216228.
[82] J.T. Xu, Q. Wang, Z.Q. Fan, Eur. Polym. J. 41 (2005) 30113017.
[83] L.M. Wei, T. Tang, B.T. Huang, J. Polym. Sci. Part A: Polym. Chem. 42 (2004)
941949.
[84] Q. Wang, Z.Y. Zhou, L.X. Song, H. Xu, L. Wang, J. Polym. Sci. Part A: Polym.
Chem. 42 (2004) 3843.
[85] D.W. Jeong, D.S. Hong, H.Y. Cho, S.I. Woo, J. Mol. Catal. A: Chem. 206 (2003)
2052011.
[86] T. Sun, J.M. Garces, Adv. Mater. 14 (2002) 128130.
[87] L.K. Johnson, C.M. Killian, M. Brookhart, J. Am. Chem. Soc. 117 (1995)
64146415.
[88] L.K. Johnson, C.M. Killian, M. Brookhart, J. Am. Chem. Soc. 118 (1996)
267268.
[89] B.L. Small, M. Brookhart, A.M.A. Bennet, J. Am. Chem. Soc. 120 (1998)
40494050.
[90] G. Leone, F. Bertini, M. Canetti, L. Boggioni, L. Canzatti, I. Tritto, J. Polym. Sci.
Part A: Polym. Chem. 47 (2009) 548564.
[91] F.A. He, L.M. Zhang, Compos. Sci. Technol. 67 (2007) 32263232.
[92] J.S. Bergman, G.W. Coates, H. Chen, E.P. Giannelis, M.G. Thomas, Chem. Commun. 21 (1999) 21792180.
[93] J. Rong, H. Li, Z. Jing, X. Hong, M. Sheng, J. Appl. Polym. Sci. 82 (2001)
18291837.
[94] D. Damyanov, M. Velikava, Eur. Polym. J. 15 (1979) 10751078.
[95] S. Abedi, M. Daftari-Besheli, S. Shaei, J. Appl. Polym. Sci. 97 (2005)
17441749.
[96] P. Moore, Polypropylene Handbook, rst ed., Hanser, Munich, 1996.
[97] K. Soga, M. Terano, Catalyst Design for Tailor-Made Polyolens, rst ed., Elsevier, Amsterdam, 1994.
[98] A. Ramazani, F. Tavakolzadeh, Macromol. Symp. 274 (2008) 6571.
[99] F. Yang, X. Zhang, H. Zhao, B. Chen, B. Huang, Z. Feng, J. Appl. Polym. Sci. 89
(2003) 36803684.
[100] C.B. Yang, H.S. Chang, K.H. Lee, WO 2005/082950 (2005).
[101] M. Ferraris, F. Rosati, US 4469648 (1984).
[102] M.C. Forte, F.M.B. Coutinho, Eur. Polym. J. 32 (1996) 223231.
[103] T. Toida, T. Shinozaki, M. Kioka, US 5887256 (1999).
[104] M. Ohgizawa, M. Koika, US 5844046 (1998).
[105] J. Rong, Z. Jing, H. Li, M. Sheng, Macromol. Rapid Commun. 22 (2001)
329334.
[106] F.A. He, L.M. Zhang, F. Yang, L.S. Chen, Q. Wu, J. Macromol. Sci. Part A: Pure
Appl. Chem. 44 (2007) 1115.
[107] L. Cui, S.I. Woo, Polym. Bull. 61 (2008) 453460.
[108] S. Abedi, M. Abdouss, M. Nekkomanesh-Haghighi, N. Sahri-Sanjani, Polym.
Bull. 70 (2013) 13131325.
[109] S. Abedi, M. Abdouss, M. Daftari-Besheli, S.M. Ghafelehbashi, F. Azadi, S.R.
Nokhbeh, A. Sanejad, Polym. Bull. 70 (2013) 27832792.
[110] L.L. Bohm, Polymer 19 (1978) 553561.
[111] J.C. Chien, T.J. Nozaki, J. Polym. Sci. Part A: Polym. Chem. 29 (1991)
505514.
[112] M.I. Nikolaeva, T.B. Mikenas, M.A. Matsko, L.G. Echevskaya, V.A. Zakharov, J.
Appl. Polym. Sci. 122 (2011) 30923101.
[113] S. Abedi, N. Hassanpour, J. Appl. Polym. Sci. 101 (2006) 14561462.
[114] K.J. Chu, J.B.P. Soares, A. Penlidis, S.K. Ihm, Eur. Polym. J. 36 (2000)
311317.
[115] H. Baniasadi, A. Ramazani, S. Javannikkhah, Mater. Des. 31 (2010) 7684.
[116] M. Inoue, J. Polym. Sci. Part A: Gen. Papers 1 (1963) 26972709.
[117] S.M. Kurtz, The UHMWPE Handbook, rst ed., Elsevier Academic Press, London, 2004.
[118] J. Ma, Z. Qi, Y. Hu, J. Appl. Polym. Sci. 82 (2001) 36113617.
[119] F.S. Dyachkovskii, L.A. Novokshonaova, Russ. Chem. Rev. 53 (1984)
117131.
[120] F.S. Dyachkovskii, Trends Polym. Sci. 1 (1993) 2741984.
[121] X. Kornmann, H. Lindberg, L.A. Berglund, Polymer 42 (2001) 13011310.
[122] H.C. Zhao, X.Q. Zhang, F. Yang, B. Chen, Y.T. Jin, G. Li, Z.L. Feng, B.T. Huang,
Chin. J. Polym. Sci. 21 (2003) 413418.
[123] H.L. Liu, S.J. Xiao, Makromol. Chem. 194 (1993) 421429.
[124] K. Du, A.H. He, X. Liu, C.C. Han, Macromol. Rapid Commun. 28 (2007)
22942299.
[125] Y. KeFang, H. YingJuan, D. JinYong, Chin. Sci. Bull. 52 (2007) 181187.
[126] M.F. Marques, J. Rosa, K. Cruz, Chem. Chem. Technol. 5 (2011) 433438.
[127] K. Suga, T. Sano, S. Ikeda, Polym. Bull. 21 (980) 817819.

S. Abedi, M. Abdouss / Applied Catalysis A: General 475 (2014) 386409


[128] M.L. Dias, R.M. Fernandes, R.H. Cunha, S. Jaconis, A.C. Silvino, Appl. Catal. A:
Gen. 403 (2011) 4857.
[129] N. Pasquini, Polypropylene Handbook, 2nd ed., Hanser Publications, Munich,
2005.
[130] G. Cecchin, E. Marchetti, G. Macromol, Chem. Phys. 202 (2001) 19871994.
[131] M. Monji, S. Abedi, S. Pormahdian, F. Afshar-Taromi, J. Appl. Polym. Sci. 112
(2009) 18631867.

409

[132] M.P. McDaniel, J. Polym. Sci. Polym. Chem. Ed. 19 (1981) 19671976.
[133] M. Monji, S. Pourmahdian, M. Vatankhah, F. Afshar-Taromi, J. Appl. Polym. Sci.
112 (2009) 36633668.
[134] M.L. Dias, R.M. Fernandes, R.H. Cunha, S. Jaconis, A.C. Silvino, Appl. Catal. A
Gen. 403 (2011) 4857.
[135] S. Abedi, M. Abdouss, M. Daftari-Besheli, A. Moghimi, S.M. Ghafelehbashi, M.A.
Pourian, J. Inorg. Org. Polym. Mat. (2013) DOI:10.1007/s10904-013-9982-y.

Vous aimerez peut-être aussi