You are on page 1of 32

The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming

Behavior

The Preparation and


Characterization of Branching
Poly(ethylene terephthalate) and its
Foaming Behavior
Haiming Liu1, Xiangdong Wang1,#, Hongfu Zhou1, Wei Liu2, and
Bengang Liu1
1School of Materials and Mechanical Engineering, Beijing Technology and Business
University, Beijing 100048, PR China
2School of Materials and Metallurgical Engineering, Guizhou Institute of Technology,

Guizhou 550001, China

Received: 15 November 2104, Accepted: 7 January 2015

SUMMARY
Chain extension was an effective method for increasing the molecular weight, the
melt strength and the foaming property of linear polymer. In this paper, pyromellitic
dianhydride (PMDA) was used as chain extender to improve these properties of
linear poly (ethylene terephthalate) (PET). The intrinsic viscosity, rheological and
thermal characterizations of various PET samples was investigated. The results
demonstrated that the increasement of the viscoelasticity at low frequencies
was correlated to the raise of the intrinsic viscosity and the formation of long
chain branching. These structural changes resulted in the decreasement of
the crystallization temperature and melt temperature as well as the increase
in the cold crystallization values with the increasing content of PMDA. The
cellular morphology and expansion ratio of CEPET foams were also obviously
improved by the introduction of PMDA. The expansion ratio of CEPET foam with
the PMDA content of 1.0 phr would reach 31.78. In addition, the effect of the
chain extension reaction time on the intrinsic viscosity, the rheological behavior,
and foaming properties of PET were also studied. The results showed that the
intrinsic viscosity, the rheological behavior, and the foamability of CEPET also
decreased gradually with increasing chain extension reaction time, which should
be attributed to the occurrence of more and more intense thermal degradation.

Keywords: Poly(ethylene terephthalate); Chain extender; Branching; Degradation; Foam

#Author to whom correspondence should be addressed. (Xiangdong Wang).


E-mail address: wangxid@th.btbu.edu.cn
©Smithers Information Ltd. 2015

Cellular Polymers, Vol. 34, No. 2, 2015 63


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

INTRODUCTION
Currently, poly (ethylene terephthalate) (PET) foams have been widely used
in construction, aerospace, modern automotive, marine, food packaging,
and other industries [1, 2, 3]. In particular, PET microcellular foams using
supercritical carbon dioxide as blowing agent are considered one of the most
promising substitute to replace polyvinyl chloride foaming materials in wind
power fields because of its recyclable, processing costs and comparable
properties [4, 5]. Preparing PET foam with excellent property has gradually
become a research hotspot.
However, common PET is still a challenge for preparing foams with low density
on the grounds that its poor melt strength associated with linear molecular
structure and low molecular weight. Therefore, some technological ways,
such as, chain extension, blending, and filling with some special fillers, were
employed to solve these problems [6, 7, 8]. Among these methods, many
researchers found that the chain extension reaction was the most effective
way. It could be easily applied to modify the molecular structure of PET and
prepare PET foams with high performance by the common melt processing
method [9, 10, 11].
S. Japon, L and Wang reported that tetraglycidyl diaminodiphenyl methane
(TGDDM) was used as chain extender (CE) with epoxy groups to modify PET
[12, 13]. The results showed that the introduction of TGDDM would significantly
increase the molecular weight and improve the rheological properties of PET.
When the addition of TGDDM was excessive, the crosslinking structure and
gel would occur. Triphenyl phosphate and diimidodiepoxides were employed
as CE to modify PET by Bikiaris et al. They found that these compounds were
highly reactive with PET as well as the molecular weight and melt strength of
PET were increased significantly [14].
During chain extension, branching or crosslinking reactions might occur,
the molecular structure and some other properties of PET would be greatly
changed. The presence of branching structure would produce a remarkable
enhancement in the molecule weight and the melt strength of chain extended
PET (CEPET). In the foaming process, PET with the branching structure
could stabilize the bubble growth and retard cell coalescence or rupture.
Furthermore, it was easy to produce crosslinking structure when the addition
of chain extender was excessive, especially for chain extender with epoxy
groups. Crosslinking polymers had some advantages such as high burst
strength, heat tolerance, and better abrasion resistance. But during the PET
foaming process, with the increasing degree of crosslinking, the diffusion of
the gas in CEPET melt would decrease [15]. For another important aspect,
the crosslinking polymer was unrecyclable.

64 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

The chain extension reaction time is another important factor for PET
foaming. Because high processing temperature, high processing pressure,
and the presence of moisture all result in the loss of intrinsic viscosity and
the decreasement of foamability during the PET processing. The degradation
of PET occurred during processing, which had attracted much attention in
the literatures [16, 17, 18, 19]. Filippo Samperia et al. investigated thermal
degradation of PET at high processing temperature. They found that the
addition of small amount of p-Toluenesulfonic acid monohydrate to PET,
heated at 270 and 285°C, induced a strong hydrolytic reaction with consequent
increasement of carboxyl terminated polyester chains [20]. Three kinds of
degradations happened in PET processing: thermal degradation, thermal-
oxidation degradation, and thermal-hydrolysis reaction. At high processing
temperature (above 280°C), hydrolysis reactions occurred between water
and PET, resulting in the shorter PET chains. Fortunately, hydrolysis reaction
could be greatly relieved by pretreating PET samples dried in the vacuum
to remove extra moisture. However, the thermal degradation reaction is
inevitable in processing of PET. The thermal cleavage of ester bond of PET
results in molecular chains with carboxyl and vinyl ester end groups [21].
In a high processing temperature, the excessive increment of reaction time
would make the thermal degradation reaction become intense, which was
disadvantage for increasing the melt strength of PET. Therefore, in order
to relieve the thermal degradation of PET during processing, controlling
chain extension reaction times are very important. Based on the above
considerations, not only the CE content but also the chain extension
reaction time are important factors for obtaining PET foam with excellent
performance. However, there were seldom work reported on the effect of
CE content and chain extension reaction time on the molecular structure
and the foaming behavior of PET, simultaneously.
This work focused on the effect of CE content and chain extension reaction
time on the foaming behavior of CEPET using PMDA as CE. The molecular
structure of PMDA was shown in Figure 1. PMDA was selected since it‘s
melting point (about 283°C) was near to the processing temperature of PET
and was suitable to react with PET, helping to obtain the structure with high
branching degree. The change of the molecular structure was expected to
improve the dynamic shear rheological properties, thermal performance, and
foaming behavior of PET.

Cellular Polymers, Vol. 34, No. 2, 2015 65


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

Figure 1. The molecular structure of PMDA

EXPERIMENTAL

Materials
PET resin (BG-03-80) with number-average molecular weight (Mn) about
20,000 g/mol was purchased from FSPG HI-TECH Co. Ltd, China. PMDA
was supplied by Aldrich Co. Ltd, USA. PMDA was a tetra-functional chain
extender that could generate branching and/or crosslinking structures.
The formula was shown in Table 1. All components were recorded using
the unit of “parts per hundred resin” (phr) which was weight based in the
formula. Samples of PET modified with 0.25, 0.5, 0.75, and 1.0 phr PMDA
were prepared, respectively.
Table 1. The weight ratio of the various CEPET
Serials No. PET/phr PMDA/phr
1# 100 0
2# 100 0.25
3# 100 0.50
4# 100 0.75
5# 100 1.00

Preparation of CEPET
The CEPET samples were prepared in a Haake internal mixer at 280°C, with
a different mixing time. Prior to the melt mixing, PET was dried 160°C for 4 h
in a vacuum oven to remove the moisture [22]. After the chain extension, the
CEPET samples were dried in the dryer for further characterization.
The PET foams were prepared by batch foaming method in autoclave using
supercritical CO2 as physical blowing agent. In order to investigate the
differences of foaming behaviors for CEPET with various PMDA contents,
the samples were prepared and foamed under the same conditions. First, the

66 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

samples were put into autoclave at a temperature of 270°C and a foaming


pressure of 20 MPa for 1 h. After the CO2 fully diffused and dissolved in
the PET melt, the samples experienced an instantaneous pressure drop by
release of CO2 from 20 MPa to 0.1 MPa, which provided a driving force for
cell nucleation and growth. The schematic diagram of autoclave batching
foaming was shown in Figure 2. At last, the foaming samples were prepared
for further characterizations.

Figure 2. Schematic diagram of autoclave batching foaming, using supercritical CO2


as physical blowing agent

Characterizations

Viscosity
Solution viscosity measurement of the various PET samples was done by an
Ubbelohde-1B viscometer in order to determine the intrinsic viscosity ([η]).
A solvent mixture composed of phenol/1,1,2,2-tetrachloroethane (m/m =
60/40) was needed to prepare the PET solutions. The various PET samples
were dried in a vacuum oven for 4 h at 160°C to avoid polymer degradation.
Then, every PET sample (0.06, 0.1, 0.14, 0.18, 0.2 g) was dissolved in the
solvent mixture (20 mL) at 110°C for 5 hours, respectively. After the complete
solubilization, the solutions were filtered and tested at 30°C in water bath,
according to standard test method ASTMD 4603 (Standard Test Method for
Determining Inherent Viscosity of PET), respectively [23]. From the flow time of
the pure solvent (t0) and solvent with the CEPET (t), it was possible to obtain
the relative (ηr), specific (ηesp), reduced (ηred), and inherent (ηiner) viscosities
by means of the following Equations (1-4) [24]:

Cellular Polymers, Vol. 34, No. 2, 2015 67


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

(1)

(2)

(3)

(4)

Where, t is the flow time of polymer solutions, t0 is the flow time of pure
solvent mixture, and c is the polymer solution concentration (g/100 ml). The
extrapolation of ηred or ηiner to concentration limit tending to zero gives intrinsic
viscosity ([η]) [25, 26]:

(5)

Five measures of flow time of PET solution with different concentration were
obtained to produce a linear regression curve of ηiner versus concentration,
the corresponding extrapolation for concentration limit tending to zero
corresponds to [η], according to Equation (5).

Fourier Transformation Infrared Spectroscopy (FTIR)


Infrared spectra of various PET samples were obtained using an FTIR (Nicolet
iS10 thermo scientific spectrometer) in transmission mode. Each spectrum
was obtained within the range of 3800 cm-1-500 cm-1 with a wavelength
resolution of 4 cm-1.

Thermal Behaviour
Crystallization and melting behaviors of unmodified PET and CEPET samples
were studied by DSC system purged with nitrogen. The samples were heated
to 300°C at rate of 10°C/min, isothermal for 3 minutes, then cooled to 40°C
at rate of 10°C/min, the crystallization and melting behaviors were measured
[27, 28]. The crystallinity of PET was computed by Equation (6) [29]:

(6)

68 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

Where, ΔHm and ΔH0m are fusion enthalpy of polymer and hypothetic fusion
enthalpy of the PET 100% crystalline, respectively. ΔHm is calculated from
area of the endothermic signal, while ΔH0m is 140 J·g-1 [30].

Dynamic Shear Rheology


Shear rheological behavior of various PET samples was observed by a strain-
controlled rheometer (MARS Rheometer, TA, USA) at 260°C, with a parallel
plates (20 mm in diameter with a gap of 1.0 mm). The frequency range was
from 0.1 to 100 rad/s, and the maximum strain was fixed at 5%, to confirm
that these conditions were within the linear viscoelastic region under nitrogen
[31, 32, 33]. The storage modulus (G’), complex viscosity (h*), and loss factor
(tan δ) were measured at various frequencies.

Cellular Morphology
The cellular morphology of foaming samples were characterized by a
scanning electron microscope (SEM, TESCAN YEGA II, TESCAN s.r.o) at an
acceleration voltage of 10 kV. The magnification ratio was 400 for observing
cell morphology. Cell density was analyzed by using software image tool and
calculated as shown in Equations (7)-(8) [34, 35]:

(7)

(8)

Where Nc is cell density (cells/cm3), n is cell amount within statistical areas,


A is statistical areas in SEM photo (cm2), φ is the foaming expansion ratio of
polymer, ρf is foamed density (g/cm3), ρu is unfoamed density (g/cm3).

Foam Density
The density of foaming samples was determined by using density tester
(UltraPyc actual density tester, Quanta chrome, America).

Cellular Polymers, Vol. 34, No. 2, 2015 69


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

RESULTS AND DISCUSSION

The Reaction Mechanism Between PET and PMDA During the


Chain Extension
During the chain extension process, several chemical reactions existed
and took place: chain scission induced by degradation, branching and
crosslinking induced by chain extension. In the typical branching reaction,
four PET molecular chains would react with one PMDA molecule, generating
the starlike architecture. When some PMDA molecules reacted with multi
PET molecules, the crosslinking structure or gel would occur. The reaction
mechanism illustration of this process was shown in Figure 3.

70 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

Figure 3. Illustration of possible chemical reactions (degradation, branching,


crosslinking) during the chain extension of PET

Efficiency of Chain Extender With PET


The torque curves of PET mixing with PMDA (0, 0.25, 0.5, 0.75, and 1.0 phr)
at 280°C for 15 min in the Haake rheometer were shown in Figure 4. The
torque curves could reflect on the torque tendency of CEPET melt and the
relationship between the torque with time, but it was not a well-defined
rheological property. The above stated three reactions (degradation, branching,
crosslinking) took place simultaneously and competed with each other. As
expected, the unmodified PET only showed a low torque, decreasing with

Cellular Polymers, Vol. 34, No. 2, 2015 71


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

the mixing time. When the content of PMDA was at 0.5-1.0 phr, there was
an increase in the torque of PET from 4 min to 10 min of mixing time, which
should be attributed to the formation of branching and crosslinking structures.
And then, the torque curves decreased from 10 min to 15 min of mixing time,
since the degradation was predominant in the competed reaction. The torque
value of samples (3-5#) reach the maximum at approximately 10 min, thus the
optimal chain extension reaction time was chosen 10 min. The new samples
(1-5#) with the same formula were processed with 10 min again, these new
samples were used to characterize and discuss subsequently. The overall
torque evolution of the CEPET strongly depended on the content of chain
extender. With the content of chain extender, the maximum value of the torque
increased, which should be attributed to the integrated consequence of the
above three reactions.

Figure 4. Torque curves of unmodified PET and CEPET

Intrinsic Viscosity
The intrinsic viscosity of the PET samples was considered as a critical parameter
in chain extension reaction because of its relationship with molecular weight.
The determination of solution intrinsic viscosity of PET samples was carried
out according to methods ASTMD 4603. The flow time measurements were
performed for at least five different concentrations for every PET sample. The
intrinsic viscosity of each PET sample (1-4#) at different solution concentration

72 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

was shown in Figure 5. The extrapolation of ηiner to concentration limit tending


to zero gave [η]. Five PET samples’ intrinsic viscosity was tested and four
PET samples’ data (1-4#) was shown in Table 2.

Figure 5. The extrapolation of ηiner to obtain the intrinsic viscosity of each PET sample

In Table 2, the intrinsic viscosity was reported for all samples (1-4#) analyzed.
We can see that the intrinsic viscosity of CEPET with the lowest amount of
PMDA (0.25 phr) was higher than that of the unmodified PET. With the PMDA
content increasing from 0.5 to 0.75 phr, the intrinsic viscosity increased
gradually. These results indicated that PMDA was an effective chain extender
in this work. Moreover, it was important to point out that some crosslinking
structure would generate in the CEPET samples when the content of the chain
extender was higher than 0.75 phr. For this reason, the Ubbelohde viscometer
data of the sample PMDA-1.00 was not reported. In fact, PMDA had four
reactive sites and could react with four PET molecular chains (in Figure 3). As
a consequence, some branching structure occurred during chain extension
of PET, resulting in the increasement of the intrinsic viscosity of the CEPET.

Cellular Polymers, Vol. 34, No. 2, 2015 73


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

Table 2. Intrinsic viscosity of various PET samples


Samples ASTMD 4603 (30°C)
1# 2# 3# 4# 5#
[η](dl/g) 0.63 0.84 1.17 1.54 --

Reaction Between PET and Chain Extender


FTIR is an important analytical method that allows evaluating possible chemical
changes occurred in the reaction of PET with PMDA. Figure 6a illustrated
the effect of PMDA content on FTIR-transmission spectra of unmodified PET
and CEPET. In the FTIR spectra, double shoulder peaks from 1250 cm-1 to
1050 cm-1 were related to the vibration of C-O from carboxyl groups, and
absorption bands at 1340 cm-1 and 1370 cm-1 were attributed to the wagging
of the ethylene units [36]. The intensive peak around 1716 cm-1 represented
C=O stretching vibrations. The absorption band at 3100 cm-1-2800 cm–1 was
related to aromatic and aliphatic C-H bond stretching of the PET that did not
change after the reaction of PET with PMDA [37], as shown in Figure 6c. The
large band at 3420 cm–1 was attributed to the absorption of hydroxyl groups
[38]. The anhydride groups on PMDA can react theoretically with the hydroxyl
groups on the end of PET chain. The integration area ratio between O-H
group and C-H bond of PET are 41.52/161.93, 36.81/157.78, 33.84/162.93,

(a)

Figure 6. (a) The FTIR spectra of various samples in wavenumber between 500 cm–1
and 3800 cm–1

74 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

(b)

(c)

Figure 6. (b) FTIR spectra of the hydroxyl region, (c) FTIR spectra of the anhydride
region

Cellular Polymers, Vol. 34, No. 2, 2015 75


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

32.81/164.94, 25.55/171.54, respectively. As we expected, with the content of


PMDA, the integration area ratio decreased gradually which could be related
to the reaction of PET and PMDA, as shown in Figure 6b. The peaks around
1860 cm-1, 1806 cm-1, and 1770 cm−1 were related to the anhydride stretching
vibration of PMDA in the spectrum of Figure 3c [39]. However, these three
peaks don’t appear in the spectrum of the samples of CEPET, indicating that
the anhydride groups on PMDA have been consumed in the reaction.

Thermal Behavior of CEPET


DSC was utilized to evaluate the crystallization and melting properties of
various PET samples. Figures 7-9 were the cooling and heating curves for
unmodified PET and CEPET samples. The corresponding parameters of various
PET samples, such as crystallization temperature (Tp), cold crystallization
temperature (Tcc), melting temperature (Tm), the crystallization enthalpy (DHc)
and the melting enthalpy (DHm), were summarized in Table 3, respectively.
It could be seen from Figures 7-8 and Table 3, the Tp and Tcc of the PET samples
were affected significantly by the addition of PMDA. The Tp of unmodified PET
sample (1#) was of around 209.07°C. The crystallization peak shifted towards
lower temperatures from 199.2°C to 193.82°C with the content of PMDA

Figure 7. Non-isothermal DSC crystallization curves for PET samples at cooling rate of
10°C/min

76 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

increasing. Moreover, the crystallization peak declined and became broader


by increasing PDMA content. The possible reason of the crystallization peak
changes was that the branching and crosslinking (PMDA-1.0) structures of
CEPET might affect the structural regularity of PET chains, which resulted in
the decreasement of crystallinity. Through the chain extension reactions, the
rigid benzene ring group was introduced into the PET molecular chains, and
the molecular chain length of PET increased. As a result, the mobility of PET
chain segment became difficult and the crystallization temperature of PET
decreased. The obvious decreasement of crystallization temperature of PET
samples could provide sufficient time for the growth and cooling stability of
PET foams in the subsequent foaming process.
In Figure 8, with the content of PMDA increasing, the cold crystallization
temperature of PET samples increased from 130.2°C to 137.8°C, and the cold
crystallization enthalpy also increased gradually from 1.16 J.g-1 to 4.01 J.g-1.
The possible reason was that the higher viscosity of CEPET samples restricted
the assembly of CEPET molecular chains into crystal lattices adequately,
when the temperature was cooling [40].

Figure 8. Non-isothermal DSC cold crystallization curves for PET samples at heating
rate of 10°C/min

As in Figure 9, DSC curve of unmodified PET possessed a melting peak at


about 257.54°C. However, with the increasing content of PMDA, the melt

Cellular Polymers, Vol. 34, No. 2, 2015 77


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

peak of CEPET samples shifted toward lower temperature from 253.35°C to


250.74°C, and the melting enthalpy also increased gradually from 38.02 J.g-1
to 34.66 J.g-1. Its low melting enthalpy results in smaller crystal size. The
change of melting temperature and enthalpy were also associated with the
molecular structure of PET.

Figure 9. Non-isothermal DSC melting curves for PET samples at heating rate of 10°C/min

As summarized in Table 3, the crystallinity of PET decreased due to the


addition of PMDA and the formation of the branching and crosslinking structure
of PET which influenced the crystallization process. The crystallinity of PET
decreased from 27.8% to 21.89%, which would influence some physical
and mechanical properties of PET foams, such as gas permeability, impact
strength, and heat deformation temperature.
Table 3. The crystallization parameters of PET and CEPET samples
Samples No. 1# 2# 3# 4# 5#
Tp (°C) 209.07 199.20 195.82 194.38 193.82
Tcc (°C) 0 130.2 132.4 135.31 137.8
Tm (°C) 257.54 253.35 252.56 251.88 250.74
ΔHcc(J.g-1) 0 1.16 2.60 3.66 4.01
ΔHm(J.g-1) 38.92 38.02 36.68 35.73 34.66
Xc (%) 27.80 26.31 24.34 22.91 21.89

78 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

Dynamic Shear Rheology of CEPET


The knowledge of the rheological behavior of PET during processing was a
key point in understanding the melt microstructure affected by the processing
conditions, determining optimal processing conditions, and better controlling
the quality of the final foaming products. The critical role of rheology in PET
processing was illustrated in Figure 10.

Figure 10. The role of rheology in PET processing. The solid line means “influences”
and the dash line means “gives information about”

Shear rheological test was employed to study the linear visco-elasticity of


polymers. Shear rheological behavior was related and sensitive to the length of
molecular chain and topological structure. When adding PMDA into PET, the
segment length and topological structure of PET chains would be changed,
and thus the viscoelasticity of PET would be changed, too.
To analyze the rheological behavior of CEPET, three techniques were used.
They were plots of the storage modulus (G’), the log complex melt viscosity
(η*) and loss factor (tan δ) versus various frequency. The improvement of melt
elasticity was reflected by the increment of G’. The bigger the G’, the better
the melt elasticity, thus the higher the melt strength. As shown in Figure 11.
At the low frequency zone, four G’-ω curves for the samples (2-5#) were higher
than unmodified PET, which implied that the melt elasticity of CEPET was
enhanced by adding PMDA. This phenomenon also indicated that CEPET had
a longer relaxing process due to the formation of branching and crosslinking
structure of PET. Hingmann and Marczinke reported the similar phenomenon
about long chain branching PP [41]. The increment of G’ and the decline of the
slope of G’ at the low frequency region were signatures of branching polymer
and the increment of melt elasticity. The shear rheological behaviors of the
samples (1-5#) suggested that the longer relaxation mechanism existed in the
low frequency region by increasing PMDA content. The melt elasticity of PET

Cellular Polymers, Vol. 34, No. 2, 2015 79


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

was improved by the introduction of branching structure and the broadening


of molecular weight distribution, which could maintain the stability of cell wall
during cell growth.

Figure 11. Relationships between storage modulus of various PET samples and
angular frequency

Figure 12 illustrated the relationship of complex melt viscosity and angular


frequency of the samples. In Figure 12, the h* of CEPET samples was obviously
higher than that of unmodified PET. Compared with other samples, the h* of
the samples (4# and 5#) had a steep slope and no Newtonian plateau at low
frequency. The increasement of the h* of CEPET could be attributed to two
reasons: the high molecular weight and the existence of branching structure.
The latter was thought to be the main reason because of flow restrictions of
CEPET chain. High h* value was expected to prevent cell rupture during cell
growing [42].
Loss factor (tan d) referred to the angle that the strain lagged behind the stress
while the PET melt was exposed in an alternating stress field. The smaller
the tan d, the faster the elastic response of the melt, thus the higher the melt
elasticity was. The tan d curves of samples (1-5#) were depicted in Figure 13.
In Figure 13, it was obvious that the tan d curves of CEPET were much smaller
than unmodified PET in low frequency zone, which indicated that the melt
elasticity of CEPET samples were superior to that of unmodified PET.

80 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

Figure 12. Relationships between complex viscosity of various PET samples and
angular frequency

Figure 13. Relationships between loss factor of various PET samples and angular
frequency

Cellular Polymers, Vol. 34, No. 2, 2015 81


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

Foaming Behavior of CEPET


In order to further study the effect of the content of PMDA and the chain
extension reaction time of CEPET on its foaming behavior, unmodified PET
and CEPET foams were prepared by batch foaming method using supercritical
CO2 as blowing agent. Figure 14 presented SEM micrographs of 1-5# foaming
samples with various PMDA contents in magnification ratio of 400x. It could
be observed that the cellular morphology of PET was improved gradually with
increasing PMDA content.

Figure 14. SEM micrographs for unmodified PET and CEPET at different chain
extender content

82 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

Unmodified PET could not be foamed due to its poor melt elasticity. As seen
from Table 4, with the PMDA content increasing from 0.25 phr to 1.0 phr,
the corresponding cell size increased gradually from 26.56 to 102.31 μm,
the expansion ratio was also increased sharply from 4.93 to 31.78, and the
cell density was decreased from 1.85×108 to 5.28×107 cells/cm3. The main
reason for that was the melt strength of samples increase gradually due to the
introduction of branching/crosslinking structure with increasing PMDA content.
For another reason, the small bubbles appeared, as they got closer to each
other, they combined together to form larger bubbles in foaming process. As
above stated, the crystallization peak shifted towards lower temperatures
with increasing PMDA content, which could provide enough time to merge
bubbles. The cellular structure of CEPET foams (2#) exhibited irregular oval
cellular structure, compared with 3-5# samples. This phenomenon should be
due to the lower branching degree of PET which could not provide enough
melt strength for cell growth. The SEM micrographs (3-5#) exhibited the
dodecahedron cellular morphology of CEPET with the PMDA content more
than 0.5 phr.

Table 4. Cell morphology data of unmodified PET and CEPET samples


Foaming samples Cell size Density Expansion Cell density
No. (μm) (g/cm3) ratio (cells/cm3)
1# - 1.38 1 -
2# 26.56 0.284 4.93 1.85×108
3# 61.51 0.097 14.35 9.67×107
4# 76.67 0.052 27.84 7.97×107
5# 102.31 0.045 31.78 5.28×107

Cell size distribution was a parameter to show the cell number of different size
range. It has been used by many researchers to investigate the relationship
between porous structure and foams properties, especially the impact strength
and thermal conductivity [43]. In Figure 15, the cell size of CEPET showed
a broad distribution. For 2# sample, the cell size of most cells was in the
region of 20-30 μm. With the increasing PMDA content, the cell size increased
gradually. Taken the PMDA content of 0.75 phr as an example, its smaller cell
size were around 40 μm and its larger cell was more than 120 μm. For the
PMDA content of 1.0 phr, the size of the smaller cell was around 60 μm and
the size of the larger cell was above 150 μm. So, the cell size had a relatively
broad cell size distribution than that of other samples.
Another aspect should be addressed that the chain extension reaction time
would affect the degradation of PET. Thus, in order to further clarify the
influence between the chain extension reaction time and the degradation of

Cellular Polymers, Vol. 34, No. 2, 2015 83


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

Figure 15. Cell size distribution of 2#-5# foams

CEPET, processing torque, the intrinsic viscosity, rheological properties, and


foaming behavior of various CEPET samples were measured at different chain
extension time. As mentioned the part of solution viscosity measurement,
some crosslinking structure would generate in the CEPET samples when the
content of the chain extender was higher than 0.75 phr. In order to avoid the
generation of crosslinking structure during the processing, the new CEPET
samples with PMDA content of 0.75 phr were studied at different chain
extension reaction time (T-10 min, T-15 min, T-30 min, and T-60 min).
The torque curves of PET mixing with PMDA-0.75 in the Haake rheometer
at different chain extension reaction time were illustrated in Figure 16. The
torque values were accomplished by the same weight of PMDA (0.75 phr)
at 280°C with a screw rotation speed of 40 rpm. As expected, there was a
decrease in the torque of PET after 10 min of mixing time, and with reaction time
increasing, the torque curves of PET decreased gradually. The decreasement
of the torque curves should due to the occurrence of degradation.

84 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

Figure 16. Torque curves of unmodified PET and CEPET

The intrinsic viscosity of various PET samples (from T-10 min to T-60 min) at
different solution concentration was shown in Figure 17. The extrapolation
of ηiner to concentration limit tending to zero gave intrinsic viscosity. It could
be seen from Table 5, the intrinsic viscosity of PET samples decreased
gradually with the increasement of the reaction time of CEPET. During the
chain extension of PET, with the increasement of the reaction time, the thermal
degradation reaction would become more and more intense, which may cause
the decreasement of the molecular weight and melt strength of CEPET.

Table 5. Intrinsic viscosity of PET samples


Samples ASTMD 4603 (30°C)
T-10 min T-15 min T-30 min T-60 min
[η] (dl/g) 1.54 1.48 1.16 0.86

Dynamic Shear Rheology of CEPET


As shown in Figure 18, at the low frequency zone, the G’-ω curves of the
sample (T-10 min) was higher than these of the samples (T-15 min, T-30 min,
and T-60 min), which implied that the melt elasticity of CEPET decreased
with the increasing chain extension reaction time of CEPET. The main reason
was that during the chain extension of PET, with the increasement of reaction

Cellular Polymers, Vol. 34, No. 2, 2015 85


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

Figure 17. The extrapolation of ηiner to obtain the intrinsic viscosity of each PET sample

Figure 18. Relationships between storage modulus of various PET samples and
angular frequency

86 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

time, the thermal degradation reaction would become more and more intense,
which caused the decreasement of the melt strength of CEPET.
In Figure 19, with the increasing reaction time, the h* of CEPET samples
decreased gradually, which meant that the sample with smaller chain extension
reaction time had better melt strength to prevent cell rupture during the bubble
growing stage.

Figure 19. Relationships between complex viscosity of various PET samples and
angular frequency

In Figure 20, it was obvious that the tan d of T-10 min sample was much
smaller than those of T-15 min, T-30 min, and T-60 min, especially in low
frequency zone, which indicated that the melt elasticity of T-10 min sample
was superior to that of T-15 min, T-30 min, and T-60 min samples, and the melt
elasticity of T-15 min, T-30 min, and T-60 min samples decreased gradually
with increasing the chain extension reaction time of CEPET.

Foaming Behavior of CEPET`


Figure 21 presented the SEM micrographs of the CEPET foams (from T-10 min
to T-60 min) at different chain extension reaction time in magnification ratio

Cellular Polymers, Vol. 34, No. 2, 2015 87


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

Figure 20. Relationships between loss factor of various PET samples and angular
frequency

Figure 21. SEM micrographs for v PET and CEPET foams at different chain extension
time

88 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

of 400x. With the chain extension reaction time increasing from 10 min to
30 min, the cell size of CEPET foam was roughly decreased from 78.1 to 32.97
μm, the cell density was increased from 8.28×107 to 4.19×108 cells/cm3. The
main reason was that the molecular weight of the PET samples decreased
gradually with increasing reaction time due to the degradation reaction,
which would induce the increasement of the crystallization temperature [44,
45]. This meant that the time for the growth of cell became less, leading to
the increasement of cell density and expansion ratio. For the sample of T-60
min, the cellular structure exhibited irregular cellular structure, bigger cell size,
and non-uniform cell size distribution, compared with the CEPET foams (T-15
min and T-30 min). This was because the molecular chain scission of CEPET
became serious with the increasement of reaction time of CEPET, leading to
very poor melt strength.
In Figure 22, for the CEPET foams (T-10 min and T-15 min), the size of most
cells was almost in the region of 40-100 μm. The cell size of the CEPET foam
(T-30 min) was a relatively narrower cell size distribution, compared with the
CEPET foams (T-10 min and T-15 min). For the higher reaction time of 60 min,

Figure 22. Cell size distribution of various foams (from T-10 min to T-60 min)

Cellular Polymers, Vol. 34, No. 2, 2015 89


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

the porous structure of the CEPET foam (T-60 min) also became non-uniform
with the smaller cell size around 60 μm and the larger cell size more than
180 μm. The foaming parameters of the CEPET samples (from T-10 min to
T-60 min) were shown in Table 6.

Table 6. Cell morphology data of CEPET samples (from T-10 min to


T-60 min)
Foaming samples Cell size (μm) Density Expansion Cell density
No. (g/cm3) ratio (cells/cm3)
T-10 min 78.10 0.049 28.11 8.28×107
T-15 min 70.98 0.065 21.23 2.99×108
T-30 min 32.97 0.105 13.64 4.19×108
T-60 min 108.96 0.608 2.27 5.45×106

CONCLUSIONS
In this paper, the effect of the CE content and the chain extension reaction
time on the intrinsic viscosity, thermal and rheological properties and foaming
behavior of PET using PMDA as CE had been studied. The results demonstrated
that CEPET possessed a branching and/or crosslinking structure, which was
beneficial to improve its melt viscoelasticity and foamability.
The thermal properties of CEPET samples were affected by the molecular
architecture of PET. The intrinsic viscosity of PET increased gradually with the
increasing content of PMDA. Compared with unmodified PET, the crystallization
temperature, the melting temperature, and crystallinity of CEPET decreased
with the content of PMDA. CEPET melt exhibited an improved storage modulus
and a faster melt elasticity response.
The cellular morphology and expansion ratio of CEPET foams were also improved
by the introduction of PMDA. Unmodified PET could not be foamed due to its
poor foam-ability. PET foam with the finest cell morphology and the highest
expansion ratio could be obtained by adding the proper content of PMDA. The
expansion ratio of CEPET with the PMDA content of 1.0 phr would reach 31.78.
With the increasement of chain extension reaction time, the thermal degradation
of PET would be more and more serious. As a result, the intrinsic viscosity
of PET decreased gradually, and the viscoelasticity and the foamability of
CEPET would become poorer by the thermal degradation reaction. The
expansion ratio of PET with the PMDA content of 0.75 phr for 10 min was the
highest of 28.11. When the chain extension reaction time reached 60 min, the
expansion ratio of CEPET was as low as 2.27, which demonstrated that the
chain extension reaction time was very important for the foam-ability of PET.

90 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

ACKNOWLEDGEMENTS
This work was supported by Chinese Ministry of Science and Technology
[grant number 2013EG111219].

REFERENCES

1. Sorrentino L., E. Di Maio, and S. Iannace. “Poly (ethylene terephthalate)


foams: correlation between the polymer properties and the foaming process”.
Journal of applied polymer science, 116(1) (2010) 27-35.

2. Guan Rong, Biqin Wang, and Deping Lu. “Preparation of microcellular poly
(ethylene terephthalate) and its properties”. Journal of applied polymer
science, 88(8) (2003) 1956-1962.

3. Japon Sonia, Yves Leterrier, and Jan-Anders E. Månson. “Recycling of poly


(ethylene terephthalate) into closed cell foams”. Polymer Engineering and
Science, 40(8) (2000) 1942-1952.

4. Larsen Kari. «Recycling wind turbine blades”. Renewable energy focus, 9(7)
(2009) 70-73.

5. Stewart, Richard. “Sandwich composites excel at cost-effective, lightweight


structures”. Reinforced Plastics, 55(4) (2011) 27-31.

6. Yavari, A., et al. “Effect of transesterification products on the miscibility and


phase behavior of poly (trimethylene terephthalate)/bisphenol A polycarbonate
blends”. European polymer journal, 41(12) (2005) 2880-2886.

7. Fraisse, Frederic, et al. “Recycling of poly (ethylene terephthalate)/


polycarbonate blends”. Polymer degradation and stability, 90(2) (2005) 250-
255.

8. Nguyen Quoc Tuan, et al. “Molecular characterization and rheological


properties of modified poly (ethylene terephthalate) obtained by reactive
extrusion”. Polymer Engineering and Science, 41(8) (2001) 1299-1309.

9. Dagli Sanjiv S., and Kunal M. Kamdar. “Effects of component addition


protocol on the reactive compatibilization of HDPE/PET blends”. Polymer
Engineering and Science, 34(23) (1994) 1709-1719.

10. Kalfoglou N.K., and D.S. Skafidas. “Compatibility of blends of poly (ethylene
terephthalate) with the ionomer of ethylene-methacrylic acid copolymer”.
European polymer journal, 30(8) (1994) 933-939.

11. Daver Fugen, Rahul Gupta, and Edward Kosior. “Rheological characterisation
of recycled poly (ethylene terephthalate) modified by reactive extrusion”.
Journal of Materials Processing Technology, 204(1) (2008) 397-402.

Cellular Polymers, Vol. 34, No. 2, 2015 91


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

12. S. Japon L. Boogh, Y. et al. “Reactive processing of poly(ethylene terephthalate)


modified with multifunctional epoxy-based additives”. Polymer, 41(15) (2000)
5809-5818.

13. Haiming Liu, Xiangdong Wang et al. “Reactive modification of poly(ethylene


terephthalate) and its foaming behavior”. Cellular polymers, 33(4) (2014)
189-212.

14. Bikiaris Demetris N., and George P. Karayannidis. “Chain extension of


polyesters PET and PBT with two new diimidodiepoxides. II.” Journal of
Polymer Science Part A: Polymer Chemistry, 34(7) (1996) 1337-1342.

15. Pauly, S., Permeability and diffusion data, in Polta, in Polymer Handbook, 4th
ed., Brandrup, J., Immergut, E.H., and Grulke, E.A., Eds., John Wiley and
Sons, New York, (1999) chap. 6.

16. Carlsson D.J., Day M., Suprunchuk T., Wiles D.M. “Pyrolysis of poly(ethylene
terephthalate) fibers: Characterization of involatile residues.” Journal of
applied polymer science, 28(2) (1983) 715-724.

17. Montaudo G., Puglisi C., Samperi F. “Chemical reactions which occur in
the thermal treatment of polycarbonate/poly(ethylene terephthalate) blends,
investigated by direct pyrolysis mass spectrometry.” Polymer Degradation
and Stability. 31(3) (1991) 291-326.

18. Montaudo G., Puglisi C., Samperi F. “Primary thermal degradation


mechanisms of PET and PBT.” Polymer Degradation and Stability, 42(1)
(1993) 13-28.

19. Botelho G., Queiro`s A., Liberal S., Gijsman P. “Studies on thermal and thermo-
oxidative degradation of poly(ethylene terephthalate) and poly(butylene
terephthalate)”. Polymer Degradation and Stability, 74(1) (2001) 39-48.

20. Filippo Samperi, Concetto Puglisi et al. “Thermal degradation of poly(ethylene


terephthalate) at the processing temperature”. Polymer Degradation and
Stability, 83(1) (2004) 3-10.

21. FirasAwaja, FugenDaver and Edward Kosior. “Recycled Poly(ethylene


terephthalate) Chain Extension by a Reactive Extrusion Process”. Polymer
Engineering and Science, 44(8) (2004) 1579-1587.

22. Milgrom J. “Polyethylene Terephthalate (PET).” Carl Hanser Verlag, Plastics


Recycling: Products and Processes (Germany), (1992) 45-72.

23. American Society for Testing and Materials-ASTM. D4603-03: Standard test
method for determining inherent viscosity of poly(ethylene terephthalate)
(PET) by glass capillary viscometer. West Conshohocken: American Society
for Testing Materials (2003).

24. American Society for Testing and Materials-ASTM. D2857: Standard practice
for dilute solution viscosity of polymer. West Conshohocken: American
Society for Testing Materials (2007).

92 Cellular Polymers, Vol. 34, No. 2, 2015


The Preparation and Characterization of Branching Poly(ethylene terephthalate) and its Foaming
Behavior

25. Mark H., Whitby G.S. eds. “Scientific Progress in the Field of Rubber and
Synthetic Elastomers. (Advances in Colloid Science).” New York: Interscience,
2 (1946) 197-251.

26. Sanches N.B., Dias M.L., Pacheco E.B.A.V. “Comparative techniques for
molecular weight evaluation of poly(ethylene terphthalate) (PET)”. Polymer
Testing, 24(6) (2005) 688-693.

27. Veleirinho B., Rei M.F., Lopes-DA-Silva J.A. “Solvent and concentration
effects on the properties of electrospun poly (ethylene terephthalate) nanofiber
mats”. Journal of Polymer Science Part B: Polymer Physics, 46(5) (2008)
460-471.

28. Incarnato L., et al. “Structure and rheology of recycled PET modified by
reactive extrusion”. Polymer, 41(18) (2000) 6825-6831.

29. Karagiannidis Panagiotis G., Anagnostis C. Stergiou, and George P.


Karayannidis. “Study of crystallinity and thermomechanical analysis of
annealed poly (ethylene terephthalate) films”. European Polymer Journal,
44(5) (2008) 1475-1486.

30. Mehta Aspy, Umesh Gaur, and Bernhard Wunderlich. “Equilibrium melting
parameters of poly (ethylene terephthalate)”. Journal of Polymer Science:
Polymer Physics Edition, 16(2) (1978) 289-296.

31. Xiangdong Wang, Wei Liu et al. “Study on the effect of dispersion phase
morphology on porous structure of poly(lactic acid)/poly(ethylene terephthalate
glycol-modified) blending foams”. Polymer, 54(21) (2013) 5839-5851.

32. Liu Jianye, et al. “Preparation and rheological characterization of long chain
branching polylactide”. Polymer, 55(10) (2014) 2472-2480.

33. Fortunato, G., et al. “Development of poly-(ethylene terephthalate)


masterbatches incorporating highly dispersed TiO2 nanoparticles:
Investigation of morphologies by optical and rheological procedures”.
European Polymer Journal, 57 (2014) 75-82.

34. Naguib Hani E., et al. “Strategies for achieving ultra low-density polypropylene
foams.” Polymer Engineering and Science, 42(7) (2002) 1481-1492.

35. Zhai Wentao, et al. “Cell structure evolution and the crystallization behavior
of polypropylene/clay nanocomposites foams blown in continuous
extrusion.” Industrial and Engineering Chemistry Research, 49(20) (2010)
9834-9845.

36. J.R. Atkinson, F. Biddlestone, J.N. Hay. “An investigation of glass formation
and physical ageing in poly(ethylene terephthalate) by FT-IR spectroscopy”.
Polymer, 41(18) (2000) 6965-6968.

37. Ziyu Chen, J.N. Hay. “The thermal analysis of poly(ethylene terephthalate) by
FTIR spectroscopy”. Thermochimica Acta, 552 (2013) 123-130.

Cellular Polymers, Vol. 34, No. 2, 2015 93


Haiming Liu, Xiangdong Wang, Hongfu Zhou, Wei Liu, and Bengang Liu

38. Breno Heins Bimestrea, Clodoaldo Saron. “Chain Extension of Poly (Ethylene
Terephthalate) by Reactive Extrusion with Secondary Stabilizer”. Materials
Research, 15(3) (2012) 467-472.

39. I. Karamancheva, V. Stefov et al. “FTIR spectroscopy and FTIR microscopy


of vacuum-evaporated polyimide thin films”. Vibrational Spectroscopy, 19(2)
(1999) 369-374.

40. Patrizio Raffa, Maria-Beatrice Coltelli, et al. “Chain extension and branching
of poly(ethylene terephthalate) (PET) with di- and multifunctional epoxy or
isocyanate additives: An experimental and modelling study”. Reactive and
Functional Polymers, 72(1) (2012) 50-60.

41. Hingmann, R., and B.L. Marczinke. “Shear and elongational flow properties
of polypropylene meltsa)”. Journal of Rheology, 38(3) (1994) 573-587.

42. Liu Wei, et al. “Effect of Chain Extension on the Rheological Property and
Thermal Behaviour of Poly (lactic acid) Foams”. Cellular Polymers, 32(6)
(2013) 343-367.

43. Li D.C., Liu T., Zhao L., Yuan W.K. “Foaming of linear isotactic polypropylene
based on its non-isothermal crystallization behaviors under compressed
CO2”. Polymer, 52(1) (2011) 89-97.

44. Gan Zhihua, Hideki Abe, and Yoshiharu Doi. “Crystallization, melting, and
enzymatic degradation of biodegradable poly (butylene succinate-co-14 mol
ethylene succinate) copolyester”. Biomacromolecules, 2(1) (2001) 313-321.

45. Qian ZhiYong, et al. “Synthesis and in vitro degradation study of poly (ethylene
terephthalate)/poly (ethylene glycol)(PET/PEG) multiblock copolymer”.
Polymer degradation and stability, 83(1) (2004) 93-100.

94 Cellular Polymers, Vol. 34, No. 2, 2015

You might also like