You are on page 1of 18

Article

Journal of Thermoplastic Composite


Materials
Composites of Recycled 25(6) 747–764
! The Author(s) 2011

PET Reinforced with Reprints and permissions:


sagepub.co.uk/journalsPermissions.nav

Short Glass Fiber DOI: 10.1177/0892705711412816


jtc.sagepub.com

NML Mondadori1, RCR Nunes2,


LB Canto3 and AJ Zattera1

Abstract
Composites of recycled poly(ethylene terephthalate) (PET) and short glass fibers (GFs)
in different compositions (0, 20, 30, and 40 wt% of GFs) with optimized microstructures
and high mechanical performance were obtained through melt processing. Composites
showed appropriate dispersion and distribution and suitable bonding of the GF through-
out the PET matrix, using either recycled bottle-grade PET in its degraded form or in a
solid-state polymerized (SSP) form and GFs treated with either aminosilane or epox-
ysilane coupling agents. The high level of reinforcement of these PET/GF composites
was confirmed by comparison of the experimental elastic modulus values of PET/GF
composites with theoretical ones obtained using the Halpin–Tsai model. An important
aspect highlighted by this study is that although it has been stated in the literature that
this is only possible with the use of twin-screw extruders, these PET/GF composites
with optimized microstructure and with high mechanical performance were com-
pounded using a single-screw extruder with a double-flight barrier screw. In general,
slightly better mechanical strengths were achieved for the composites based on the SSP
recycled PET, which may be associated with its highly entangled amorphous phase
arising from its higher molecular weight.

Keywords
PET, recycling, composite, short glass fiber

1
Chemical Engineering Department (DENQ), Caxias do Sul University (UCS), Caxias do Sul, RS 95070-560,
Brazil
2
Institute of Macromolecules (IMA), Federal University of Rio de Janeiro (UFRJ), Rio de Janeiro, RJ 21945-970,
Brazil
3
Materials Engineering Department (DEMa), Federal University of São Carlos (UFSCar), São Carlos,
SP 13565-905, Brazil

Corresponding author:
AJ Zattera, Chemical Engineering Department (DENQ), Caxias do Sul University (UCS), Caxias do Sul, RS
95070-560, Brazil
Email: ajzatter@ucs.br

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


748 Journal of Thermoplastic Composite Materials 25(6)

Introduction
The potential for the application of recycled bottle-grade poly(ethylene terephthal-
ate) (PET) is often limited due to the deterioration of its properties caused by
thermomechanical degradation during its reprocessing.1
Solid-state polymerization (SSP) has been used as an alternative to overcome
this problem and broaden the possibilities for the use of recycled PET. This process
consists of heating PET granules at temperatures around 220 C–230 C (between its
glass transition temperature and its crystalline melting temperature) for 10–30 h
under low pressure under the agitation of the solid in a hot inert gas environment.
In these conditions, the PET molecular weight is increased through transesterifica-
tion/polycondensation and esterification reactions, with a consequent improvement
in the PET properties.1–3
The use of short glass fibers (GFs) for the preparation of composites based on
recycled PET is another important aspect that places the recycled PET on the
market as molded artifacts with greater properties as compared to the recycled
PET without reinforcement.4–13 There is a very close and interdependent relation-
ship between the processing conditions, the microstructure, and the final properties
of these composites. The microstructure, and thus the final properties, of PET/GF
composites is governed by the physicochemical characteristics of both the PET and
GF in addition to their processing history. Compounding of recycled PET with
GFs has been successfully achieved using both single-screw4 and twin-screw4–6
extruders followed by injection molding. A previous study4 demonstrated that a
single-screw extruder with a double-flight barrier screw can be used to produce
composites of recycled bottle-grade PET containing GFs with mechanical and
thermomechanical performance similar to composites processed in an intermesh
co-rotating twin-screw extruder. Moreover, our previous study showed that the
properties of PET/GF composites are highly dependent on the injection molding
conditions used. The degree of reinforcement of the PET/GF composites is depen-
dent on the aspect ratio (fiber length to diameter, l/d) and volume fraction of the
GFs in the composite, fiber orientation, and distribution through the polymeric
matrix, and the adhesion between the GFs and the polymeric bulk, which is depen-
dent on the fiber–matrix interfacial chemistry. GFs surface functionalized with
amine groups are most commonly used for the preparation of PET/GF compos-
ites.4,10 Another important aspect governing the reinforcement of PET/GF com-
posites is the degree of crystallinity of the PET, which is influenced by the molding
conditions.4
The aim of this study was to develop composites of recycled PET with short GFs
in different compositions (0, 20, 30, and 40 wt% of GFs) through compounding in a
single-screw extruder with a double-flight barrier screw followed by injection mold-
ing. A recycled bottle-grade PET, a SSP recycled bottle-grade PET, and two types of
GFs with different chemical surface functionalities (amine and epoxy) were used. The
effects of these parameters on the microstructure and mechanical properties (tensile,
flexural, and impact strength (IS)) of PET/GF composites were evaluated.

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


Mondadori et al. 749

Experimental
Materials
A recycled PET in flake form (PET-flake), which is a washed and crushed waste
originating from soft drink bottles, was supplied by Sul PET Company (Brazil).
The PET-flakes, as received, were irregular in shape and had low bulk density
(0.289 g cm3), presenting nonfreeflowing behavior in the extruder feed hopper.
For this reason, the PET-flakes were agglutinated prior to the extrusion processing,
which converted them into crystallized granules with a bulk density of 0.443 g cm3
and with a free-flowing behavior in the extruder feed hopper. The SSP recycled
PET in crystallized pellet form (PET-ssp), also originating from soft drink bottles,
was supplied by Bahia PET Company (Brazil) and used as received.
The intrinsic viscosities (IVs) of the PET samples (Table 1) were determined by
the dilute solution viscosity method (ASTM D2857) using an Ubbelohde 1C vis-
cometer in a solution of phenol and 1,1,2,2-tetrachloroethane (60/40 wt/wt%) at
30 C. IVs were approximated from the data obtained at a single solution concen-
tration (0.5 g cm3) by the following equation: IV ¼ 0:25ðrel  1 þ 3 ln rel Þ=c,
where rel is the relative viscosity and c is the solution concentration. The number
average molecular weights (Mn) of the PET samples were calculated based on the
IV values through the following relationship: Mn ¼ 3:29  104 ðIVÞ1:54 .
The short GF compounds with two different chemical treatments were supplied
by Owens Corning Vetrotex Company, Brazil, under codes EC 983 and EC 952.
These are E-glass roving of 4.5-mm chopped filaments with fiber diameter of 10 mm
and were designed with different sizing (coupling agents): EC 983 is a GF treated
with an aminosilane coupling agent, whereas EC 952 is treated with an epoxysilane
coupling agent. In this study, these materials were named GF-amine and GF-
epoxy, respectively. Scheme 1 shows how GF surfaces are typically functionalized
with amine or epoxy groups through reaction with aminotrialkoxysilane or epox-
ytrialkoxysilane coupling agents, respectively. Functionalization comprises hydro-
lysis of alkoxy groups and subsequent condensation with silanol groups occurring
on the glass surface to form siloxane bond between the GF and coupling agent, so
that the desired functional group is attached on the GF surface.

Table 1. IV and number average molecular weight (Mn) of as-received and processed PET
samples.

Sample IV (dL g1) Mn (g mol1)

PET-flake 0.74 20,706


PET-flake processed 0.58 14,262
PET-ssp 0.81 23,954
PET-ssp processed 0.69 18,493

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


750 Journal of Thermoplastic Composite Materials 25(6)

functionalized RO Si OR
trialkoxysilano

OR

+ 3 H2O - 3 ROH hydrolysis

functionalized HO Si OH
silanol

OH

OH
silanol group on
glass fiber surface

condensation /
- H2O
functionalization

functional group on
HO Si OH
glass fiber surface

O
Functionalities (F): NH2 (amine) CH CH2 (epoxy)

Scheme 1. Schematic drawing showing the functionalization of GF surfaces with amine or epoxy
groups through reaction with functionalized trialkoxysilane coupling agents.

Processing
Prior to each processing step, the materials were dried in an air-circulating oven at
80 C for 24 h.
PET/GF composites with different compositions (0, 20, 30, and 40 wt% of GFs)
were compounded in a single-screw extruder with a double-flight barrier screw
(L/D = 32 and D = 35 mm). The single-screw configuration is shown in

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


Mondadori et al. 751

Figure 1. Single-screw configuration.

Figure 1. The extruder was operated with a barrel temperature profile of 250 C–
275 C (from hopper to dye) and screw speed of 50 rpm. The materials were formu-
lated by introducing all of the components simultaneously into the extruder hopper
during a single processing step. The flow rate was around 8 kg h1 under these
conditions. The extruded strands were quenched in a water bath and pelleted.
Specimens for mechanical tests (ASTM standard bars) were injection molded
from the extruded pellets using a HIMACO LH 150-80 machine at a barrel tem-
perature of 280 C, holding pressure of 650 bar for 4 s, mold cooling time of 45 s,
and mold temperature of 120 C. The dosage and plasticating step was conducted
under mild conditions, that is, at low-screw rotation speed and low-screw back
pressure, in order to minimize GF breakage and thus to preserve the microstructure
of the composites generated by extrusion.

Characterization
The lengths of the GF in the molded PET/GF composites were determined by
optical microscopy using image analyzer software. The fibers were recovered
from the injection-molded impact bars by burning off the PET at 600 C for 3 h.
The fibers were distributed over a glass plate with the aid of a 1:1 solution of
distilled water and ethanol and then left to stand until the total evaporation of
the solution. Images were collected at a magnification of 50 using a Nikon optical
microscope, model Epihot 200, coupled to a Nikon digital camera DMX 1200 F.
Around 300 fibers were counted for each composite using Image-Pro Plus software
specifically designed for the analysis of images. The data were compiled into his-
tograms in order to determine the number average fiber length (Ln), weight average
fiber length (Lw), and the polydispersity (P). The calculations were carried out
using Equations (1), (2), and (3), where Ni is the number of fibers with length i.
The critical fiber length for composite reinforcement, (l/d)c, was deduced using
Equation (4), which expresses the balance of force between the tensile strength
of the fiber ( f) and the fiber–matrix interfacial shear strength ( int):14

Ni Li
Ln ¼ ð1Þ
Ni

Ni L2i
Lw ¼ ð2Þ
Ni Li

Lw
P¼ ð3Þ
Ln

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


752 Journal of Thermoplastic Composite Materials 25(6)

 
l f
¼ ð4Þ
d c 2int

The microstructure of the PET/GF composites was investigated using a Jeol JSM
6060 scanning electron microscope (SEM), examining the top of cryofractured
injection-molded impact bars and the top of the fractured surfaces after tensile
tests. The fractured surfaces were coated with gold using a diode sputtering coater
before observation by SEM.
Tensile tests were performed on type I ASTM bars using an EMIC DL 2000 uni-
versal testing machine at a crosshead speed of 50 mm min1, according to ASTM D
638. The reinforcement degree achieved was evaluated through comparative analysis
between the experimental results for the elastic modulus and the theoretical values
obtained using the Halpin–Tsai model.15 According to the Halpin–Tsai model, the
theoretical tensile modulus of a fiber-reinforced composite can be expressed in terms of
the corresponding properties of the matrix and the fiber phase together with their
proportions and the fiber geometry, using Equations (5) and (6):

Ec 1 þ f
Er ¼ ¼  ð5Þ
Em 1  f

Ef =Em  1
¼  ð6Þ
Ef =Em þ 

In Equations (5) and (6), Er is the relative modulus and Ec, Em, and Ef are the
moduli of the composite, matrix, and fiber, respectively; f is the fiber volume
fraction. The factor  is an empirical constant that describes the influence of the
fiber geometry. For oriented short fibers with aspect ratios higher than the critical
value (l/d)c, the factor  is two times the critical value for fibers oriented parallel to
the mechanical loading and was equal two for fibers oriented perpendicular to the
mechanical loading. Therefore, the modulus is maximum when the fibers are ori-
ented parallel to the mechanical loading and attain the minimum value when the
fibers are oriented perpendicularly.
Flexural tests (three-point bending) were carried out according to ASTM D 790,
with an EMIC DL 2000 analyzer, at a crosshead speed of 2 mm min1.
Izod IS was measured with a CEAST Resil 25 pendulum using notched speci-
mens, according to ASTM D 256.
The bars for the mechanical tests were conditioned at 23 C and 50% relative
humidity for 48 h prior to the tests, which were all conducted under the same
climatic conditions. Each mechanical test value was calculated as the average of
at least five-independent measurements. The standard deviations of each value
were calculated and are shown as error bars in the plots.
The degrees of PET’s crystallinity in the injection-molded bars were measured
through differential scanning calorimetry (DSC). Scans from 20 C to 300 C with a

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


Mondadori et al. 753

heating rate of 10 C min1 and under N2 atmosphere (50 mL min1) were carried
out in a Shimadzu DSC-50 calorimeter. Samples (ca. 10 mg) were taken from the
inner part of the injection-molded Izod bars. The degrees of PET’s crystallinity (Xc)
was calculated as follows:

Hm
Xc ¼  100% ð7Þ
H0m 

where  is the PET weight fraction in the composites, Hm is the melting enthalpy,
which was calculated from the DSC curves, and is the heat of fusion of 100%
crystalline PET (120 J g1).16

Results and Discussion


The values for the IVs and the number average molecular weights (Mn) for the
as-received PET samples (PET-flake and PET-ssp) and those melt-processed PET
samples (extrusion followed by injection molding) under the conditions adopted
in this study are given in Table 1. The results show that the molecular weight of
as-received PET-ssp is greater (16%) than that of the as-received PET-flake, as
a result of the SSP process to which the first was submitted. The molecular
weight of as-received PET-ssp is similar to the commercial bottle-grade virgin
PET.1 After the melt processing, there was a decrease in the molecular weights of
both samples, with reductions of 23% for the PET-ssp and 31% for the PET-
flake. Furthermore, the molecular weight of the processed PET-ssp was only 10%
smaller than that of the as-received PET-flake. This shows that the levels of
PET’s degradation under the processing conditions adopted in this study are
within acceptable values, especially if one considers that the thermomechanical
degradation of PET is pronounced when it is processed in the molten state.17
Moreover, the use of PET-ssp can be considered a way to circumvent the deg-
radation that normally occurs when bottle-grade recycled PET is reprocessed in
the molten state.
Table 2 shows the average lengths of the GFs in the molded composites,
obtained by optical microscopy and with the aid of an image analyzer.
Composites with increasing GF content present shorter fiber lengths. The breaking
up of the GFs during melt processing with polymers occurs basically through two
types of interaction: strain induced on the fibers through shear flow of the molten
polymer and strain resulting from the fiber–fiber interaction (collisions and fric-
tion).13 Thus, the higher viscosity and the higher probability of fiber–fiber collisions
for the composites with increasing GF content resulted in more fiber break up and
shorter fiber lengths. The length of the fibers in the composites varied between 90
and 700 mm, resulting in Ln values between 215 and 304 mm and Lw values between
275 and 384 mm, which corresponds to aspect ratios (l/d) of 21.5–30.4, based on the
Ln values, and of 27.5–38.4 based on the Lw values, considering that the GFs are 10
mm in diameter.

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


754 Journal of Thermoplastic Composite Materials 25(6)

Table 2. Short GF length distributions of injection molded PET/GF composites.

Number average Weight average Polydispersity,


Sample GF length, Ln (mm) GF length, Lw (mm) Lw/Ln

PET-flake/GF-amine 80/20 304 384 1.27


PET-flake/GF-amine 70/30 241 306 1.27
PET-flake/GF-amine 60/40 221 265 1.20
PET-ssp/GF-amine 80/20 369 457 1.24
PET-ssp/GF-amine 70/30 297 329 1.11
PET-ssp/GF-amine 60/40 223 266 1.19
PET-flake/GF-epoxy 80/20 260 331 1.27
PET-flake/GF-epoxy 70/30 226 291 1.29
PET-flake/GF-epoxy 60/40 215 275 1.28

The critical aspect ratio (l/d)c of the GF for effective reinforcement of the PET/
GF composites was calculated using Equation (4). A GF length above the critical
value (l/d)c is essential for ensuring effective stress transfer from the PET matrix to
the reinforcement as the composites are mechanically overloaded. Assuming a
perfect bonding between the GF and the PET matrix, the interfacial shear strength
is that of the PET matrix, which is 45 MPa.18 Taking a typical value for the GF
strength of 1500 MPa19 and considering that the GFs are 10 mm in diameter, a
critical fiber length of 167 mm is obtained for the PET reinforcement.
Thus, the average GF lengths (Table 2) of the PET/GF composites under study,
compounded in a single-screw extruder with a double-flight barrier screw followed
by injection molding, are around 1.3–1.8 times and 1.7–2.3 times the critical value
for PET, based on the Ln and Lw values, respectively. These are typical values for
composites of recycled PET with GFs compounded in a twin-screw extruder fol-
lowed by injection molding.4–6
The SEM technique was used to determine the distribution of GFs in the PET
matrix. The images for the composites PET-flake/GF-amine, PET-ssp/GF-amine,
and PET-flake/GF-epoxy with 30 wt% of GFs are shown in Figure 2(a–c), respec-
tively. SEM images of all composites show that the GFs are homogeneously dis-
tributed throughout the PET matrixes with no evidence of bundles. In general, it
was not possible to observe significant differences between the microstructures of
the composites prepared with the different types of PET and GFs. Homogeneous
mixing of the GFs in the PET matrix is another characteristic required to obtain
satisfactory mechanical properties for the composites.
The high degree of dispersion and even distribution observed for the mixing of
GFs into PET matrices evidence the high performance of the process used in this
study to prepare PET/GF composites, in agreement with our previous findings.4
The double-flight barrier divides the main screw channel of the extruder, so that the
melt pool is continuously separated from the solid bed along the screw length. This

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


Mondadori et al. 755

Figure 2. Scanning electron micrographs of cryofractured surfaces of injection-molded PET/GF


composites with 30 wt% of GFs: (a) PET-flake/GF-amine, (b) PET-ssp/GF-amine, and (c) PET-flake/
GF-epoxy.

melt separation enlarges the area of heat conduction between the solid bed and the
barrel wall, which increases the overall melting capacity and the melt temperature
homogeneity. This increases the residence time in which GFs remain in mixture
with molten PET along the screw length. Besides, the melt pool can flow over the
barrier flight through the space (clearance) between the barrier tip and the barrel
surface. All these mixing mechanisms improve the dispersion and distribution of
GFs in the PET matrix during compounding in the single-screw extruder with a
double-flight barrier.
Another aspect revealed by the SEM images in Figure 2(a–c) is the composite
anisotropy, that is, the GFs are mainly oriented in the direction parallel to the flow
of the mold filling, since the SEM images were taken of surfaces fractured perpen-
dicular to the mold-filling direction.
The degrees of crystallinity (Xc) of PET’s composites are shown in Figure 3.
Small increases in these values were observed for composites with increasing
amounts of GF, suggesting that GF acts as a nucleating agent for PET, in addition
to its role in reinforcement. The nucleation effect of the GFs on the PET crystal-
lization in the composites is noteworthy, since recycled PET has a low crystalliza-
tion rate.20 For all of the PET/GF composites with different GF contents, the
degree of crystallinity of the PET was not affected by either the PET molecular
weight or the GF surface functionality. The invariance of the degree of crystallinity

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


756 Journal of Thermoplastic Composite Materials 25(6)

Figure 3. Degree of crystallinity of PET/GF composites.

of PET in the composites despite the differences in molecular weight (PET-flake vs.
PET-ssp) is in agreement with reports in the literature, which states that the degree
of crystallinity does not have a strong influence on the molecular weight for the
magnitude of PET molecular weights used in this study.21
Figure 4 shows the tensile modulus values for the PET/GF composites. The
tensile modulus showed a linear increase with GF content in the composites. For
similar GF content in the composites, practically no difference in the tensile mod-
ulus values was found with regard to the PET molecular weight and the GF surface
functionality. The flexural modulus values (Figure 5) also followed the same trend.
The relative tensile modulus values for the PET/GF composites under study are
shown in Figure 6(a–c), along with the theoretical values obtained according to
the Halpin–Tsai model (Equations (5) and (6)), assuming the tensile modulus of the
GF to be 70 GPa.19 The experimental relative modulus represents the gain in the
composite properties in relation to the polymeric matrix. As can be seen in Figure 6
(a–c), the experimental relative tensile modulus values for all of the composites are
very close to the theoretical maximum value, in which the fibers are considered to
be oriented parallel to the mechanical loading. This mechanical performance sug-
gests a high level of GF orientation in the molded parts and strong bonding
between the GF and the PET matrix. These data show an effective reinforcement
in all the composites under study, which is in agreement with the good distribution
of the GFs in the composites (Figure 2) and the optimized GF size distributions in
the composites (Table 2).
Since the modulus is almost an elastic property, it is not expected to be depen-
dent on the interfacial strength unless there is a significant slippage at the fiber–

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


Mondadori et al. 757

Figure 4. Tensile modulus of PET/GF composites.

Figure 5. Flexural modulus of PET/GF composites.

matrix interface. Thus, in this study, the modulus is practically unaffected by the
PET molecular weight and the GF functionality. Similar behavior in relation to the
differences in PET molecular weight has been found for short GF-reinforced PET
composites.10 Also, a similar effect, based on fiber functionalities, has been found
for short GF-reinforced polyamide 6 composites.22

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


758 Journal of Thermoplastic Composite Materials 25(6)

Figure 6. Relative tensile modulus of PET/GF composites along with the predicted values
obtained according to the Halpin–Tsai model: (a) PET-flake/GF-amine, (b) PET-ssp/GF-amine,
and (c) PET-flake/GF-epoxy.

The values for the tensile and flexural strength are shown in Figures 7 and 8,
respectively. It can be observed that for the neat PET samples and composites
containing GF-amine, these values are slightly higher for PET-ssp-based compos-
ites. The effect of the GF functionality in the composite with PET-flake is almost
negligible, although higher values for these properties were noted for the GF-con-
taining epoxy groups (GF-epoxy), particularly in the flexural strength values.
Figure 9 shows SEM micrographs of the fracture surfaces obtained after tensile
tests for the composites PET-flake/GF-amine (Figure 9(a)) and PET-flake/GF-
epoxy (Figure 9(b)) with 30 wt% of GFs. It can be seen for both composites that
the GFs are well bonded to the surrounding PET matrix and that fiber breakage
dominates over fiber pullout in the tensile tests. This reveals suitable interfacial
adhesion in the composites using both GF functionalities, that is, amine and epoxy,
which is in agreement with their mechanical performance. The effectiveness of the
GFs containing amine functionalities in promoting good adhesion in composites
with PET prepared by melt processing has been previously studied.4,23 This has
been attributed to the grafting of PET chains onto GF surfaces during processing
by chemical reaction of the PET functionalities (carboxyl end groups) and the

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


Mondadori et al. 759

Figure 7. Tensile strength of PET/GF composites.

Figure 8. Flexural strength of PET/GF composites.

amine groups of the GF. On the other hand, the reaction between epoxy groups of
the GF and PET end groups has been shown only for PET composites with long
GFs prepared by in situ solid-state polycondensation.24 In this study, we provide
evidence that short GFs functionalized with epoxy groups are also efficient in
promoting suitable interfacial adhesion with the recycled PET matrix in composites

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


760 Journal of Thermoplastic Composite Materials 25(6)

Figure 9. Scanning electron micrographs of fractured surfaces obtained after tensile tests for
PET/GF composites with 30 wt% of GFs: (a) PET-flake/GF-amine and (b) PET-flake/GF-epoxy.

prepared by melt processing. Scheme 2 shows the grafting reactions between PET
carboxyl end groups and the amine or epoxy functions occurring on the GF sur-
face, which can occur during PET/GF melt processing. These reactions create a
chemical bond between the PET matrix and the GF substrate resulting in strong
physical interaction.
Notched Izod ISs of the composites are shown in Figure 10. The IS increases
with GF content. This increase is due to fiber-related energy dissipation mecha-
nisms, such as fiber debonding, pullout, bridging, and fracture, which induce plas-
tic deformation of the polymeric matrix before failure. Bridging and fiber fracture
are likely to occur as a consequence of a set of GFs with length longer than the
critical value for effective reinforcement (Equation (4)), while debonding and fiber

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


Mondadori et al. 761

O O

PET C OH H 2N PET C NH

O O O OH
PET C OH CH2 CH PET C O CH2 CH

Scheme 2. Schematic drawing showing grafting reactions between PET carboxyl end groups and
the amine or epoxy functions occurring on the GF surface.

Figure 10. Notched Izod IS of PET/GF composites.

pullout are expected to occur as the result of a set of GFs with length shorter than
the critical value (Table 2). Composites based on PET-ssp with higher molecular
weight showed slightly higher values for IS. Pegoretti and Penati10 have shown the
same trend for PET/GF composites based on PET matrices with different molec-
ular weights as a result of hygrothermal aging. Conversely, there were no changes
in this property as a function of the different GF surface chemical treatments.
The composite strength and impact values show notable viscoelastic properties,
that is, the PET matrix undergoes elastic and plastic deformation during the tests.
Therefore, these properties are expected to be dependent on the PET matrix yield

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


762 Journal of Thermoplastic Composite Materials 25(6)

behavior as well as the composite interfacial strength. Since all PET-/GF-molded


parts showed similar microstructures (Table 2; Figure 2) and degrees of crystallin-
ity (Figure 3), as well as suitable interfacial adhesion (Figure 9), the PET-ssp may
have given the best results due to its highly entangled amorphous phase arising
from its higher molecular weight after melt processing (PET’s molecular weights
are shown in Table 1). Thus, these composites can undergo greater deformation as
they are mechanically overloaded, resulting in better properties of tensile, flexural,
and IS.
The values for the modulus (Figures 4, 5, and 6), tensile strength (Figure 7),
flexural strength (Figure 8), and IS (Figure 10) for the composites based on
recycled PET developed in this study represent a set of properties related to stiff-
ness, strength, and toughness. In general, these values are of the same magnitude as
those reported in the literature for composites of virgin PET reinforced with con-
ventional short GFs treated with aminosilane coupling agent compounded in a
twin-screw extruder.25 It is interesting to note that we have obtained PET/GF
composites with optimized microstructures and with high-mechanical performance
using a single-screw extruder, with a suitable screw configuration. This contradicts
literature reports, which have stated that such composites can only be obtained
through the use of twin-screw extruders. Furthermore, we show that the use of
recycled PET in the SSP form can lead to slightly better strength values for the
PET/GF composites. In addition, it was shown that GFs treated with epoxysilane
coupling agents are as effective as conventional aminosilane-treated GFs for the
mechanical reinforcement of recycled PET matrices.

Conclusions
Composites made of recycled PET and short GFs in different compositions (0, 20,
30, and 40 wt% of GFs) with optimized microstructures and high mechanical per-
formance were obtained in this study. The procedure involved compounding in a
single-screw extruder with a double-flight barrier followed by injection molding,
using either recycled bottle-grade PET in its degraded form or in a SSP form and
GFs treated with either aminosilane or epoxysilane coupling agents. The addition
of this reinforcement led to the optimized microstructure of the composites, with
appropriate dispersion and distribution of the GFs throughout the PET matrix,
and with suitable degree of interfacial adhesion between the PET matrix and the
GFs. The high level of reinforcement of the recycled PET afforded by the incor-
poration of GFs was confirmed by comparison of the experimental elastic modulus
values for the PET/GF composites with theoretical values obtained using the
Halpin–Tsai model. An important aspect highlighted by this study is that these
PET/GF composites, with optimized microstructures and good mechanical perfor-
mance, may be compounded using a single-screw extruder with a suitable screw
configuration, even though it is stated in the literature that this is only possible
employing twin-screw extruders. In general, slightly better mechanical strength was
achieved for the composites based on SSP recycled PET, which may be associated

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


Mondadori et al. 763

with its highly entangled amorphous phase arising from its higher molecular
weight. Both types of GF functionalities (amine and epoxy) were found to be
effective in ensuring a suitable degree of interfacial adhesion between the PET
matrix and the GFs providing composites with good mechanical performance.

Acknowledgment
The authors thank FAPERGS-Brazil for the grant awarded to N.M.L. Mondadori.

References
1. Awaja F and Pavel D. Recycling of PET. Eur. Polym. J. 2005; 41: 1453–1477.
2. Pang K, Kotek R and Tonelli A. Review of conventional and novel polymerization
processes for polyesters. Prog. Polym. Sci. 2006; 31: 1009–1037.
3. Zhao J, Xiao H, Qiu G, Zhang Y, Huang N and Tang Z. Solid-State Polycondensation
of Poly(Ethylene Terephthalate) Modified with Isophthalic Acid: Kinetics and Simulation.
Polymer 2005; 46: 7309–7316.
4. Mondadori NML, Nunes RCR, Zattera AJ, Oliveira RVB and Canto LB. Relationship
Between Processing Method and Microstructural and Mechanical Properties of
Poly(Ethylene Terephthalate)/Short Glass Fiber Composites. J. Appl. Polym. Sci.
2008; 109: 3266–3274.
5. Giraldi ALFM, Bartoli JR, Velasco JI and Mei LHI. Glass Fibre Recycled Poly(ehylene
terephthalate) Composites: Mechanical and Thermal Properties. Polym. Test. 2005; 24:
507–512.
6. Giraldi ALFM, Jesus RC and Mei LHI. The influence of extrusion variables on the
interfacial adhesion and mechanical properties of recycled PET composites. J. Mater.
Process. Technol. 2005; 162/163: 90–95.
7. Cantwell WJ. The Facture Behavior of Glass Fiber/Recycled PET Composites, Journal of
Reinforced Plastics and Composites. J. Reinf. Plast. Compos. 1999; 18: 373–387.
8. Ronkay F and Czigany T. Development of Composites with Recycled PET Matrix.
Polym. Adv. Technol. 2006; 17: 830–834.
9. Fung KL and Li RKY. Mechanical Properties of Short Glass Fibre Reinforced and
Functionalized Rubber-toughened PET blends. Polym. Test. 2006; 25: 923–931.
10. Pegoretti A and Penati A. Recycled poly(ethylene terephthalate) and its short glass fibres
composites: effects of hygrothermal aging on the thermo-mechanical behaviour. Polymer.
2004; 45: 7995–8004.
11. Takahashi K and Choi NS. Influence of Fibre Weigh Fraction on Failure Mechanisms of
Poly(Ethylene Terephthalate) Reinforced by Short-Glass-Fibers. J. Mater. Sci. 1991; 26:
4648–4656.
12. Kwok KW, Choy CL and Lau FP. Elastic Moduli of Injection-Molded Short-Glass-
Fiber-Reinforced Poly(Ethylene Terephthalate). J. Reinf. Plast. Compos. 1997; 16:
290–305.
13. Folkes MJ. Short Fibre Reinforced Thermoplastics. New York: Research Studies Press,
1985.
14. Kelly A and Tyson WR. Tensile properties of fibre-reinforced metals: cooper/tungsten and
copper/molybdenum. J. Mech. Phys. Solids 1965; 13: 329–350.
15. Halpin JC and Kardos JL. The Halpin-Tsai Equations: A Review.
Polym. Eng. Sci. 1976; 16: 344–352.

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016


764 Journal of Thermoplastic Composite Materials 25(6)

16. Metha A and Wunderlich B. J. Equilibrium melting parameters of poly(ethylene ter-


ephthalate). Polym. Sci. Polym. Phys. 1978; 16: 289–296.
17. Spinace MAS and De Paoli MA. Characterization of poly(ethylene terephtalate) after
multiple processing cycles. J. Appl. Polym. Sci. 2001; 80: 20–25.
18. Luethi B, Reber R, Mayer J, Wintermantel E, Janczak-Rusch J and Rohr L. An energy-
based analytical push-out model applied to characterize the interfacial properties of knitted
glass fibre reinforced PET. Compos. Part A. 1988; 29(A): 1553–1562.
19. Chawla KK. Composite Material Science and Engineering, 2nd edn. New York:
Springer-Verlag, 1998.
20. Viana JC, Alves NM and Mano JF. Morphology and mechanical properties of injection
molded poly(ethylene terephthalate) Polym. Eng. Sci. 2004; 44: 2174–2184.
21. Loyens W and Groeninckx G. Rubber toughened semicrystalline PET: influence of the
matrix properties and test temperature. Polymer. 2003; 44: 123–136.
22. Laura DM, Keskkula H, Barlow JW and Paul DR. Effect of glass fiber surface chemistry
on the mechanical properties of glass fiber reinforced, rubber-toughened nylon 6. Polymer.
2002; 43: 4673–4687.
23. Cheng H, Tian M and Zhang L. Toughening of recycled poly(ethylene terephthalate)/
glass fiber blends with ethylene-butyl acrylate-glycidyl methacrylate copolymer and maleic
anhydride grafted polyethylene-octene rubber. J. Appl. Polym. Sci. 2008; 109: 2795–2801.
24. Yan W, Han K, Zhou H and Yu M. Study on the grafting of PET onto the glass fiber
surface during in situ solid-state polycondensation. J. Appl. Polym. Sci. 2006; 99: 775–781.
25. Arencon D and Velasco JI. The Influence of Injection-Molding Variables and Nucleating
Additives on Thermal and Mechanical Properties of Short Glass Fiber/PET Composites. J.
Thermoplast. Compos. Mater. 2002; 15: 317–336.

Downloaded from jtc.sagepub.com at PENNSYLVANIA STATE UNIV on September 16, 2016

You might also like