ArticlePDF Available

Rheological/thermal properties of poly(ethylene terephthalate) modified by chain extenders of pyromellitic dianhydride and pentaerythritol

Authors:

Abstract and Figures

Due to low molecular weight and wide molecular weight distribution, polyethylene terephthalate (PET) shows weak melt strength properties. In this study, the synergistic effect of using different types of chain extenders and catalyst on rheological behavior of PET has been investigated. Long‐chain branching is known as a suitable method for developing the structure of PET during reactive melt processing. Thus, pyromellitic dianhydride (PMDA) and pentaerythritol (PENTA) were added to the fiber grade PET. The best formulation was determined based on rheological results, which revealed an improvement in both storage modulus and complex viscosity of PMDA‐modified samples. Samples containing 1.5% PMDA and 0.5% PENTA exhibited the best rheological properties. Also, dibutyltin dilaurate (DBTDL) acted as an accelerator for chain extension reaction during reactive melt blending. Subsequently, the rheological properties were improved by increasing the chain extending rate. Moreover, thermal properties such as crystallization and melting temperatures and the degree of crystallinity for modified PET were investigated by differential scanning calorimetry.
This content is subject to copyright. Terms and conditions apply.
ARTICLE
Rheological/thermal properties of poly(ethylene
terephthalate) modified by chain extenders of pyromellitic
dianhydride and pentaerythritol
Saeed Dolatshah | Shervin Ahmadi | Amir Ershad-Langroudi | Hedieh Jashni
Iran Polymer and Petrochemical Institute,
Tehran, Iran
Correspondence
*Shervin Ahmadi, Iran Polymer and
Petrochemical Institute, P.O. Box
14965-115, Tehran, Iran.
Email: sh.ahmadi@ippi.ac.ir
Funding information
Iran Polymer and Petrochemical Institute
Abstract
Due to low molecular weight and wide molecular weight distribution, polyeth-
ylene terephthalate (PET) shows weak melt strength properties. In this study,
the synergistic effect of using different types of chain extenders and catalyst on
rheological behavior of PET has been investigated. Long-chain branching is
known as a suitable method for developing the structure of PET during reac-
tive melt processing. Thus, pyromellitic dianhydride (PMDA) and pen-
taerythritol (PENTA) were added to the fiber grade PET. The best formulation
was determined based on rheological results, which revealed an improvement
in both storage modulus and complex viscosity of PMDA-modified samples.
Samples containing 1.5% PMDA and 0.5% PENTA exhibited the best rheologi-
cal properties. Also, dibutyltin dilaurate (DBTDL) acted as an accelerator for
chain extension reaction during reactive melt blending. Subsequently, the rhe-
ological properties were improved by increasing the chain extending rate.
Moreover, thermal properties such as crystallization and melting temperatures
and the degree of crystallinity for modified PET were investigated by differen-
tial scanning calorimetry.
KEYWORDS
applications, crystallization, differential scanning calorimetry (DSC), glass transition, rheology
1|INTRODUCTION
Polyethylene terephthalate (PET) is an engineering plas-
tic, which is a semi-crystalline thermoplastic polymer
14
with high mechanical properties, appropriate thermal
and chemical resistance, and ease of processing. Its
proper price has caused it to be used in a variety of appli-
cations, such as food and beverage packages, films, fibers,
tapes and batteries.
58
Along with the advantages of PET,
this polymer has also several handicaps. In fact, the lin-
ear structure, low molecular weight, narrow molecular
weight range, and low crystallization rate, have caused
PET to be restricted in processes such as blow molding,
film-blowing, foaming and molding.
7,911
On the other
hand, condensation polymers such as PET undergo
hydrolysis and thermal degradation during the process
(especially recycling). The occurrence of this reaction
reduces the chain's length and generates smaller chains.
In the reaction of thermal degradation, the formed small
chains have end-groups of vinyl ester and carboxyl,
whereas, in the case of hydrolysis, the formed chains pos-
sess hydroxyl ester and carboxylic end groups. It is obvi-
ous that the chains' breakage results in a molecular
weight drop and consequently reduces the viscosity and
applications.
1214
Increasing the molecular weight and
molecular weight range by creating long branches is a
Received: 28 March 2020 Revised: 24 August 2020 Accepted: 30 August 2020
DOI: 10.1002/app.49917
J Appl Polym Sci. 2020;e49917. wileyonlinelibrary.com/journal/app © 2020 Wiley Periodicals LLC 1of15
https://doi.org/10.1002/app.49917
suitable method for increasing both melt strength and
viscosity. There are three main approaches to achieve this
aim: reactor polycondensation, solid-state polycondensa-
tion (SSP), and reactive extrusion (REX). In the first
method, for the creation of a low-branched structure,
multifunctional monomers such as glycerol and pen-
taerythritol (PENTA) are directly added to the PET dur-
ing synthesis. Comparing the two methods of REX and
SSP, the SSP method has the advantage of occurring at a
low temperature, so it will not be degraded. However,
this method has two major problems; One is the long
reaction time due to the control of the reaction by the
penetration phenomenon, and the other problem is the
requirement of special materials and equipment,
while the REX method has higher performance and
speed.
9,10,13
In the REX method, a multifunctional mono-
mer is utilized as a chain extender which is able to react
with the end groups of PET, creating a branched struc-
ture and increasing the length of the chain.
4,15
Several
types of chain extenders used for PET include
epoxides,
1624
anhydrides,
13,2529
isocyanates,
18,3032
bis-
oxazoline,
30,3335
and triphenyl phosphite.
36
PMDA is a
common chain extender agent for PET. Draver and
coworkers carried out the process of branching and chain
extension of PET by PMDA in a Twin-Screw Reactive
Extrusion. Modified PET exhibited a definite improve-
ment in rheological properties. They proved that by
increasing the amount of PMDA, the storage modulus
and melt viscosity have increased, indicating an incre-
ment in the molecular weight.
13
Xia and coworkers also
used PMDA to improve the properties of foamable PET
and stated that by increasing the amount of PMDA, the
molecular weight, molecular weight range, melt elasticity
and melt strength of PET were improved. Also, the
apparent viscosity of the modified PET sample was
enhanced. They claimed that the best PMDA amount to
reach the highest viscosity value was 0.8%.
25
In addition,
some studies have focused on the point that a limited
number of chain extenders have the ability to be used
with PMDA together.
3740
Forsythe and coworkers used
different proportions of PMDA and PENTA to modify
bottle grade PET. They showed that if these two agents
were employed together, melt flow index (MFI) reduction
and melt viscosity and apparent viscosity improvement
were significantly higher than when just PMDA was
used. Based on the results of the Post Extrusion Time
Sweep, they claimed that the branching was raised with
increasing PMDA and PENTA amounts. Eventually, they
pointed that the modified PET structure is
hyperbranched with very long branches.
41
In another
study, Hanley and coworkers also achieved similar
results. They suggested that by increasing the chain
extender agent, the number average molecular weight
(M
n
) was approximately constant, however, the weight
average molecular weight (M
w
), molecular weight range,
and the apparent viscosity were significantly increased
and MFI was decreased. In the case of samples modified
with PMDA and PENTA synergy, the improvement in
properties was more definite rather than the PMDA-
modified sample. They further revealed that by increas-
ing branching and increasing molecular weight, the crys-
tallization rate is enhanced.
42
This point could be a
resolver of one of the main weaknesses of the PET.
Because the tough structure of the PET can cause a limi-
tation in mobility, this can lead to prolongation of its
molding time.
7
Dibutyltin dilaurate (DBTDL) is an
organotin catalyst which is often soluble and applicable
in broad-temperature ranges. It is a widely used catalyst
in the synthesis of polyurethanes and silicon vulcaniza-
tion. It is also utilized in the industry as a polyvinyl chlo-
ride stabilizer. Recently, there have been many reports
based on the application of DBTDL as a catalyst for ester-
ification and transesterification reactions.
4345
Einloft
et al. used various compounds of tin and sulfuric acid as
catalysts in the esterification reaction between rice bran
oil and methanol, and eventually reported that DBDTL
had better efficiency than other catalysts and in lower
contents of DBDTL, high amounts of products were
accessible.
46
In this study, the reactive melt extrusion method was
utilized for chain extension of PET. In this regard, differ-
ent ratios of the PMDA and PENTA synergy were used
compared to PMDA alone. In addition, the effect of the
presence of DBTDL as a catalyst on the branching and
chain extension process was investigated. Given the reac-
tions of PMDA and PENTA with PET chains are of the
stratification and/or transesterification types, it is
expected that the use of DBTDL could lead to an increase
in the reaction rate of the chain extension, which is an
important economic factor. Finally, thermal properties
such as the amount and rate of crystallinity were evalu-
ated by the differential scanning calorimetry (DSC) test.
2|EXPERIMENTAL
2.1 |Materials
Fiber grade PET (intrinsic viscosity of 0.64 dl/g) was pur-
chased from Tondgooyan petrochemical Co, Iran. It has a
glass transition at 82C and a melting point at 258C.
PMDA as chain extender and PENTA were used to pro-
mote the branching reaction in PET. They were provided
by Merck Co. According to the supplier datasheet, the
purity and melting point of PMDA and PENTA are 97%,
280C and 99%, 255C, respectively. Also, DBTDL (95%)
2of15 DOLATSHAH ET AL.
was supplied by Merck Co, and used to improve the rate
of chain extension reaction in PET. In order to avoid
hydrolysis of PET during mixing, the PET granules were
dried for 24 h at 90C in a vacuum oven prior to mixing.
Drying was repeated for PMDA and PENTA for 12 hr
at 80C.
2.2 |Method
A Brabender type internal mixer was used for the reactive
melt blending process. In the first step, PET granules
were added to the mixer and melted for 5 min at a rota-
tion speed of 60 rpm and temperature of 270C. Then,
various amounts of PMDA, PENTA, and DBTDL were
loaded to the chamber according to formulations shown
in Table 1 for 10 min.
2.3 |Rheological measurement
The storage modulus and complex viscosity of pure PET
and the extending samples were measured using an oscil-
latory rheometer (physica Anton paar, MCR-30, Ger-
many) with a parallel plate 25 mm in diameter and a gap
of 1 mm under nitrogen atmosphere. The test was carried
out in the shear rate range of 0.1100 S
1
at 280Cata
fixed strain of 10% for all the samples, to ensure that the
measurement was performed within the linear viscoelas-
tic region.
2.4 |Melt flow index (MFI)
The melt flow index of the samples was examined by
GottfertMI-4 type at 285C with an overhead weight of
2.16 kg according to ASTM D1238-13. In order to ensure
the accuracy of the results, measurements were repeated
three times for any samples. Prior to the MFI test, the
samples were dried at 120C for 24 h in a vacuum oven.
2.5 |Gel content
The crosslinking reaction was assessed by measurement
of the gel content. At first, the samples were solved in a
solution of 50/50 wt% phenol/tetrachloroethane for 24 h,
followed by filtering the unsolved parts via drying in a
vacuum oven for 24 h at 120C.
47,48
2.6 |Thermal analysis
Thermal behavior of the samples was analyzed by a
Netzsch 200F3 type DSC instrument in Al crucible. The
DSC equipment was purged with a liquid nitrogen
cooling system. To measure the glass transition, crystalli-
zation and melting temperatures and the degree of crys-
tallinity for pure PET and modified PET, 5 mg of the
samples were initially heated to 300C at a rate of
20C/min to remove the heating history, followed by
cooling to room temperature at a rate of 20C/min. Then,
the second heating run was carried out by the previous
rate until 300C. All steps were carried out under nitro-
gen atmosphere. Prior to testing, the samples were dried
in a 120C vacuum oven for 24 min.
3|RESULTS AND DISCUSSIONS
3.1 |Complex viscosity
The complex viscosity curve in terms of frequency for the
nonmodified sample is shown in Figure 1. By increasing
TABLE 1 Formulations of the
samples Sample code PMDA (wt%) PENTA (wt%) DBTDL (wt%)
PET –– –
PM0.5 0.5 ––
PM1.0 1.0 ––
PM1.5 1.5 ––
PM0.5/PE0.5 0.5 0.5
PM1.0/PE0.5 1.0 0.5
PM1.5/PE0.5 1.5 0.5
PM1.5/PE0.5/D0.1 1.5 0.5 0.1
PM1.5/PE0.5/D0.4 1.5 0.5 0.4
Abbreviations: DBTDL, dibutyltin dilaurate; PET, polyethylene terephthalate; PENTA, pen-
taerythritol; PMDA, pyromellitic dianhydride.
DOLATSHAH ET AL.3of15
the frequency, no obvious changes were observed in
viscosity and a flat region was seen in the range of
100 rad/s. In other words, by raising the frequency in the
neat sample, the shear thinning behavior was not appre-
ciable. The viscosity dependence on shear for the
branched polymer is very different from that of linear
polymer and varies depending on the structure. In lower
shears, branched chains are able to show the complex vis-
cosity of about 100 times higher than linear polymers
with the same molecular weight. On the other hand, in
spite of the shear thinning behavior at high shear rates, a
less complex viscosity compared to the linear mode may
be seen.
49
PMDA-modified samples have displayed higher vis-
cosity in comparison with linear samples, which is due to
the presence of long branches. The viscosity increment in
low frequencies in PMDA-modified samples had a direct
relation with the increase in the content of the chain
extender agent. However, by increasing the frequency,
due to the shear thinning behavior, an intense drop was
observed in complex viscosity.
10
The reason for this phe-
nomenon can be the changes in the entanglement of
polymer chains and the enhancement of the relaxation
mechanism.
50
On the other hand, the Newtonian flat
area was removed or shortened by increasing the amount
of branching, and rheological behavior was directed
toward non-Newtonian.
51
In samples modified with both
PMDA and PENTA, the increment in complex viscosity
at lower frequencies was much more evident than treat-
ment with PMDA, indicating the presence of longer
branches as well as more entanglement. The manner of
chain extending and branching in these specimens seems
to be unlike the samples that have been modified individ-
ually with PMDA, which has led to an increase in entan-
glements. During PET processing, the chain breakage
results in the formation of carboxyl and hydroxyl groups.
The carboxylic groups in PMDA can only do esterifica-
tion reactions with hydroxyl groups of the chain
(Figure 2(a)).
41
But the presence of the PENTA hydroxyl
group carrier provides the possibility of reaction with car-
boxylic terminated chains (Figure 2(b)).
41
In such cases,
there is the possibility of reaction for both terminal
groups of PET chain. Therefore, the feasibility of chain
length increment, development of long branches and
increase in molecular weight in these conditions are
much promising rather than the case of a single chain
extender (Figure 2(c)).
3941
For samples that were modi-
fied in the presence of the catalyst, the amount of com-
plex viscosity at low frequencies was increased. In fact,
the presence of catalyst seems to have a significant role
in increasing the trend of chain branching and extension.
Indeed, DBTDL acts as a catalyst for the reactions of
esterification and transesterification and increases the
speed and rate of branching.
44
3.2 |Storage modulus
The storage modulus curves versus frequency for modi-
fied and nonmodified PET are shown in Figure 3. In the
FIGURE 1 The complex viscosity of samples versus frequency at 280C
4of15 DOLATSHAH ET AL.
unmodified sample, the storage modulus has increased
with frequency, since at low frequencies there is enough
time for relaxation of the chains and the storage modulus
comprises fewer values, while at higher frequencies, the
chains do not have the opportunity for relaxation, and
hence, the storage modulus is increased. In the case of
PMDA-modified samples, the storage modulus was much
more than the linear one; also storage modulus variation
versus frequency was observed with lower slope, indicat-
ing an increase in relaxation time and elasticity due to
the enhancement in chain length and entangle-
ment.
11,26,40,52
Also, by increasing the number of entan-
glements, the chains are more capable of transferring
external forces; thereby the storage modulus is increased
by improving the melt elasticity.
53
For concurrent modifi-
cation of samples with PMDA and PENTA, the increase
in storage modulus at all frequencies was more pro-
nounced, indicating the presence of longer chains and
creation of longer branches.
As a matter of fact, the storage modulus in term of
frequency for modified samples was changed with a low
slope which is the resultant of increment in chain length
and entanglement.
41
For modified samples in the pres-
ence of catalyst, the storage modulus curve had the
broadest plateau region, also the storage modulus showed
the least frequency dependence in this sample. Specifi-
cally, the presence of the DBTDL catalyst increased the
modulus of the sample by increasing the esterification
FIGURE 2 Reaction of
PET chains with (a) PMDA,
(b) PENTA, and (c) PMDA and
PENTA. PET, polyethylene
terephthalate; PMDA,
pyromellitic dianhydride;
PENTA, pentaerythritol
DOLATSHAH ET AL.5of15
and transesterification reaction rates of branching and
chain development.
44,54
Indeed, the increment in the
branching trend and propagation in quantity of entangle-
ments generated a clear ascending trend in the amount
of storage modulus.
51
Specifications of the crossover point of the storage
and loss moduli can provide precious information about
M
W
, MWD, and melt elasticity. The crossover is the point
at which liquid-like behavior tends toward solid-like
behavior and dominant-elastic behavior, and at higher
frequencies the melt elasticity overcomes the loss behav-
ior. According to the Figure 4, the crossover point is not
observed in the frequency range (0.1100 rad/s) for the
PET sample and appears to occur at frequencies above
100 rad/s. In PMDA-modified samples (i.e., PM1.5), the
crossover frequencies were shifted to lower values, indi-
cating an increase in molecular weight and elasticity after
the reaction between PET and the chain extender. Fur-
thermore, the crossover modulus decreased following
PMDA-modification, which causes the MWD to expand.
For the samples simultaneously modified with PMDA
and PENTA, as well as samples modified in the presence
of the catalyst, the crossover point was transferred to the
frequencies below 0.1 rad/s, implying a significant
increase in elasticity and molecular weight, but making a
comment about the changes in the crossover modulus is
not certified.
2,55
3.3 |ColeCole
An effective method to assess the elastic behavior in vari-
ous structures of a polymer is Cole-Cole diagram which is
strongly dependent on the molecular weight distribution
and branching, while is independent of molecular weight.
In Cole-Cole diagram, the criterion of melt elasticity is the
status of the data relative to the G" = G' line.
24
In Figure 5, the Cole-Cole diagram is presented for
modified and unmodified samples. The results indicated
that the Cole-Cole diagram for modified and unmodified
sample were very disparate. In the constant values of the
loss modulus, the storage modulus increased when the
PMDA content was augmented. It means that by increas-
ing the chain extender contents, the melt elasticity,
molecular weight and molecular weight distribution were
raised. For samples modified with PMDA and PENTA,
the amount of curve transfer toward storage modulus
was more apparent. The augmentation of melt elasticity,
molecular weight, and molecular weight distribution
were validated by further shift for co-addition modified
samples. This diversity was due to the changes in struc-
ture and chain length.
52
The chains in former state pos-
sess the comb-like structure
39
and in co-addition state
have a hyperbranched structure.
41
Moreover, in samples modified in the presence of cat-
alyst, with increasing branching and chain extension, the
FIGURE 3 The storage modulus of samples versus frequency at 280C
6of15 DOLATSHAH ET AL.
FIGURE 5 Cole-Cole plots of samples at 280C
FIGURE 4 The storage and loss moduli of samples versus frequency at 280C
DOLATSHAH ET AL.7of15
enhancement of storage modulus in constant values of
loss modulus was very impressive.
24,56
3.4 |Tan δ
The phase difference in terms of frequency for modified
specimens and nonmodified sample is demonstrated in
Figure 6. For linear polymers, the phase discrepancy in
lower frequencies showed higher values, indicating the
viscous behavior. As the frequency was enhanced, Tan δ
was changed with a relatively sharp slope, while in modi-
fied sample containing chain extender agent, Tan δgraph
was transferred to much lower values, which showed an
increase in elastic behavior. This behavior reveals an
increment in both molecular weight and molecular
weight distribution due to lengthening of the chains and
creation of long branches. Tan δgraph for the samples
modified with PMDA and PENTA was appeared at a
lower level, indicating a higher melt viscosity along with
longer branches. In the samples modified in attendance
of the catalyst, the reduction of Tan δwas much higher,
and in a wider range of frequencies, Tan δshowed a flat
area, which is a proof on augmentation of long branches,
entanglements and the like-gel behavior.
28,50
. It should
be noticed in the samples containing PENTA, at high fre-
quencies Tan δshowed an insignificant increase. As
shown in Figure 2, PENTA may perform a series of
alcoholysis reactions with the ester group and form some
short chains. There appears to be some residual PENTA
in the system, which at high frequencies causes to gener-
ate a few short chains, thus slightly increasing the viscos-
ity and consequently tan δ.
3.5 |Van Gurp-Palmen plot
Another graph that is plotted based on the linear visco-
elastic data is the Van Gurp-Palmen diagram. This dia-
gram is utilized for two polymers of different chemical
nature or rheological identification of long side chains.
56
This diagram is severely dependent on molecular weight
and molecular weight distribution. In this graph, the
values of δare plotted in term of G*. As much the δvalue
is closer to 90, indicating the viscous behavior and the
linearity of the polymer.
55,57
As can be seen in Figure 7,
for the nonmodified sample, the content of δin the high
range of G* is close to 90, which is a linear representa-
tion of the sample. The value of δin the modified sample
is significantly decreased compared to the unmodified
sample, which could be due to the chain extension and
increase in molecular weight distribution. By increasing
the number of co-addition samples, the rate of δ-drop
was significantly noticeable. Consequently, the improve-
ment of branching reaction and chain development in
co-addition form are concluded of these plots. In samples
containing catalyst, the values of δwere reduced com-
pared to samples free of it. By increasing the catalyst con-
tent, the δdiminish was obvious and showed a much
elastic behavior in samples containing 0.1 or 0.4%
FIGURE 6 Tan δplots of samples versus frequency at 280C
8of15 DOLATSHAH ET AL.
catalyst.
15,58
It should be attended that increasing the
molecular weight distribution can increase the elastic
behavior and consequently decrease the amount of δ, but
cannot justify the concavity in the diagram. In fact, the
concavity appears to be a reason to prove the presence of
long chains.
50
3.6 |Zero shear viscosity
The zero shear viscosity has an intense dependence on
molecular weight and the presence of long branches, so
the study of zero shear viscosity can provide important
information on the quality of branching. The method of
data extrapolation was used to calculate the zero shear
viscosity. For this purpose, Cross and Carreau-Yasuda
models were employed. The results of these two models
are presented in the Table 2. It seems that the Carreau-
Yasuda model is more suitable than Cross model in pre-
diction of zero shear viscosity in the desired frequency
range. The results of zero shear viscosity of all samples
calculated according to Carreau-Yasuda model are given
in Table 3, which is clearly consistent with the complex
viscosity results. In fact, the zero shear viscosity incre-
ment in the modified samples confirmed an increase in
mw and MWD in all modified samples compared to the
PET sample. Concerning the samples modified with
PMDA and PENTA co-addition, as well the samples mod-
ified in the presence of a catalyst, the obvious spreading
of MWD can be expressed with confidence, while for
these samples, the crossover point has certainly not
shown MWD alteration.
2,39
3.7 |Molecular weight
One of the most important characteristics of PET is its
intrinsic viscosity, which is one of the specifications in
classification of this polymer. Sanchez and his colleagues
obtained the M
n
and M
w
from viscosity using the Mark-
Houwink equation (Equation (1)).
59
FIGURE 7 Van gurp-Palmen plots of samples at 280C
TABLE 2 The correlation ratio of cross and Carreau-Yasuda
models of samples
Sample code
Correlation ratio (R)
Carreau - Yasuda Cross
PET 0.98 0.92
PM0.5 0.92 0.91
PM1.0 0.99
PM1.5 0.98
PM0.5/PE0.5 0.86 0.74
PM1.0/PE0.5 0.88 0.83
PM1.5/PE0.5 0.91 0.90
PM1.5/PE0.5/D0.1 0.86 0.81
PM1.5/PE0.5/D0.4 0.87 0.86
Abbreviation: PET, polyethylene terephthalate.
DOLATSHAH ET AL.9of15
η0=KM
wα
:ð1Þ
where αparameter is dependent to the critical molar
mass and is between 3.4 to 3.6 for polymers, while the
value of K parameter depends on the temperature and
polymer type. In addition, Härth and colleagues believed
that in a nonlinear polymer with long chains compared
to a linear one with the same molecular weight, the
hydrodynamics radius is decreased. This issue made it
difficult to measure the intrinsic viscosity using
Ubbelohde viscometer. Moreover, they claimed that by
adding PMDA as the chain extender agent, the number
average molecular weight remains almost constant, while
the weight average molecular weight increases exponen-
tially. Formerly the graph of zero shear viscosity was
plotted as a function of molecular weight, therefore the
amounts of K and αin Mark-Houwink equation were
3.2 ×10
14
and 3.5, respectively. In our study, the above
mentioned equation was applied to calculate the weight
average molecular weight, and the values are given in the
Table 3. The results revealed that by increasing the chain
extender agent and amount of branching, the molecular
weight was enhanced,
15
while a gradual increase in M
n
occurred and as a result, a massive increase in molecular
weight distribution was happened.
58
3.8 |Melt flow index (MFI)
MFI is an important parameter in investigation of poly-
mer behavior, especially the viscosity of polymer. The
MFI values are dependent on the molecular weight and
broadness of molecular weight distribution.
41,60
In
Table 4, the MFI values are given for all samples. It was
observed that PMDA-modified specimens followed MFI
reductions at low PMDA contents, indicating increased
branching, development of long branches, and increased
molecular weight, resulting in penetrated
(entanglemented) networks and reduced chain mobility,
which leads to increase in melt strength. By increasing
the PMDA amount, MFI was decreased. In the PMDA-
PENTA-modified samples, a drastic decrease in MFI was
observed compared to those modified with PMDA alone,
indicating an increase in melt strength. In these samples,
by increasing the PENTA content, it is possible that the
chains also react with the carboxyl end group. Anyway,
PET chains have the opportunity to react from both sides;
from one head with carboxyl groups of PMDA and at the
other side with hydroxyl groups of PENTA (Figure 2).
Therefore, the chain extension and branching would be
more improved rather than utilizing the alone PMDA. In
fact, it seems that in this state, the resulted structure is
hyperberanched with low distance between branches.
The increment in chain length, the number, and length
of branches are the main reasons behind MFI reduction
in these samples. For the samples containing catalyst, the
MFI was slightly reduced compared to the catalyst-free
specimens, which is due to increase in branching.
41
3.9 |Gel content
The rheological behavior of modified specimens, espe-
cially in higher amount of chain extenders, is various
from the nonmodified sample. Storage modulus values
for some of the samples were independent of frequency,
and it almost presented elastic behavior in all frequen-
cies. On the Cole-Cole diagram, several samples crossed
the line G" = G', which indicates change in the chain
structure, such as the formation of a long side chain and
even formation of gel.
58
Also, in the tan δdiagram of the
TABLE 3 Zero shear viscosity and weight-average molecular
weight of samples
Sample code
Zero shear
viscosity
(pa.s)
Weight-average
molecular weight
(kg/mole)
PET 20 17
PM0.5 219 33
PM1.0 230 34
PM1.5 277 36
PM0.5/PE0.5 19,727 121
PM1.0/PE0.5 31,756 139
PM1.5/PE0.5 45,261 153
PM1.5/PE0.5/D0.1 48,345 156
PM1.5/PE0.5/D0.4 64,927 170
Abbreviation: PET, polyethylene terephthalate.
TABLE 4 Melt flow index of samples
Sample code MFI (g/10 min)
PET 30 ± 0.7
PM0.5 28 ± 1.1
PM1.0 28 ± 1.0
PM1.5 38 ± 1.4
PM0.5/PE0.5 18 ± 0.8
PM1.0/PE0.5 18 ± 0.7
PM1.5/PE0.5 19 ± 0.8
PM1.5/PE0.5/D0.1 20 ± 1.6
PM1.5/PE0.5/D0.4 26 ± 0.9
Abbreviation: PET, polyethylene terephthalate.
10 of 15 DOLATSHAH ET AL.
modified samples, the frequency dependence was
reduced, which can be because of network structure for-
mation.
61
The results of this test revealed that gel content
for all the samples was almost zero, which means even
with the highest percentage of chain extender agent, net-
work creation was not observed, indicating either that
the samples did not contain the gel or that the gel parti-
cles were infinitesimal.
15
3.10 |Differential scanning
Calorimetry (DSC)
Besides the influence of branching on rheological proper-
ties such as viscosity and shear thinning behavior, it also
affects thermal properties such as melting temperature
(T
m
), glass transition temperature (T
g
) and crystallization
temperature (T
c
).
55
In Table 5, the thermal properties
and T
c
of all samples are tabulated. The polymer glass
transition temperature has a definite dependence on the
free volume fraction; therefore, any factor that can affect
the free volume also changes the glass transition temper-
ature. Side chain is one of the factors, which affects the
temperature of the glass transition by changing the free
volume. The presence of side chains, in addition to
increasing free volume, by restricting chain movement,
may have a conflicted effect by increasing free volume.
Based on the length of created side chains and the bra-
nches, these two factors can reduce or increase the glass
transition temperature.
62
As the results represent, the
glass transition temperature for sample PM0.5 was only
one degree lower than the nonmodified sample which is
owing to the presence of short side chains in the polymer
main chains which decreases the chain movement due to
the augment in the entanglement amounts, while
increases the free volume around the chain. As the free
volume and entanglement alterations are in contrast, the
T
g
variations were insignificant. However, as the number
of chain extender agents' content is increased, the glass
transition temperature was increased again due to the
increment in side chains, entanglements and the con-
straint in chains movement. On the other hand, the side
chains growth is accompanied by an increase in the OH
end groups that are capable of hydrogen bonding, and
this alone can have a stronger effect than the presence of
side chains and elevates the glass transition tempera-
ture.
63
These outcomes were also repeated for modified
samples in the presence of a catalyst, and the glass transi-
tion temperature increased compared to the nonmodified
sample.
3.11 |Crystallization temperature
PET as a polymer with symmetrical structure in its chain
is known as a semi-crystalline polymer with a low crys-
tallization rate, thus in order to increase the rate of pro-
duction and increase its consumption as an engineering
polymer, increasing the rate of crystallization is required.
Increasing the crystallization speed causes the defect in
the use of PET in the injection process to be
eliminated,
7,64
but it can cause problem in fiber spinning.
The crystallization peak of various samples is
depicted in Figure 8 and the values of the crystalline tem-
perature are given in Table 5. The results revealed that by
increasing the PMDA content, the crystallization temper-
ature was directed toward higher values. This trend was
maintained for samples that were modified with the addi-
tion of PMDA and PENTA. It seems that by increasing
the chain extender agents, the branch structures act as
homogeneous nucleation sites, which increases the initial
rate of crystallization as well as the crystallization
TABLE 5 Thermal properties of the samples
Sample code T
g
(C) T
c
(C)(peak) T
m
(C)(end set) ΔH
m
(j/g) X
c
(%) L ( ×10
10
;m)
PET 82 168 259 44.4 32 89.01
PM0.5 81 168 255 40.6 29 83.03
PM1.0 84 172 255 36.5 26 83.03
PM1.5 84 176 251 43.5 31 77.72
PM0.5/PE0.5 84 172 255 40.5 29 83.03
PM1.0/PE0.5 84 173 255 42.1 30 83.03
PM1.5/PE0.5 84 176 253 42.0 30 80.30
PM1.5/PE0.5/D0.1 83 177 249 43.4 31 75.29
PM1.5/PE0.5/D0.4 83 180 249 43.4 31 75.29
Abbreviation: PET, polyethylene terephthalate.
DOLATSHAH ET AL.11 of 15
temperature.
2,53
Another point that is obvious at the crys-
talline peaks is the increase in width at half height of the
peak. It can be claimed that in the presence of branching
agents, by decreasing the structural order, the polydisper-
sity of the chains is increased.
18,53
3.12 |Cold crystallization
In the curve of Figure 9, it is clear that only the sample
PM0.5 showed a cold crystallization peak. It seems that
in this sample the cooling rate was higher than the
required time for the mobility of adjacent chains, there-
fore, chains did not have enough time to enter the crystal,
which is due to the low crystallinity rate of PET. By
reheating at 140C, enough time was provided for adja-
cent chain to move inside the crystal and at a specified
temperature, the cold crystallization was appeared. In
fact, it seems that reheating allows a low percentage of
branching by increasing the movement of the chains
(due to the reduction of the packing), and the chains are
able to go into the crystals. On the other hand, by aug-
mentation, the branching, entanglement increment has a
deterrent effect on the mobility of the chains, which can
be a reductive factor in the crystallization behavior,
18
therefore; cold crystallization was not observed in the
other samples.
3.13 |Melting point
By adding the chain extender agent and conversion of the
side chain of carbon-zero to a carbon, the steric hin-
drance is enhanced. In these samples, by increasing the
chain extender agents, a decrease in the melting point
was observed (Figure 9). When the side chain is extended
beyond one carbon, it increases the flexibility and free
volume since the side chains prevent packing of the main
chain. On the other hand, by increasing the amount of
chain extender, as well as the number and length of bra-
nches, the flexibility is decreased due to raising the
amount of entanglements. Furthermore, it should be
noted that the long chains are obstacles to packing,
which lead to the creation of defects in the crystal forma-
tion, thus, the size of the crystals is changed. Since the
melting temperature depends on the size of the
crystal, the melting temperature decreases with increas-
ing branching.
53,65
The crystal thickness (L) of a polymer
could be obtained by GibbsThomson equation
(Equation (2))
66
;
Tm=T
m1
2σe
ΔH
mL

:ð2Þ
where, T
m
is the melting point (K) and T
m
is the equilib-
rium melting temperature of an infinite crystal. Also,
FIGURE 8 DSC plots (cooling run) of samples. DSC, differential scanning calorimetry
12 of 15 DOLATSHAH ET AL.
ΔH
m
and σ
e
are the melting enthalpy of an ideal poly-
mer crystal per unit volume which is equal to
171 j/cm
3
,
67
and the surface free energy per unit area,
correspondingly. Huyskens and coworkers rewritten
Equation (2 for semi-crystalline PET as Equation ((3.
68
According to the obtained results of this equation
(Table 5), the thickness of crystal decreased with increas-
ing the content of chain extender.
1
Tm
=1
954:5+0:001759
L:ð3Þ
In addition, it was observed that by increasing the
branching quantity, the peak of melting became wider,
which could be due to the formation of several types of
crystals; in fact, the type and amount of the modifier
agent could affect the structure type of the created crys-
tal.
2
According to the literature,
69
by reducing the
heating rate in the DSC test, two melting peaks can be
seen which is because of different crystal structures.
3.14 |Degree of crystallinity
The percentage of crystallinity is another important char-
acteristic of the crystalline engineering polymers, which
is affected by various factors including nucleation and
molecular structure transformation. By changing these
two factors, the crystallinity could be partially improved.
The crystallinity values for the various samples are
reported in Table 5. The degree of crystallinity was calcu-
lated using the Equation (4.
Xc =ΔHmΔHccðÞ
ΔH×100:ð4Þ
where ΔH
m
and ΔH
represent the melting enthalpy of
the sample and 100% crystalline PET (reported as 140 j/g),
respectively, and ΔH
cc
is the cold crystallization
enthalpy.
3
In the lower amount of chain extender agent,
the degree of crystallinity was decreased gradually due to
the defects and reduction of structural order in the crystal,
which was caused by side chains.
53,70
Also, it should be
noted that as the number of entanglements is increased,
the folding of the chain to enter the crystal is decreased
and becomes more difficult, which causes a reduction in
the degree of crystallinity.
71
As a matter of fact, via an
increase in chain length and molecular weight, the diffu-
sion criterion was taking place with a slower rate rather
than linear state. Hence, the crystal growth in the
branched state is fewer and finally, the size of the crystals
becomes smaller.
72
By increasing the content of chain
FIGURE 9 DSC plots (second heating) of samples. DSC, differential scanning calorimetry
DOLATSHAH ET AL.13 of 15
extender agents and formation of long branches, the pos-
sibility of nucleation by these branches is provided,
resulting in a higher percentage of crystallinity. Obvi-
ously, the addition of chain extender agents reduced the
crystallinity relative to the unmodified sample.
4|CONCLUSIONS
The reaction between PET and PMDA had a significant
effect on increasing its rheological properties. Cole-Cole,
Van Gorp-Palman, storage modulus, and complex viscos-
ity diagrams showed that by increasing PMDA, viscosity
and shear thinning as well as the elastic behavior were
improved. The amount of increment in viscosity, modu-
lus, and shear thinning and decrement in MFI for the
combination of PMDA and PENTA were significantly
more pronounced than in the case of PMDA alone. These
results revealed that the long-chain branch structure sub-
sequently increased with increasing the chain extender
agent. Elastic behavior and improved rheological proper-
ties of the modified specimens were more evident in the
presence of DBTDL accelerator, which was related to the
molecular structure of these specimens, indeed, Long-
chain branching was more advanced in these specimens.
Moreover, the DSC diagrams depicted that the modified
PET had a lower melting temperature, while the low
crystallinity rate of PET was improved by increasing the
chain extender agents and the crystallization temperature
was shifted to higher values.
ORCID
Shervin Ahmadi https://orcid.org/0000-0003-1038-5146
REFERENCES
[1] N. Hudson, W. A. MacDonald, A. Neilson, R. W. Richards,
D. C. Sherrington, Macromolecules 2000,33, 9255.
[2] Y. Jahani, M. Ghetmiri, M. R. Vaseghi, RSC Adv. 2015,5,
21620.
[3] A. Ghanbari, M. C. Heuzey, P. J. Carreau, M. T. Ton-That,
Polymer (Guildf). 2013,54, 1361.
[4] S. Mandal, A. Dey, Recycling of Polyethylene Terephthalate Bot-
tles, Elsevier, The Boulevard, UK 2019,p.1.
[5] X. Xu, Y. Ding, Z. Qian, F. Wang, B. Wen, H. Zhou, S. Zhang,
M. Yang, Polym. Degrad. Stab. 2009,94, 113.
[6] Y. M. Chen, D. Zhao, L. Y. Wang, Q. C. Wu, Adv. Mater. Res.
2013,815, 738.
[7] S. Dong, Y. Jia, X. Xu, J. Luo, J. Han, X. Sun, J. Colloid Inter-
face Sci. 2019,539, 54.
[8] K. Bocz, B. Molnár, G. Marosi, F. Ronkay, J. Polym. Environ.
2019,27, 343.
[9] C. J. Tsenoglou, P. Kiliaris, C. D. Papaspyrides, Macromol.
Mater. Eng. 2011,296, 630.
[10] H. ZHONG, Z. XI, T. LIU, L. ZHAO, J. Chinese, Chem. Eng.
2013,21, 1410.
[11] L. Wang, T. Hao, L. Huang, J. Huang, R. Huang, J. Wuhan
Univ. Technol. Sci. Ed. 2019,34, 1228.
[12] B. H. Bimestre, C. Saron, Mater. Res. 2012,15, 467.
[13] F. Daver, R. Gupta, E. Kosior, J. Mater. Process. Technol. 2008,
204, 397.
[14] S. Makkam, W. Harnnarongchai, Energy Procedia 2014,
56, 547.
[15] M. Härth, J. Kaschta, D. W. Schubert, Macromolecules 2014,
47, 4471.
[16] R. Dhavalikar, M. Yamaguchi, M. Xanthos, J. Polym. Sci. Part
A Polym. Chem. 2003,41, 958.
[17] L. Xiao, H. Wang, Q. Qian, X. Jiang, X. Liu, B. Huang, Q.
Chen, Polym. Eng. Sci. 2012,52, 2127.
[18] P. Raffa, M.-B. Coltelli, S. Savi, S. Bianchi, V. Castelvetro,
React. Funct. Polym. 2012,72, 50.
[19] P. Raffa, M. B. Coltelli, V. Castelvetro, J. Appl. Polym. Sci.
2014,131, 40881.
[20] M. Xanthos, C. Wan, R. Dhavalikar, G. Karayannidis, D.
Bikiaris, Polym. Int. 2004,53, 1161.
[21] R. Dhavalikar, M. Xanthos, Polym. Eng. Sci. 2004,44, 474.
[22] A. A. Haralabakopoulos, D. Tsiourvas, C. M. Paleos, J. Appl.
Polym. Sci. 1999,71, 2121.
[23] Japon, S.; Boogh, L.; Leterrier, Y.; Ma, J. E. 2000,41, 5809.
[24] K. Wang, J. Qian, F. Lou, W. Yan, G. Wu, W. Guo, Polym. Bull.
2017,74, 2479.
[25] T. Xia, Z. Xi, T. Liu, X. Pan, C. Fan, L. Zhao, Polym. Eng. Sci.
2015,55, 1528.
[26] Incarnato, L.; Scarfato, P.; Maio, L. Di; Acierno, D. 2000,41,
6825.
[27] M. Ju, F. Chang, J. Appl. Polym. Sci. 1999,73, 2029.
[28] H. Yan, H. Yuan, F. Gao, L. Zhao, T. Liu, J. Appl. Polym. Sci.
2015,132, 42708.
[29] A. E. Keyfoglu, U. Yilmazer, MRS Proc. 2004,856, BB3.6.
[30] N. Torres, J. J. Robin, B. Boutevin, J. Appl. Polym. Sci. 2001,
79, 1816.
[31] Y. Zhang, W. Guo, H. Zhang, C. Wu, Polym. Degrad. Stab.
2009,94, 1135.
[32] Y. Zhang, C. Zhang, H. Li, Z. Du, C. Li, J. Appl. Polym. Sci.
2010,117, 2003.
[33] G. P. Karayannidis, E. A. Psalida, J. Appl. Polym. Sci. 2000,77,
2206.
[34] N. Cardi, R. Po, G. Giannotta, E. Occhiello, F. Garbassi, G.
Messina, J. Appl. Polym. Sci. 1993,50, 1501.
[35] C. R. Nascimento, C. Azuma, R. Bretas, M. Farah, M. L. Dias,
J. Appl. Polym. Sci. 2010,115, 3177.
[36] F. N. Cavalcanti, E. T. Teófilo, M. S. Rabello, S. M. L. Silva,
Polym. Eng. Sci. 2007,47, 2155.
[37] Shea, S. O.; Au, V.; Dean, R.; Au, V.; Simon, D. Us Pat.
2006,1.
[38] Diepen, V. Us Pat. 2002,1.
[39] M. Kruse, P. Wang, R. S. Shah, M. H. Wagner, Polym. Eng. Sci.
2019,59, 396.
[40] J. Liu, L. Ye, X. Zhao, Polym. Eng. Sci. 2019,59, 1190.
[41] J. S. Forsythe, K. Cheah, D. R. Nisbet, R. K. Gupta, A. Lau,
A. R. Donovan, M. S. O'Shea, G. Moad, J. Appl. Polym. Sci.
2006,100, 3646.
14 of 15 DOLATSHAH ET AL.
[42] T. Hanley, D. Sutton, E. Heeley, G. Moad, R. Knott, J. Appl.
Crystallogr. 2007,40, s599.
[43] A. G. Davies, H. Yamamoto Eds., Hydrogen-Transfer Reactions,
Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany
2007, p. 1560.
[44] A. B. Ferreira, A. Lemos Cardoso, M. J. da Silva, ISRN Renew.
Energy 2012,2012,1.
[45] D. R. de Mendonça, J. P. V. da Silva, R. M. de Almeida, C. R.
Wolf, M. R. Meneghetti, S. M. P. Meneghetti, Appl. Catal. A
Gen. 2009,365, 105.
[46] S. Einloft, T. O. Magalh~
aes, A. Donato, J. Dullius, R. Ligabue,
Energy Fuels 2008,22, 671.
[47] D. Bikiaris, G. Karayannidis, Polym. Int. 2003,52, 1230.
[48] S. Zhu, M. Shi, M. Tian, L. Qu, G. Chen, J. Text. Inst. 2018,
109, 294.
[49] P. M. Wood-Adams, J. M. Dealy, A. W. DeGroot, O. D.
Redwine, Macromolecules 2000,33, 7489.
[50] Z. Yang, C. Xin, W. Mughal, X. Li, Y. He, J. Appl. Polym. Sci.
2018,135, 45805.
[51] D. Berg, K. Schaefer, M. Moeller, Macromol. Symp. 2017,375,
1600180.
[52] J. Chen, W. Wei, Q. Qian, L. Xiao, X. Liu, J. Xu, B. Huang, Q.
Chen, Rheol. Acta 2014,53, 67.
[53] Y. Liu, A. Wirasaputra, Z. Jiang, S. Liu, J. Zhao, Y. Fu,
J. Therm. Anal. Calorim. 2018,133, 1447.
[54] S. M. Hong, H. O. Yoo, S. S. Hwang, K. J. Ihn, C. H. Lee,
Polym. J. 2001,33, 457.
[55] B. Khaledi, N. Golshan Ebrahimi, J. Appl. Polym. Sci. 2019,
136, 46896.
[56] F. Daver, F. Awaja, E. Kosior, R. Gupta, F. Cser, V. R, R.
Crescent, Rheology 2004,1, 981.
[57] Z. Qian, G. B. McKenna, Polymer (Guildf). 2018,155, 208.
[58] M. Kruse, M. H. Wagner, Rheol. Acta 2017,56, 887.
[59] N. B. Sanches, M. L. Dias, E. B. A. V. Pacheco, Polym. Test.
2005,24, 688.
[60] M. Nofar, H. O
guz, J. Polym. Environ. 2019,27, 1404.
[61] J. J. Benkoski, G. H. Fredrickson, E. J. Kramer, J. Polym. Sci.
Part B-Polymer Phys. 2001,39, 2363.
[62] J. L. Lee, E. M. Pearce, T. K. Kwei, Macromolecules 1997,30,
6877.
[63] A. Khalyavina, L. Häußler, A. Lederer, Polymer (Guildf). 2012,
53, 1049.
[64] S. M. A. Jafari, R. Khajavi, V. Goodarzi, M. R. Kalaee, H. A.
Khonakdar, J. Appl. Polym. Sci. 2019,136, 47569.
[65] J. Li, S. Tang, Z. Wu, A. Zheng, Y. Guan, D. Wei, Polym. Sci.
Ser. B 2017,59, 164.
[66] A. Hedir, M. Moudoud, O. Lamrous, S. Rondot, O. Jbara, P.
Dony, J. Appl. Polym. Sci. 2020,137,1.
[67] M. H. Chen, C. C. Lai, H. L. Chen, C. H. Lin, H. T. Hsiao,
L. C. Liu, C. M. Chen, J. Phys. Chem. Solids 2019,129, 354.
[68] P. Huyskens, G. Groeninckx, P. Vandevyvere, Bull. des Sociétés
Chim. Belges 2010,99, 1011.
[69] S. Tzavalas, V. Drakonakis, D. E. Mouzakis, D. Fischer, V. G.
Gregoriou, Macromolecules 2006,39, 9150.
[70] Y. Srithep, D. Pholharn, A. Dassakorn, J. Morris, IOP Conf.
Series: Mater. Sci. Eng. 2017,213, 012008.
[71] N. G. Karsli, J. Thermoplast. Compos. Mater. 2017,30, 1157.
[72] L. Mandelkern, Crystallization Polym. Equilibrium Concepts
2011,1, 214.
How to cite this article: Dolatshah S, Ahmadi S,
Ershad-Langroudi A, Jashni H. Rheological/
thermal properties of poly(ethylene terephthalate)
modified by chain extenders of pyromellitic
dianhydride and pentaerythritol. J Appl Polym Sci.
2020;e49917. https://doi.org/10.1002/app.49917
DOLATSHAH ET AL.15 of 15
... As mentioned above, it is difficult to measure the elastic plateau region of PET in frequency sweep experiments because of the relatively low molecular weight and short relaxation time [66]. For this reason, many studies have focused on the low-G part of the curve [40,47,48,52,61,66,102,103]. Several authors used Cole-Cole plots to estimate the effects of various chain extenders on linear PET systems. ...
... Several authors used Cole-Cole plots to estimate the effects of various chain extenders on linear PET systems. Daver [47] evaluated the efficiency of PMDA; Wang [61] studied a step-chain extension process with two different additives, SAG-008 and TGDDM; Dolatshah [48] examined the effect of a catalyst, DBTDL, on the chain-extension reaction; and Liu [40] analyzed the combination of two different chain extenders, PMDA and TADE. Cole-Cole plots were reported to be sensitive to polydispersity and the degree of branching but independent of molecular weight [104]. ...
... In general, virgin PET samples are above the line G = G . The main outcome of Cole-Cole analysis in the literature on PET is that a higher degree of branching shifts the curve towards higher G values with constant G values [48,58,61], as shown in Figure 7. It can also be noted that a higher degree of branching induces a shoulder in the plot at high G values. ...
Article
Full-text available
Polyethylene terephthalate (PET) is a thermoplastic material that is widely used in many application fields, such as packaging, construction and household products. Due to the relevant contribution of PET to global yearly solid waste, the recycling of such material has become an important issue. Disposed PET does not maintain the mechanical properties of virgin material, as exposure to water and other substances can cause multiple chain scissions, with subsequent degradation of the viscoelastic properties. For this reason, chain extension is needed to improve the final properties of the recycled product. Chain extension is generally performed through reactive extrusion. As the latter involves structural modification and flow of PET molecules, rheology is a relevant asset for understanding the process and tailoring the mechanical properties of the final products. This paper briefly reviews relevant rheological studies associated with the recycling of polyethylene terephthalate through the reactive extrusion process.
Article
Full-text available
Polyethylene terephthalate (PET) is used in a variety of packaging applications, such as bottles, films, fibers, for its cost‐effective and excellent physical properties. However, the increase in plastic usage and packaging waste has become major global problems. Therefore, the importance of plastic recycling is being emphasized. In this study, the bottle‐grade PET was repeatedly melt processed up to three times to evaluate of PET recyclability. Intrinsic viscosity (IV) and mechanical properties have significantly decreased due to chain scission during the melt reprocessing. In particular, the IV of 3rP (reprocessing three times) decreased to 0.582 dL/g, and the mechanical properties decreased more than twice compared with 0rP. To enhance the physical properties of reprocessed PET (rP‐PET), it was modified using two chain extenders. The IV of the modified 3rP‐PET increased from 0.645 to 0.737 dL/g. And the tensile strength was improved up to approximately 93% of the 0rP properties, confirming the chain extension effect. Additionally, a biaxial orientation process was applied to confirm the feasibility of modified rP‐PET to various industrial packaging applications. Consequently, as the stretching ratio increased, the physical properties of modified rP‐PET improved, confirming the feasibility of rP‐PET in various packaging fields.
Article
A series of modified aromatic oxadiazole polymers were synthesized via condensation reactions of terephthalic dihydrazide (TPH) and varying amounts of Trimesic acid (H 3 BTC) in polyphosphoric acid (PPA). Pentaerythritol was introduced to construct intermolecular hydrogen bonds among polymeric chains, resulting in a three-dimensional network structure, and correlations between H-bonds and mechanic properties are discussed. Structures and mechanic properties of the polymers were characterized and analyzed. By investigation of molecular structure and membrane formation process of chemically modified and hydrogen bond-regulated polymer membranes, it was found that the modified polymer membranes retain the unique thermal stability, chemical stability, and good hydrophobicity of the POD rigid polymer. Additionally, hydrogen bonds not only significantly improve the overall quality of the membrane, but also play an important role in enhancing the mechanical properties of the membrane, as demonstrated in the highest observed of up to a maximum of 167.19 MPa. Furthermore, impact of H 3 BCT modification on light absorption and fluorescence performance of T-PODx is also analyzed and discussed in this paper.
Article
Plastics were developed to change our world for the better. However, plastic pollution has become a serious global environmental crisis. Thermoplastic polyesters and polyolefins are among the most abundant plastic waste. This work presents an in-depth non-isothermal crystallization kinetics analysis of recycled post-consumer poly(ethylene terephthalate) (rPET) and recycled polypropylene (rPP) blends prepared through reactive compounding. The effect of pyromellitic dianhydride (PMDA) on crystallization kinetics and phase morphology of rPET/rPP blends was investigated by differential scanning calorimetry (DSC) and microscopy techniques. DSC results showed that increasing rPP content accelerated rPET crystallization while reducing crystallinity, which indicates the nucleation effect of the rPP phase in blends. Further, it was found that the incorporation of PMDA increased the degree of crystallinity during non-isothermal crystallization, even though the rate of crystallinity decreased slightly due to its restriction effects. The non-isothermal crystallization kinetics was analyzed based on the theoretical models developed by Jeziorny, Ozawa, Mo, and Tobin. The activation energy of the crystallization process derived from Kissinger, Takhor, and Augis-Bennett models was found to increase in rPET/rPP blends with increasing PMDA due to hindered dynamics of the system. Rheological measurements revealed that rPET melt viscosity is remarkably increased in the presence of PMDA and reactive blending with rPP relevant for processing. Moreover, nanomechanical mapping of the rPP phase dispersed in the rPET matrix demonstrated the broadening of the interfacial domains after reactive blending due to the branching effect of PMDA. Findings from this study are essential for the recycling/upcycling thermoplastics through non-isothermal fabrication processes, such as extrusion and injection molding, to mitigate the lack of sorting options.
Article
In this paper the effect of different organosepiolites on PET matrix was evaluated through the analyses of Melt Flow Index (MFI) and Intrinsic Viscosity (IV) of the nanocomposites. PET/sepiolite nanocomposites were produced through a two-stage process. The masterbatch production was done in a melt compounding extruder, and then these were taken to an injection machine, where different specimens were obtained for characterization purposes. Two types of organomodifiers were tested, silanes in one hand and a quaternary ammonium salt on the other. Thermal analyses were done to determine thermal stability of the organomodifiers. Then, different experiments were carried out increasing the concentration of coupling agents to analyse its effect on PET matrix degradation. As results indicate a high degree of degradation of the nanocomposites matrix, pyromellitic dianhydride (PMDA) was added as chain extender to react with PET molecular chains, with the aim of increasing its molecular weight (Mw).
Article
Full-text available
Thermal degradation of polyethylene terephthalate (PET)/carbon nanotube (CNT) nanocomposites is a serious issue in the manufacturing process of this nanocomposite that can limit the applications of the nanocomposite. In this study, the effects of two kinds of chain extenders, pyromellitic dianhydride (PMDA) and pentaerythritol (PENTA), on the rheological behavior, thermal characteristics, and crystallinity of the PET were investigated. The results of shear rheology revealed improvement of the storage modulus and complex viscosity, which indicated the recoupling of broken chains, increasing the chain length, and the generation of long branches was the reason for chain extension. Also, differential scanning calorimetry (DSC) results clarified an increase in PET crystallization rate in the presence of CNT and chain extenders. Furthermore, SEM was performed to detect CNT dispersion in the nanocomposite, and four‐probe technique test was utilized to determine the influence of CNT on the electrical properties of PET. The effect of presence of CNT, CE, solely and simultaneously on rheological properties.
Article
Full-text available
In this study, PBT/recycled-PET blends were developed using a twin-screw extruder. A commercial chain extender (Joncryl ADR 4468) was also used to melt mix with PBT/recycled-PET blend systems. Firstly, Joncryl at different loadings was extruded with recycled-PET to explore the influence of branching on melt behavior and crystallization of PET. The effect of blending recycled-PET with PBT on the final properties was then explored at different blending ratios (25w/75w, 50w/50w and 75w/25w). Similar blends were subsequently prepared while incorporating chain extender. Melt behavior, phase miscibility, crystallization behavior, solid viscoelastic properties, tensile and impact properties of the blends were eventually analyzed using differential scanning calorimeter (DSC), melt flow indexer (MFI), dynamic mechanical analyzer (DMA), and tensile and Izod notched impact testing, respectively. The results showed that the addition of chain extender increased the melt viscosity of PET and at the low contents enhanced the PET’s crystallization rate. On the other hand, the blends of PBT/recycled-PET are fully miscible in the amorphous region whereas the crystalline phases are immiscible subsequent to a fast cooling. The PBT and PET molecules could also co-crystallize and be fully miscible in crystalline phases upon slow cooling of the melt. Blending recycled-PET with PBT didn’t suppress the tensile properties of PBT, however it could enhance the PBT’s ductility and reduce its impact strength. The chain extender didn’t influence the mechanical properties much. Graphical Abstract
Article
Full-text available
This article deals within the study of the effect of artificial radiations on physical and chemical properties of the crosslinked polyethylene (XLPE) material, widely used for manufacturing high‐voltage cables. Within this framework, several experimental tests, using essential characterization techniques, were performed to study XLPE behavior under ultraviolet (UV) aging. Attenuated total reflection Fourier transform infrared spectroscopy, differential scanning calorimetry, thermogravimetric analysis, X‐ray diffraction, and scanning electron microscopy were thus carried out to identify the main structure changes of the material before and after exposure to UV. In addition, appearance changes and DC (Direct Current) volume resistivity were evaluated. The obtained results showed that UV radiation has a great effect on the physicochemical properties of XLPE cable insulation. © 2019 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2020, 137, 48575.
Article
Full-text available
Nonisothermal crystallization kinetics of poly(ethylene terephthalate)/polylactic acid (PET/PLA) (90/10 and 75/25 wt %/wt %) blend prepared by a melt mixing process in the presence of graphene oxide (GO) and exfoliated graphite nanoplatelets (xGnP) were investigated. The transmission electron microscope (TEM) results revealed that both the GO and xGnP were not inclined to the PLA phase, and the distribution of GO in the blend matrix was better than that of xGnP. The scanning electron microscope (SEM) results demonstrated that both the fillers showed compatibilizing effect and refined the morphology of PLA dispersed phase. The crystallinity and crystallization behavior of samples were studied by X‐ray diffraction (XRD) and multiple cooling rate Differential scanning calorimetry (DSC), respectively. The results showed that the degree of crystallinity and crystallization rate of PET was reduced by blending with PLA. Inclusion of the fillers to the blend further reduced the crystallinity while the crystallization rate was increased. Analysis based on Hoffman‐Lauritzen theory revealed that, compared to xGnP, the GO as a better nucleating agent could more effectively reduce lamella thickness and surface folding free energy of the nanocomposites. Crystallization kinetic studies by Jeziorny model revealed two levels of primary and secondary crystallization mechanisms for all samples. Moreover, activation energy of crystallization of the GO‐containing nanocomposites was lower than that of xGnP‐based systems. All the theoretical treatment and experimental results confirm that, compared to xGnP, the GO is more suitable filler and has stronger effect on enhancing the crystallization kinetic of the PET/PLA blends. © 2019 Wiley Periodicals, Inc. J. Appl. Polym. Sci. 2019, 136, 47569. HIGHLIGHTS Studied crystallization kinetics of PET/PLA blend containing xGnP and GO nanoplatelets Compared the role of xGnP and GO in crystallization kinetics of the system Analyzed the crystallization kinetics of the system by various models Determined the surface folding free energy of the developed systems GO‐based systems exhibited lower crystallization activation energy than that of the xGnP‐based systems
Article
Full-text available
Recycled polyethylene-terephthalate (RPET) bottle regrinds were physically foamed after two types of industrially feasible molecular weight increasing processes. Intrinsic viscosity (IV) of initial waste (0.71 dL/g) increased both after reactive extrusion carried out using a multifunctional epoxy-based chain extender (0.74 dL/g) and after solid state polycondensation (SSP) (0.78 dL/g), while capillary rheometry revealed higher degree of branching in the chain extended PET material. Despite the relatively low IV values (below 0.80 dL/g), physical foaming, a mild and cost-efficient way, was successful in both cases, uniform microcellular foam structures with void fractions ranging between 75 and 83% were achieved. During the experiments morphology change in the materials was tracked by differential scanning calorimetry (DSC) besides recording IV values. The IV drop during foaming was between 0.03 and 0.10 depending on the pre-processing technology. Structure of foams produced from the two different modified RPET materials was compared with each other based on scanning electron microscopic imaging of cryogenic fracture surfaces. The average cell diameters were measured to be 213 and 360 µm in the case of chain extended and SSP-modified materials, respectively.
Article
To decrease the modulus and increase the mechanical properties of recycled PET by introducing the flexible segments, the recycled polyethylene-terephthalate (PET) was modified by the chain extender produced with 2, 4, 6, 8-tetramethyl-2, 4, 6, 8-tetra (2, 3-epoxypropoxy) propylcyclotetrasiloxane by the hydrosilylation reaction. Different percentages of chain extender were used in order to investigate the effect of chain extender content on the modulus of PET. The dynamic mechanical analysis was performed on PET samples, which showed that Tg of PET was firstly raised by the addition of chain extender and then decreased. Small amplitude shearing test performed on PET samples confirmed that the addition of the flexible segments could remarkably decrease the modulus of PET, which would be helpful for the extrusion process. The entanglements from branches were also influenced by the addition of extender.
Article
Poly(ethylene terephthalate) (PET) was long‐chain branched (LCB) by ring‐opening reaction with both pyromellitic dianhydride and tetrahydrophthalic acid diglycidyl ester as chain extenders through reactive melt processing. It was found that with the increase of chain extenders dosage, the intrinsic viscosity of PET increased and melt index decreased greatly, while both the tensile strength and impact strength of PET were remarkably improved. The elastic modulus (G′) and viscous modulus (G″) were enhanced by chain branching. Compared with PET, the complex viscosities of LCB‐PET were much higher at full frequency range, and obvious shear thinning was presented. The Cole–Cole curve deviated from the semicircular shape and the curve end was inclined to upward in high viscosity region, indicating the formation of the multiple hierarchical structures. The molecular weight of the branch (MB) was much greater than critical entanglement molecular weight (Me), which essentially confirmed the existence of LCB structure and fairly strong molecular entanglement in the LCB‐PET molecular chain. When subjected to external force, the entanglement point, acting as physical crosslinking point between the molecules, was in favor of increasing the molecular interaction, reducing the molecular slippage, and bearing a large deformation. POLYM. ENG. SCI., 2019. © 2019 Society of Plastics Engineers
Article
Long-chain branched polyethylene terephthalates (PETs) composed of ethylene glycol, terephthalic acid, trimesic acid, and their biaxially stretching films with superior crystal sizes and inferior crystal growth rates, melting points, equilibrium melting points, fold surface free energies, and orientation parameters compared to common PET were manufactured by T-Die extrusion and biaxially stretching treatment. In order to evaluate the hydrolysis resistance for biaxially stretching films of lab-made long-chain branched PETs, we examined their half-life period of elongation, defined as the time for which the films could retain their elongation at 50% during the hydrolysis procedure under conditions of 100% humidity, 1.5 atm pressure and 120 °C. Experimental results revealed that biaxially stretching film of long-chain branched PET II had high potential as an optical substrate for optoelectronic devices because its half-life period of elongation was 58 compared to 21 h for common PET.
Article
Poly(ethylene terephthalate) (PET) generally suffers from low crystallization rate and long molding duration, which as a result limit its application as engineering plastics. To overcome these drawbacks, series of PET/layered double hydroxide (LDH) nanocomposites were prepared by a solution blending process. The effect of metal composition (MgAl and CaAl) and organo-modification (stearic acid intercalated) for LDH fillers on the crystallization behavior of the nanocomposites was investigated. It was revealed that, compared with PET/CaAl-LDH, the PET/MgAl-LDH nanocomposite exhibits a higher crystallization temperature and faster crystallization rate, which is associated with the superior nucleation ability of MgAl-LDH. The nucleation mechanism of PET induced by LDHs was explored by means of Avrami equation and theory of Hoffman-Lauritzen, pointing out that the incorporation of LDHs reduce the free energy of nucleation and the fold surface free energy of PET. In order to improve the compatibility between LDH and PET, stearic acid (SA) intercalated MgAl-LDH was prepared and filled into PET matrix. The resultant PET/MgAl-LDH-SA shows a further enhanced crystallization temperature and accelerated crystallization rate, in comparison with PET/MgAl-LDH nanocomposites. In addition, the thermal stability, gas barrier and mechanical properties of PET/LDH composites were improved upon incorporation of LDH fillers.