You are on page 1of 10

Morphology and Thermal Properties of Poly(ethylene

terephthalate)-Modified Kaolin Nanocomposites

Khalil Shahverdi–Shahraki,1 Tamal Ghosh,2 Kamal Mahajan,2 Abdellah Ajji,1 Pierre J. Carreau1
1
CREPEC, Department of Chemical Engineering, Polytechnique Montreal, Montreal, Quebec,
Canada H3C 3A7

2
PepsiCo Advanced Research—Beverage Packaging, Hawthorne, New York 10532

Kaolin particles were dispersed in a poly(ethylene ter- parison with other engineering materials such as metals,
ephthalate) (PET) matrix using a displacement proce- different type of fillers have been incorporated to polymer
dure in which kaolin was initially treated with potassium
acetate and, subsequently, melt-blended with PEO and matrices to overcome these shortcomings. The sizes of
PET. The disappearance of characteristic peaks in XRD these traditional fillers are usually in the range of micro-
patterns of chemically treated particles revealed that meters [1]. With the development of nanoparticles, many
the crystalline form and layered structure of kaolin par- researchers have considered using them as nanofillers to
ticles were mainly destroyed. Scanning electron micros-
reinforce polymer composites. Nanoparticles can provide
copy and transmission electron microscopy showed
that the thickness of dispersed particles were generally a large contact area between different phases in the com-
in the range of 10–100 nm. It was also observed that the posite, which may result in a significant reinforcement
dispersion level of the chemically treated kaolin was effect on polymers [2–10].
much better than that of “as received” kaolin particles. Kaolin is a layered aluminosilicate in which each layer
Rheological studies showed the formation of a network-
like structure in samples containing the chemically in the structure comprises two sublayers: an alumina
treated kaolin as a result of improved dispersion of par- octahedral sheet and a silica tetrahedral sheet that share a
ticles. It was also observed that both the kaolin particles common plane of oxygen atoms. Regarding the asymmet-
and modifiers could accelerate the degradation process ric structure of kaolin layers, which creates large super-
and lower the molecular weight of the PET. The intro- posed dipoles, and hydrogen bonds between oxygen
duction of kaolin particles shifted the crystallization
temperatures of PET to higher temperatures. It was atoms on one side of each layer and hydroxyl groups on
observed that the incorporation of a chain extender sig- the other side of adjacent layers, there will be a large
nificantly restricted the crystallization process. The cohesive energy between the layers [11]. As a conse-
Avrami–Jeziorny analysis confirmed the alteration of the quence of this cohesive energy, only some limited
crystalline structure of the filled polymer. Based on TGA
thermograms the decomposition temperature of PET–
organic molecules, such as potassium acetate and
kaolin composites was slightly lower than that of the dimethyl sulfoxide can intercalate the space between the
neat PET. POLYM. COMPOS., 37:1443–1452, 2016. V C 2014 Soci- layers of kaolin [12–20].
ety of Plastics Engineers Recently, a few investigations on the incorporation of
organic modified kaolin in polymers have been reported
[21–25]. Komori et al. [26, 27] intercalated poly(vinyl
INTRODUCTION pyrrolidone) into kaolin interlayer spaces at room temper-
Nanocomposites are a new class of engineering materi- ature via a solution method. Liu and coworkers [28]
als, which have found many applications in various indus- reported the intercalation of poly(styrene/maleic anhy-
trial fields such as automotive, construction, and dride) into kaolin via in situ polymerization using kaolin-
packaging due to their good properties, low cost and DMSO (dimethyl sulfoxide) as a starting material. They
weight. Based on the fact that polymeric materials suffer achieved a partially exfoliated structure when the kaolin
from lack of thermal stability and low modulus in com- content was in the range of 1–10 wt%. Turhan et al. [29]
prepared a PVC-kaolin nanocomposite via a solution
method; at first the interlamellar spaces were expanded
Correspondence to: A. Ajji; e-mail: abdellah.ajji@polymtl.ca by DMSO and then PVC chains were intercalated into the
Contract grant sponsor: PepsiCo.
DOI 10.1002/pc.23313
galleries resulting in a complete exfoliated structure.
Published online in Wiley Online Library (wileyonlinelibrary.com). Polystyrene was intercalated into kaolin using kaolin-
C 2014 Society of Plastics Engineers
V DMSO intermediate [30]. Tunney and coworkers [31]

POLYMER COMPOSITES—2016
prepared poly(ethylene glycol)/kaolin intercalates using obtained using a Hitachi S4700 instrument with a cold
kaolin-dimethyl sulfoxide (kao-DMSO) and reported an field emission gun under an acceleration voltage of 2 kV.
expansion in interlayer galleries of kaolin. Itagaki et al. A JOEL JEM-2100F transmission electron microscopy
[32] produced nylon 6/kaolin composites via melt mixing (TEM) operating at 200 kV was used to evaluate the tac-
in a twin-screw extruder. They prepared an intercalated toid size, presence of agglomerates, and to fully charac-
compound by the polymerization of 6-aminohexanoic terize the morphology of the filler. The samples were
acid (AHA). The results indicated that mechanical proper- microtomed into 50- to 80-nm thick slices using an
ties had been improved. Sunitha George et al. [33] Ultracut FC microtome (Leica, Germany) with a diamond
reported the production of PS/HDPE-nanokaolin compo- knife. Rheological measurements were conducted using a
sites; however, a high concentration of compatibilizer parallel plate stress-controlled rheometer (Gemini of Mal-
(10–15%) was needed to obtain acceptable properties vern) with a gap size of 1 mm and a plate diameter of
Numerous studies have focused on the development 25 mm. Time and frequency sweeps in small-amplitude
and on the properties of PET-MMT composites; however, oscillatory shear were carried out at 270 C. The rheome-
very few reports can be found in the literature about the ter was equipped with a convection oven to control the
dispersion of other types of clay mineral in a PET matrix. temperature. To minimise the thermal degradation all the
In this work, the fabrication of PET/kaolin nanocomposite measurements were performed under nitrogen atmosphere
is reported. The influence of chemical treatment of kaolin in the linear viscoelastic regime. Melting and crystalliza-
particles with potassium acetate (KAc) and poly (ethylene tion characteristics of the samples were determined by
oxide) on the morphology and crystallization behavior of differential scanning calorimetry (DSC) using a
PET/kaolin nanocomposites is investigated. The crystalli- DSCQ1000 TA Instrument under helium atmosphere. A
zation kinetics were studied in nonisothermal conditions thermogravimeter TGA500 TA Instrument was used to
in order to determine the effect of the treated particles on study the thermal properties. About 10 mg of the samples
the nucleation and growth of crystals in the crystallization were heated at 10 C min21 from 30 to 700 C under nitro-
of PET. gen atmosphere.

EXPERIMENTAL Sample Preparation


A stepwise procedure was followed in order to interca-
Materials late the PET resin into the kaolin clay. In the first step
kaolin was intercalated by KAc, which is capable of pen-
A bottle grade polyethylene terephthalate (PET) resin
R etrating between the layers and expanding the structure of
Laser1V 7000 was supplied by DAK Americas LLC. It
kaolin. This was the kao–KAc precursor. In the next step
had an intrinsic viscosity of 0.84 dL g21, a melting point
this precursor was blended with the PEO and finally
of 242 C and 35% crystallinity.
intercalation of PET chains proceeded by displacing of
A commercial grade kaolin (denoted hereafter as kao),
KAc and PEO molecules.
hydrous aluminosilicate, with a density of 2.58 g cm23 at
The kaolin was intercalated with KAc via a solution
25 C (particle size: D90 5 2 lm) was supplied by BASF
method in which 60 g kaolin was mixed with a KAc solu-
Corporation. poly(ethylene oxide), PEO, with a molecular
tion (50 g KAc dissolved in 120 g water) and stirred at
weight of 100,000 g mol21 and potassium acetate (KAc)
room temperature for 48 h. The treated particles were
were purchased from Sigma–Aldrich and used as
separated from the mixture by centrifugation at 5,000 rpm
received. A 30 wt% masterbatch of PET chain extender
for 5 min and dried in a vacuum oven at 70 C for 24 h.
(denoted hereafter as ch) supplied by Polyvel was used to
A PEO–kaolin masterbatch (denoted hereafter as kao-
raise the intrinsic viscosity and increase the melt strength
KAc-PEO) was prepared in a Brabender internal mixer.
of the PET resin during the melt blending. This product
Treated kaolin and PEO were fed into the chamber at a
has full FDA compliance for food contact applications. It
weight ratio of 0.33 and the mixing process was done at
contains JoncrylV R ADR 4368–F, which is a styrene-
90 C for 10 min with a rotating speed of 100 rpm.
acrylic multifunctional epoxy-based chain extender with a
PET granules were ground in liquid nitrogen to obtain
molecular weight (Mw) of 6,800 g mol21.
a fine powder of PET. The PET powder was manually
mixed with a kaolin powder (or ground PEO masterbatch)
for a few minutes and then the mixture was melt-blended
Characterization and Testing
by extrusion using an 18 mm diameter twin screw
A Philips X’Pert diffractometer (Cu Ka radiation, extruder (LEISTRITZ Extruder, L/D 5 40) to obtain a
k 5 1.54056 A ), operating at a voltage of 50 kV and cur- masterbatch containing 20 wt % kaolin particles. The
rent of 40 mA, was used for wide angle X-ray diffraction temperature profile was set in the range of 250–275 C
(WAXD) measurements. The scanning rate was 0.02 s21 and the screw rotating speed was 150 rpm. The master-
and the 2u ranged from 2 to 15 . Scanning electron batch was eventually diluted with the neat PET to yield
microscopy microphotographs of the samples were the final composites containing 1 wt% filler loadings; the

1444 POLYMER COMPOSITES—2016 DOI 10.1002/pc


FIG. 1. XRD patterns: kaolin particles before and after treatment (a) and neat PET and PET/kaolin composites (b).

chain extender (1 wt%) was added to the composites at in the pattern. This could be a sign of partial delamina-
this stage if needed. Because the presence of ppm tion and destruction of crystalline structure of kaolin
amounts of moisture in PET during the melt compound- layers in the PET matrix. After melt-mixing with PET,
ing leads to hydrolysis reactions, all the materials were the kaolin stacks which had been already intercalated by
well dried in a vacuum oven at 110 C for 24 h before KAc and PEO were delaminated, and therefore the dif-
being processed. fraction peaks at 0.98 and 1.38 nm (as shown in Fig. 1a)
were not detected for PET-(kao-KAc-PEO). The existence
of the characteristic peak of kaolin with very low inten-
RESULTS AND DISCUSSION sity in the XRD patterns of all nanocomposites reveals
that even after chemical treatment and compounding there
X-ray Diffraction Results are still a fraction of particles, which did not undergone
any intercalation and retained their initial structure.
Figure 1a presents the X-ray diffraction (XRD) pat- The addition of chain extender did not have a remark-
terns of the pristine kaolin, kao-KAc precursor, and kao- able effect on morphology of the composite. As shown in
KAc-PEO masterbatch. The KAc molecules were able to Fig. 1b, the XRD pattern of PET-(kao-KAc-PEO-ch) is
penetrate between the layers of kaolin and increase the almost the same as PET-(kao-KAc-PEO). The chain
basal spacing of layers from 0.72 nm up to 1.4 nm corre- extender molecules are supposed to react with PET end
sponding to 2u equal to 6.40 . It can be seen that by groups and connect the chains together in order to
blending the kao-KAc precursor with PEO, PEO mole- increase the molecular weight, but they are not likely to
cules intercalated the kaolin galleries so that the intensity have any interactions with kaolin particles.
of the characteristic peak at 0.72 nm (2u 5 12.46 ) signif-
icantly decreased and, furthermore, two new peaks
appeared at 2u of 9.02 and 6.40 , which correspond to a
SEM and TEM Images
d-spacing of 0.98 and 1.38 nm, respectively. Both PEO
and kaolin are known as hydrophilic materials, which SEM images of the “as-received” kaolin powder, kao-
possess large quantities of oxygen and hydroxyl groups in KAc-PEO masterbatch, PET-kao, and PET-(kao-KAc-
their structure. The existence of hydroxyl groups can PEO) composites are shown in Fig. 2. The kaolin par-
increase the tendency of PEO chains to interact with kao- ticles tend to flocculate and form very big agglomerates,
lin layers during the melt blending. Figure 1b presents the which are 5-20 lm in size. However, when kaolin was
XRD patterns of different PET composites. When PET mixed with PET, due to de high shear applied during the
was compounded with the untreated kaolin (as received) extrusion the flocculated particles partially broke up and
there was no change in the XRD pattern and the intense formed some aggregates in the range of a few microns
peak at 0.72 nm was clearly observed. This indicates that (Fig. 2b). As the kao-KAc-PEO masterbatch has a very
no intercalation and/or delamination occurred. On the high filler content (33 wt%) it is no surprise to observe
other hand when PET was blended with KAc-treated par- large agglomerates in the SEM image of Fig. 2c. A com-
ticles, PET-(kao-KAc), the peaks at 10 and 7.5 were parison between the images of the microtomed PET-kao
totally disappeared and the intensity of the characteristic and PET-(kao-KAc-PEO) shows that the treatment of
peak (0.72 nm) was much lower compared to as received kaolin with KAc and blending with PEO could promote
kaolin. Eventually when the kao-KAc-PEO masterbatch the dispersion of particles in the PET matrix; in many
was diluted with the neat PET the peaks at 0.98 and parts of the PET-Kao composite image of Fig. 2b no kao-
1.38 nm disappear and only the characteristic peak of lin particles are found at all while well-dispersed particles
kaolin (0.72 nm) with a very weak intensity was detected are easily observable (small white spots) in the

DOI 10.1002/pc POLYMER COMPOSITES—2016 1445


FIG. 2. SEM images: as received kaolin (a), PET-kao composite (b), kao-KAc-PEO masterbatch (c), PET-
(kao-Ac-PEO) nanocomposite (d).

micrograph of PET-(kao-KAc-PEO) in Fig. 2d. The mor- Because XRD is not reliable as a stand-alone tech-
phology of these two composites is totally different: the nique for characterising polymer nanocomposites, TEM
compound with the untreated kaolin contains a few big images were prepared to support the results of the mor-
agglomerates in different positions of the matrix, whereas phological analysis. TEM images of cross-section of the
numerous fine kaolin particles can be found in the com- PET-(kao-KAc-PEO) composite are presented in Fig. 3.
pound with the treated kaolin. Few agglomerates were They show that tactoids of several layers of kaolin were
still observed in PET-(kao-KAc-PEO), however the distributed in the PET matrix whereas no large-scale
majority of particles were uniformly dispersed. This could aggregates were observed. The length of these tactoids
be explained to the delamination of kaolin particles or was in the rage of 100-200 nm with an aspect ratio
breaking-down of the micron-size aggregates under high around 10. However, many particles with higher aspect
shear rates, which eventually lead to the dispersion of ratio can be found. It should be noted that, due to the spe-
kaolin particles in the PET matrix. cific structure of kaolin it is not categorized as

FIG. 3. TEM images of PET-(kao-KAc-PEO) nanocomposite at different magnifications.

1446 POLYMER COMPOSITES—2016 DOI 10.1002/pc


reduced the viscosity with respect to the neat matrix; this
effect has been shown elsewhere [34]. The viscosity loss is
a result of large decrease in the molecular weight of the
resin. There are two possible reasons for this reduction in
viscosity: (i) degradation of PET during the melt process-
ing, (ii) apparent slip effect at the walls of the rheometer
plates. To check if the latter has any influence on the vis-
cosity, two sets of experiments at two different gap heights
(1 and 2 mm) were performed and result confirmed that
there was no slip during the measurements. Hence, the deg-
radation should be the only reason for the severe viscosity
reduction. The degradation is induced by the thermal
decomposition of the modifiers, followed by the hydrolysis
of PET. The existence of hydroxyl groups in the structure
and on the edges of the kaolin particles has been consid-
ered as the main factor that causes the polymer degradation
during the processing at high temperature. It seems that the
hydroxyl groups in the structure of kaolin act as Brønsted
acidic sites and play the most important role in the degra-
dation of the PET chains [35, 36]. The effect of the chemi-
cal treatment on the shear viscosity of the samples is also
displayed in this figure. The viscosity of samples contain-
ing the treated fillers was drastically lower than that of the
untreated filler. The stability of the samples with time is
another point that can be concluded from Fig. 4a. The com-
plex viscosity of all samples nearly remained constant dur-
ing the time of experiment (20 min), except the composite
containing the chain extender for which the complex vis-
cosity increases significantly during the rheological meas-
urements due the reaction between the chain extender and
end groups of the PET molecules. [37]
Frequently sweep measurements on the neat PET and
PET/kaolin composites were carried out to study the
interactions between the polymer matrix and the filler
particles. The results presented in Fig. 4b and c indicates
that the treatment of the kaolin particles with KAc and
PEO improved the dispersion of kaolin particles in the
matrix. The strong shear-thinning behavior of the com-
plex viscosity for PET-(kao-KAc-PEO) and PET-(kao-
KAc-PEO-ch) observed in Fig. 4b indicated the formation
of a network-like structure due the particle-particle inter-
FIG. 4. Rheological measurements: complex viscosity vs. time (a), action, whereas the other samples exhibited a Newtonian
complex viscosity vs. frequency (b), elastic modulus vs. frequency (c) behavior up to frequencies around 100 rad s21. The com-
(test conditions: temperature 5 270 C, stress amplitude 5 15 Pa). plex viscosity of PET-(kao-KAc-PEO) nanocomposite at
higher frequencies decreased below the value of the PET-
expandable clay minerals such as MMT; therefore the dis- kao due to the degradation mechanism described befor-
persion level of kaolin should not be compared to those e.The storage (G0 ) modulus of the neat PET and PET/kao-
of PET-MMT nanocomposites, which are extensively lin composites are presented in Fig. 4c. It is shown than
reported in the literature. the composite containing treated particles had increased
melt elasticity at lower frequencies. Storage modulus is
very sensitive towards the morphology of the dispersed
Rheological Properties
phase in the polymer melt. The slope of storage modulus
To investigate the effect of chemical treatment and vs. frequency decreased from 1.60 to 0.58 after the intro-
blending, rheological experiments were conducted and the duction of treated kaolin particles indicating that the
results are presented in Fig. 4. Figure 4a shows the isother- dependence of G’ on x decreased. The significant change
mal time sweep of the melt viscosity for the composites. in the slope of G0 is typical of a network-like structure of
The addition of the kaolin particles to the PET matrix dispersed particles in nanocomposites. The data revealed

DOI 10.1002/pc POLYMER COMPOSITES—2016 1447


FIG. 5. Crystallization (a) and melting (b) behavior of different composites at heating/cooling rate of 10 C min21.

that the addition of chain extender did not have any influ- explained by heterogeneous nucleation effects of the kao-
ence on the rheological behavior of PET but shifting the lin particles in the PET matrix, which accelerates the
curves to higher values. This indicates that the chain crystallization process. It was observed that the sample
extender did not change the morphology of the particles containing chain extender crystallized at temperatures
while reacting with PET chains and increasing the viscos- even lower than that of the neat PET. The chain extender
ity of the matrix. acts as a cross linking bridge between the polymer chains
and prevent the rearrangement of PET chains in a crystal-
line structure.
Crystallization/Melting Behavior
The total crystallinity of the composite containing
The nonisothermal crystallization curves of the neat untreated kaolin is slightly lower than that of the neat
PET and PET/kaolin composites are displayed in Fig. 5. PET, as observed in Table 1. This effect was also
The onset melting temperature, Tm,o, the onset crystalliza- reported for other type of fillers [39]. Although the filler
tion temperature, Tc,o, the temperature of maximum crys- particles can promote the development of nucleation sites,
tallization rate, Tc,max, and the degree of crystallinity, Xc, the mobility of PET chains, on the other hand, is
are listed in Table 1. Xc was calculated using the follow- restricted due to the presence of the particles [40]. More-
ing equations: over, filler particles can also play the role of physical bar-
riers against the crystal growth.
DHm Figure 6 displays the crystallization and melting
Xc 5 (1)
ð12Wf 2Wp Þ3DH0 behavior of PET-kao-KAc-PEO nanocomposite at differ-
ent cooling/heating rates. The crystallization curves
where DHm is the melting enthalpy of sample, DH0 is the become broader and shift to lower temperatures with
melting heat of 100% crystalline PET which is taken 140 increasing the cooling rate. At lower cooling rates sam-
J g21 according to the literature [38], Wf and Wp denote ples have a longer period of time to form the nuclei;
the filler and PEO weight fractions respectively. It should therefore the crystallization occurs at higher temperatures.
be noticed that melting of PEO occurs at temperatures It has been suggested that at higher cooling rates the
(66-75 C) much lower than that of PET (245 C). There- motion of PET chains cannot follow the cooling tempera-
fore PEO is very unlikely to make any contributions in ture and consequently more supercooling is required to
the melting enthalpy of PET. initiate crystallization [41, 42].
As shown in Fig. 5a, for the samples containing kaolin Figure 5b shows a multiple melting behavior for all
particles the crystallization exotherm becomes narrower samples; both neat PET and PET composites displayed
and shifts to the higher temperatures. This can be two major melting peaks at two different temperatures.
This existence of multiple melting peaks was frequently
TABLE 1. Characteristic data for nonisothermal crystallization of PET reported for isothermal crystallization of PET and other
and PET/kaolin composites.
semi-crystalline polymers [43]. The first (low tempera-
u ( C Tc,o Tc,max Tm,o Xc ture) and second (high temperature) melting peaks are
min21) ( C) ( C) ( C) (%) attributed to the fusion of secondary and primary crystals
respectively [44]. Because the crystals formed by second-
Neat PET 10 205.5 192.3 193.2 29.6 ary crystallization are smaller than those formed by pri-
PET–kao 10 208.3 197.7 200.1 28.1
mary crystallization, they will melt at lower temperatures.
PET–kao–KAc–PEO 10 208.9 199.0 199.6 29.2
PET–kao–KAc–PEO–ch 10 202.3 187.4 182.0 25.2 The occurrence of multiple melting was more pronounced
at higher cooling rates. It was observed that in case of

1448 POLYMER COMPOSITES—2016 DOI 10.1002/pc


FIG. 6. Crystallization (a) and melting (b) behavior of PET-(kao-KAc-PEO) at different cooling rates.

neat PET the second peak was predominant for all heat- method showed large deviations from experimental data
ing rates; in fact the contribution of the first peak became and therefore only the results of the Jeziorny analysis is
much smaller at higher heating rates. However, as shown reported.
in Fig. 6b for PET-kao-KAc-PEO at heating rates smaller The following equation was used to calculate the
than 20 C min21 the first peak was dominating but at relative crystallinity as a function of temperature:
higher heating rates (>20 C min21) the second peak
ðT
showed higher intensities. This behavior could be due to
the change in the crystalline structure at higher heating ðdH=dTÞdT
T
rates. Another reason could be the existence of crystallites XðTÞ5 ð Tc;o
1
(2)
with different sizes. ðdH=dTÞdT
A number of models have been proposed to study the Tc;o
crystallization of polymer materials; the Ozawa [45] and
Jeziorny [46] methods are more frequently applied to where Tc,o and T1 denote the onset and end temperatures
study the non-isothermal crystallization process. In this of crystallization, respectively, and H is the heat flow.
work both methods were applied; however, the Ozawa This equation can be written as a function of time, X(t),

FIG. 7. Relative crystallinity versus temperature and time at different cooling rates: neat PET (a, b), PET-
(kao-KAc-PEO) (c, d).

DOI 10.1002/pc POLYMER COMPOSITES—2016 1449


TABLE 2. Result of the Avrami–Jeziorny analysis for nonisothermal results for all the composites are summarized in Table 2.
crystallization of neat PET and PET composites. The neat PET had a n value >2 regardless of the heating
rate. However, for composites n depended on the heating
u Tc,max t1/2
( C min21) ( C) (min) n Zc 3 102 rate; for lower heating rates n was in the range of 1.5-2
while it increased to values above 2 for higher heating
Neat PET 2.5 206.5 4.56 2.31 21.24 rates. n values in the range of 2-3 can be interpreted as
5 194.8 2.38 2.59 59.29 the three-dimensional growth of crystals, and the shift to
10 192.3 1.54 2.73 88.70
values below 2 implies that the mode of crystallization
20 182.6 0.92 2.76 99.25
40 169.1 0.62 2.49 102.04 was probably changed to two-dimensional growth with
PET–kao–KAc–PEO 2.5 216.4 3.69 1.84 32.99 heterogeneous nucleation. At a given cooling rate, the
5 207.2 2.11 1.73 71.75 PET-kao composite had a larger Zc compared to the neat
10 199.0 1.48 2.00 89.14 PET; larger values of Zc imply that the crystallization
20 189.6 0.79 2.10 101.06
was faster in the presence of kaolin particles.
40 177.7 0.51 2.39 103.14
PET–kao 10 197.7 1.17 1.89 93.58 The temperature of maximum crystallization rate,
PET–kao–KAc–PEO–ch 10 187.4 1.65 1.87 87.79 Tc,max, and crystallization half-time (the time taken for
the relative crystallinity of the sample to reach the value
of 50%), t1/2, are two main parameters that characterize
considering the relationship between time and temperature the rate of crystallization. The higher Tc,max and smaller
in nonisothermal crystallization: t1/2 indicate a higher crystallization rate. The t1/2 values
of composites were generally smaller than that of the neat
Tc;o 2T PET; the only exception was PET-kao-KAc-PEO-ch.
t5 (3) PET-Kao exhibited the smallest t1/2 for all cooling rates.
a
However, when the particles were treated with KAc and
where t, T, and a denote temperature, time, and cooling PEO, the values of t1/2 slightly increased. In spite of the
rate, respectively. fact that the growth rates were lower at higher tempera-
The Avrami equation [38] is widely used to describe tures, the high number of nuclei provided by the clay par-
the isothermal crystallization of polymers: ticles generated a large quantity of crystallites that grew
concurrently, which compensated the effect of lower
12XðtÞ5exp ð2Zt 3tn Þ (4) growth rates. This indicates that the filler particles act as
where X(t) is the relative crystallinity at time t, n is the a heterogeneous nucleation agent and induce the forma-
Avrami index, and Zt is the kinetic crystallization rate. tion of crystals, which subsequently increases the crystal-
Jeziorny [46] tried to modify this method for nonisother- lization rate of the sample.
mal crystallization at constant cooling rate by defining
the following parameter: Thermal Stability

ln ZðtÞ The TGA thermograms of the neat PET along with those
ln Zc 5 (5) of other composites are presented in Fig. 8. The onset tem-
u
perature of degradation (Tonset), measured as the tempera-
where u is the cooling rate, Zc is the nonisothermal crys- ture required for 5% weight loss, for the neat PET, PET-
tallization kinetic rate constant. kao, and PET-(kao-KAc-PEO) were 418, 407, and 397 C,
Figure 7 shows the plots of relative crystallinity versus respectively. In spite of the fact that Tonset of PET was
temperature and time for two typical samples and the slightly shifted to lower temperatures as kaolin particles

FIG. 8. Thermal decomposition temperature of neat PET and its composites: weight loss (a), derivative
weight loss (b).

1450 POLYMER COMPOSITES—2016 DOI 10.1002/pc


were added to the resin, it seems that the main degradation 3. Y. Imai, S. Nishimura, E. Abe, H. Tateyama, A. Abiko, A.
was caused by the chemical modifiers and not the kaolin Yamaguchi, T. Aoyama, and H. Taguchi, Chem. Mater., 14,
particles. The kaolin particle and chemical modifiers can 477 (2002).
exert influence not only on the Tonset, but also on the tem- 4. J.-H. Chang, S. J. Kim, Y. L. Joo, and S. Im, Polymer, 45,
perature at which the maximum mass loss occurs (Tmax), as 919 (2004).
shown in Fig. 8b. The inclusion of the untreated kaolin had 5. T. Wu and Y. Ke, Polym. Degrad. Stabil., 91, 2205 (2006).
a small effect on Tmax of the composites; however, the 6. D.W. Chae and B.C. Kim, Compos. Sci. Technol., 67, 1348
TGA data showed that for samples treated with KAc and (2007).
PEO the maximum mass loss occured at temperatures 7. T.M.B. Maria, L.F.D.M.G. Andre, M.D.C. Cesar, I.V. Jose,
lower than those of the neat PET. It seems that the decom- A.D.A. Marcos, and H.I.M. Lucia, J. Appl. Polym. Sci.,
position of kao-KAc is the main reason for the shift of Tmax 104, 1839 (2007).
to lower temperatures. The decomposition of kao-KAc 8. H.-U. Kim, Y. H. Bang, S.M. Choi, and K.H. Yoon, Com-
took place at 455 C and subsequently the decomposition pos. Sci. Technol., 68, 2739 (2008).
products gave rise to an accelerated degradation of the PET 9. G. Choudalakis and A.D. Gotsis, Eur. Polym. J., 45, 967
resin. It is also noteworthy that blending with PET (2009).
improved the thermal stability of PEO, as the Tmax of PEO 10. Y. Wang, J. Gao, Y. Ma, and U.S. Agarwal, Compos. B, 37,
is 262 C while no mass loss is observed for PET-(kao- 399 (2006).
KAc-PEO) at this temperature. 11. F. Bergaya, B.K. Theng, and G. Lagaly, Handbook of Clay
TGA data showed that the addition of chain extender Science, Elsevier Science, Amsterdam (2011).
to the system could slightly improve the thermal stability, 12. J.E. Gardolinski, L.P. Ramos, G.P. de Souza, and F.
as Tmax was shifted to 473 C for PET-(kao-KAc-PEO-ch) Wypych, J. Colloid Interface Sci., 221, 284 (2000).
which was 15 C higher than maximum decomposition 13. R.L. Frost, J. Kristof, G.N. Paroz, and J.T. Kloprogge, Phys.
temperature of PET-(kao-KAc-PEO), however, it was still Chem. Miner., 26, 257 (1999).
lower than that of the neat PET. 14. W.N. Martens, R.L. Frost, J. Kristof, and E. Horvath, J.
Phys. Chem. B, 106, 4162 (2002).
CONCLUSIONS 15. K. Tsunematsu and H. Tateyama, J. Am. Ceram. Soc., 82,
1589 (1999).
XRD and TEM indicated that PET/kaolin nanocompo-
16. X. Zhang and Z. Xu, Mater. Lett., 61, 1478 (2007).
site was achieved via melt blending. It was shown that
the treatment of kaolin with KAc and blending with PEO 17. Y. Komori, H. Enoto, R. Takenawa, S. Hayashi, Y.
Sugahara, and K. Kuroda, Langmuir, 16, 5506 (2000).
improved its dispersion in a PET matrix. However, a low
level of exfoliation was achieved. The size of nanoscale 18. J. Murakami, T. Itagaki, and K. Kuroda, Solid State Ionics,
172, 279 (2004).
particles ranged from 100 to 200 nm in length and 10-
50 nm in thickness in the final nanocomposite. Because 19. S. Letaief, T.A. Elbokl, and C. Detellier, J. Colloid Inter-
face Sci., 302, 254 (2006).
of the existence of hydroxyl groups in the structure of
kaolin particles, a remarkable degradation of PET was 20. S. Letaief and C. Detellier, J. Mater. Chem., 15, 4734
observed during the melt mixing step and this effect was (2005).
more pronounced for chemically treated particles. The 21. D.M. Ansari and G.J. Price, Polymer, 45, 3663 (2004).
kaolin particles, as nucleating agent, affected the crystalli- 22. M. Buggy, G. Bradley, and A. Sullivan, Compos. A., 36,
zation/melting behavior of PET and increased the crystal- 437 (2005).
lization rate of the PET. It was also concluded that 23. Q. Liu, Y. Zhang, and H. Xu, Appl. Clay Sci., 42, 232
chemical treatment of particles with KAc and PEO would (2008).
slightly decrease the thermal stability of nanocomposite. 24. R. Sukumar and A.R.R. Menon, J. Appl. Polym. Sci., 107,
3476 (2008).
ACKNOWLEDGMENTS 25. L.E. Yahaya, K.O. Adebowale, and A.R.R. Menon, Appl.
Clay Sci., 46, 283 (2009).
The authors gratefully acknowledge PepsiCo, Inc., for 26. T. Itagaki, Y. Komori, Y. Sugahara, and K. Kuroda, J.
the support of this work. Dr. Tamal Ghosh and Dr. Kamal Mater. Chem., 11, 3291 (2001).
Mahajan are employees of PepsiCo, Inc. The views 27. Y. Komori, Y. Sugahara, and K. Kuroda, Chem. Mater., 11,
expressed in this publication are those of the authors and do 3 (1998).
not necessarily reflect the position or policy of PepsiCo, Inc. 28. X. Liu, H. Zhang, Z. Yang, and C. Ha, Chin. Sci. Bull., 50,
1320 (2005).
REFERENCES
29. Y. Turhan, M. Dogan, and M. Alkan, Indus. Eng. Chem.
Res., 49, 1503 (2010).
1. K. Friedrich, S. Fakirov and Z. Zhang, Polymer Composites 30. T.A. Elbokl and C. Detellier, J. Phys. Chem. Solids, 67, 950
From Nano- to Macro-Scale, Springer, New York (2005). (2006).
2. B.D. Beake and G.J. Leggett, Polymer, 43, 319 (2002). 31. J.J. Tunney and C. Detellier, Chem. Mater., 8, 927 (1996).

DOI 10.1002/pc POLYMER COMPOSITES—2016 1451


32. T. Itagaki, A. Matsumura, M. Kato, A. Usuki, and K. 39. F. Xin, L. Li, S.H. Chan, and J. Zhao, J. Compos. Mater.,
Kuroda, J. Mater. Sci. Lett., 20, 1483 (2001). 46, 1091 (2012).
33. T.S. George, A. Krishnan, N. Joseph, R. Anjana, and K.E. 40. C. Ou, M. Ho, and J. Lin, J. Polym. Res., 10, 127
George, Polym. Compos., 33, 1465 (2012). (2003).
34. A. Sanchez-Solis, A. Garcia-Rejon, and O. Manero, Macro- 41. W. Xu, M. Ge, and P. He, J. Polym. Sci. Part B: Polym.
mol. Symp., 192, 281 (2003). Phys., 40, 408 (2002).
35. X. Xu, Y. Ding, Z. Qian, F. Wang, B. Wen, H. Zhou, S. 42. J.Y. Kim, H.S. Park, and S.H. Kim, Polymer, 47, 1379
Zhang, and M. Yang, Polym. Degrad. Stabil., 94, 113 (2009). (2006).
36. H. Qin, S. Zhang, H. Liu, S. Xie, M. Yang, and D. Shen, 43. B. Lee, T.J. Shin, S.W. Lee, J. Yoon, J. Kim, and M. Ree,
Polymer, 46, 3149 (2005). Macromolecules, 37, 4174 (2004).
37. A. Ghanbari, M.C. Heuzey, P.J. Carreau, and M.T. Ton- 44. G.Z. Papageorgiou and G.P. Karayannidis, Polymer, 42,
That, Polymer, 54, 1361 (2013). 2637 (2001).
38. J. Brandrup, E.H. Immergut, E.A. Grulke, A. Abe, and D.R. 45. T. Ozawa, Polymer, 12, 150 (1971).
Bloch, Polymer Handbook, Wiley, New York (1999). 46. A. Jeziorny, Polymer, 19, 1142 (1978).

1452 POLYMER COMPOSITES—2016 DOI 10.1002/pc

You might also like