Você está na página 1de 69

PREPARATION AND LONG-TERM

PERFORMANCE OF POLY(ETHYLENE-CO-
BUTYL ACRYLATE) NANOCOMPOSITES AND
POLYETHYLENE

Sohail Nawaz
Doctoral Thesis


Kungliga Tekniska Hgskolan, Stockholm 2012



AKADEMISK AVHANDLING


Som med tillstnd av Kungliga Tekniska Hgskolan i Stockholm framlgges till
offentlig granskning fr avlggande av teknisk doktorsexamen fredagen den 30
november 2012, kl. 13.00 i sal F3, Lindstedtsvgen 23, KTH, Stockholm.
Avhandlingen frsvaras p engelska. Opponent: Professor Mikael Skrifvars,
University of Bors.

























Copyright 2012 Sohail Nawaz
All rights reserved
Paper I 2012, Journal of Applied Polymer Science
Paper II 2012, Polymer Degradation and Stability
Paper III submitted, 2012, Polymer Degradation and Stability
Paper IV submitted, 2012, Polymers from Renewable Resources
Paper V manuscript, to be submitted to Polymer Testing


TRITA-CHE Report 2012:44
ISSN 1654-1081
ISBN 978-91-7501-491-3


























I feel honored to be considered as a Ph.D. candidate in the department of
Fibre and Polymer Technology at KTH, where I got the opportunity to work
with both academia and industry.




Sohail Nawaz
November 1, 2012- Stockholm
KTH, Royal Institute of Technology

ABSTRACT

The current study discusses the preparation and long-term performance of
polymer composites used for various purposes under different ageing conditions.

The first part deals with the preparation and characterization of polymer
nanocomposites based on poly(ethylene-co-butyl acrylate) (EBA13 and EBA28
with 13 and 28 wt % butyl acrylate, respectively) and 212 wt % (0.53 vol %) of
aluminum oxide nanoparticles (two types with different specific surface areas and
different hydroxyl-group concentrations; uncoated and coated with, respectively,
octyltriethoxysilane and aminopropyltriethoxysilane). The nanocomposite with
EBA13 showed better overall nanoparticle dispersion while EBA28 resulted in
poor dispersion, probably due to insufficiently high shear forces acting during
extrusion mixing which were unable to break down nanoparticle agglomerates.

The activity of hindered phenolic antioxidant (0.2 wt%) in all EBA
nanocomposites was assessed by determining the oxidation induction time using
DSC. The composites containing uncoated aluminium oxide nanoparticles showed
a much shorter initial OIT than the pristine polymer with the same initial
concentration of antioxidant, indicating adsorption of antioxidant onto the
nanoparticle surfaces. Composites containing coated nanoparticles showed a
significantly smaller decrease in the initial OIT, suggesting the replacement of
hydroxyl groups with organic silane tails, decreasing the concentration of
available adsorption sites on the nanoparticle surfaces. The decrease in OIT with
increasing ageing time in dry air at 90 C of the nanocomposites was slower than
that of the unfilled pristine polymer, suggesting a slow release of antioxidant from
adsorption sites.

The EBA nanocomposites exposed to liquid water at 90C showed faster decrease
of OIT than samples exposed to dry or humid air. The migration rate of
antioxidant was controlled by the boundary conditions in the case of ageing in
humid air and liquid water. The antioxidant diffusivity was lower for the
composites containing uncoated ND than for the composites containing ND
coated with octyltriethoxysilane or aminopropyltriethoxysilane.

The migration and chemical consumption of deltamethrin DM, (synthetic
pyrethroid) and synergist piperonyl butoxide from molded polyethylene sheets
was also studied. Deltamethrin and piperonyl butoxide are often used for food

storage and insect control purposes. DM showed no signs of crystallization and


remained in a liquid state after being cooled to room temperature. Exposure of
polyethylene compound sheets to liquid water (at 80 & 95 C), caused degradation
and hydrolysis of the ester bond in the DM, present in the prepared material, and
generated species containing hydroxyl groups. Liquid chromatography and
infrared spectroscopy showed a significant migration of the active species in
liquid water, whereas in air at 80 C (60 and 80 %RH) the loss of DM and PBO was
negligible over 30 days.

The long-term performance of medium-density polyethylene stabilized with six
different phenolic antioxidants (0.1 wt%) in aqueous chlorinated media at 70 C
was studied. The results were compared with data for previously studied
solutions of antioxidants in squalane (a liquid, low molar mass analogue of
polyethylene). A linear relationship was established between the time to reach
antioxidant depletion in polyethylene tape samples and the time in squalane
samples. Infrared spectroscopy and scanning electron microscopy of drawn
samples revealed the onset of surface oxidation and surface embrittlement in tape
samples exposed beyond the time for antioxidant depletion.

Keywords: Al2O3, EBA, Polymer nanocomposites, Ageing, Antioxidant, OIT, Migration
of stabilizer, Deltamethrin, Piperonyl butoxide, Pyrethroid, Polyethylene, Chlorine
dioxide, long-term performance















LIST OF PAPERS
This thesis is a summary of the following papers:

I.

II.
III.

IV.

Preparation and Characterization of Aluminum OxidePoly(ethylene-


co-butyl acrylate) Nanocomposites P. Nordell, S. Nawaz, B. Azhdar, H.
Hillborg and U. W. Gedde, Journal of Applied Polymer Science, Vol 125,
975-983, 2012.
Antioxidant Activity in Aluminium OxidePoly(ethylene-co-butyl
acrylate) Nanocomposites S. Nawaz, P. Nordell, H. Hillborg and U.W.
Gedde, Polymer Degradation and Stability, Vol 97, 1017-1025, 2012.

Migration of Phenolic Antioxidant from Aluminium Oxide
Poly(ethylene-co-butylacrylate) Nanocomposites in Aqueous Media S.
Nawaz, H. Hillborg, M.S. Hedenqvist and U.W. Gedde, submitted to
Polymer Degradation and Stability.
Migration and Chemical Consumption of Deltamethrin and Piperonyl
Butoxide from Polyethylene in Aqueous Media Sohail Nawaz, Nazanin
Alipour, Mikael S. Hedenqvist and Ulf W. Gedde, submitted to Polymers
from Renewable Resources.

V.





Assessing the Long-term Performance of Polyethylene Stabilised With


Phenolic Antioxidants Exposed to Water Containing Chlorine Dioxide
W. Yu, E. Sedghi, S. Nawaz, T. Hjertberg, J. Oderkerk, F.R. Costa,
U.W.Gedde, to be submitted to Polymer Testing.

Scientific papers also published during this period, which are not included in the
thesis:

VI.

Dielectric Properties of Alumina-filled Poly(ethylene-co-butyl acrylate)


Nanocomposites, part I - dry studies Nadja Jverberg, Hans Edin,
Patricia Nordell, Sohail Nawaz, Bruska Azhdar, Ulf Gedde, IEEE
Transactions on Dielectrics and Electrical Insulation, Vol. 19, Issue 2,
383-390, 2012.

VII. Dielectric Properties of Alumina-filled Poly(ethylene-co-butyl acrylate)

Nanocomposites, part II - wet studies Nadja Jverberg, Hans Edin,


Patricia Nordell, Sohail Nawaz, Bruska Azhdar, Ulf Gedde, IEEE
Transactions on Dielectrics and Electrical Insulation, Vol. 19, Issue 2,
391-399, 2012.


ABBREVIATIONS

EBA
AO
NDU
NDA
NDO
NAU
NAA
NAO
DM
PBO
DSC
FTIR
HDPE
LDPE
HPLC
SEM
XPS

Poly(ethylene-co-butyl acrylate), (13 & 28 wt% butyl acrylate)


Antioxidant
Nanodur Uncoated (nanoparticles)
Nanodur Aminopropyltriethoxysilane-coated
Nanodur Octyltriethoxysilane-coated
Nanoamor Uncoated
Nanoamor Aminopropyltriethoxysilane-coated
Nanoamor Octyltriethoxysilane-coated
Deltamethrin
Piperonyl Butoxide
Differential Scanning Calorimetry
Fourier Transform Infrared
High Density Polyethylene
Low Density Polyethylene
High Pressure Liquid Chromatography
Scanning Electron Micrography
X-ray photoelectron spectroscopy


TABLE OF CONTENTS

PURPOSE OF THE STUDY .............................................................................. 1
1 INTRODUCTION ........................................................................................ 2
1.1 POLYMER NANOCOMPOSITES BACKGROUND ...................................................... 2
1.2 POLY(ETHYLENE-CO-BUTYL ACRYLATE) ............................................................... 2
1.3 ALUMINIUM OXIDE NANOPARTICLES .................................................................... 3
1.4 SURFACE TREATMENT OF NANOPARTICLES .......................................................... 3
1.5 LONG-TERM PERFORMANCE OF POLYMER NANOCOMPOSITES ............................. 4
1.6 MIGRATION OF STABILIZER .................................................................................... 5
1.7 LOSS OF SYTHETIC INSECTICIDE FROM POLYETHYLENE ........................................ 6
1.8 LONG-TERM PERFORMANCE OF POLYETHYLENE IN AQUEOUS CHLORINE
DIOXIDE ................................................................................................................ 7

2 EXPERIMENTAL .......................................................................................... 9
2.1 MATERIALS ............................................................................................................. 9
2.1.1 Poly(ethylene-co-butyl acrylate) ...................................................................... 9
2.1.2 Aluminium oxide nanoparticles ...................................................................... 9
2.1.3 Chemicals & materials ................................................................................... 10

2.2 INSTRUMENTATION .............................................................................................. 11


2.3 SAMPLE PREPARATION AND PROCEDURES .......................................................... 12
2.3.1 Surface treatment of nanoparticles ................................................................ 12
2.3.2 Preparation of nanocomposite materials ....................................................... 13
2.3.3 Preparation of sandwiched composites .......................................................... 14
2.3.4 Preparation of polyethylene compounds ........................................................ 14
2.3.5 Sample extraction for HPLC ......................................................................... 14

2.4 THERMAL AGEING OR EXPOSURE CONDITIONS .................................................. 15


2.4.1 Al2O3/Poly(ethylene-co-butyl acrylate) ....................................................... 15
2.4.2 Polyethylene sheet compounds ...................................................................... 15

2.4.3 Exposure of squalane or polyethylene samples to chlorinated media ............ 15

3 RESULTS AND DISCUSSION ................................................................ 17


3.1 PREPARATION OF ALUMINUM OXIDE POLY(ETHYLENE-CO-BUTYL ACRYLATE)
NANOCOMPOSITES. ............................................................................................ 17
3.1.1 General characteristics of Al2O3 nanoparticles ............................................. 17
3.1.2 Dispersion of Al2O3 nanoparticles in EBA .................................................. 19

3.2 ANTIOXIDANT ACTIVITY IN ALUMINIUM OXIDE POLY(ETHYLENE-CO-BUTYL


ACRYLATE) NANOCOMPOSITES .......................................................................... 23
3.2.1 Adsorption of antioxidant onto Al2O3 nanoparticles ................................... 23
3.2.2 Concentration of effective antioxidant in EBA composites ........................... 24

3.3 MIGRATION OF STABILIZER FROM POLYMER NANOCOMPOSITES IN AQUEOUS


MEDIA ................................................................................................................. 30
3.4 LOSS OF DM AND PBO FROM POLYETHYLENE IN AQUEOUS MEDIA ........... 37
3.4.1 Physical behaviour/properties of DM and PBO ............................................ 37
3.4.2 Evaporation of DM and PBO from polyethylene .......................................... 39
3.4.3 Degradation of DM and PBO in PE compound in aqueous medium ........... 40
3.4.4 Migration of DM and PBO in PE compound ............................................... 42

3.5 LONG-TERM PERFORMANCE OF STABILIZED POLYETHYLENE IN WATER


CONTAINING CHLORINE DIOXIDE ...................................................................... 45
3.5.1 Consumption of antioxidant in squalane and polyethylene .......................... 45
3.5.2 Degradation of polyethylene .......................................................................... 47

4 CONCLUSIONS ......................................................................................... 50
5 ACKNOWLEDGEMENTS ........................................................................ 52
6 REFERENCES .............................................................................................. 54

Purpose of the Study

PURPOSE OF THE STUDY


The overall purpose of the work described in this thesis was to develop different
polymers based composite materials and to investigate their long-term
performance under different operating conditions for different applications. The
goals of the work were to elaborate and discuss:
effective organic surface treatment of Al2O3 nanoparticles to make them
compatible with the matrix polymer and also to prevent the nanoparticles
from aggregating and instead becoming uniformly distributed.
the preparation of stabilized aluminium oxide - poly(ethylene-co-butyl
acrylate) nanocomposites.
the stability of these novel materials under ageing to assess the long-term
performance as insulation material for high-voltage power cables.
the migration of stabilizer from polymer nanocomposites in aqueous
conditions, keeping the end application in mind where these materials
will be subjected to moist conditions.
the preparation of polyethylene-based materials containing synthetic
pyrethroid insecticide (deltamethrin, DM) and synergist (piperonyl
butoxide, PBO), to be used for killing ticks, bugs and insects.
the degradation and migration of DM and PBO from polyethylene in
aqueous conditions.
the assessment of the long-term performance of stabilized polyethylene in
water containing chlorine dioxide.

Introduction

1 INTRODUCTION
1.1

POLYMER NANOCOMPOSITES BACKGROUND

Polymer composites based on nanometer-sized inorganic filler particles have


developed strongly since the work done at the Toyota Company, Japan, which led to
the commercialization of polyamide-6/clay nanocomposites in 1990 [1]. Polymer
nanocomposite materials have lately received great attention in both academia and
industry. Small additions of inorganic, usually surface-treated, nanoparticles have a
very positive effect on the electrical properties of insulating polymers: enhanced
dielectric strength and resistance towards partial discharges and a reduction of space
charges [2-4]. Polymers such as polyamide, polyimide, epoxy and silicone rubber, with
the addition of inorganic nanofiller have a great resistance towards partial discharges
[5-9]. Studies also show that both thermoplastics, such as polypropylene, ethylene-
vinyl acetate copolymer and low-density polyethylene, and thermosets such as epoxy
filled with different nanoparticles show a higher breakdown strength, and/ or voltage
endurance at moderate field strengths than polymers with either no or micro-sized
filler [10-16]. Oxides of silicon, aluminium and titanium have attracted attention as in
certain polymer matrices they provide a package of attractive thermal, mechanical and
dielectric properties [1719]. The large surface-to-volume ratio of the nanofillers
suggests that the particles may affect the segmental mobility of the polymer phase,
provided that the polymer molecules are efficiently attached to the filler particles in a
fashion similar to that of amorphous chain segments to crystals in semi-crystalline
polymers.

1.2

POLY(ETHYLENE-CO-BUTYL ACRYLATE)

Poly(ethylene-co-butyl acrylate), often referred to as EBA, is manufactured by


copolymerization under high hydrostatic pressure in the same reactors as those used

Introduction

to make low-density polyethylene. The acrylic acid units reduce the crystallinity and
make the polymer more polar than low-density polyethylene. EBA is currently used as
semi-conductive material in electrical insulation systems. A poly(ethylene-co-butyl
acrylate) copolymer has been used as a host matrix in this study. The butyl acrylate is
intended to increase the polarity of the matrix and improve the particle dispersion,
compared to that of a pure polyethylene.


Figure 1: Chemical formula of EBA.

1.3

ALUMINIUM OXIDE NANOPARTICLES

Aluminium oxide nanoparticles are found in various crystallographic forms (, , and


phase), with a consequent variation in material properties because of the following
inherent properties: moderate thermal conductivity, high melting temperature, low
thermal expansion coefficient, chemical inertness, non-toxicity, hardness and
toughness etc. Al2O3 has been widely used as filler material in high-performance
materials to increase ductility, scratch resistance and toughness. Aluminium oxide is
also used for crucibles and thermocouples, as an absorbent material or for chemical
synthesis. It has good dielectrical properties, i.e. low dielectric constant at 20 C and 1
GHz (10) and high electrical resistivity (>1014 cm). Together with its large specific
area, it shows very good potential for creating an interesting and highly modifiable
polymer-nanofiller interphase [20,21].

1.4

SURFACE TREATMENT OF NANOPARTICLES

The nanoparticles used in the composites have to be coated with an organic layer to
make them compatible with the matrix polymer and also to prevent the nanoparticles
from aggregating. The interface, i.e., the boundary zone between filler and
surrounding polymer is also critical since it contributes to the structural integrity and

Introduction

long-term performance of the nanocomposites. It is therefore important that the


nanoparticles are coated efficiently and well dispersed in the polymer matrix.
Silanization in one of the standard methods used to overcome the chemical
incompatibility between inorganic and organic materials [20-24]. Wide ranges of silane
compounds are commercially available and they can be covalently bonded to those
nanoparticle types with hydroxyl groups. These compounds also offer many
possibilities including reactive functional groups. Two different silanes,
aminopropyltriethoxysilane and octyltriethoxysilane, were used in this study. The
silane molecules hydrolyze with water or alcohol to form reactive silanols, after which
they condense with the hydroxyl groups on the filler surface. Surface treatment of the
filler can also be achieved by other methods: nitroxide-mediated radical
polymerization [25], reversible addition-fragmentation-chain-transfer polymerization
[26] and atom transfer radical polymerization [27]. It is however fair to state that the
conversion of hydroxyl groups is never complete using any of these method.

1.5

LONG-TERM PERFORMANCE OF POLYMER


NANOCOMPOSITES

Ageing of insulating materials is one of the most important problems in electrical


systems. A number of chemical and physical processes contribute to reducing the
lifetime of the polymeric insulating materials, i.e., thermal oxidation, photo-oxidation,
thermal degradation, and hydrolysis of stabilizer. These processes ultimately lead to a
loss of structural integrity of the insulator, significantly reducing its performance.
Ageing also causes significant changes in the dielectric constant, dielectric loss and
dielectric strength of the insulating materials. Electrical ageing of the insulator may
also occur. Water treeing may also occur in insulations in wet conditions, which may
lead to electrical treeing and failure. The long-term performance of polymeric materials
is governed by the degradation reactions that may occur, and by the activity of the
stabilizer, i.e. antioxidant. Air, sunlight, water vapor, and various atmospheric
pollutant gases may contribute to the degradation of a polymer. Thermal and photo-
oxidative induced degradation are the most common types of degradation [28]. When
the degradation is dominated by oxidation reactions, it can be assessed by oxygen
uptake measurements [29,30], and by infrared spectroscopy [31,32]. In order to inhibit
thermal oxidation during compounding and subsequent service, stabilizers are usually
added to the polymers. Once the stabilizer is consumed, a gradual oxidation of the
polymer will start. It is thus critical to assess the amount of stabilizer left in the system
in order to predict the lifetime. The amount of stabilizer left is assessed by differential

Introduction

scanning calorimetry, by determining the oxidation induction time of the stabilized


polymer. Due to the large surface-to-volume ratio of nanoparticles, interaction
between the stabilizers and the nanoparticles is expected. It is possible that the
antioxidant molecules may be adsorbed onto the particle surfaces, which leads to a
reduction in their efficiency. However, careful surface treatment of the nanoparticles
results in slow release of stabilizer molecules with time, and this contributes to an
extended extension of the polymer matix lifetime.

1.6

MIGRATION OF STABILIZER

The migration of active substances (antioxidants, insecticides or flame-retardants etc.)


is controlled either by diffusion or by the boundary conditions, i.e. evaporation to a
surrounding gas phase or dissolution in a liquid medium. Antioxidant efficiency in
polymers can also be lost by internal precipitation of excess antioxidant, or by the
oxidation process itself. A depletion of antioxidant in the boundary regions resulting
in a parabolic antioxidant concentration profile is characteristic of diffusion-controlled
migration [33,34]. A flat antioxidant concentration profile is characteristic of boundary-
controlled migration [35,36]. In the case of high initial concentration of antioxidant,
internal precipitation of antioxidant is also expected to occur [37]. Deactivation of
antioxidant by adsorption onto fillers is also a recognized phenomenon. Obviously, the
relative importance of this mechanism increases with increasing specific filler surface
area. For example, nanometer-sized titanium oxide pigment has been reported to
induce more oxidative degradation of polyethylene during processing and thermal
ageing than the corresponding m-sized pigment [38]. The hindered phenolic-type
stabilizer Irganox 1010 (Fig. 2) has a potential for multi-point adsorption between the
ester groups, as well as via the phenolic hydroxyl groups, and the surface hydroxyl
groups of metal oxides [39,40]. By using additives that sacrificially adsorb on to the
filler surfaces, the adsorption can be blocked. The adsorption of antioxidant onto a
nanofiller surface is not necessarily a negative phenomenon; adsorbed antioxidant can
potentially provide an antioxidant reservoir for slow release to the polymer matrix
[41,42].

Introduction


Figure 2: Chemical formula of Irganox 1010.



Surface coating (section 1.4) is another efficient way to control the concentration of
accessible hydroxyl groups available for antioxidants to adsorb, by reacting silane
compounds with the hydroxyl groups present on the nanofiller surfaces. High surface
area fillers, surface functionalized with active groups, are an appealing approach for
obtaining a controlled release or preferential adsorption of antioxidants [42].
Antioxidants can also be immobilized onto the surface of nanoparticles via silane
coupling agents. It has been shown that antioxidants covalently bonded to nano-silica
in low-density polyethylene retained excellent stability in water [43].

1.7

LOSS OF SYTHETIC INSECTICIDE FROM POLYETHYLENE

Deltamethrin (DM) is an important insecticide and is a member of one of the safest


classes of pesticides (synthetic pyrethroids) [44]. It is popular not only because of its
effectiveness but also because it works efficiently even at a very low concentration.
Piperonyl butoxide (PBO) is an effective synergist with DM against all of the species
investigated [45]. DM in combination with synergist PBO is used for many purposes
from storage of food commodities to bugs control [4648].

DM is often impregnated in polymers for different applications including collars for
dogs to protect them from bites of flies and bugs [49,50], where a controlled rate of
release of insecticides from a polymer matrix is regarded as a difficult issue. A release
that is too rapid exhausts the insecticidal effect of the matrix quickly, and too slow a
release does not have the desired efficiency killing the insects. DM has a tendency to
degrade in water forming decamethrinic acid [44], and the rate of conversion depends
on the mechanism of introduction into water. Generally, the loss of additive

Introduction

functionality can be due to migration to the surrounding media or to a chemical


reaction converting the additive to a form of little value. This study dealt with the loss
mechanism of DM and PBO from polyethylene sheets when exposed to aqueous
media.



(a)

(b)

(c)



Figure 3: Chemical structures of (a) DM, (b) PBO and (c) decamethrinic acid, a degradation
product of DM.

1.8

LONG-TERM PERFORMANCE OF POLYETHYLENE IN


AQUEOUS CHLORINE DIOXIDE

Plastic pipe systems have received great attention over the past 50 years and have
replaced the traditional copper, steel, clay and iron pipes for hot and cold water as well
as for gas distribution applications. During water distribution, chlorinated compounds
(e.g. chlorine dioxide) are added to water as disinfectants in order to prevent the
spread of infection. These chlorinated substances are however strong oxidants that

Introduction

cause degradation which shortens the service lifetime of structural components such as
pipes and containers [5154]. Polyethylene pipes, when exposed to water containing
chlorine dioxide (4 ppm, pH = 6.8 and 90 C), rapidly lose their antioxidant protection
far into the pipe wall, and the unprotected polymer degrades extensively by a reaction
confined to the immediate surface [55]. The cracks formed in the brittle surface layer
due to exposure to a chlorinated environment extend to the fresh material beneath,
and further crack propagation is assisted by degradation of the material at the crack tip
(degradation-assisted crack growth) until a crack of critical size is formed leading to
pipe failure [55]. Chlorine dioxide selectively attacks phenolic antioxidants by a one-
electron transfer from the phenolic unit to chlorine dioxide and by further reactions
yielding coloured products of little value as antioxidants [55,56].
Efficient methods are required to assess the efficiency of antioxidants in polyethylene
exposed to chlorinated aqueous media. Azhdar et al. [57] presented a method using a
low molecular analogue (squalane) instead of polyethylene. The stability of phenolic
antioxidants dissolved in squalane (typical antioxidant concentrations were 0.1 wt.%)
was assessed during 300 min exposures by determining the oxidation induction time
using differential scanning calorimetry [57]. This method is applicable to phenolic
antioxidants, because the oxidation induction time is proportional to the concentration
of efficient antioxidant [58,59]. The assessment of concentration of other antioxidants,
such as hindered amines and secondary antioxidants, requires more time-consuming
methods such as extraction followed by liquid chromatography. The conditions
prevailing in the squalane phase are different from those in polyethylene: (i) Stirring
causes convective transport within the squalane phase that largely eliminates
concentration gradients of chlorine dioxide and antioxidant within this phase. If
polyethylene is exposed to water containing chlorine dioxide, concentration gradients
for both chlorine dioxide and antioxidant develop [55]. (ii) At the testing temperature,
squalane is an amorphous liquid whereas polyethylene is semicrystalline; the polymer
used in the present study has a mass crystallinity at room temperature close to 60 %.












Experimental

2 EXPERIMENTAL
2.1

MATERIALS

2.1.1

Poly(ethylene-co-butyl acrylate)

Poly(ethylene-co-butyl acrylate)s containing either 13 or 28 wt % of butyl acrylate were


supplied by Borealis AB, Stenungsund, Sweden. The densities were 924 kg m3 (EBA-
13) and 926 kg m-3 (EBA-28). The melt flow indices according to ISO 1133 (190 C, 2.16
kg weight) were 1.1 g/ 10 min (EBA-13) and 4 g/ 10 min (EBA-28).

2.1.2

Aluminium oxide nanoparticles

Two types of aluminum oxide nanoparticles were used; Nanodur (ND) supplied by
Nanophase Inc., USA, consisting of almost spherical particles (70 wt% d-phase and 30
wt% g-phase) with an average particle diameter of 45 nm and a specific surface area of
45 m2 g-1 (Fig. 4a) and Nanoamor (NA) supplied by Nanostructured and Amorphous
Materials Inc., USA, with an average particle diameter of 25 nm and a specific surface
area of 190 m2 g-1 (Fig. 4b). The manufacturers provided all data concerning the
aluminum oxide grades.








Experimental

(a)

(b)





Figure 4: Scanning electron micrographs of uncoated aluminum oxide nanoparticles showing
(a) Nanodur and (b) Nanoamor

2.1.3

Chemicals & materials

The stabilizer Irganox 1010 was provided by Ciba Specialty Chemicals, Switzerland.
The silane coupling agents, aminopropyltriethoxysilane by Sigma Aldrich and
octyltriethoxysilane by Fluka, were used as received.
Deltamethrin (DM) was supplied by Bayer Crop Science AG, Germany in powder form
and the synergist piperonyl butoxide (PBO) by Endura SpA, Bologna, Italy as a slightly
yellow liquid. High-density polyethylene (HDPE) and low-density polyethylene
(LDPE) supplied by Indothene, Indian Petrochemicals Corporation Ltd., India were
used to manufacture the polyethylene compound. The densities at 23 C were 955 kg
m3 (HDPE) and 922 kg m3 (LDPE). n-Hexane HPLC grade (CAS Number 110-54-3)
supplied by Scharlab SL, Spain and 1,4-dioxane HPLC grade supplied by VWR
international AB, Stockholm were used as mobile phase in liquid chromatography.
Dibutyl phthalate supplied by Sigma Aldrich and xylene technical grade supplied by
VWR international SAS, France, were used for extraction of the polyethylene
compound sheet samples. MilliQ grade water was provided by Millipore AB, Solna,
Sweden and Diiodomethane was supplied by Sigma Aldrich (CAS Number 75-11-6).



10

Experimental

2.2

INSTRUMENTATION

Differential Scanning Calorimetry (DSC)


Differential Scanning Calorimetry was used to assess the thermal characteristics of
nanocomposites with Mettler Toledo STARe software V9.2. Samples of 15 to 30 mg
weight were heated in aluminum sample holders from 80 to 170C, cooled from 170
to 80C and finally heated from 80 to 170C at a scanning rate of 10C min-1 in
nitrogen with a flow rate of 50 mL min-1.
Oxidation induction time (OIT) method
The concentration of effective antioxidant was assessed by determining the oxidation
induction time (OIT) at 200C in a temperature- and energy-calibrated Mettler-Toledo
DSC820. Each sample with a weight of 233 mg was enclosed in a 100 ml standard
aluminium crucible with three holes (diameter=1 mm) in the cover, heated from 25C
to 200C at a rate of 10C min-1 in nitrogen with a flow rate of 50 mL min-1, and allowed
to rest for 5 min before the atmosphere was switched to pure oxygen, also at a flow
rate of 50 mL min-1. The samples were maintained at constant temperature, and the
exothermal heat associated with oxidation was recorded. The oxidation induction time
was determined as the time to the intersection between the isothermal baseline and the
tangent at 0.2Wg-1 deviation from the baseline. The standard deviation in the OIT data
was estimated by analysis of five individual samples from three different composites
containing 2 wt% ND filler: 481 min (uncoated ND), 631 min (ND treated with
octyltriethoxysilane) and 702 min (ND treated with aminopropyltriethoxysilane).
Since the standard deviation was so small, each data point in the present work is the
average of only two independent measurements.
Scanning electron microscopy (SEM)
The particle dispersion in the polymer matrix was studied and quantified from freeze-
fractured surfaces in a Hitachi S-4800 field emission SEM (FE-SEM). The samples were
sputtered with a 510 nm conductive coating layer in an Agar high-resolution sputter
coater (Model 208RH) prior to the study.
Fourier Transform Infrared (FT-IR)
FTIR spectra were obtained using a Perkin Elmer FT-IR Spectrometer 2000 equipped
with a single reflection ATR accessory from Graseby Specac Ltd., UK.




11

Experimental

X-ray photoelectron spectroscopy (XPS)


XPS spectra were recorded on powder samples using an AXIS Ultra X-ray
photoelectron spectrometer (Kratos Analytical Ltd., Manchester, UK). The samples
were analysed using a monochromatic aluminium X-ray source under ultra-high-
vacuum conditions. Both uncoated and coated nanoparticle samples were studied
either after drying or after being soaked in a 0.01 wt% solution of Irganox 1010 in n-
hexane for 12 h at 50C, cooled to 25C, recovered by decantation of the clear solutions,
rinsed with an excess volume of n-hexane at 25C and finally dried to constant weight
at 25C under reduced pressure (0.5 kPa).

Thermogravimetric analysis (TGA)
Thermogravimetric analysis were performed on a Mettler-Toledo SDTA/TGA 851e
thermo-balance to study the evaporation rates from pristine DM and PBO, from a
solution of DM and PBO (11 wt.% DM) and from the polyethylene sheets at different
temperatures between 110 and 180C in nitrogen with a flow rate of 50 mL min-1. The
initial sample masses were18.0 0.5 mg.

High-pressure liquid chromatography (HPLC)
HPLC analysis was performed on a HP Agilent 1100 system consisting of a degasser
(DEGASYS DG-2), a binary pump, an auto sampler and UV detector. Separation was
achieved on a Lichrosorb Si 60 (100 mm x 4.6 mm, 5m) column, with an injection
volume of 10 L and a flow rate of 1.3 mL min1. The mobile phase consisted of 95
vol.% n-hexane and 5 vol.% 1,4-dioxane. The chromatograms were recorded with
Chemstation (Agilent Technologies Inc, 1994-2003, USA) and the wavelength was set
to 236 nm.

Contact angle measurements
The contact angle measurements were made on a CAM 200 instrument from KSV
Instrument Ltd. (Helsinki, Finland) using MilliQ water and Diiodomethane. 3 l
droplets of the two liquids were studied on each sample.

2.3

SAMPLE PREPARATION AND PROCEDURES

2.3.1

Surface treatment of nanoparticles

The aluminum oxide nanoparticles were vacuum dried (0.5 kPa) for 24 h at 190 C to
remove adsorbed water before surface treatment. The silanization procedure was as

12

Experimental

follows: (i) addition of 80 g of dried nanoparticles to 4000 mL of methanol/water (25 :


75 vol %), (ii) mechanical stirring for 15 min (iii), dropwise addition of silane under
ultrasonication using a wand (300 W; Sonics Vibra Cell VCX 750) for 10 min, (iv)
vigorous mechanical stirring for 4 h. For silanization of ND, 130 mL of each silane was
added to 80 g of particles. In the case of NA, 300 mL of each silane was added due to
the higher specific surface area of this nanofiller. To promote hydrolysis of
octyltriethoxysilane, the pH was adjusted to 4.5 by addition of acetic acid in
combination with an extension of the reaction time from 4 to 24 h. After this reaction
time, the particles were collected by centrifugation in a Hettich bench centrifuge at
4500 rpm. The nanoparticles were washed in ethanol and isopropanol before final
drying. The coated nanoparticles were dried at 110C for 24 h, ground with pestle and
mortar, and finally vacuum dried at 60C for 24 h (paper I).

2.3.2

Preparation of nanocomposite materials

Poly(ethylene-co-butyl acrylate) pellets were cryo-ground to a particle size of 0.5 mm.


Batches of 300 g were produced as follows: 6, 18 and 36 g (yielding compounds
containing 2, 6 and 12 wt% nanofiller) of aluminium oxide nanoparticles and 0.2 wt%
(based on the final weight of the entire compound) of Irganox 1010 were mixed with 30
mL isopropanol. The slurry was ultra-sonicated using a wand (Sonics Vibra-Cell VCX
750, Sonic & Materials Inc., USA) for 30 min according to the following cycle program:
9 s at 300W followed by 1 s rest. After ultrasonication, the cryo-ground EBA powder
was added and the slurries were mixed using a three-dimensional ultra-mixer
(Turbula Shaker Mixer Type T2F, WAG, Switzerland) at 25C for 3 h. The mixture was
then dried at reduced pressure at 70C for 30 min, after which it was again mixed in
the three-dimensional ultra-mixer for 2 h. All batches were finally dried in a vacuum
oven (pressure=0.5 kPa) at 50C for 24 h. The powder obtained was melt compounded
in a Prism Eurolab 16 XL twin-screw extruder (Thermo Electron Corporation, USA).
The extruded strand was cooled in a water bath and thereafter pelletized using a Prism
Eurolab strand plasticizer. Finally, the pellets were extruded in a Plasticorder PL2000
extruder (Brabender OHG, Germany) to tape-shaped films using a ribbon diehead (100
mmx1.5 mm). The extruder was equipped with a screw diameter of 19 mm and
L/D=25. The temperature profile ranged from 150160C (feeder) to 170C (die), and
the screw rotational speed was 40 rpm. All the tape films were finally dried under
reduced pressure (0.5 kPa) at 50C for 24 h. The extruded tapes had, according to
differential scanning calorimetry, the following crystallinities (based on the pristine
polymer component) and melting peak temperatures obtained from a 10C min-1
heating scan [60]: 432 wt% and 104.50.5C (EBA-13) and 223 wt% and 951C (EBA-
28) (papers I, II & III).

13

Experimental

2.3.3

Preparation of sandwiched composites

Sandwiched EBA13 nanocomposite materials were prepared for diffusion studies


with EBA13 containing 6 wt.% ND on one side and pristine EBA13 on the other side,
by compression molding. Dried pristine EBA13 powder was compression molded to
1.5 mm thick sheets in a TP400 Fontijne Press (Netherlands) at 170 C by applying a
compressive load of 300 kN for 5 min and finally cooled at a rate of 10 C/min to room
temperature while maintaining the compressive load. Finally, 1.5mm thick films of
pristine EBA13 were compression molded with the EBA13 nanocomposite
containing 6 wt.% uncoated, aminopropyltriethoxysilane and octyltriethoxysilane
coated ND nanoparticles. The total thickness of the films was 3 mm (paper III).

2.3.4

Preparation of polyethylene compounds

HDPE and LDPE pellets were cryo-ground to 0.5 mm particle size in a ZM200 Retsch
grinder (Germany). An amount weighing 500 g was produced according to the
following recipe: 431 g (86.2 wt.%) HDPE, 48 g (9.6 wt.%) LDPE, 2.25 g (0.45 wt.%) DM,
17.5 g (3.5 wt.%) PBO. A solution of DM and PBO (DM is soluble in PBO at room
temperature) was added to the cryo-ground HDPE and LDPE powders. This slurry
was homogenized using a three-dimensional ultra mixer (Turbula Shaker Mixer Type
T2F, WAG, Switzerland) at 25C for 2 h. The mixture was the dried under vacuum at
50C for 30 min, after which it was again mixed in the three-dimensional ultra mixer
for 2 h. The mixture was finally dried in a vacuum oven at 40C for 24 h. The dried
material obtained was compression molded to 1 mm thick sheets in a TP400 Fontijne
Press (Netherlands) at 180C by applying a compressive load of 300 kN for 5 min and
the cooling the material at a rate of 10C/min to room temperature while maintaining
the compressive load (paper IV).

2.3.5

Sample extraction for HPLC

The DM and PBO contents in the polyethylene compound sheets were determined by
HPLC analysis of extracts obtained from samples of these sheets. The extract solution
consisted of 95 vol.% n-hexane and 5 vol.% 1,4-dioxane. An internal standard solution
containing 15 mg (mL)-1 of dibutyl phthalate in 95 vol.% n-hexane/5 vol.% 1,4dioxane
was prepared. The polyethylene sheets were cut into small pieces (0.5 g of material per
sample) and transferred to a 100 mL thermo-glass vessel that contained 19 mL xylene
and 1 mL of the internal standard solution. The vessel was tightly capped. Extraction
was carried at 150 C for 45 min while the solution was stirred. The solution was then

14

Experimental

cooled to room temperature and filtered through a 0.45 m filter, after which 0.1 mL of
the filtrate was taken in a 1 mL vial and the xylene present was completely evaporated
in nitrogen gas. The residual left in the vial was filled with 1 mL of the concentrated
extract solution and the sample was injected into the HPLC column. Standard
solutions of DM and PBO were prepared for calibration at four different
concentrations, covering the active ingredient concentration range of DM and PBO in
the polyethylene sheet samples (paper IV).

2.4

THERMAL AGEING OR EXPOSURE CONDITIONS

2.4.1

Al2O3/Poly(ethylene-co-butyl acrylate)

The tape samples (ca. 1 mm thick) were aged in an ULE-699 ventilated oven
(Memmert, Germany) at 901C with dry air as the surrounding medium. The
maximum ageing time was 720 h (paper II).
The samples were aged in air at RH 100% and in liquid water in an ULE-699 ventilated
oven (Memmert, Germany) at 901 C. Sandwiched EBA13 nanocomposite materials
were aged only in dry air and in air at 100% RH. Each sample was microtomed at 0.8,
1.3, 1.8, 2.4 and 3 mm from the surface (pristine EBA-13) and run in DSC to check the
OIT and study the depth profiles (paper III).

2.4.2

Polyethylene sheet compounds

The polyethylene compound sheets were aged under four different conditions: in
water at 80 2 C and 95 2C, and in air at relative humidities of 60 % and 80 % at 80
2C in an ULE-699 ventilated oven (Memmert, Germany). A saturated solution of
sodium sulfate in water was prepared in a desiccator to give a relative humidity of 80
% RH. A saturated solution of potassium iodide was used to give 60 % RH (paper IV).

2.4.3

Exposure of squalane or polyethylene samples to chlorinated


media

The experimental set-up included a bottom heater, an enclosing heater controlled by a


Eurotherm 914 PID device, a condenser (10 C) and a test tube equipped with a
magnetic stirrer containing the aqueous solution of chlorine dioxide (volume = 50 mL;
Milli-Q water containing 10 ppm chlorine dioxide with pH = 6.8 0.1) and the
squalane-antioxidant solution (volume = 5 mL) or the polyethylene tape sample. The

15

Experimental

liquids were stirred at 250 rpm using a PTFE-coated magnet bar. The temperature was
controlled at 70 3 C. Azhdar et al. [57] have presented further details about the
squalane testing. The chlorine dioxide solution was renewed by two different methods:
(i) 25 mL of the aqueous phase was replaced every 30 min, which is referred to as the
intermittent method, as described by Azhdar et al. [57]. (ii) Fresh 10 ppm chlorine
dioxide solution at pH = 6.8 was added continuously to the test tube which was
equipped with an overflow system (flow rate = 1 mL min1; L/S PTFE Tubing Pump
equipped with a variable speed drive model 7524-40, a pump head model 77800-60
and PTFE tubing (96412-13) supplied by Cole-Palmer, USA). The continuous process
was used only for the exposure of the polyethylene tape specimens (paper V).












16

Results and Discussion

3 RESULTS AND DISCUSSION


3.1

PREPARATION OF ALUMINUM OXIDE POLY(ETHYLENE-CO-


BUTYL ACRYLATE) NANOCOMPOSITES.

3.1.1

General characteristics of Al2O3 nanoparticles

The specific surface area of the Al2O3 nanoparticles was determined by nitrogen
desorption measurements on BET to confirm the value provided by the suppliers. The
specific surface area values obtained for two different batches of uncoated ND were: 36
1 and 42 1 m2 g-1. The value for ND provided by the manufacturer was 40 m2 g-1.
The specific surface area value obtained for uncoated NA was 174 5 m2 g-1, which
was close to the value (180 m2 g-1) provided by the manufacturers. The coated ND
nanoparticles showed slightly lower values, i.e., octyltriethoxysilane-coated ND
showed 25 1 m2 g-1 and aminopropyltriethoxysilane-coated ND showed 31 1 m2 g-1,
suggesting some agglomeration during the coating process. These agglomerates are
however easy to break during melt compounding.
The nanoparticles surfaces contain hydroxyl groups, which resulted in moisture
adsorption due to interaction with these accessible hydroxyl groups. As these hydroxyl
groups take part in surface coating (silanization), it was therefore important to remove
the adsorbed water. The nanoparticles were therefore dried according to the method
described by Schadler et al [61]. The two different types of aluminium oxide
nanoparticles used in this study were different in particle size and specific surface
area, 45 and 190 m2 g-1 for ND and NA respectively. The moisture sorption of dried
nanoparticles at 20 C and 50% RH showed a weight gain of 0.4% for ND and 7.9% for
NA. The nanoparticles (NA) with a high specific surface area adsorbed moisture
almost five times more than the ND particles. Since the amount of adsorbed moisture
was related to the number of available surface hydroxyl groups, this indicated that NA
contained about five times more hydroxyl groups per unit surface area than ND.

17

Results and Discussion

The efficiency of surface treatment by silanization was assessed from weight loss data
by thermogravimetry. The loss in weight of uncoated and coated nanoparticles at 800
C and surface coverage of silane is shown in Table 1. It was assumed that the silane
hydrolyzed completely and that the difference in weight loss at 800C between coated
and uncoated particles was completely due to elimination of the hydrocarbon
component of the silanes. It was also assumed that the inorganic part of the silane (Si
O) was oxidized to silica.
The graft density of aminopropyltriethoxysilane-coated ND, 2.2 mol m-2, was lower
than the reported value of 8.3 mol m-2 [62], and for octyltriethoxysilane-coated ND, it
was even lower, 0.7 mol m-2. The data indicate that ND particles were only partially
covered with silane, which was confirmed by transmission electron microscopy that
revealed weak cluster-like silane structures on the nanoparticle surfaces (Fig. 5).
NA showed a greater degree of surface coverage than ND, for both
octyltriethoxysilane and aminopropyltriethoxysilane, 2.3 and 4.2 mol m-2
respectively.



mol
Silane coverage

-
2
Nanoparticle
Mass loss (%)
Silane (m )
(nm-2)
NDU

0.1

NDO

0.5

0.7

0.4

NDA

0.7

2.2

1.4

NAU

8.9

NAO

13.6

2.3

1.4

NAA

13.3

4.2

2.5


Table 1: Mass loss and silane coverage of nanoparticles




18

Results and Discussion

50 nm



Figure 5: Transmission electron micrograph of aminopropyltriethoxysilane-coated ND
nanoparticles. The arrows indicate silane-protruding structure.
Higher graft density of NA was due to high hydroxyl group concentration. For
comparison, 4.87.5 mol m-2 of 3-(trimetoxysilyl) propyl methacrylate was grafted
onto a different aluminium oxide [63,64].

3.1.2

Dispersion of Al2O3 nanoparticles in EBA

The dispersion of different filler loadings (2, 6 and 12 wt %) of aluminum oxide


nanoparticles in EBA was studied.
Figures 6 is a SEM micrograph of fractured surfaces of EBA13/ ND-composite
obtained after immersion of the samples in liquid nitrogen. The nanoparticles are
observed as spherical objects sitting in wholes with a diameter greater than the
nanoparticles (Figs. 6ac).

19

Results and Discussion




Figure 6: Scanning electron micrographs of fracture surfaces of composites based on EBA13
and octyltriethoxysilane ND with the following nanoparticle contents, (a) 2wt%; (b) 6 wt%; (c)
12 wt%.

The composites containing 2 and 6 wt% octyltriethoxysilane-coated ND displayed
good dispersion with mostly solitary single particles and only a few small
agglomerates of several particles. The composite with 12 wt% octyltriethoxysilane-
coated ND showed several two to three particle agglomerates (Fig. 6c). In general,
octyltriethoxysilane-coated ND was uniformly dispersed in EBA13. The dispersion of
uncoated ND in EBA13 was also good (Fig. 7a), but a large number of small
agglomerates were observed in contrast to the composites with octyltriethoxysilane-
coated ND.

Figure 7: Scanning electron micrographs of fracture surfaces of composites based on EBA13


composites containing: (a) 12wt% uncoated ND; (b) 12 wt% aminopropyltriethoxysilane-
coated ND.

20

Results and Discussion

The composites with aminopropyltriethoxysilane-coated ND nanoparticles showed a


good degree of dispersion intermediate between those of octyltriethoxysilane-coated
ND and uncoated ND (Fig. 7b). The fibers seen on the fractured surfaces were due to
overheating of samples prior to cracking (Fig. 7b). These fibers were not seen in the
samples containing a lower content of nanofillers, which further confirms the
assumption that these fibers are artificial structures.
Scanning electron micrographs with a total surface area of ca. 2 mm2 were screened to
investigate further the nanoparticles dispersions and the presence of agglomerate. The
average numbers of particles in the small agglomerates in the different EBA13
composites are presented in Table 2. The composite materials based on
octyltriethoxysilane-coated ND showed the fewest particles in small agglomerates, and
thus
the
best
dispersion.
The
composite
materials
containing
aminopropyltriethoxysilanecoated ND showed a larger average particle size,

Nanoparticle

2 wt %

6 wt %

12 wt %

NDU

3.4

2.6

3.5

NDO

2.1

1.5

2.6

NDA

2.5

2.4

4.1

NAU

15

21

18

NAO

3.2

3.1

3.1

NAA

3.7

13

8.8


Table 2: Average number of particles in small agglomerates in EBA13 composites.
whereas uncoated ND-containing composites showed an even more significant
increase in the average particle size/number. The NA composite exhibited a larger
number of small agglomerates than the ND composites. The composite with uncoated
NA displayed the largest number of agglomerates and was the least well-dispersed
composite.
The composite based on EBA28 showed a poor dispersion of the nanoparticles. A
possible explanation could be the low melt viscosity of EBA28, so that the shear forces

21

Results and Discussion

during the extrusion mixing were not sufficient to break the nanoparticle
agglomerates.
The crystallinity of pristine EBA13 was calculated to be 42 wt %, and the EBA13
composite showed a scatter in crystallinity 42 wt % regardless of the filler type and
content. The standard deviation for all the EBA13 composites was 23 wt %. A similar
scatter was observed in the EBA28 composites, with pristine EBA28 displaying a
crystallinity of 21 wt %. It is believed that when nanoparticles, especially ND were
added into EBA28, they acted as nucleation agents leading to an increase in
crystallization temperature.












22

Results and Discussion

3.2

ANTIOXIDANT ACTIVITY IN ALUMINIUM OXIDE


POLY(ETHYLENE-CO-BUTYL ACRYLATE) NANOCOMPOSITES

3.2.1

Adsorption of antioxidant onto Al2O3 nanoparticles

The atomic surface composition of the outermost 210 nm of the ND particles was
analyzed by X-ray photoelectron spectroscopy and the data is presented in Table 3.
Uncoated ND contained a low concentration of carbon from impurities together with
aluminium and oxygen (Table 3). The calculated values of 35 at% Al and 65 at% O
deviated slightly (5%) from the theoretical contents of 40 at% Al and 60 at% O. The
aminopropyltriethoxysilanecoated ND surfaces showed a higher carbon content (11.8
at%) and almost three per cent of both silicon and nitrogen (Table 3). The ratios based
on XPS data of added amounts carbon to silicon and nitrogen respectively were 3.5:1
and 3.5:1, which is close to the theoretical ratio (3:1 and 3:1) for bound and fully
hydrolyzed aminopropyltriethoxysilane.


Atomic surface composition (mole%)

Al

Si

2.2

63.7

34.1

NDU

8.9

57.8

32.9

0.3

NDA

11.8

53.6

29.1

2.8

2.7

NDAb

17.8

49.8

26.1

3.9

2.4

NDOa

13.1

54.6

31.2

1.1

NDOb

20.9

50.4

27.1

1.6

NDU


Table 3: Atomic surface composition of ND particles according to X-ray photoelectron
spectroscopy.
Nanoparticles after drying.
Nanoparticles after immersion in 0.01 wt.% Irganox-n-hexane solution followed by rinsing in
pure n-hexane.

a

The octyltriethoxysilanecoated ND displayed an even greater increase in carbon


content (13.1 at%) and a lower concentration of silicon (1.1 at%). The ratio of carbon
added to silicon was 9.9:1, which is 20 % higher than the theoretical value (8:1) for fully
bound and hydrolyzed octyltriethoxysilane. The increase in carbon content after
treatment of particles with carbon-rich silane (octyltriethoxysilane) was only slightly

23

Results and Discussion

greater than after treatment with aminopropyltriethoxysilane (containing fewer carbon


atoms), 11 at% and 10 at% respectively. These results were in agreement with the
finding that there was a lower silane group coverage after octyltriethoxysilane
treatment (0.7 mol silane m-2) than after aminopropyltriethoxysilane treatment (2.2
mol silane m-2).
After antioxidant solution treatment followed by the removal of any loosely bound
antioxidant by rinsing with pure solvent, the immediate surface of the particles
contained a higher concentration of carbon. On average, an increase of 7 at% for the
three nanoparticles was observed (Table 3), suggesting that antioxidant was adsorbed
on the particles surfaces.

3.2.2

Concentration of effective antioxidant in EBA composites

The concentration of effective antioxidant in nanocomposites was assessed by


determining the oxidation induction time (OIT), which is proportional to the
concentration of accessible/effective phenolic antioxidants such as Irganox 1010 [31,32].
The data presented in Fig. 8 show the depth profiles (OIT) through a 1.4 mm thick film
during ageing. The antioxidant profile remained flat on extended ageing, indicating
that the loss of stability was controlled by boundary conditions, i.e., evaporation of
antioxidant. Similar observations have been reported in the literature for polyethylene
stabilized with Irganox 1010 [35,36].


Figure 8: OIT as a function of depth for aged samples (ageing times displayed in each rectangle)
for samples based on EBA28 containing 6 wt% uncoated NA.

24

Results and Discussion


Fig. 9a shows OIT data for EBA13 composites with different contents of uncoated ND.
A decrease in initial OIT from 70 min for pristine EBA13 containing 0.2 wt% of
Irganox 1010 to between 40 and 50 min for the composite materials with 2 and 6 wt%
ND and only 10 min for the composite containing 12 wt% was observed. The decrease
in initial OIT for the composite with coated nanoparticles was significantly less (Figs.
9ac). In the case of the composites with 12 wt% of ND, the decrease in initial OIT with
respect to that of pristine EBA13 with the same content of antioxidant was 60 min
(uncoated ND), 18 min (octyltriethoxysilanecoated ND) and 27 min
(aminopropyltriethoxysilanecoated ND). A possible reason could be that the
antioxidant was adsorbed onto the surfaces of the nanoparticles and this leade to a
decrease in OIT, which was confirmed by X-ray photoelectron spectroscopy. These
observations are in accordance with the study reported by Allen et al. [38,39] on silica-
polypropylene and titanium dioxide-polypropylene nanocomposites. It was therefore
believed that 86% (uncoated ND), 26% (octyltriethoxysilanecoated ND) and 39%
(aminopropyltriethoxysilanecoated ND) of the antioxidant was adsorbed on the
surfaces of the nanoparticles in the composites containing 12 wt% nanoparticles. The
strength of the bonds between adsorbed antioxidant and nanoparticles was tested at
200C over an extended isothermal period from 5 to 30 min in an inert atmosphere
prior to the switch to pure oxygen. No change in the OIT was observed and it was
found that the OIT data was independent of the length of this conditioning period.

(a)

(b)

25

Results and Discussion

(c)


Figure 9: OIT as a function of square root of ageing time at 90 C for EBA-13 containing ND:
(a) uncoated; (b) octyltriethoxysilanecoated; (c) aminopropyltriethoxysilanecoated, 2 wt%
(), 6 wt% (), 12 wt% ( ) and pristine EBA-13 with antioxidant ().

The data presented in Fig. 10 shows the decrease in initial OIT with respect to
pristine EBA13 with 0.2 wt% Irganox 1010, as a function of filler content. The
composites containing uncoated and octyltriethoxysilanecoated ND, showed a
linear (with some scatter in the data for the composite with 2 wt% filler) increase in
the OIT reduction with increasing filler content. Interestingly, the composites
containing aminopropyltriethoxysilanecoated ND (2 & 6 wt%) showed a small
increase in the initial OIT (12 min) with respect to the pristine polymer (Figs. 9c &
10). The overall trend in the reduction vs. filler content deviated from a linear
relationship indicating another effect of nanofiller than antioxidant adsorption on
the filler surfaces. It has been reported that there is a synergistic stabilizing effect
between phenolic antioxidants and aliphatic amines in polypropylene [65]. A
synergistic effect on the thermal stability of polypropylene, by a combination of
hindered phenolic and amine antioxidants has also been reported by Desai et al.
[66]. However, this has not always been the case, and opposite results have also
been reported [67].

26

Results and Discussion

The fact that some of the composites showed an initial OIT 0 min indicated that
the antioxidant was not effectively protective towards the polymer oxidation when
adsorbed on the filler surfaces.



Figure 10: Reduction in initial OIT of the EBA13/ND composite with respect to that of
pristine EBA13 with antioxidant as a function of filler content: uncoated (line a);
octyltriethoxysilanecoated ND (line b); aminopropyltriethoxysilanecoated ND (line
c).
The OIT of most of the composites containing ND filler, including the pristine
polymer, decreased with increasing ageing time and the decrease followed a linear
trend (Figs. 9a-c). This behaviour could have been the result of a diffusion-
controlled loss of antioxidant to the surrounding medium, but the data for the OIT
profile (Fig. 8) indicated that the migration of antioxidant was controlled by the
boundary conditions, i.e. by evaporation. The evaporative loss of antioxidant
should, as a first approximation, be proportional to the boundary concentration of
antioxidant. The loss of antioxidant is thus expected to lead to a gradual decrease in
the antioxidant loss rate, which is in agreement with the observed OIT vs. time1/2
trend. The majority of the nanocomposites showed OIT vs. time1/2 lines with a less
steep negative slope than that of the pristine polymer (Figs. 9a-c). There were only
two exceptions to this tendency, i.e., 12 wt% uncoated ND and 12 wt%
aminopropyltriethoxysilane-coated ND. The lower rate of loss of antioxidant in the
composites can be due to a slower release of antioxidant from the nanoparticle
surfaces to the surrounding polymer phase. The effect on the overall loss rate of a
possible decrease in antioxidant diffusivity due to trapping at the interfaces should

27

Results and Discussion

be marginal, since the migration is largely controlled by the evaporative loss of


antioxidant.
Figs. 11a and b show the OIT vs. time1/2 trend for EBA13 composite containing
uncoated and aminopropyltriethoxysilane-coated NA filler. These composites
showed very low initial OIT values (a few minutes) and the reduction in the initial
OIT was much more greater than that in composites with ND.


(a)


(b)



Figure 11: OIT as a function of square root of ageing time at 90 C for EBA-13 containing NA:
(a) uncoated; (b) aminopropyltriethoxysilanecoated, 2 wt% (), 6 wt% (), 12 wt% ( ) and
pristine EBA-13 with antioxidant ().
Hence, 90100% of the antioxidant was adsorbed onto the nanoparticles. The
composites with aminopropyltriethoxysilane-coated NA showed a non-linear OIT
reduction with increasing filler content, probably due to the synergistic stabilizing
effect of the aliphatic aminephenolic antioxidant combination [6567]. The composite
with 2 wt% nanofiller showed an even longer initial OIT (80 min) than the pristine
polymer (70 min), whereas the composites with higher filler loadings had relatively
short initial OITs of ca. 20 min (6 wt%) and 8min (12wt%). The specific surface area of
NA was four times greater than that of ND. Furthermore, water sorption
measurements indicated that the number of hydroxyl groups per unit surface area was
five times greater for NA than for ND [60]. Thus, the strong adsorbing capacity of NA
can be attributed to its high specific surface area and the high concentration of
hydroxyl groups.
Fig. 12a shows the effect of the difference in specific area between ND and NA on the
OIT reduction, where the abscissa is the product of filler content and specific surface
area. The data for the uncoated fillers are to the left in the region of increase whereas

28

Results and Discussion

the data for coated fillers are on the right, indicating that a coating lowers the
concentration of adsorption sites for the antioxidant molecules. Particularly interesting
are the three data points associated with an increased OIT (i.e. negative values of
((OIT))), which can be attributed to the synergistic stabilizing effect of the
combination of phenolic antioxidant and aliphatic amine [6567].

(a)

(b)

Figure 12: Reduction in OIT of (a) EBA-13 and (b) EBA28 composites with respect to that of
pristine polymers with antioxidant as a function of the product of filler content and specific
surface area (SSA): uncoated ND ; ND coated with octyltriethoxysilane ; ND coated
with aminopropyltriethoxysilane ; uncoated NA ; ND coated with octyltriethoxysilane
n ; ND coated with aminopropyltriethoxysilane . The horizontal line indicates antioxidant
depletion.
Fig. 12b shows the reduction in OIT with respect to that of the pristine EBA28 as a
function of the product of filler content and specific surface area (FC x SSA). The
general trend is similar to that of the EBA13 composites with a pronounced increase
in (OIT) at low FC x SSA-values, reached saturation at higher FC x SSA-values. The
EBA28 composites showed a weaker increase in (OIT) with increasing FC x SSA
than the EBA13 and also considerably more scatter in the data. Only one of the EBA
28 composites showed OIT-values close to zero, whereas three of the EBA13
composites showed such low OIT-values. The poor dispersion of the nanoparticles in
the EBA28 composites, i.e. the presence of a significant fraction of particles in large
agglomerates, was probably the main reason for the relatively low (OIT) values; this
being due to a lower tendency for antioxidant adsorption in these composites. These

29

Results and Discussion

data thus suggest that the large agglomerates are inaccessible to the antioxidant
molecules.

3.3

MIGRATION OF STABILIZER FROM POLYMER


NANOCOMPOSITES IN AQUEOUS MEDIA

The oxidation induction time (OIT) data obtained for composite samples after
exposure to aqueous conditions were compared with the data for the samples aged in
dry air in the previous study [68], to assess the concentration of effective antioxidant.
The composites containing 2 wt.% nanofiller showed a faster decrease in OIT with
increasing time of exposure to liquid water than on exposure to dry air or air at 100%
RH (Figs. 13a,b). The data showed a linear decrease with the square root of time.
About 40% of the initial OIT remained after 700 h of exposure to liquid water, whereas
more than 80% remained after exposure to humid or dry air. The difference in loss rate
was small for samples exposed to dry or humid air.


Figure 13a: Normalized OIT (OIT (t = 0), was 48 min) for EBA composite containing 2 wt.%
uncoated ND after ageing at 90 C as a function of square root of ageing time in the following
surrounding media: dry air (; line a), air at 100 RH% (;line b), and liquid water (: line
c). The data for samples aged in dry air are from Nawaz et al. [68].



30

Results and Discussion














Figure 13b: Normalized OIT (OIT (t = 0), was 72 min) for EBA composite containing 6 wt.%
ND coated with aminopropyltriethoxysilane after ageing at 90 C as a function of square root of
ageing time in dry air (; line a), air at 100 RH% (; line b) and liquid water (; line c). The
data for samples aged in dry air are from Nawaz et al. [68].
A different behaviour was observed for composite samples containing 6 wt.%
nanofiller; the fastest decrease in OIT occurred for the composites aged in water and
the slowest for the samples aged in dry air (Figs. 14a,b). Importantly, the samples
exposed to humid air showed a rate of OIT decrease intermediate between those of the
samples exposed to dry air and those exposed to liquid water (Figs. 14a,b). Similar
observations were made with regard to the composites containing 12 wt.% nanofiller.


31

Results and Discussion


Figure 14a: Normalized OIT (OIT (t = 0), was 43 min) for EBA composite containing 6 wt.%
uncoated ND after ageing at 90 C as a function of square root of ageing time in dry air (;
line a), air with 100 RH% (, line b) and liquid water (: line c). The data for samples aged in
dry air are from Nawaz et al. [68].

Figure 14b: Normalized OIT (OIT (t = 0), was 72 min) for EBA composite containing 6 wt.%
ND coated with aminopropyltriethoxysilane after ageing at 90 C as a function of square root of
ageing time in dry air (; line a), air with 100 RH% (, line b) and liquid water (: line c).
The data for samples aged in dry air are from Nawaz et al. [68].

32

Results and Discussion

The migration rates for composites with different filler loadings under the different
conditions are summarized in Fig. 15. The loss rates from the different composite
samples were on average 4 to 5 times higher in liquid water than in dry air. The loss
rate in humid air was 2 to 3 times greater than in dry air. It should be noted that none
of the composites containing 12 wt.% nanofiller were exposed to the humid air
conditions (Fig. 15).

Figure 15: Antioxidant loss rate (average value during the first 225 h; expressed reduction in
OIT minutes per hour of exposure) plotted as a function of filler content for the following
ageing conditions: dry air at 90 C (), humid air (100% RH) at 90 C () and liquid water at
90 C (). The OIT data for samples aged in dry air are from Nawaz et al. [14]. The data point
indicated with an arrow marked A is for a composite sample containing 12 wt.% uncoated ND
with a very low initial stability (OIT = 10 min).
The faster migration from the composite sample to liquid water than to air (dry or
humid) can be explained by the boundary conditions; the dissolution rate into a water
phase was greater than the evaporation rate to a gas phase. In fact, similar
observations have been made with regard to other phenolic antioxidants [35,36]. The
impact of water present in the composite samples is studied by comparing the
migration rates in dry and in humid air. It may be assumed that the boundary
conditions, i.e. the evaporation rates, should be similar in these two cases. The
composites with the higher filler loadings (6 and 12 wt.%) showed higher migration
rates to humid air than to dry air. This suggests that the internal structure was
modified by the moisture in a way that makes diffusion faster in the moist state than in
the dry state. This may be due to a competition between water and antioxidant

33

Results and Discussion

molecules at the adsorption sites (hydroxyl groups) on the nanoparticle surfaces.


However, the AO concentration profiles were flat after ageing in dry air, which means
that the migration was controlled by the boundary conditions [68]. If the diffusion rate
increased due to the presence of moisture in the composite, this would mean that
boundary conditions would be even more dominant. Thus, the faster migration rate in
humid air than in dry air must be due a change in the boundary conditions, i.e., a
faster evaporation to humid air than to dry air. The diffusion rate cannot be calculated
in this experiment because the migration rates were controlled by the boundary
conditions.

The OIT profiles for samples exposed to liquid water at 90 C, appeared flat even after
extended periods of exposure to water (720 h) (Fig. 16). This proves that the migration
was also controlled by the boundary conditions in the case of exposure to water. As it
was shown in an earlier paper [68] that the migration to dry air was controlled by the
boundary conditions, i.e., by evaporation, it is thus clear hat the loss of Irganox 1010 to
a variety of surrounding media (dry air, humid air and liquid water) was controlled by
the loss rate at the boundary and not by the rate of diffusion.

Figure 16: OIT profiles in three composite samples (EBA containing 6 wt.% ND coated with
aminopropyltriethoxysilane) aged in liquid water at 90 C for different periods of time. The
lines are linear fits to the experimental data.
The data for the diffusivity of antioxidant were obtained by analysis of the OIT-profiles
in sandwich samples after exposure to either dry air or humid air. Half of the
sandwich samples, contained 0.2 wt.% of phenolic antioxidant (Irganox 1010) in a
composite layer with 6 wt.% nanofiller and the remaining part contained pristine EBA
without antioxidant. Fig. 17 presents a series of profiles after different exposure times.

34

Results and Discussion

The solid lines in this graph have been obtained by fitting two coupled Fickian
equations (Eq. 1 & 2, one for each layer). The adjustable parameters were the
antioxidant diffusivities in the two layers, D1 (EBA-layer) and D2 (composite layer).
The initial state (time = 0) showed flat OIT profiles in the two layers: 27 min in the
composite layer and 0 min in the pristine EBA layer. The aged sandwich samples
showed positive gradients in both of the two layers, which was an expected effect of
diffusion from the antioxidant-rich composite layer towards the antioxidant-free EBA
layer. Note that the fitting included all profiles (obtained after exposure for different
periods of time; only half of them are displayed in Fig. 17).

Figure 17: OIT profiles in sandwich samples (the noncomposite layer contained 6 wt.%
uncoated ND) after different periods of ageing in dry air: 0 h (; line a), 24 h (; line b), 144 h
(; line c), 360 h (; line d) and 720 h (; line e).

! !"! $
= !! ## " &&########## !"##$ < ! < !%&
!!
" !! %

!!

(Eq. 1)

! !!! $
= !! ## ! &&"""""""""" !"#!1.5 < x < 3
!!
" !! %

!!

35

(Eq. 2)

Results and Discussion

where x is the distance from the outer boundary of the pristine EBA layer. Suitable
boundary conditions were used to achieve an adequate numerical treatment.

The antioxidant diffusivity data is presented in Table 4. The antioxidant diffusivity
was lowest for the composites containing uncoated ND, 2 to 3 times lower than that of
the composites containing ND coated with octyltriethoxysilane or
aminopropyltriethoxysilane. This difference in diffusivity can be explained by the
higher concentration of hydroxyl groups on the uncoated ND particle surfaces (with
respect to those of the coated ND surfaces), which will have a retarding effect on the
diffusion of the antioxidant molecules. The difference between the sandwich samples
exposed to dry and humid air was smaller, the antioxidant diffusivities were 10 to 50%
greater in humid air than in dry air. It seems that the interaction between antioxidant
molecules and hydroxyl groups bound to the nanoparticle surfaces was stronger than
the interaction between the antioxidant and water bound to the hydroxyl groups.

Sample/mediuma D1b (cm2 s1)

D2c (cm2 s1)

NDUd/dry air

1.5 ! 108

1.1 ! 108

NDUd/humid air

3.6 ! 108

1.6 ! 108

NDOe/dry air

0.8 ! 108

3.7 ! 108

NDOe/humid air

1.1 ! 108

4.2 ! 108

NDAf/dry air

0.7 ! 108

3.4 ! 108

NDAf/humid air

0.8 ! 108

3.8 ! 108


Table 4: Fitted diffusivity parameters of sandwich samples.


Nanocomposite layer consisted of 6 wt % filler. Conditions for ageing: dry air (< 5 %RH) at
90 C and humid air (100 % RH) at 90C.
b Fitted value of the antioxidant diffusivity in the pristine EBA layer.
c Fitted value of the antioxidant diffusivity in the nanocomposite layer.
d NDU = composite containing uncoated ND.
e NDO = composite containing ND coated with octyltriethoxysilane.
f NDA = composite containing ND coated with aminopropyltriethoxysilane.
a

36

Results and Discussion

3.4

LOSS OF DM AND PBO FROM POLYETHYLENE IN AQUEOUS


MEDIA

3.4.1

Physical behaviour/properties of DM and PBO

Pristine DM (powder as received from the supplier) showed a narrow melting peak at
101C. Unexpectedly, no crystallization exotherm was recorded on cooling. Different
ways were adopted to promote crystallization of the DM: very slow cooling, heating
just above the melting temperature range, and possible nucleation from different
surfaces including polyethylene. However, none of these attempts led to any
detectable crystallization. Hence, the conclusion is that once DM is melted, it will
remain in an amorphous liquid or glassy state depending on the temperature.
The data presented in Fig. 18 show the mass loss for DM, PBO and the DM/PBO
(11 wt.% DM) solution at 180C. The pure compounds showed a linear decrease in
mass with time except for an initial period of 10 min during which the mass showed a
non-linear decrease with time. The rate of evaporation of PBO was almost 10 times
greater than that of DM. The curvature of the graph for the DM/PBO solution may
therefore be the result of the gradual increase in DM content in the solution. An
interesting observation is that the rate of loss of mass was greater from the solution
than from pristine PBO (curves b and c in Fig. 18). The rate of evaporation was
calculated from the average slope between 10 and 120 min of isothermal time
considering the sample area in contact with the surrounding gas phase. The
evaporation rate data are shown in an Arrhenius diagram in Fig. 19. For all three
samples, the data followed a linear trend, permitting a reliable calculation of the
activation energy for each system: 82 kJ mol-1 (DM), 66 kJ mol-1 (PBO) and 55 kJ mol-1
(DM/PBO solution). The results are partly expected; DM had a lower volatility than
PBO (Fig. 10), which is in accordance with the higher activation energy of DM. More
unexpectedly, the rate of evaporation from the solution (presumably dominated by the
evaporation of PBO) was at all the temperatures greater than that from pristine PBO;
the difference increasing with decreasing temperature. It may be assumed that the
equilibrium vapor pressure of PBO above the solution at these temperatures is greater
than that predicted by Raoults law, which in turn suggests that the intermolecular
forces between PBO and DM molecules are weaker than that between PBO molecules
as indicated by the low activation energy for evaporation from the solution
(55 kJ mol1).

37

Results and Discussion


100

Normalized mass

95

90

85
b
c

80

20

40

60

80

100

120

140

Time (min)


Figure 18: Sample mass as a function of time (isotherms; mass normalized to 100 at the start of
the isothermal period) obtained by thermogravimetry at 180C for: (a) pristine DM; (b) pristine
PBO; (c) DM/PBO solution (11.1 wt.% DM).


Figure 19: Arrhenius diagram based on the isothermal rate of mass loss for: (a) pristine DM;
(b) pristine PBO; (c) DM/PBO solution (11.1 wt.% DM). Note that the ordinate scale is
logarithmic.

38

Results and Discussion

3.4.2

Evaporation of DM and PBO from polyethylene

The mass loss data for the polyethylene compound sheets in nitrogen at four different
temperatures are presented in Fig. 20. The initial concentrations of DM and PBO were
0.45 wt.% and 3.5 wt.%, respectively. The total initial concentration of volatiles was
thus ca. 4 wt.%. The isotherm at 180 C indicates that approximately half the volatiles
had left the sample after 120 min. The mass loss versus time curves at 150 and 180 C
showed a pronounced curvature. These mass loss rate data are included in an
Arrhenius diagram in Fig. 21 together with the mass loss rate data for the DM/PBO
solution. The activation energy for the mass loss rate from the polyethylene compound
sheets is 92 kJ mol1, which is higher than the activation energies for evaporative loss
from pristine PBO (66 kJ mol1) and from the DM/PBO solution (55 kJ mol1). The
evaporative mass loss from the polyethylene compound sheets is predicted to be small
at typical service temperatures, 104 g m2 h1 (30 C) and 3 ! 104 g m2h1 (40 C).



100
120C

110C

Normalized mass

150C

99

180C

98

97

20

40

60

80

Time (min)

100

120

140


Figure 20: Sample mass as a function of time (mass normalized to 100 at the start of the
isothermal period) obtained by TG at temperatures between 110 and 180 C for the PE
compound sheets.

39

Results and Discussion



Figure 21: Arrhenius diagram based on the isothermal rate of mass loss for the polyethylene
compound sheet samples: initial loss rate; loss rate at 3 wt.% concentration of volatiles.
Mass loss rate for the DM/PBO solution (11.1 wt.% DM) are presented for comparison;
indicated by .

3.4.3

Degradation of DM and PBO in PE compound in aqueous


medium

Fig. 22 presents the IR spectra for the unexposed polyethylene compound, showing
strong absorption peaks at 2950, 1470 and 720 cm-1 that were assigned to the methylene
units in the polymer [69,70]. An absorption peak at 1740 cm-1 was assigned to the ester
carbonyl groups present in DM and Irganox 1010 [71,72], and weaker absorption peaks
at 1040 and 1150 cm-1 were assigned to ether groups in PBO [7375].
The polyethylene compound sheets exposed to humid air at 80 C for 30 days showed
an IR spectrum no different from that of the unexposed sheet, whereas samples
exposed to liquid water at elevated temperatures showed significant differences in
their IR spectra. The spectral differences increased with increasing exposure time (Fig.
22). Twin sharp absorption bands at 1540 and 1575 cm-1 observed after three days,
exposure or longer were assigned to the formation of hydroxyl groups [75] due to
reactions between water and DM [76,77]. Hydrolysis of the ester group in DM led to
the formation of decamethrinic acid and an alcohol, both containing hydroxyl groups.
The nitrile group may also undergo hydrolysis, forming species containing hydroxyl
groups. The normalized absorbance of the 1540 and 1575 cm-1 bands increased with
increasing time of exposure to water, finally reaching saturation after 10 to 15 days at

40

Results and Discussion

80 C (Fig. 23a). The growth of the 1540 and 1575 cm-1 bands was faster at 95 C,
leading to even higher relative absorbance values after 27 days of exposure (Fig. 23b).
Fig. 22 also shows a pronounced increase in the ether group absorption band (1040 cm
1; originating from PBO) with increasing exposure time at 80 C. The same tendency
was observed after exposure to water at 95 C. There was a gradual increase in the
1040 cm1 absorption with increasing exposure time at both temperatures; in fact the
rate of increase in absorbance was practically the same at the two temperatures. The
concentration profiles of both DM and PBO through the cross-section of the sheets
were completely flat after exposure to water. The gradual increase in PBO
concentration in the surface layer as indicated by the ATR-IR spectroscopy method
suggests that a thin layer rich in PBO was formed towards the water phase. Similar
profiles were obtained for sheets aged in humid air. Hence, it can be concluded that
the migration of PBO and DM to both liquid water and humid air was controlled by
the boundary conditions and not by diffusion.




Figure 22: IR spectra of polyethylene compound samples exposed to liquid water at 80 C for
different periods of time.



41

Results and Discussion

(a)

(b)



Figure 23: Relative absorbance values (normalized with respect to the methylene absorption
band at 2950 cm) for the 1540 cm1 (); and 1575 cm1 (); bands for the polyethylene
compound samples exposed to liquid water plotted as a function of exposure time at: a) 80C; b)
95C.

3.4.4

Migration of DM and PBO in PE compound

Liquid chromatography was used to determine the concentrations of DM and PBO in


the polyethylene compound samples exposed to different media. Fig. 24 shows
chromatograms for a series of samples exposed to liquid water at 80 C. PBO and DM
showed significantly different retention times, 4.0 0.1 min (PBO) and 4.8 0.1 min
(DM) and both the DM and PBO peaks showed a gradual decrease in intensity with
increasing exposure time. A new peak at 4.4 min retention time appeared after 3 days
of exposure and the intensity of this peak increased with increasing exposure time (Fig.
24). It is suggested that this 4.4 min peak was due to degradation products originating
from DM. The lack of standard solutions makes conversion to an absolute
concentration impossible, but the intensity of the 4.4 min peak increased more rapidly
at the higher test temperature (95 C).

42

Results and Discussion

PBO Degradation peak

Absorbance (mAU)

DM

30 days

15 days

3 days

Unaged

3.6

4.4

4.8

5.2

Retention time (min)


Figure 24: Liquid chromatograms of extracts from polyethylene compound samples exposed to
liquid water at 80 C for different periods of time. The assignments of the different peaks (PBO,
DM and degradation products of DM) are shown in the graph.

For DM and PBO, on the other hand, standard solutions were available and the
concentrations could be calculated and followed as a function of exposure time (Figs
25 and 26). There is a considerable scatter in the data, but the general trends are clear.
The scatter in the data makes a reliable assessment of the activation energy difficult.
Clearly DM reached a state of depletion faster than PBO. One factor affecting this is the
lower initial concentration of DM, only 1/8 of the initial PBO concentration. In
addition, the loss of DM was due to two mechanisms; (i) migration to the aqueous
phase and (ii) hydrolysis (chemical consumption).
Polyethylene compound samples exposed to humid air (60 and 80 % RH) at 80 C
showed no statistically significant decrease in DM and PBO concentration, in
agreement with the evaporation. A calculation of the expected loss of DM and PBO
based on these data (high-temperature data obtained by thermogravimetry) yielded
values smaller than the observed scatter in the data obtained by liquid
chromatography.

43

Results and Discussion

12

DM concentration (mg/l)

10
8
80C

6
95C

4
2
0

10

15
Time

1/2

20

25

30

(h)



Figure 25: Concentration of DM in polyethylene compound samples exposed to liquid water
plotted as a function of the square root of the exposure time at: , 80 C and , 95 C.

PBO concentration (mg/l)

100





80

80C

95C
60



40



20


0

0
5
10
15
20
25
30




1/2

Time (h)
Figure 26: Concentration of PBO in polyethylene compound samples exposed to liquid water
plotted as a function of the square root of the exposure time at: , 80 C and , 95 C.

44

Results and Discussion

3.5

LONG-TERM PERFORMANCE OF STABILIZED


POLYETHYLENE IN WATER CONTAINING CHLORINE
DIOXIDE

3.5.1

Consumption of antioxidant in squalane and polyethylene


Fig. 27 shows the oxidation induction time for the stabilised squalane and
polyethylene (with different phenolic antioxidants) before exposure to the chlorine
dioxide solution. It has been known for many years that the oxidation induction time is
proportional to the concentration of effective (phenolic) antioxidant in polyolefins
[58,59], and the same relationship applies to phenolic antioxidants in squalane [57].
The oxidation induction time of the polyethylene samples was linearly related to the
oxidation induction time of the squalane samples. The mean error of estimate of the
oxidation induction time of the polyethylene samples was 15 %. The antioxidant
concentration in the samples showed individual variations within 10 % of the mean
values (0.1 wt.%). The difference in oxidation induction times between the different
samples (with different antioxidants) was partly due to differences in the molar
concentration of the phenolic groups per gram sample (COH); the antioxidants with low
COH generally showed short oxidation induction times.


Figure 27: Initial oxidation induction time of PE tape samples plotted as a function of initial
oxidation induction time of squalane for the six different antioxidants. The line is a linear fit to
the experimental data with the constraint that the line should pass through the origin.

45

Results and Discussion

On exposure to aqueous chlorine dioxide, the oxidation induction time for the
squalane samples showed a linear decrease with increasing exposure time. On the
other hand, the polyethylene tape samples showed an almost linear decrease in
oxidation induction time until 60 to 75 % of the initial antioxidant has been lost and
thereafter a significantly slower and non-linear decrease until complete depletion.
A linear relationship was found between the time to reach antioxidant depletion in the
polyethylene samples and the time to reach antioxidant depletion in the squalane
samples (Fig. 28). The slope of the line is 6.6, i.e. the data showed a better resolution
in the polyethylene samples than in the squalane samples. The data for the samples
based on AO5 were not included since they showed a significant deviation from the
regression line. The longest antioxidant lifetimes were obtained for polyethylene
samples containing AO3 (ca. 2500 min) and AO2 (ca. 2000 min).



Figure 28: Time to reach depletion of the antioxidant from the polyethylene test plotted versus
the time to reach depletion of the antioxidant in the squalane. The line is a linear fit to the
experimental points.

The initial slopes of the OIT vs. exposure time plots were converted into rates of
antioxidant consumption (since it was known that the initial concentration in all the
systems studied was 0.1 wt.%). The rates of antioxidant consumption in polyethylene
samples were calculated on the basis of the initial linear region. The relationship
between the antioxidant consumption rates in the two media (squalane and
polyethylene) is presented in Fig. 29. The data for systems stabilized with AO5 are not

46

Results and Discussion

included for the same reasons as in Fig. 28. It is thus possible to rank different
antioxidant systems with regard to their efficiency in these chlorinated aqueous media
based on short-term data (<300 min) by both the squalane test and the polyethylene
tape test.



Figure 29: Rate of consumption of antioxidant (note the factor of 10 000) in polyethylene
(obtained from the OIT-slope during the first linear regime) plotted versus the rate of
consumption of antioxidant in squalane.

3.5.2

Degradation of polyethylene

To study the degradation of polymer, scanning electron micrographs of the unexposed


and exposed samples after drawing beyond necking were studied. The unexposed
samples and the samples exposed for short periods of time (less than the time to reach
antioxidant depletion) showed a smooth and fibrous surface texture. The samples
exposed for longer periods of time, i.e. significantly longer than the time to reach
antioxidant depletion, displayed characteristic island shaped surface crack pattern
(Figs. 30a,b). Between the cracks, the surface texture resembled that of the fresh
samples (Fig. 30b), suggesting that the surface cracks stopped beneath the border
between highly degraded and fresh polymer. The samples examined in the scanning
electron microscope were tilted at an angle of 60 with respect to the electron beam,
making it possible to measure the crack depth in the electron micrographs.


47

Results and Discussion


(a)

(b)




Figure 30: Scanning electron micrographs of the upper surface of drawn specimens of tapes
stabilised with AO3 exposed to water containing chlorine dioxide for (a) 5400 min and (b) 7560
min.


Figure 31: Crack depth as a function of exposure time for tapes of polyethylene stabilised with
AO3. The arrow indicates the time to reach antioxidant depletion as assessed by oxidation
induction time measurements.
The data presented in Fig. 31 shows the crack depth versus exposure time for the tape
samples stabilized with AO3. The surface embrittlement occurred after 4000 to
4500 min of exposure to the chlorine dioxide solution, compared with an antioxidant

48

Results and Discussion

depletion time of 2500 min. The increase in crack depth for this material was of the
order of 5 m per 1000 min exposure (Fig. 31). The tape samples based on the other
antioxidants showed a similar behaviour; no surface cracking was observed until well
beyond the antioxidant depletion time, whereas after longer exposure times samples
displayed surface cracking after necking.


Figure 32: Carbonyl index (A1710/A1465) plotted as a function of exposure time for tapes of
polyethylene stabilised with AO3.

The carbonyl index (i.e. the ratio of carbonyl absorbance at 1710 cm1 to that of the
methylene band at 1465 cm1) is plotted as a function of chlorine dioxide exposure time
shown in Fig. 32. Until antioxidant depletion (on the basis of DSC-OIT measurements
on 0.3 mm thick samples) the spectrum displayed no carbonyl absorbance peak, but
samples exposed for only marginally longer periods of time showed a gradually
increasing intensity of the carbonyl band (Fig. 32). The upturn in the curve of the
carbonyl index versus exposure time occurred immediately after antioxidant depletion
but prior to the start of marked surface embrittlement. All the tape samples reaching
the state of antioxidant depletion (oxidation induction time = 0) were studied by
infrared spectroscopy and none of them displayed any surface oxidation.

49

Conclusions

4 CONCLUSIONS
Polymer nanocomposites based on poly(ethylene-co-butyl acrylate) with 13 wt %
butyl acrylate and a high melt viscosity showed good dispersion of the
nanoparticles. The nanoparticle agglomerates present after surface treatment were
fragmented by shearing of the polymer melt during extrusion compounding. The
other poly(ethylene-co-butyl acrylate) with 28 wt % butyl acrylate and a low melt
viscosity was not able to break down the nanoparticles agglomeration due to
insufficient shear stresses. The nanoparticles with the higher specific surface area
and higher specific hydroxyl group concentration more easily formed
agglomerates; the attractive forces between the nanoparticles must be greater for
this nanofiller than for the other nanofillers with a lower specific surface area and
a lower hydroxyl group concentration. The surface treatment of these
nanoparticles however reduced these attractive forces between the nanoparticles
and gave composites with a better nanoparticle dispersion.

The concentration of antioxidant in the poly(ethylene-co-butyl acrylate)/Al2O3
nanocomposites was assessed by oxidation induction time (OIT) measurements.
X-ray photoelectron spectroscopy and OIT showed that the antioxidants were
adsorbed onto aluminium oxide nanoparticles in well-dispersed nanocomposites.
Surface treatment of the nanoparticles with either octyltriethoxysilane or
aminopropyltriethoxysilane resulted in a significant reduction in antioxidant
adsorption. It was observed that the adsorbed antioxidant was not effective in
inhibiting polymer oxidation in this particular state whereas there were
indications of a slow release of antioxidant from the adsorbed state on prolonged
ageing at 90C.

The EBA nanocomposites exposed to liquid water showed a faster reduction in the
OIT than the samples exposed to both dry & humid air. The migration of
antioxidant was controlled by boundary conditions during ageing in humid air
and liquid water. The antioxidant diffusivity was lower in the composites
containing uncoated ND than in those containing ND coated with

50

Conclusions

octyltriethoxysilane or aminopropyltriethoxysilane. This difference in diffusivity


was due to the higher concentration of hydroxyl groups on the uncoated ND
particle surfaces than on those of the coated ND surfaces, which retard the
diffusion of the antioxidant molecules.

It was observed that deltamethrin (DM), which is a highly crystalline substance
with a melting point of 101 C, never regained its crystallinity after being melted
but remained in an amorphous liquid state at typical service temperatures.
Infrared microscopy showed that the loss of DM and PBO from the polyethylene
compound was controlled by the boundary conditions, i.e. by evaporation to a
surrounding gas phase or dissolution into an aqueous phase. Rates of evaporation
of DM and PBO in a variety of systems were determined by thermogravimetric
analysis at high temperatures, and their temperature dependence was well
described by the Arrhenius equation. Loss of DM and PBO to liquid water is a
more serious problem from a practical point of view. Infrared spectroscopy
showed that DM was hydrolyzed in contact with water.

Polyethylene tape samples stabilized with different phenolic antioxidants (0.1 wt
%) were exposed to water containing 10 ppm chlorine dioxide at 70 C. A linear
relationship was established between the time to reach antioxidant depletion in
the polyethylene tape samples and the corresponding time in samples of squalane
containing the same antioxidant. Surface oxidation and surface embrittlement
were detected by infrared spectroscopy and scanning electron microscopy of the
tape samples after they had been stretched beyond necking. The idea was to
propose a method that can efficiently assess different aspects of the long-term
performance of polyethylene materials and their stability towards water
containing chlorine dioxide.

51

Acknowledgements

5 ACKNOWLEDGEMENTS
First and foremost I would like to express my deepest Gratitude (with a capital
and bold G) to my supervisor, Professor Ulf Gedde, for accepting me as a PhD
student, and for his outstanding guidance and endless support throughout my
work. It has been an inspirational journey (that goes back to 2007, when I started
as a masters thesis student) and I have learnt a lot from you Ulf. Secondly I would
like to express my sincere gratitude to Professor Mikael Hedenqvist for all the
support, discussions (both scientific & otherwise), and for being friendly and very
cooperative.
I would also like to thank Henrik Hillborg (my co-supervisor), your cooperation
and guidance have always been of great value. I am very thankful to the members
(Bruska Azhdar, Patricia Nordell, Nadja Jverberg) of the nano-project for the
discussions and being helpful. The members of the reference group in the nano-
project, Claire Pitois (ABB), Sven Jansson (Elforsk), Eva Malmstrm (KTH), Hans
Edin (KTH), Christian Andersson (Ericsson), are also thanked for support and
feedback. I would like to mention and thank the members of the Borealis project,
Thomas Hjertberg, Jeroen Oderkerk (Brealis, Stenungsund) and Francis Costa
(Borealis, Linz). Matthieu Zellweger is also thanked for helpful discussions and
support.
Elforsk, the Swedish Research Council, ABB, EIT KIC InnoEnergy, Vestergaard
and Borealis are thanked for the financial support.
I am very thankful to Hasan & Peter (I wish I could use that Turkish word, that
you taught me/us) for being such great friends and the awesome time we spent,
especially the afterwork (thank you for everything bros). Very special thanks to
Nazanin (my nice office fellow), Erik (i know its Linde not Linda), Wenbin (good
luck with your thesis miss), Dahlia, Sevil, Emma, Thorsak, Richard Andersson,
Chung, Thomas, Fritjof (the modeling king), Sung Woo, Nima, Irfan, Fei, Nils,
Qiong, Love, Amir, Dongming, for being good friends, colleagues and making the
work place so lively. Richard Olsson and Fredrik Svensson are also thanked for
being very supportive and friendly.

52

Acknowledgements

I want to thank everyone in the surface coating division, polymer technology and
fibre technology.
I would also like to mention Gareth Bayley (aka Professor Bayley) here, thank you
for the great time we spent and I hope you still remember TGI, in here its always
Friday. Stefano Farris, my Italian friend from the lovely city of Alghero, thank
you for being such a nice host and for those discussions (scientific and otherwise)
during espresso breaks. Thank you also to all the friends outside of KTH.
Last but certainly not the least; a big thank you to my mom and dad for always
being there for me, to my brothers (Fakhar & Ali) and also my sister, for their love
and continuous support, you guys mean a lot to me.



53

References

6 REFERENCES
1.

Okada O, Kawakami M, Kojima Y, Kurauchi T, and Kamigaito O,


Synthesis and Properties of Nylon-6/Clay Hybrids, Proc. MRS Symp.
1990;171:45.
Tanaka T, Kindersberger J, Frchette M, et al. Polymer nanocomposites:
fundamentals and possible applications in power sector, CIGR Publication
451; 2011.
Ma D, Siegel RW, Hong J-I, Schadler LS. J Mater Res 2004;19:857.
Perrin-Sarazin F, Sepher M, Bouricha S, Denault J. Polym Eng Sci
2009;49:651.
Kozaka M, Fuse N, Ohki Y, Okamoto T, Tanaka T. IEEE Trans Dielectr
Electr Insul 2005;11:833.
Fuse N, Kozaka M, Tanaka T, Murase S, Ohki Y. Ann Report IEEE-CEIDP
2004;4-4:322.
Kozaka M, Kido R, Fuse N, Ohki Y, Okamoto T, Tanaka T. Ann Report
IEEE-CEIDP 2004;5A-15:398.
Kozako M, Yamano S, Kido R, Ohki Y, Kohtoh M, Okabe S. Proc ISEIM
2005, 231.
Kozako M, Kido R, Imai T, Ozaki T, Shinizu T, Tanaka T. Proc ISEIM 2005,
661.
Nelson JK, Hu Y. Ann Report IEEE-CEIDP 2003;8-2:719.
Nelson JK, Hu Y. Ann Report IEEE-CEIDP 2004;7P-10:832.
Zilg C, Kaempfer D, Mlhaupt R, Montanari GC. Ann Report IEEE-CEIDP
2003;6-5:546.
Imai T, Sawa F, Ozaki T, Nakano T, Shimizu T, Yoshimitsu T. IEEE Trans
A 2004;A-124, 11, 1064.
Ding HZ, Varlow BR. Ann Report IEEE-CEIDP 2004, 4-6, 332.
Ma D, Siegel RW, Hong JI, Schadler LS, Mrtensson E, nneby C. J Mater
Res 2004;19:857.
Roy M, Nelson JK, MacCrone RK, Schadler LS. IEEE Trans Dielectr Electr
Insul 2005;12:629.

2.

3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.

54

References

17.
18.
19.
20.

Chang CC, Chen WC, Chem Mater 2002;14:4242.


Li H, Liu G, Liu B, Chen W, Chen S. Mater Lett 2007;61:1507.
Tsai MH, Liu SJ, Chiang PC. Thin Solid Films 2006;515:1126.
Abboud M, Turner M, Duguet E, Fontanille M, J. Mater. Chem 1997;7:
1527.
Guo Z, Pereira T, Choi O, Wang Y, Hahn H. Thomas. J. Mater. Chem
2006;16:2800.
Witucki GL. J Coating Technol 1993;65:57.
Posthumus W, Magusin PCMM, Brokken-Zijp JCM, Tinnemans AHA,
Van der Linde R. J Colloid Interf Sci 2004;269:109.
Olsson RT, Hedenqvist MS, Strm V, Deng J, Savage SJ, Gedde UW.
Polym Eng Sci 2011;51:862.
Hawker CJ, Bosman AW, Harth, E. Chem Reviews 2001;101:3661.
Moad G, Chong YK, Postma A, Rizzardo E, Thang SH. Polymer
2005;46:8458.
Edmondson S, Osborne VL, Huck WTS. Chem Soc Rev 2004;33:14.
Allara David L, Environmental Health Perspectives, Vol. 11, pp 29-33,
1975.
Wise J, Gillen KT, Clough RL. Polymer Degrad Stab 1995;49:403.
Gillen KT, Bernstein R, Clough RL, Celina M. Polym Degrad Stab
2006;91:2146.
Yang CQ, Martin LK. J Appl Polym Sci 1994;51:389.
Brll R, Geertz G, Kothe H, Macko T, Rudschuck M, Wenzel M. Macromol
Mater Eng 2008;293:400.
Karlsson K, Smith GD, Gedde UW. Polym Eng Sci 1992;32:649.
Smith GD, Karlsson K, Gedde UW. Polym Eng Sci 1992;32:658.
Lundbck M, Strandberg C, Hedenqvist MS, Albertsson AC, Gedde UW.
Polym Degrad Stab 2006;91:1071.
Lundbck M, Hedenqvist MS, Mattozzi A, Gedde UW. Polym Degrad
Stab 2006;91:1571.
Viebke J, Gedde UW. Polym Eng Sci 1997;37:896.
Allen NS, Edge M, Ortega A, Sandoval G, Liauw CM, Stratton J. Polym
Degrad Stab 2004;85:927.
Allen NS, Edge M, Ortega A, Liauw CM, Stratton J, McIntyre RB. Dyes
and Pigments 2006;70:192.
Crank J. ''The mathematics of diffusion''; Oxford: Clarendon Press; 1986.
Allen NS, Edge M, Corrales T, Childs A, Liauw C, Catalina F, et al. Polym
Degrad Stab 1997;65:125.

21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.

55

References

42.

Liauw CM, Childs A, Allen NS, Edge M, Franklin KR, Collopy DG. Polym
Degrad Stab 1999;65:207.
Gao X, Hu G, Qian Z, Ding Y, Zhang S, Wang D, et al. Polymer
2007;48:7315.
Erstfeld KM. Chemosphere 1999;39:1737.
Kovacic T, Mrklic Z, Thermochim. Acta 1998;322:129.
Aruthur FH, J. Stored Products Res. 2004;40:317.
Kljajic P, Peric I. J. Stored Products Res. 2007;43:523.
Rodriguez DG, Torrijos RC, Ferreira RAL, Diaz AMC. Food Chem.
2012;135:259.
Lakshmanan LCW. Adv. Drug Delivery Reviews 1999;38:113.
Kendrick RK, Kendrick MK, Focheux C, Dereure J, Puech MP, Cadiergues
MC, Medical Veterinary Entomology 1997;11:105.
Ifwarson M, Aoyama K. Proceedings of the Plastic Pipes X Conference,
Gothenburg; 1998.
Bradley SW, Bradley WL. Proceedings of the Conference of Deformation,
Yield and Fracture, Cambrige, UK; 1997.
Dear JP, Mason NS. Polym Comp 2001;9:1.
Hassinen J, Lundbck M, Ifwarson M, Gedde UW. Polym Degrad Stab
2004;84:261.
Yu W, Azhdar B, Andersson D, Reitberger T, Hassinen J, Hjertberg T,
Gedde UW. Polym Degrad Stab 2011;96:790.
Yu W, Reitberger T, Hjertberg T, Oderkerk J, Costas FR, Gedde UW.
Polym Degrad Stab 2012;97:2370.
Azhdar B, Yu W, Reitberger T, Gedde UW. Polym Testing 2009;28:661.
Billingham NC, Bott DC, Manke AS. In: Grassie N, editor. Developments
in polymer degradation3, London: Applied Science Publishers, 1981,
p.63.
Howard JB. Polym Eng Sci 1973;13:429.
Nordell P, Nawaz S, Azhdar B, Hillborg H, Gedde UW. J Appl Polym Sci
2012;125:975.
Ma D, Siegel RW, Hong J-II, Schadler LS. J Mater Res 2004;19:857.
Zhao S, Auletta T, Hillborg H, Schadler LS. Comp Sci Tech 2008;68:2965.
Baumgarten E, Wagner R, Lentes-Wagner C, Fresenius Z. Anal Chem
1989;334:246.
Hall WK, Lutinski FE, Gerberich HR. J Catal 1964;3:512.
Gijsman P, Gittan-Chevalier M. Polym Degrad Stab 2003;81:483.
Desai J, Pendyala VNS, Xavier SF, Misra AN, Nair KB. J Appl Polym Sci
2004;91:1097.

43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.

59.
60.
61.
62.
63.
64.
65.
66.

56

References

67.

Balint G, Halasz A, Iring D, TdoNs F. Angew Macromol Chemie


1983;114:151.
Nawaz S, Nordell P, Hillborg H, Gedde UW. Polym Degrad Stab
2012;97:1017.
Fox JJ, Martin AR. Proc. Roy. Soc. 1940;A175:208.
Nielsen JR, Woollett AH. J. Chem. Phys. 1957;26:1391.
Luongo JP. J. Polym. Sci. 1960;42:139.
Grafmller F, Husemann E. Macromol. Chem. 1960;40:161 and 170.
Nair RK, Kadam MK, Sawant MR, J. Dispersion Sci. Technol. 2007;27:1015.
Ibrahim M, Nada A, Kamal DE. Indian J. Pure Appl. Phys. 2005;43:911.
Montes JC, Cadoux D, Creus J, Touzain S, Maurin EG, Correc O. Polym.
Degrad. Stab 2012;97:149.
Lutnicka H, Bogacka T, Wolska L. Water Res 1999;33:3441.
Anand SS, Bruckner JV, Haines WT, Muralidhara S, Fisher JW, Padilla S.
Toxicol. Appl. Pharmacol. 2006;212:156.

68.
69.
70.
71.
72.
73.
74.
75.
76.
77.

57

Você também pode gostar