Sie sind auf Seite 1von 13

Journal of Polymers and the Environment (2019) 27:37–49

https://doi.org/10.1007/s10924-018-1320-6

ORIGINAL PAPER

Recycled Poly(Ethylene Terephthalate)/Clay Nanocomposites:


Rheology, Thermal and Mechanical Properties
Ravindra Reddy Chowreddy1   · Katrin Nord‑Varhaug1 · Florian Rapp2

Published online: 19 October 2018


© Springer Science+Business Media, LLC, part of Springer Nature 2018

Abstract
Polymer nanocomposites containing recycled poly(ethylene terephthalate) (r-PET) as a polymer matrix and ­Cloisite® 10A
as a reinforcement were prepared through melt compounding. First, a masterbatch containing 20 wt% of C ­ loisite® 10A clay
was prepared and later diluted with neat r-PET to obtain nanocomposites containing 1, 2, 4 and 6 wt% of clay. The rheologi-
cal, thermal, mechanical and morphological properties of the PET–clay nanocomposites were characterized. The complex
viscosity of the nanocomposites gradually increases with increase in clay content. The storage modulus and loss modulus of
nanocomposite containing 1 wt% of clay was similar to neat r-PET but increases with increase in clay content. Incorporation
of clay into a r-PET slightly increases the crystallisation temperature and degree of crystallinity due to the heterogenous
nucleating effect of clay. Thermal stability and glass transition temperature of nanocomposite containing ­Cloisite® 10A
clay at lower loadings was similar to r-PET and gradually decreases with increase in clay content. The tensile properties of
PET–clay nanocomposites increased gradually with clay content. The impact strength of nanocomposites was not altered
until 4 wt% and decreased at 6 wt%. Morphological investigations indicated homogeneous dispersion of clay in r-PET.

Keywords  Polyethylene terephthalate · Recycled · Cloisite® 10A · Clay · Polymer nanocomposites · Rheology · Tensile
properties

Introduction The increased consumption of PET, especially in the


packaging sector is creating serious environmental problems
Poly(ethylene terephthalate), (PET), is an important poly- as the vast amount of plastic waste is generated every year
ester which may exist both as an amorphous and semi-crys- due to the short life of packaging. PET is a non-biodegrada-
talline polymer. PET possess excellent properties such as ble plastic material which has high resistance to weathering
tensile and impact strength, chemical resistance, process- and biological agents under normal conditions. Disposal of
ability, transparency, and appropriate thermal stability [1] the waste to the landfill is becoming undesirable due to leg-
and it is widely utilized in industrial sectors such as food islation pressures, rising costs and the poor biodegradability
and beverages packaging, textiles, photographic films, X-ray of polymers [2]. Therefore, other ways of PET waste treat-
films, audio video tapes, etc. ment, such as recycling is highly desired.
Recycling processes are the best way to economically
reduce PET waste [3, 4] and PET is one of the most exten-
* Ravindra Reddy Chowreddy sively recycled polymeric materials [5]. There are three
ravindra.chowreddy@norner.no distinct approaches to the recycling of post-consumer plas-
Katrin Nord‑Varhaug tic packaging materials. Primary recycling means that the
katrin.nord‑varhaug@norner.no recyclable material/product is recovered and reused without
Florian Rapp being changed in any way and usually for the very same pur-
florian.rapp@ict.fraunhofer.de pose. Secondary recycling, also defined as physical, mechan-
ical or material recycling, means that the material/product
1
Norner Research AS, Asdalstrand 291, 3962 Stathelle, is reused in some other way without reprocessing, while
Norway
tertiary recycling refers to a process that involves chemical
2
Fraunhofer Institute for Chemical Technology ICT, altering of the material/product in order to make it reusable.
Joseph‑von‑Fraunhofer‑Straße 7, 76327 Pfinztal, Germany

13
Vol.:(0123456789)

38 Journal of Polymers and the Environment (2019) 27:37–49

Physical recycling involves several treatments or opera- Materials and Methods


tions such as waste collection, grinding, melting, and reform-
ing plastic material. Prior to melting and reprocessing, the Materials
ground, flaked, or pelletized polymer is generally washed to
remove contaminants. Recycled PET is more sensitive to ther- Recycled poly(ethylene terephthalate) (rPET), TEXPET R
mal and hydrolysis degradation. Recycling of PET gives rise 760 was purchased from Texplast GmbH, Germany. The
to a decrease in the melt viscosity, average molecular weight, TEXPET R 760 was received in the form of pellets and
thermal and mechanical properties of the material due to the was a regranulate from the post-consumer PET bottles
hydrolytic chain scission and thermomechanical degradation used in the beverage industry. The intrinsic viscosity (IV)
during material processing [6]. To produce a plastic with the of the material was 0.80 ± 0.02 dL/g.
desired performance characteristics, small amounts of virgin The clay utilized in the study was C­ loisite® 10A, an
PET resin and/or additives such as chain extenders, antioxi- organically modified bentonite and was received from
dants, processing aids or fillers are added to the recycled resin. BYK Additives, Germany. The organic modifier present
Currently physical recycled PET is being used to make food in ­Cloisite® 10A is dimethyl benzyl hydrogenated tallow, a
and non-food containers, straps, sheeting, and textile fibres. quaternary ammonium species. The quaternary ammonium
Incorporation of small quantities of nano-fillers into pol- modifier is incorporated into bentonite by cation exchange
ymeric matrices is known to improve the mechanical, elec- process. These quaternary ammoniums modified benton-
trical, barrier and thermal properties [7–10]. The improved ites are known to be suitable as fillers in thermoplastic
properties in polymeric matrices with nano-fillers is due to applications.
the efficient reinforcing of nano-filler in polymer matrix. Vari-
ous nano-fillers, like clays [11–14], carbon nanotubes, CNTs
[15–19], graphene [20–22], silica [23–25] have been explored Methods
in reinforcement of PET. Incorporation of nano-fillers in recy-
cled PET could increase the viscosity, thermal stability and Production of PET–Cloisite Nanocomposites
mechanical properties. Incorporation of inexpensive clay into
recycled PET is an efficient route to improve the processability First, a PET–clay masterbatch containing 20 wt% of
and performance. The literature available in this area is how- ­Cloisite® 10A was made by melt compounding. Required
ever limited. In a study, C­ loisite® Na+ and organo-modified amounts of milled recycled PET and ­Cloisite® 10A were
® mixed well using a kitchen blender and dried under vac-
clay, ­Cloisite 25A have been investigated as fillers in recycled
PET [26]. The authors concluded that C ­ loisite® 25A is a bet- uum at 80 °C for 2 days. The mixture was melt extruded
­ loisite® Na+. In another study, the
ter reinforcing filler than C using twin screw Prism 24 extruder with L/D is equal to
authors have shown that the silane modification of C ­ loisite® 30. The extrusion was performed at 270 °C and a screw
clays makes them efficient reinforcing fillers for recycled PET speed of 300 rpm. After compounding, the polymer string
[27]. was cooled under water bath and pelletized.
In the present study, we are reporting the preparation and The PET nanocomposites containing 1.0, 2.0, 4.0 and
characterization of recycled PET nanocomposites containing 6.0 wt% of ­Cloisite® 10A were produced by an extrusion
­Cloisite® 10A as a reinforcement. The ­Cloisite® 10A clay is an dilution technique. Like the masterbatch production, first
organically modified bentonite, treated with dimethyl benzyl required amounts of masterbatch and milled recycled PET
hydrogenated tallow, a quaternary ammonium species by ion- (neat) were mixed well using a kitchen blender and dried
exchange. The C ­ loisite® 10A clay was chosen in the present overnight at 120 °C. Later, the dried mixture was melt
study due to the better compatibility with aromatic polyester. extruded using a twin-screw Prism 16 extruder with L/D
The nanocomposites containing 1.0–6.0 wt% C ­ loisite® 10A is equal to 25. The extrusion was performed at 260 °C
clay were produced by conventional melt compounding by and a screw speed of 500 rpm. After compounding, the
using twin-screw extruder via masterbatch dilution. The rheo- polymer string was cooled under water bath and pelletized.
logical, thermal and mechanical properties, and the morphol- Composition of the different PET–clay nanocomposites
ogy of nanocomposites have been investigated. are presented in Table 1.

13
Journal of Polymers and the Environment (2019) 27:37–49 39

Table 1  Composition of PET– Sl. no. Sample designation Cloisite content in the Content of recy- Content of PET–
cloisite nanocomposites nanocomposite (wt%) cled PET (wt%) cloisite masterbatch
(wt%)

1 Ref. PET 0.0 100 0


2 PET–clay (1.0) 1.0 95 5
3 PET–clay (2.0) 2.0 90 10
4 PET–clay (4.0) 4.0 80 20
5 PET–clay (6.0) 6.0 70 30

Characterization of Nanocomposites temperature and 750 °C with heating rate of 10°C/min in


both air and nitrogen atmospheres.
Rheological Measurements
Dynamic Mechanical Analysis (DMA)
Dynamic rheology measurements on the nanocompos-
ites were performed on Dynamic Analyser RDA II from Dynamic mechanical analysis (DMA) on nanocomposites
Rheometrics. The measurements were carried out with was performed on DMA Q800 from TA Instruments using
parallel plate flow geometry using 25 mm diameter plates single cantilever in bending mode. Analysis was performed
at 270 °C. Steady shear viscosity measurements were made on bar-like cut specimens from compression moulded plate.
in a shear rate range varying from 0.1 to 300 rad/s. For The specimen had length, width and thickness respectively
rheology measurement, circular specimens with 30 mm 40 mm, 5.0 mm and 2.0 mm. Experiments were performed
diameter and 1.5 mm thickness were made by compression between room temperature and 190 °C at the heating rate
moulding. Prior to making the specimens by compression of 3°C/min. The frequency of bending was fixed at 1 Hz.
moulding, the nanocomposite pellets were dried well to DMA analysis was performed to evaluate the dynamic
avoid any degradation during compression moulding. The mechanical properties of nanocomposites and glass transi-
storage modulus, loss modulus and complex viscosity was tion temperature.
measured.
Tensile Testing
Differential Scanning Calorimetry (DSC)
Tensile testing on the nanocomposites was performed on
Differential scanning calorimetry (DSC) analysis on nano- Zwick Z010 according to ISO 527. A multipurpose type
composites was performed using Netzsch 204-F1 instru- 1B injection moulded dog-bone specimens were utilized for
ment. About 10–15 mg of pellet sample was encapsulated testing. The measurement was carried out at a crosshead
into sealed aluminium pans and analysis. The analysis rate of 5 mm/min. Five parallel measurements were made
involved three cycles of heating and cooling (1) heating the at room temperature and average values are reported. Ten-
sample from room temperature to 280 °C at a heating rate sile properties such as tensile strength, modulus, elonga-
of 10°C/min and held for 5 min, (2) cooling from 280 to tion at yield and elongation break were measured from the
30 °C at a cooling rate of 10°C/min and held for 5 min and stress–strain curves.
(3) heating the sample from 30 to 280 °C at a heating rate of
10°C/min. The glass transition temperature, melt tempera- Charpy Impact
ture, crystallization temperature, crystallinity and enthalp-
ies of melting and crystallization of nanocomposites was Charpy impact strength on the nanocomposite samples was
determined from DSC analysis. measured on Resilvis 125 Cryo according to ISO 179-1.
A notched specimen was made from the injection moulded
Thermogravimetric Analysis (TGA) bars and utilized in testing. Five parallel measurements were
made at room temperature. The amount of energy absorbed
Thermogravimetric analysis (TGA) on nanocomposites sam- by the sample before it breaks was measured.
ples was carried out using Q500 from TA instruments. Ther-
mograms obtained from TGA analysis provides the infor- Gel Permeation Chromatography (GPC) Analysis
mation related to the thermal stability of material. About
10–15 mg of the sample was subjected for TGA analysis. Gel permeation chromatography (GPC) analysis on the ref-
TGA analysis was performed on the samples between room erence PET and nanocomposites provides the information

13

40 Journal of Polymers and the Environment (2019) 27:37–49

about the average molecular weights, ­Mn and ­Mw, and the presented in the Fig. 1a. The complex viscosity of refer-
molecular weight distribution M­ w/Mn. The GPC analysis was ence PET is independent of frequency, indicating a New-
performed on Viscotek TDA 301 from Malvern. GPC anal- tonian behaviour in the investigated frequency range. The
ysis was performed on the samples using 1,1,1,3,3,3-hex- nanocomposite sample containing 1 wt% of clay presented
afluoro-2-propanol (HFIP) as solvent. Chromatograms were Newtonian behaviour similar to reference PET. However, the
obtained by using a refractive index detector, equipped with nanocomposite samples containing 2 wt% of clay and above,
a PL HFIPgel guard plus 2 × PL HFIPgel 300 × 7.5 mm, exhibited a pronounced shear thinning response. The shear
9 µm columns at a flow rate of 0.8 mL/min and a nominal thinning behaviour for the nanocomposites increases with
temperature of 40 °C. Poly(methyl methacrylate) (PMMA) increase in clay concentration, and the shear thinning effect
was utilized to calibrate the GPC system and the results are was most pronounced at the low frequencies. The complex
expressed in “PMMA equivalent” molecular weights, suit- viscosity is increasing with higher loading of clay, and the
able for comparative purpose among the samples. increase is most pronounced for a clay content of 2 wt% and
above. The complex viscosity of the nanocomposite sam-
Light Microscopy (LM) ples decreased at higher frequency due to the shear thinning
effect. Similar observations were made in the literature [32,
In order to see how well the clay is dispersed in the nano- 33].
composites, light microscopy (LM) investigations was per- The variation of storage modulus (Gʹ) and the loss modu-
formed. A specimen with 20 µm cross sections was prepared lus (G″) of reference PET and nanocomposites as a func-
from injection moulded dog-bone samples by microtome tion of frequency respectively is presented in Fig. 1b, c. The
cutting by using Leica M2500 microtome cutter. A transmis- storage modulus for reference PET and nanocomposites
sion light/light field technique was utilized for microscopic increases with frequency and clay content. The magnitude
investigation. Microscopy analysis was carried-out using of storage modulus increase with frequency is higher for ref-
Zeiss Axiophot microscope. erence PET and PET–clay (1.0) compared to other nanocom-
posite samples. The dependence of storage modulus with
Scanning Electron Microscopy (SEM) frequency for both reference PET and PET–clay (1.0) was
similar indicating the samples possess similar viscoelastic
Scanning electron microscopic (SEM) investigation was properties. However, for nanocomposite samples containing
investigated using Philips XL-30 ESEM. For SEM analysis, 2 wt% of clay and above, less pronounced dependency of
a specimen was prepared by etching. First, a small piece of storage modulus with frequency is observed. Especially, the
injection moulded specimen was embedded into an epoxy nanocomposite samples containing 4 wt% and 6 wt% of clay
resin, and polished after curing. The polished specimen was presents nearly frequency-independent storage modulus. In
exposed to 4M aqueous potassium hydroxide (KOH) solu- addition, the storage modulus curves exhibit a plateau at
tion at around 80 °C to remove the surface PET resin by the lower frequencies (Fig. 1b) which is attributed to the
etching. The etched specimen was washed with water, dried delaminated structures clay in polymer matrix [33]. Such
and a thin layer of gold was sputter coated for SEM analysis. interactions indicate a percolation network structures in the
The SEM analysis was performed using secondary electron nanocomposite samples. Similar observations are reported
(SE) detector at an accelerating voltage of 30 kV. in the literature for other polymer composites [34–36].
The trend in the variation of loss modulus with frequency
for nanocomposites (Fig. 1c) is similar to the storage modu-
Results and Discussion lus (Fig. 1b). However, the nanocomposite samples with
higher clay content (4 and 6 wt%) presented plateau at lower
Rheological Characterization frequencies in loss modulus curves compared to storage
modulus curves. In addition, the increase in loss modulus
Rheological characterization of polymer nanocomposites with increase in clay content is less pronounced compared
is important to understand the effect of nano-fillers on the to storage modulus.
structure and the processing characteristics. In addition, The variation of storage modulus (Gʹ) and the loss
rheological characterization is extensively used to probe dis- modulus (G″) with frequency for nanocomposite sam-
persion states of organoclay in polymer matrices [28–30]. ples containing different amounts of clay are presented in
Rheological investigation of polymer nanocomposites pro- Fig. 2a–d. These plots can provide better insight into the
vides the information related to percolation threshold of the changes in the microstructure and hence the viscosity of
nano-filler in the resin [31]. the systems. It is clear from the Fig. 2a, G″ is higher than
Variations of complex viscosity of the reference PET and Gʹ in the frequency range investigated, indicating normal
PET–clay nanocomposites as a function of frequency are viscoelastic response. However, for the nanocomposite

13
Journal of Polymers and the Environment (2019) 27:37–49 41

Fig. 1  Variation of a complex
viscosity, b storage modulus
and (C) loss modulus of PET–
clay nanocomposites

sample containing 2 wt% clay, Gʹ becomes higher than TGA Analysis


G″ (cross over) at lower frequency. As the clay content
in PET matrix is increased, solid-like behaviour becomes In order to investigate the thermal stability of the
more pronounced as Gʹ dominates G″ in the entire fre- PET–cloisite composites, thermogravimetric analysis (TGA)
quency range investigated. The nanocomposite sample was performed. The TGA was performed under both nitro-
with 6 wt% of clay presented higher Gʹ compared to G″ gen and air atmospheres. The TGA of samples under nitro-
and almost no cross over is observed indicating that the gen and air atmospheres respectively represent pyrolytic and
sample has transformed from the liquid-like behaviour to thermo-oxidative conditions. The TGA weight loss curves
more elastic behaviour over whole frequency range. This and first derivative TGA curves for reference PET and the
further suggests a stronger network structure between the nanocomposite samples under air atmosphere are presented
delaminated clay particles in the PET matrix. in Fig. 3. The first derivative TGA curves present variations

13

42 Journal of Polymers and the Environment (2019) 27:37–49

Fig. 2  Variation of Gʹ and G″ with frequency for PET–clay nanocomposites. a PET–clay (1.0), b PET–clay (2.0), c PET–clay (4.0) and d PET–
clay (6.0)

in the thermal stability among the samples. It is clear from The TGA weight loss curves and first derivative TGA
Fig. 3 that the samples exhibit a two-step decomposition curves for reference PET and the nanocomposite samples
during TGA analysis under air, an oxidative environment. under nitrogen atmosphere (pyrolytic environment) are
According to literature, the first step of decomposition presented in Fig. 4. Unlike TGA analysis under air atmos-
for reference PET is due to overlapping of two decomposi- phere, in nitrogen atmosphere the reference PET and the
tion processes [37]. The first process is mainly due to the PET nanocomposites exhibited only one step decomposition.
degradation of polymer chain through end group initiated The main degradation process during TGA analysis under
mechanism and the second degradation process is due to nitrogen atmosphere is due to the combined degradation
the thermal degradation of the products formed during first processes—polymer chain degradation through end group
decomposition process. The second degradation step for ref- initiated mechanism and the thermal degradation of the
erence PET in the temperature range 520–640 °C is due to products formed during polymer chain degradation. Under
the decomposition of thermally stable species formed dur- nitrogen environment, the thermal decomposition of poly-
ing the first degradation step. The thermally stable species mer is leading to thermally stable cross-lined carbonaceous
formed during first degradation step could be cross-linked species which are not undergoing any further decomposition
carbonaceous structures. In case of PET nanocomposite due to the presence of inert nitrogen atmosphere and hence
samples, the first step of decomposition took place similar the large amounts of reside at the end of TGA analysis.
to reference PET, where the decomposition is due to both The results of TGA analysis such as temperature at 1%
PET resin and organic modifier present in the clay. The weight loss, temperature at 5% weight loss, onset of decom-
second decomposition step for PET nanocomposites was position, and the amount of residue at end of TGA analysis
broader and the broadness increased with increase in clay (at 750 °C) are summarized in the Table 2. It is clear from
content. The second decomposition step took place in the the results that the temperatures at 1% and 5% weight loss
temperature range 520–720 °C for PET–cloisite (6%). This as well as the onset of decomposition is lower for the nano-
observation could be due to the formation of the reasonably composites compared to reference PET. This observation
large amounts of cross-linked carbonaceous species which could be due to the degradation of alkyl ammonium modi-
undergo slow decomposition in a wide temperature range. fier which is present in Cloisite 10A. The alkyl ammonium

13
Journal of Polymers and the Environment (2019) 27:37–49 43

Fig. 3  a TGA weight loss curves and b first derivative TGA curves Fig. 4  a TGA weight loss curves and b first derivative TGA curves
for PET–clay nanocomposites during TGA analysis under air atmos- for PET–clay nanocomposites during TGA analysis under nitrogen
phere atmosphere

present in the Cloisite is known to undergo Hoffman degra- It is clear from the results in the Table 3 that incorpora-
dation around 200 °C [38, 39]. The amount of residue left tion of clay into the PET matrix reduces the ­Tg of the PET
over on TGA analysis (at 750 °C) under nitrogen atmosphere matrix. The reason for the reduction in ­Tg with incorporation
is higher than that of the TGA analysis under air atmosphere. of clay could be due to the higher polymer chain mobility
in presence of clay. In addition, the modifier present in the
clay could act as a plasticizer and increase polymer chain
DSC Analysis motion. Similar reduction in ­Tg is reported in the literature
for nanocomposites based on PET and organo-modified clay
The influence of clay on thermal properties and crystalliza- [42–44]. Addition of clay into PET has very little influence
tion behaviour of recycled PET was investigated by using on melting temperature, T ­ m2 of nanocomposite samples.
differential scanning calorimetry (DSC). Thermal properties However, incorporation of clay into PET slightly increases
of PET–clay nanocomposites, such as glass transition tem- the crystallization temperature, ­Tc of nanocomposites. With
perature ­(Tg) crystallization temperatures ­(Tc), melting tem- increase in clay loading, the T ­ c of nanocomposites gradu-
peratures ­(Tm), enthalpies of melting (ΔHm), and the degree ally increases. In addition, the degree of crystallinity of the
of crystallinity were obtained from the DSC analysis. The PET–clay nanocomposites slightly increase with increase in
results of the DSC analysis are presented in the Table 3. The clay content. This increase in crystallization temperature and
degree of crystallinity (Xc) was determined from the melting the degree of crystallinity in the nanocomposites could be
enthalpy values using Eq. (1) [40]. due to the nucleation effect of clay platelets in PET matrix.
𝛥Hm Presence of nano-reinforcements, such as clays and CNTs
Xc = × 100 (1) are known to exhibit nucleating effect in polymer matrices
(1 − φ)𝛥H om
[8, 45, 46].
where 𝛥Hm is second melting enthalpy of the samples (J/g), The DSC second heating curves and cooling curves of
𝛥H om is the enthalpy value of melting of a 100% crystalline PET nanocomposites containing different amounts of clay
form of PET, 117.6 J/g [41], and ϕ is the weight fraction are presented in Fig. 5a. A double melting peak appeared for
of filler. both reference PET and the nanocomposites. This behaviour

13

44 Journal of Polymers and the Environment (2019) 27:37–49

Table 2  The results of TGA Sample Temp. at 1% Temp. at 5% Onset of Wt. loss during Residual
analysis on reference PET and weight loss weight loss decomposition 2nd decomp. mass at
PET–clay nanocomposites (%) 750 °C (%)
Air N2 Air N2 Air N2 Air N2 Air N2

Ref. PET 369 391 407 412 429 429 14.3 NA 0.1 6.7
PET–clay (1.0) 372 378 408 405 430 425 15.2 NA 1.0 10.0
PET–clay (2.0) 370 368 401 400 426 424 15.8 NA 1.4 12.5
PET–clay (4.0) 364 363 399 398 424 424 17.1 NA 2.7 13.2
PET–clay (6.0) 352 352 400 399 425 425 17.3 NA 3.8 17.1

Table 3  Results of DSC Sample Tg (°C) Tm1 (°C) Tm2 (°C) Tc (°C) Enthalpy of Crystallinity (%)
analysis—Tg, ­Tm1, ­Tm2, ­Tc, melting (J/g)
enthalpy of melting and degree
of crystallinity Ref. PET 79.5 242 249 196 44.0 37.4
PET–clay (1.0) 79.5 242 249 197 43.5 37.0
PET–clay (2.0) 78.8 242 249 199 46.5 39.5
PET–clay (4.0) 78.5 242 250 200 47.0 40.0
PET–clay (6.0) 76.0 242 251 200 45.0 38.3

peaks, indicating that added clay has no influence on the


crystal morphology in the PET–clay nanocomposites. As
seen from the thermograms, the position of the first melt-
ing ­(Tm1) is unaltered and the second melting peak tem-
perature ­(Tm2) is slightly affected by presence of clay filler.
The low temperature first melting peak is observed at about
242 °C while the high temperature main peak position var-
ies between 249 and 251 °C depending on the clay content.
In addition, with increase in clay loading, the first melting
peak shoulder is less pronounced, being obvious for the
nanocomposite containing 6 wt% of clay. This is probably
because clay at higher loadings starts to influence crystal-
lization behaviour and morphology.
The DSC crystallization endotherms of PET and
PET–clay nanocomposites are presented in Fig. 5b. A slight
shift in the crystallization peak position towards the higher
temperature is observed for the nanocomposites containing
clay. This means that the nanocomposite samples crystallize
at higher temperature compared to the reference PET, as
also seen from the corresponding T ­ c values in Table 3. The
shift in crystallization peak towards higher temperature is
due to the heterogenous nucleating effect of dispersed clay
in PET matrix.

Fig. 5  DSC melting endotherms (a) and crystallization exotherms (b) GPC Analysis
for reference PET and PET–clay nanocomposites
In order to understand the influence of clay on the molecular
weight of PET during the processing of nanocomposites,
could be due to the presence of a dual lamella thickness dis- GPC analysis was performed on PET–clay nanocomposites.
tribution which is formed during crystallization of PET [47]. Thermal processing of PET in presence of clay could lead
The nanocomposite samples also presented double melting to thermal, hydrolytic or oxidative degradation resulting in

13
Journal of Polymers and the Environment (2019) 27:37–49 45

Table 4  The GPC analysis results of PET and PET–clay nanocom- decreases gradually with increase in clay content. This
posites observation indicates that melt processing of PET in pres-
Sample Mw Mn PDI ­(Mw/Mn) ence of clay leads to degradation of polymer, hence a reduc-
tion in molecular weight. The polydispersity index values
As received PET 34,350 7460 4.6
for the resin in nanocomposite samples is of similar order.
Ref. PET 31,950 8775 3.7
PET–clay (1.0) 31,150 8545 3.7
Dynamic Mechanical Analysis (DMA)
PET–clay (2.0) 31,150 8265 3.8
PET–clay (4.0) 29,000 8015 3.7
Dynamic mechanical analysis (DMA) was carried-out on
PET–clay (6.0) 27,750 7805 3.6
PET–clay nanocomposites to understand the influence of
clay on viscoelastic behaviour of the materials. The vari-
ation of the storage and loss moduli with temperature for
decrease of molecular weight. The properties of PET are PET–clay nanocomposites are presented in the Fig. 6a, b,
strongly linked to its molecular weight. For meaningful respectively. The storage modulus behaviour in Fig. 6a shows
comparison of properties among the nanocomposites, it is that the PET–clay nanocomposites possessed slightly higher
important to measure the molecular weight of PET matrix storage modulus compared to the reference PET below the
in different composite samples. The results of GPC analysis, Tg. A gradual increase in storage modulus is observed with
weight average molecular weight M ­ w, the number average increase in clay content in PET matrix. Among the nano-
molecular weight ­Mn and the molecular weight distribution composite samples, PET–clay (6.0) had the highest storage
­Mw/Mn for reference PET and PET–clay nanocomposites modulus (about 17%) compared to reference PET and other
are presented in Table 4. The reference PET and PET–clay nanocomposite samples. The improved storage modulus in
nanocomposites showed slightly lower molecular weights PET–clay nanocomposites is due to the reinforcement effect
compared to the as received PET material. The reference of clay in PET matrix. A sharp decrease in the moduli was
material and PET–clay nanocomposites have undergone observed corresponding to glass–rubber transition at around
similar thermal processing. Among the nanocomposite 80 °C. All nanocomposite samples had slightly higher stor-
samples, the molecular weight (both ­Mw and ­Mn) of PET age modulus at 30 °C and slightly lower storage modulus at
80 °C compared to reference PET.
Variation of loss modulus with temperature for PET–clay
nanocomposites is presented in Fig. 6b. Compared to PET,
the nanocomposites had the maxima of peak at lower tem-
peratures. The glass transition, Tg was estimated based on
the temperature maxima in the loss modulus curves. The
reference PET sample showed a Tg of 79.6 °C while the
nanocomposites had lower Tg values, the lowest Tg being for
the nanocomposite with highest amount of clay. A similar
trend in Tg is observed from DSC analysis.

Fig. 6  Variation of storage modulus (a) and loss modulus (b) with


temperature for PET–clay nanocomposites Fig. 7  Variation of impact strength among PET–clay nanocomposites

13

46 Journal of Polymers and the Environment (2019) 27:37–49

Mechanical Properties of Nanocomposites PET matrix. The decrease in tensile strength was within
the standard deviation. The higher standard deviation in
Impact Properties PET–clay (1.0) could be due to inhomogeneity of the sample
compared to others. The nanocomposite sample containing
The effect of clay incorporation on impact strength of 6.0 wt% of clay [PET–clay (6.0)], presented increased tensile
PET–clay nanocomposites was investigated. The variation modulus and tensile strength by about 28% and 10%, respec-
of impact strength among the nanocomposites containing tively. The increased tensile properties are due to the rein-
different amounts of clay is presented in Fig. 7. Incorpora- forcement effect of homogeneously dispersed delaminated
tion of clay into the PET resin at low loadings, up to 4 wt% clay platelets in the PET matrix. An even higher amount
did not alter the impact strength. However, increasing the of clay in the matrix could improve the tensile properties
clay loading beyond 4 wt%, for example, the nanocomposite even more. However, processing of PET matrix with high
containing 6 wt% of clay [PET–clay (6.0)] presented 28% loadings of clay may adversely affect the impact properties.
reduction in impact strength compared to the reference PET. The variation of elongation at yield and break in nano-
In addition, the standard deviation for the same is high. Such composites is presented in Fig. 8b. The elongation at yield
reduction in impact strength in clay reinforced polymeric for PET–clay nanocomposites decreases gradually with
materials is reported in the literature [7]. increase in clay loading, from 3.7 to 3.3%. The reference
PET and nanocomposite material containing 1 and 2 wt% of
Tensile Properties clay presented an elongation at break in the order of 220%.
With higher loadings (4 and 6 wt% of clay) the elongation
The influence of clay on tensile properties of PET–clay at break decreased significantly to 30 and 3% respectively.
nanocomposites was also investigated. The variations of the Such reduction in elongation at break is common in clay
tensile strength and tensile modulus of the PET–clay nano- reinforced polymeric materials and similar observations are
composites are presented in Fig. 8a. The tensile strength and reported in the literature [7, 27].
modulus increases gradually with increase in clay content
in nanocomposites. In case of PET–clay (1.0), the average Morphology Investigation
tensile strength was slightly lower compared to reference
Scanning Electron Microscopy

The morphology of the samples was investigated by using


SEM. The SEM analysis was performed on the etched sur-
faces of the injection moulded specimens. The SEM images
of the PET–clay (4.0) and PET–clay (6.0) nanocomposite
samples are respectively presented in Fig. 9a, b. Both the
nanocomposite samples presented no visible agglomerates
of clay platelets. However, the dispersed clay platelets in
PET matrix are not clearly visible in the microscopic images,
which could be caused by the coating of etched PET resin on
the clay platelets. An attempt was made to wash-off the PET
resin with alcohol during the etching process for the SEM
preparation, but this attempt was not successful. In addition,
an attempt to analyse the freeze fractured injection moulded
specimen surface by SEM did not provide useful information
regarding clay dispersion in the composite materials.

Light Microscopy

In order to see the micro-dispersion quality of clay in


PET–clay nanocomposites, light microscopy (LM) inves-
tigations was performed. Light microscopy images of
PET–clay (4.0) and PET–clay (6.0) nanocomposites are
respectively presented in the Fig. 10a, b. It is clear from the
Fig. 8  Tensile strength and tensile modulus (a) and elongation at microscopic images that the dispersion quality of clay par-
yield and elongation at break for of PET–CNT nanocomposites (b) ticles in PET matrix is reasonably good and no microscopic

13
Journal of Polymers and the Environment (2019) 27:37–49 47

Fig. 9  SEM images of a PET–clay (4.0) and b PET–clay (6.0)

Fig. 10  Light microscopy images of a PET–clay (4.0) and b PET–clay (6.0)

agglomerates were observed. Since the specimens for LM crystallisation temperature and degree of crystallinity com-
was made from injection moulded samples, material flow pared to reference PET. The increased crystallization tem-
lines were observed in the LM images. perature and degree of crystallinity of nanocomposites is
due to a heterogenous nucleating effect of clay. The thermal
stability and glass transition temperature of nanocomposite
Conclusion at lower clay loadings are similar to r-PET but gradually
decrease with increase in clay content. The tensile proper-
Recycled poly(ethylene terephthalate) nanocomposites con- ties of PET–clay nanocomposites increased gradually with
taining ­Cloisite® 10A as a reinforcement were prepared by clay content. The impact strength of nanocomposites was
conventional melt compounding method via masterbatch not altered until 4 wt% of clay and decreased at 6 wt% clay
dilution. The rheological, thermal, mechanical and morpho- loading. Morphological investigations did not provide the
logical properties were investigated in the nanocomposites. nano-scale dispersion quality of clay in r-PET matrix. The
Incorporation of clay into recycled PET greatly influence future study in this direction will focus on EDAX, TEM,
the rheological properties. The complex viscosity, storage XRD, WAXS, etc. characterization of the nanocomposites
and loss modulus of the nanocomposites gradually increases produced in this work.
with increase in clay content. In addition, rheological inves- The incorporation of clay into recycled PET increases
tigation indicates that complete percolation occurs at 6 the viscosity and mechanical properties and makes the
wt% of clay loading. The nanocomposites a slightly higher material suitable for extrusion and injection moulding

13

48 Journal of Polymers and the Environment (2019) 27:37–49

applications. The recycled PET–clay nanocomposites 19. Anand K, Agarwal U, Joseph R (2007) Carbon nanotubes-rein-
could find applications in value-added technical products. forced PET nanocomposite by melt-compounding. J Appl Polym
Sci 104(5):3090–3095
20. Feng R et al (2011) In situ synthesis of poly (ethylene terephtha-
Acknowledgements  The research leading to these results has received late)/graphene composites using a catalyst supported on graphite
funding from the European Union’s Seventh Framework Programme oxide. J Mater Chem 21(11):3931–3939
(FP7) under Grant agreement no. 309985. The authors would like to 21. Aoyama S et al (2014) Melt crystallization of poly (ethylene tere-
thank the Smithers Rapra, UK for their support with GPC analysis on phthalate): comparing addition of graphene vs. carbon nanotubes.
the samples. Polymer 55(8):2077–2085
22. Istrate OM et al (2014) Reinforcement in melt-processed polymer–
graphene composites at extremely low graphene loading level.
Carbon 78:243–249
23. He J-P et al (2006) In situ preparation of poly (ethylene tereph-
References thalate)–SiO2 nanocomposites. Eur Polym J 42(5):1128–1134
24. Lu H et al (2007) Hybrid poly (ethylene terephthalate)/silica nano-
1. Karayannidis GP et al (2006) Chemical recycling of PET by composites prepared by in-situ polymerization. Polym Compos
glycolysis: polymerization and characterization of the dimeth- 28(1):42–46
acrylated glycolysate. Macromol Mater Eng 291(11):1338–1347 25. Ji Q et al (2009) Characterization of poly (ethylene terephthalate)/
2. Achilias D et al (2007) Chemical recycling of plastic wastes SiO2 nanocomposites prepared by Sol–Gel method. Compos Part
made from polyethylene (LDPE and HDPE) and polypropylene A Appl Sci Manuf 40(6):878–882
(PP). J Hazard Mater 149(3):536–542 26. Pegoretti A et al (2004) Recycled poly (ethylene terephthalate)/
3. Farahat MS, Abdel-Azim A-AA, Abdel-Raowf ME (2000) layered silicate nanocomposites: morphology and tensile mechani-
Modified unsaturated polyester resins synthesized from poly cal properties. Polymer 45(8):2751–2759
(ethylene terephthalate) waste, 1. Synthesis and curing charac- 27. Kráčalík M et al (2007) Recycled PET nanocomposites improved
teristics. Macromol Mater Eng 283(1):1–6 by silanization of organoclays. J Appl Polym Sci 106(2):926–937
4. Throne JL (1987) Effect of recycle on properties and profits: 28. Krishnamoorti R, Vaia RA, Giannelis EP (1996) Structure and
algorithms. Adv Polym Technol 7(4):347–360 dynamics of polymer-layered silicate nanocomposites. Chem
5. Polk MB (2003) Depolymerization and recycling. Synth Meth- Mater 8(8):1728–1734
ods Step Growth Polym 2003:527–574 29. Krishnamoorti R, Giannelis EP (1997) Rheology of end-teth-
6. Awaja F, Pavel D (2005) Recycling of PET. Eur Polym J ered polymer layered silicate nanocomposites. Macromolecules
41(7):1453–1477 30(14):4097–4102
7. Okada A, Usuki A (2006) Twenty years of polymer–clay nano- 30. Ren J, Silva AS, Krishnamoorti R (2000) Linear viscoelasticity of
composites. Macromol Mater Eng 291(12):1449–1476 disordered polystyrene–polyisoprene block copolymer based lay-
8. Paul D, Robeson LM (2008) Polymer nanotechnology: nano- ered-silicate nanocomposites. Macromolecules 33(10):3739–3746
composites. Polymer 49(15):3187–3204 31. Hassanabadi HM, Wilhelm M, Rodrigue D (2014) A rheological
9. Moniruzzaman M, Winey KI (2006) Polymer nanocom- criterion to determine the percolation threshold in polymer nano-
posites containing carbon nanotubes. Macromolecules composites. Rheol Acta 53(10–11):869–882
39(16):5194–5205 32. Ghanbari A et al (2013) Morphological and rheological properties
10. Kim H, Abdala AA, Macosko CW (2010) Graphene/polymer of PET/clay nanocomposites. Rheol Acta 52(1):59–74
nanocomposites. Macromolecules 43(16):6515–6530 33. Kráčalík M et al (2007) Recycled PET–organoclay nanocompos-
11. Gökkurt T et al (2013) Investigation of thermal, rheological, ites with enhanced processing properties and thermal stability. J
and physical properties of amorphous poly (ethylene tereph- Appl Polym Sci 106(3):2092–2100
thalate)/organoclay nanocomposite films. J Appl Polym Sci 34. Wu D et al (2005) Study on rheological behaviour of poly (butyl-
129(5):2490–2501 ene terephthalate)/montmorillonite nanocomposites. Eur Polym J
12. Ou CF (2003) Nanocomposites of poly (trimethylene terephtha- 41(9):2199–2207
late) with organoclay. J Appl Polym Sci 89(12):3315–3322 35. Seo M-K, Park S-J (2004) Electrical resistivity and rheological
13. Costache MC et  al (2006) Preparation and characterization behaviors of carbon nanotubes-filled polypropylene composites.
of poly (ethylene terephthalate)/clay nanocomposites by melt Chem Phys Lett 395(1):44–48
blending using thermally stable surfactants. Poly Adv Technol 36. Du F et  al (2004) Nanotube networks in polymer nanocom-
17(9–10):764–771 posites: rheology and electrical conductivity. Macromolecules
14. Scaffaro R et al (2011) Effect of kind and content of organo-mod- 37(24):9048–9055
ified clay on properties of PET nanocomposites. J Appl Polym Sci 37. Wang X-S, Li X-G, Yan D (2000) Thermal decomposition kinet-
122(1):384–392 ics of poly (trimethylene terephthalate). Polym Degrad Stabil
15. Logakis E et al (2010) Low electrical percolation threshold in poly 69(3):361–372
(ethylene terephthalate)/multi-walled carbon nanotube nanocom- 38. Ray SS, Bousmina M, Okamoto K (2005) Structure and properties
posites. Eur Polym J 46(5):928–936 of nanocomposites based on poly (butylene succinate-co-adipate)
16. Kim JY, Park HS, Kim SH (2007) Multiwall-carbon-nanotube- and organically modified montmorillonite. Macromol Mater Eng
reinforced poly (ethylene terephthalate) nanocomposites by melt 290(8):759–768
compounding. J Appl Polym Sci 103(3):1450–1457 39. Xie W et al (2001) Thermal degradation chemistry of alkyl quater-
17. Yesil S, Bayram G (2011) Poly (ethylene terephthalate)/carbon nary ammonium montmorillonite. Chem Mater 13(9):2979–2990
nanotube composites prepared with chemically treated carbon 40. Durmus A et al (2009) Nonisothermal crystallization kinetics of
nanotubes. Polym Eng Sci 51(7):1286–1300 poly(ethylene terephthalate)/clay nanocomposites prepared by
18. Antoniadis G et al (2009) Melt-crystallization mechanism of melt processing. Polymers 31:1056–1066
poly (ethylene terephthalate)/multi-walled carbon nanotubes pre- 41. Imai Y, Inukai Y, Tateyama H (2003) Properties of poly (ethylene
pared by in situ polymerization. J Polym Sci Part B Polym Phys terephthalate)/layered silicate nanocomposites prepared by two-
47(15):1452–1466 step polymerization procedure. Polym J Tokyo 35(3):230–235

13
Journal of Polymers and the Environment (2019) 27:37–49 49

42. Greco A et al (2010) Analysis of the structure and mass transport 45. Ou CF, Ho MT, Lin JR (2003) The nucleating effect of montmoril-
properties of clay nanocomposites based on amorphous PET. J lonite on crystallization of PET/montmorillonite nanocomposite.
Appl Polym Sci 118(6):3666–3672 J Polym Res 10(2):127–132
43. Greco A, Gennaro R, Rizzo M (2012) Glass transition and coop- 46. Manchado ML et al (2005) Thermal and mechanical properties of
erative rearranging regions in amorphous thermoplastic nanocom- single-walled carbon nanotubes–polypropylene composites pre-
posites. Polym Int 61(8):1326–1333 pared by melt processing. Carbon 43(7):1499–1505
44. Corcione CE, Frigione M (2012) Characterization of nanocom- 47. Kong Y, Hay J (2003) Multiple melting behaviour of poly (ethyl-
posites by thermal analysis. Materials 5(12):2960–2980 ene terephthalate). Polymer 44(3):623–633

13

Das könnte Ihnen auch gefallen