You are on page 1of 14

Fabrication of Long Glass Fiber Reinforced Polyacetal

Composites: Mechanical Performance, Microstructures,


and Isothermal Crystallization Kinetics

Zhaozeng Tao,1 Yatao Wang,1,2 Jianhua Li,2 Xiaodong Wang,1 Dezhen Wu1
1
State Key Laboratory of Organic-Inorganic Composites, Beijing University of Chemical Technology, Beijing
100029, China
2

Hebei Provincial Engineering Technology Research Center of Coal-based Materials and Chemicals,
Tangshan, Hebei 063018, China

A thermoplastic pultrusion was carried out to prepare


the long fiber reinforced thermoplastic (LFT) composites based on polyacetal (POM) matrix on the customdesigned pultrusion equipment. The investigation on
mechanical performance revealed that the POM-based
LFT composites achieved much higher tensile, flexural,
and impact strength than the short glass fiber reinforced ones at the same fiber loadings. Such a promising reinforcement effect is attributed to the feature
that the residual fiber length in the injection-molded
LFT products is greatly superior to that in short fiber
reinforced ones. This takes full advantage of the
strength of the reinforcing fiber itself. The scanning
electronic microscopy demonstrated that the fiber
fracture and fiber pull-out concurred on the tensile and
impact fracture surfaces, and the former preceded the
latter. The isothermal crystallization kinetics of the
POM-based LFT composites was also intensively studied, and the results indicated that the crystallinity of
POM domain was enhanced by the heterogeneous
nucleation of glass fiber, but the crystallization rate
was postponed due to the interspace restriction
toward crystalline growth caused by long glass fiber.
These kinetic parameters provided information on the
processing conditions of POM-based LFT composites
for the injection and compression molding. POLYM.
C
COMPOS., 00:000000, 2014. V
2014 Society of Plastics
Engineers

INTRODUCTION
Short glass fiber reinforced thermoplastic composites
as one of the most important organicinorganic compo-

Correspondence to: Xiaodong Wang; e-mail: wangxdfox@aliyun.com


Contract grant sponsor: National Natural Science Foundation of China;
contract grant number: 51173010; contract grant sponsor: National Key
Basic Research Program; contract grant number: 2012CB720304.
DOI 10.1002/pc.23090
Published online in Wiley Online Library (wileyonlinelibrary.com).
C 2014 Society of Plastics Engineers
V

POLYMER COMPOSITES2014

sites have been available for many years. Such a type of


composite material can offer many desirable properties
such as the high rigidity and high strength to weight ratio
in comparison to metals, improved creep resistance, low
mold shrinkage, reduced coefficient of thermal expansion,
superior wear resistance, good damping capacity, and
easy processability to fabricate any complex 3D shapes
[1, 2]. Although short glass fibers are predominantly used
as reinforcement in thermoplastics, and however, the use
of short fiber reinforced thermoplastics in industrial applications is frequently hindered by their rather low mechanical performances due to the low residual fiber length
after processing [3]. In this case, the full strength of the
reinforcing short fibers is not realized due to their low
fiber aspect ratio. Long fiber reinforced thermoplastic
(LFT) technologies have become well developed over the
past decade and consequently resulted in the establishment of a class of high-performance engineering materials
for structural applications [4]. This is ascribed to the fact
that the aspect ratio [ratio of fiber length (l) to diameter
(d)] of fibers in LFT composites is an order of magnitude
greater than that of a short fiber, often exceeding l/d of
2,000 and, thus, takes full advantage of the strength of
the reinforcing fiber [5]. Currently, the LFT materials
have been used especially in the automotive industry to
reduce both costs and vehicle weight, because LFTs can
significantly reduce weight compared to metal, improving
fuel efficiency and reducing emissions [6]. Using LFTs to
replace metal allows suppliers to integrate parts, thus
greatly reducing assembly costs in parts. Moreover, with
the processability of the injection and compression molding, LFTs also show a great superiority in both mechanical strength and design flexibility over the thermoset
composites like sheet molding compound and bulk molding compound [7]. Apart from their excellent mechanical
performance, good economic benefits, and low weight,

the biggest advantage of the LFT composites lies in a


high level of productivity in processing, which is a result
of the short cycle times attainable with thermoplastic
matrix systems [8]. Other positive aspects include high
plant availability, consistently high quality, the wide
range of possibilities for integration of different functions
in a single component, and the fact that the system costs
can also be lowered [9]. These benefits have resulted in a
prompt and constant growth for LFTs at a much higher
rate than with conventional composite materials.
The LFT technologies could be classified as direct
LFT processes (LFT-D) and LFT granules (LFT-G). The
former is the in-line or direct compounding of LFTs, in
which the processor compounds the glass fibers, polymers, and other additives in-line with the molding or part
extrusion process. The latter is the processing of semifinished LFT materials in granule or pellet form, which are
provided as bundles of fibers preimpregnated with a
matrix polymer [10]. With the combined advantages of
LFT-D and short fiber reinforced thermoplastics, the
LFT-G represents the state-of-the-art and is suitable for
both the classic injection molding process and the injection compression molding process to fabricate the parts
with complicated shape and structure. Such a sort of LFT
composites have recently received much attention and are
finding ever-growing applications due to their excellent
short- and long-term mechanical performances compared
to their challengers [11, 12]. Long glass fibers are commonly used as the reinforcing materials for the fabrication
of LFT-G composites, although other long fibers such as
carbon, aramids, and natural fibers are used in specialty
applications [13, 14]. These composites are usually manufactured through a pultrusion process, where continuous
filaments of glass fiber rovings pass through a thermoplastic resin impregnation unit. Thus, the glass fibers are
completely enveloped by this material and then subsequently cut into pellets of a certain desired length around
1025 mm [15]. In the view of the commercial point, the
thermoplastics for the process of LFT-G are most commonly limited to polypropylene (PP) and nylon 6 and 66
[16, 17]. Other applicable thermoplastics include polyethylene terephthalate (PET), poly(butylene terephthalate),
acrylonitrile-butadiene-styrene copolymer (ABS), highdensity polyethylene, polyphenylene sulphide, and thermoplastic polyurethane (TPU) [18]. Globally, about 65%
of the LFTs market is PP-based, and nylon 6 and 66 have
about 20% of market share, while other resins such as
PET, ABS, and high temperature polyamides comprise
the remaining 15% [3].
The scientific literature shows that there are numerous
studies given to the LFT materials on their fabrication
with a variety of polymeric matrices, and more investigations focus on the effects of fiber length and content,
processing condition, and thermoplastic species on the
mechanical performance and microstructures of the LFTG injection mould. Thomason [19] investigated the influence of fiber length and concentration on the properties
2 POLYMER COMPOSITES2014

of long glass fiber reinforced PP and found that the


mechanical strength and impact resistance exhibited a
maximum in performance in the 4050 wt% of fiber content range. They also reported the study on the influence
of fiber length, diameter, and concentration on the impact
performance of long glass fiber reinforced nylon 66 [20].
Teixeira et al. [21] investigated the effect of flow restriction on the microstructure and mechanical properties of
long glass fiber reinforced polyamide 66 composites for
automotive applications. Carlsson and Astrom [22]
designed a special machine for pultrusion of thermoplastic composites and investigated the pultrusion of long
glass fiber reinforced PP composites. Kim et al. [23]
established a model for thermoplastic pultrusion process
for LFT composites and found that the experiment data
for PP/long glass fiber composites is in good agreement
with the prediction by this model. Chen and Ma [24]
studied the processability and mechanical properties of
pultruded LFT composites based on poly(methyl methacrylate)/TPU interpenetrating systems. Cilleruelo et al. [25]
reported the processing method of long glass fiber reinforced PET composites and highlighted the effects of the
processing conditions and additives on their mechanical
properties. Ramani et al. [26] also reported the compression molding and pultrusion of LFT poly(ether ketone
ketone)-based composites through a powder impregnation
technique. Miller et al. [27] and Luisier et al. [28]
reported the processing and molding method of the LFT
composites using nylon 12 as a polymeric matrix. It is no
doubt that there are a great deal of potential work dealing
with the LFT composites due to the variables and characteristics of polymeric matrix species, which can lead to a
possibility to achieve a series of high performance.
Polyacetal (POM) is considered as one of the most
important engineering thermoplastics due to its excellent surface lubrication, outstanding antifatigue performance, high electrical insulation, and good chemical
and weathering resistance [29]. It can also be widely
used in injection moulded and extruded parts and products for automotive, mechanical, and electronic applications. With a rapid growth of POM usage in the
worldwide thermoplastic marketplace, the development
of high performance POM-based composites and compounds has received a great interest [30, 31]. It is
believed that the LFT technologies performing upon
POM can result in much better reinforcing effect in
comparison with the short fiber reinforcement and thus
can greatly extend the applications of the POM-based
composites. In this work, we make an attempt to fabricate the LFT POM-based composites and investigate
the mechanical performance and microstructures of
these composites under the reinforcement of long glass
fibers, and the effect of long fibers on the crystallization kinetics of POM was also studied intensively in
the isothermal processes. The aim of this work is to
develop a sort of novel high-performance POM-based
composites via LFT technology.
DOI 10.1002/pc

FIG. 1. Scheme of thermoplastic pultrusion for long glass fiber reinforced POM composites. [Color figure
can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

EXPERIMENTAL
Materials
The POM resin (commercial grade: BS270) used in
this work was commercially obtained from Shanghai
Bluestar New Chemical Materials. This product is a lowviscosity typed POM copolymer with a number average
molecular weight of 22,500 g/mol and a melt flow index
of 27.0 g/10 min. Direct-drawn glass fiber roving was
kindly supplied by Chongqing Polycamp International,
China. This product has a linear density of 2,400 tex and
a filament diameter of 14 lm, and the fibers have been
surface-treated with 1,2-ethylenebis(trimethoxysilane)
coupling agent. Tetrakis[methylene-b-(3,5,-di-tert-butyl-4hydroxyphenyl) propionate]methane and tris-(2,4-di-tertbutylphenyl)phosphite as antioxidant and thermal stabilizer were commercially obtained from Beijing Additive
Institute, China.
Processing of LFT Composites
The pultrusion equipment used in this work was
custom-designed and consisted of a fiber creel, a preheating chamber, an impregnation apparatus connecting with
a screw extruder, a wind cooler, a pulling machine, and a
custom tailored pelletizer as illustrated in Fig. 1. The
glass fiber roving was pulled by the pulling machine
through a preheating chamber and then pulled into the
impregnation apparatus over a number of rollers, and an
extruder provides a molten POM into the chamber of the
impregnation apparatus. The temperature in the chamber
of the impregnation machine was set to 175180 C, and
the pulling speed of the rovings was limited in the range
of 1015 m/s. Impregnation occurs during contact with
the rollers, where the fiber roving is more or less spread
out and open up to receive the molten POM. Another
important effect is that pulling the fiber rovings over the
roll, in the presence of the viscous molten POM, causes
DOI 10.1002/pc

local high shear rates between the bundle and the bar,
resulting in high shear forces. These forces oppose the
moving direction of the rovings and are perceived as a
viscous drag on the rovings. The perceived total viscous
drag increases with increasing bundle speed and with an
increasing number of rollers and of course equalizes the
pulling force on the fiber rovings. After a wind cooling,
the pultruded braces were cut into pellets at a length of
1012 mm by a custom-tailored pelletizer. The resulted
pellets were dried under vacuum at 80 C overnight and
then were stored in the sealed aluminum foil bags.
Characterization
Measurement of Mechanical Performance. All of the
pelletized samples were dried at 80 C in a vacuum oven
for 8 h prior to injection molding, and then were
injection-molded into the test bars with the different
shapes required for mechanical and heat-resistant measurements. Notched Izod impact strength was measured on
a SANS ZBC-1400A impact tester equipped with a pendulum of 2.75 J according to ISO 180 standard. The
impact test bars were notched on a milling machine to
achieve a V type notch in a depth of 2 6 0.2 mm. The
tensile and flexible properties were measured with a
SANS CMT-4104 universal testing instrument using a 10
kN load transducer according to ISO 527 and ISO 178
standards, respectively. The dumbbell tensile specimens
were fabricated with a dimension of 150 mm 3 10 mm
3 4 mm, and the tensile speed was set to 5 mm/min. All
the measurements were performed at a constant temperature of 23 C, and the reported values reflected an average
from five tests.
Polarized Optical Microscopy. Polarized optical
microscopy was performed to observe the distribution of
glass fibers and the crystalline morphology of LFT POMbased composites on an Olympus BX51 polarizing microscope equipped with a Linkam THMS 600 temperature
POLYMER COMPOSITES2014 3

controller and a Sony CCD-IRIS digital camera. For the


observation of fiber distribution, the fiber specimens
obtained from the injection molded bars were calcined in
the furnace at a temperature about 500 C for 5 h to
remove the POM matrix. The remained glass fibers were
measured to determine the fiber length distribution as
well as the actual weight fraction of fiber in the composite samples. The recovered fibers were dispersed in an
aqueous solution of ethanol and then were transferred to a
cover glass for microscopic observation. The fiber length
was determined by image analysis. As for the observation
of spherical morphology, the specimen was heated to
200 C on the hot-stage, held at this temperature for 3
min, and then cooled to a temperature of 168 C, where
the growing of spherulites started. The samples were held
at this temperature for 1 h so as to perform the isothermal
crystallization.

t 


dHc
dt
dt
0


Xt 5 1
3100%
dHc
dt
dt
0

(1)

where Hc is the crystallization enthalpy during the infinitesimal time interval dt.

Thermogravimetric Analysis. Thermogravimetric analysis (TGA) was performed in a nitrogen atmosphere using
a TA Instruments Q50 thermal gravimetric analyzer. Samples were placed in a platinum crucible and ramped from
room temperature to 750 C at a heating rate of 10 C/min
while a flow of nitrogen was maintained at 50 mL/min.

RESULTS AND DISCUSSION


Scanning Electron Microscopy. Scanning electron
microscopy (SEM) images for the fractography of LFT
POM-based composites were taken on a Hitachi H-4700
scanning electron microscope. The fracture surfaces
obtained from impact test bars after impact measurement
were made electrically conductive by sputter coating with
a thin layer of gold-palladium alloy. The images were
taken in high vacuum mode with 20 kV acceleration voltage and a medium spot size.

Differential Scanning Calorimetry. The crystallization


kinetics of LFT POM-based composites was studied in
terms of both nonisothermal and isothermal crystallization
behaviors using a TA instruments Q20 differential scanning calorimeter equipped with a Universal Analysis 2000
data station. All operations were performed under a nitrogen flow of 50 mL/min with a sample weight around 5
7 mg. All the samples were first heated to 180 C and
held at this temperature for 5 min to eliminate the effect
of the thermal and processing history. The differential
scanning calorimetry (DSC) scans for the isothermal crystallization of long glass fiber reinforced POM composites
were performed at a given crystallization temperature (Tc)
ranging between 145 and 151 C with an equal temperature interval of 2 C. The samples were heated to 180 C
at a rate of 10 C/min and rapidly cooled to a desired
crystallization temperature, and then, they were kept at
the same temperature until the crystallization finished.
The crystallization peak and melting temperatures were
directly read from the cooling and heating thermograms,
respectively. The curve integral was run to calculate the
enthalpies of overall crystallization behavior. The relative
degree of crystallinity (Xt) at the crystallization time (t)
(normalized with respect to the degree of crystallinity at
t 5 1) could be determined from the ratio of the exothermic peak area at t to the total area of this exothermic
peak through the following equation:
4 POLYMER COMPOSITES2014

Mechanical Performance
Figure 2 shows the mechanical properties of the
injection-molded specimens derived from POM-based LFT
composites, and the experimental data from short fiber reinforced POM composites are also presented in the figure as
a reference. It appears that these mechanical properties
achieved a significant improvement as a result of long glass
fiber reinforcement on POM. The tensile strength is shown
to increase rapidly with increasing the glass fiber content
and obtains an increment of 172% compared to pure POM
when the fiber content is increased to 40 wt%. Analog
trends are observed in the flexural strength and flexural
modulus, which are also found to increase by 325% and
266%, respectively, for the samples containing 40 wt% of
glass fibers. Although the fiber content of long glass fiber
reinforced thermoplastic composites could be increased over
60 wt% through the melt pultrusion process, the specimens
for mechanical tests must be injection-molded by the
injection-molding machine specially designed. In this case,
no further mechanical data were provided. However, it was
reported that the higher long fiber content could cause a further improvement in mechanical performance of the long
glass fiber reinforced thermoplastic composites [19], and the
POM-based LFT composites may follow this trend. It is
also notable from Fig. 2a and b that the POM/short glass
fiber composites present much lower tensile and flexural
properties in comparison with the POM/long glass fiber
ones, and the tensile strength of the composites containing
30 wt% of short glass fiber only increases by 72% compared to pure POM. This indicates the reinforcing effect of
short fiber on POM is inferior to that of long glass fiber. It
seems that the POM composites almost could not be processed through melt extrusion with addition of 40 wt% short
glass fiber because of the highly thermal sensitivity of POM
resin, which may result in a thermal degradation during the
melt process of its composites with high fiber content.
Therefore, the mechanical data were unavailable for the
DOI 10.1002/pc

FIG. 2. Mechanical parameters of POM-based composites as a function of fiber content: (a) tensile
strength, (b) flexural strength, (c) flexural modulus, and (d) notched Izod impact strength. [Color figure can
be viewed in the online issue, which is available at wileyonlinelibrary.com.]

composites containing the short fiber content higher than


30 wt%.
It is generally accepted that the reinforcement effect of
glass fiber reinforced thermoplastics is determined by fiber
content, fiber strength, fiber length, fiber orientation, interface
adhesion, and polymeric matrix nature. In this work, the fiber
length apparently plays a key role in the reinforced POM composites, and therefore, it is of great significance to know the
status of fiber length distribution in the injection-molded
specimens for the POM samples reinforced by the long glass
fiber and the short one. Figure 3 shows digital optical photographs of the features of the POM/glass fiber composites from
original pellets to injection-molded bars. It is interestingly
observed that, for the specimen injection-molded from the
POM/long glass fiber composite pellets, the residual glass
fibers almost retained the initial framework of the bar after
ashing in the furnace. However, the injection-molded bar
made from POM/short glass fiber compounding pellets shows
some powder-like residues after ashing. These results indicate
DOI 10.1002/pc

that the residual fiber length of the long glass fiber reinforced
POM composites is much longer than that of the short fiber
ones. The residual fiber length was also evaluated by observing on optical microscope and then was calculated statistically
by the Nano-Measurer software. Figure 4 shows the fiber
length distributions of these two injection-molded bars with
the corresponding optical images as insets. It should be noted
that the residual fiber lengths in the POM-based LFT composite are still longer than those in the short fiber reinforced one
after injection molding process. As observed from the long
glass fiber reinforced POM specimen, the most glass fibers
show the lengths ranging from 2.5 to 6.5 mm, which seems to
be reduced from 10 to 12 mm of the feedstock during the
injection molding on account of the strong shear effect of
injection screw at the plasticization stage of the composites.
As a reference, the short glass fiber reinforced POM specimen
injection molded with the same processing parameters only
shows a fiber length mostly distributed in the range of 0.1
0.4 mm. The number average fiber length could be calculated
POLYMER COMPOSITES2014 5

It is evident that the tensile strength of the POM/short


glass fiber composites results from the contribution of the
matrix and fibers, whose length is lower and closer to the
critical fiber length. However, for the POM-based LFT
composites, the number average fiber length is significantly greater than the critical fiber length, and thus, these
longer fibers as well as the matrix made contribution to
the composites tensile strength. Such a contribution from
the fibers with super critical fiber length is very prominent and results in a much higher reinforcing efficiency.
Usually, the interfacial shear stress between the fiber and
matrix is lower than the shear stress of matrix, and therefore, the POM composite system requires a much longer
critical fiber length to achieve an expected reinforcement
effect. In this case, the reinforcement derived from long
glass fiber toward POM resin is especially significant in
comparison with the short glass fiber.
It is noteworthy from the impact testing data in Fig. 2d
that, apart from the excellent reinforcement effect, the
long glass fiber exhibited a remarkable toughening effect
on POM, while the POM/short glass fiber composites

FIG. 3. Digital optical photographs of the features of the original pellets, injection-molded bars, and ashing specimens of (a) POM-based
LFT composite and (b) short glass fiber reinforced POM composite.
[Color figure can be viewed in the online issue, which is available at
wileyonlinelibrary.com.]

on the basis of the distribution diagram, and the results show


that the injection-molded specimen from the POM-based LFT
composite has a much longer average fiber length of 4.2 mm,
which is almost 20 times of the length for the short glass fiber
reinforced one. The KellyTyson model established a relationship between the ultimate tensile stress and fiber length for the
fiber-reinforced thermoplastic composites by the following
equation [32]:
ruc 5



X sli Vi X
lc
1rum 12Vf
1
rf Vj 12
df l >l
2lj
l <l
i

(2)

where rum and rf are the ultimate tensile strengths of


composite and fiber respectively, and ruc the matrix stress
at the failure of composite, Vf is the volume fraction of
the reinforcement, lc the critical fiber length, df the diameter of fiber, and s the interfacial shear stress between the
fiber and matrix in composites. According to the Kelly
Tyson model, the critical fiber length could be calculated
by the following equation [33]:
lc 5

rf df
2s

(3)

If the s is set to the shear stress of POM matrix as the


upper limit, the lc is calculated to be 0.224 mm according
to the parameters of glass fiber used in the current work.
6 POLYMER COMPOSITES2014

FIG. 4. Fiber length distribution curves of (a) POM-based LFT composite and (b) short glass fiber reinforced POM composite at the same
fiber contents of 30 wt%. Inset: optical micrographs of residual fibers.
[Color figure can be viewed in the online issue, which is available at
wileyonlinelibrary.com.]

DOI 10.1002/pc

FIG. 5. SEM micrographs of tensile fracture surfaces for POM-based LFT composites with (a, b) 10 wt%,
(c, d) 20 wt%, and (e, f) 40 wt% of long glass fiber and (g, h) for the POM composites containing 30 wt%
of short glass fiber.

only achieved a slight increase in notched Izod impact


strength. The impact strength of the composite with 40
wt% long glass fiber is found to improve by 110% comDOI 10.1002/pc

pared to pure POM. Such an evident toughness enhancement with respect to the fiber length and content may be
attributed to the fracture-transformation mechanism. It
POLYMER COMPOSITES2014 7

FIG. 6. SEM micrographs of impact fracture surfaces for POM-based LFT composites with (a, b) 20 wt%
and (c, d) 30 wt% of glass fiber.

was reported that the impact energy dissipation during the


fracture of the fiber reinforced thermoplastic composites
was generally dominated by several factors, including
matrix fraction, fiber debonding, friction between the
interfaces of matrix and fibers, fiber pullout, and fiber
breakage [34]. The Cottrell model can predict the impact
energy (Uc) for a fiber reinforced composite depending
on whether the fiber length (l) is subcritical or supercritical through the following equations [35, 36].
Uc 5Um 12Vf 1


  2 
Vf lUd
Vf l sf
1
;
df
6df

when

l < lc

(4)

  2   3 
Vf l2lc Uf
Vf lc Ud
Vf lc sf
Uc 5Um 12Vf 1
1
1
;
l
df l
6df l
when l > lc
(5)
where Um, Uf, and Ud are the fracture energy for matrix,
fibers, and interfaces, respectively, and sf is the interfacial
friction during fiber pull-out. For the composite systems
with a fiber length lower or close to the critical fiber
length like short fiber reinforced thermoplastic composites, the impact energy dissipation is estimated by Eq. 4,
where the three terms cover matrix fracture, fiber-matrix
debonding, and fiber pull-out. The fiber pull-out is the
8 POLYMER COMPOSITES2014

most important energy dissipation mechanism. As for the


LFT composite systems with a super critical fiber length,
the impact energy dissipation could be determined by
Eq. 5, where the four terms account for matrix fracture,
fiber debonding, fiber fracture, and fiber pull-out limited
to the critical fiber length. If the fiber is above critical
length, fiber breakage must occur for fiber pull-out. The
contribution of energy dissipation from matrix is typically
small in contrast to the reinforcement. Considering the
fact that the residual fiber length is far over the critical
fiber length for the long glass fiber reinforced POM composites prepared in this work, the fiber breakage should
occur for the fiber to be pulled out. The contribution of
energy dissipation from fiber fracture is typically greater
than that from the fiber pull-out in contrast to the reinforcement. In this case, the long glass fiber not only supplies a prominent reinforcement but also evidently gives
its greatest benefits to notched impact in the POM-based
LFT composites. The investigation on morphology of
impact fracture surface will provide an evidence of microscopic images in the latter section.

Morphology of Fracture Surface


The tensile and impact fracture surfaces of POM-based
LFT composites were investigated with SEM, and the
DOI 10.1002/pc

FIG. 7. Development of the relative degree of crystallinity as a function of crystallization time for (a) pure
POM and its LFT composites with (b) 10 wt%, (c) 20 wt%, and (d) 40 wt% of glass fiber in the isothermal
process. [Color figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

relevant micrographs are demonstrated in Figs. 5 and 6. It


is notable from the tensile fracture surfaces on Fig. 5 that
the fracture of the tensile specimens is predominantly due
to the fiber fracture and fiber pull-out. There are a lot of
fibers remaining on surfaces with the length exceeding
the aforementioned critical length for POM/glass fiber
composites, indicating that the fiber fracture precedes the
fiber pull-out. From the magnified SEM micrographs, the
debonded fiber surface looks rough and is clung with
some resin. This phenomenon is attributed to good fiber
wetting and coupling at the fibermatrix interface. Moreover, some resin was stripped from the fiber surface, indicating a high adhesion level for the matrix and fibers,
which enhanced the stress transfer as a result of the fiber
debonding through frictional forces along the interfaces
[37]. As for the POM composite containing 30 wt% of
short glass fiber, its SEM micrographs clearly reveal that
the fiber pull-out dominated the fracture, and few fiber
fractures could be observed as shown in Fig. 5g and h.
These results suggest that the reinforcing capability of
long glass fiber is superior to that of the short one, thus
causing a more significant reinforcing efficiency for POM
resin.
DOI 10.1002/pc

An analogical morphology is observed on the SEM


micrographs of the impact fracture surfaces of the POMbased LFT composites as shown by Fig. 6. These micrographs exhibit a feature of major fiber breakage with few
fibers pulled out. The fibers are homogeneously distributed in the matrix (see Fig. 6c), and these long fibers are
still embedded in the matrix when fiber breakage occurs.
The morphology of impact fracture surfaces further confirmed the contribution of the fiber fracture to the dissipation impact energy. Furthermore, the SEM micrographs of
fracture surfaces also show an evidence of POM matrix
cohesion to the fibers (see Fig. 6b and d), confirming a
good interfacial adhesion between the matrix and fibers.
These phenomena indicate that a mechanical interlocking
has been established between fibers and matrix, and thus,
the better stress transfer could be gained. The long fibers
may fracture if the fiber stress level exceeds the local
fiber strength. The fibers that have fractured away from
the crack interface are pulled out of the matrix, which
may also involve energy dissipation. In this case, the long
glass fiber reinforced POM composites show a much better impact toughness than the conventional short glass
fiber reinforced ones.
POLYMER COMPOSITES2014 9

FIG. 8. Avrami plots of log[ln(1 Xt)] vs. log t for the isothermal crystallization of (a) pure POM and its
LFT composites with (b) 10 wt%, (c) 20 wt%, and (d) 40 wt% of glass fiber in the isothermal process.
[Color figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

Isothermal Crystallization Kinetics


POM is a semicrystalline thermoplastic polymer with
a relatively high degree of crystallinity, and its processing and molding are actually influenced by its crystallization behavior, thus determining the properties of the
final products to some extent. Meanwhile, the introduction of alien matters also affects the crystallinity of such
a semicrystalline polymer in most cases. Especially, the
development of LFT composites concerns the manufacturing and molding processes, which are involved in the
solidification of POM from melt to crystals. Therefore,
it is of great importance to investigate the crystallization
kinetics of POM within its LFT composites, because the
industrial processing conditions strongly depend on its
isothermal crystallization behaviors with technological
importance in the processability of POM. During the solidification processing from the melt, POM forms a complex heterogeneous system composed of amorphous and
crystalline phases of a different order and hierarchical
crystalline structure; therefore, the mechanical performance of POM-based LFT composites is also sensitive to
the molding process, especially pronounced in injectionmolded devices [38]. The effects of long glass fiber on
10 POLYMER COMPOSITES2014

the crystallization behaviors of POM were quantitatively


analyzed through isothermal DSC dynamic scans. The
plots of the Xt versus the crystallization time (t) derived
from the DSC scans are shown in Fig. 7. It is clearly
observed that the incorporation of 10 wt% glass fiber
fairly accelerated the crystallization of POM domain and
thus shortened the crystallization time at the selected
temperatures, indicating that the heterogeneous nucleation of long glass fiber made the POM domain crystallize faster. However, the crystallization rate does not
visibly show a further improvement with increasing the
fiber content. It seems that the other factors dominated
the crystallization of POM domain at high fiber loading
apart from the nucleating effect. Furthermore, the isothermal crystallization rate also depends on the crystallization temperature and presents a lower value at a
higher crystallization temperature, indicating that the
crystalline growth of POM domain is mainly dominated
by nucleation [39].
It has been well known that the Avrami model makes
a convenient approach to analyze the isothermal crystallization kinetics for most of semicrystalline polymers,
because it establishes the relationship between the
DOI 10.1002/pc

TABLE 1. The Avrami parameters derived from the dynamic DSC


scans for the isothermal crystallization of POM-based LFT composites.

Samples
Pure POM

POM 1 10 wt%
glass fiber

POM 1 20 wt%
glass fiber

POM 1 30 wt%
glass fiber

POM 1 40 wt%
glass fiber

Tc
( C)

k
(1/minn)

t0.5
(min)

1/t0.5
(1/min)

134
138
142
144
146
134
138
142
144
146
134
138
142
144
146
134
138
142
144
146
134
138
142
144
146

2.02
2.04
1.74
1.48
1.73
1.76
1.74
2.02
1.93
1.78
1.77
1.66
1.51
1.49
1.73
2.19
2.05
2.13
1.44
1.68
1.78
1.64
1.43
1.41
1.34

8.90
6.76
3.43
1.54
0.31
8.13
5.49
4.78
2.23
0.47
7.13
4.71
3.14
1.81
0.42
10.28
5.37
3.00
1.43
0.63
5.60
3.14
3.12
1.85
0.47

0.28
0.3
0.37
0.48
0.56
0.25
0.29
0.32
0.41
0.43
0.27
0.29
0.37
0.35
0.49
0.26
0.25
0.26
0.39
0.63
0.33
0.31
0.32
0.33
0.46

3.57
3.33
2.70
2.08
1.79
4.19
3.45
3.125
2.44
2.33
3.70
3.45
2.70
2.86
2.04
3.85
4.05
3.85
2.56
1.59
3.03
3.23
3.13
3.03
2.17

development of relative degree of crystallinity and crystallization time as expressed by [40, 41]:
12Xt 5exp 2ktn

(6)

where Xt is the relative degree of crystallinity at the crystallization time t, n the Avrami exponent depending on
the nature of nucleation and the growth geometry of the
crystals, and k the crystallization kinetic constant involving both nucleation and growth rate parameters. Figure 8
shows the typical Avrami plots of pure POM and its LFT
composites. These plots do not show a well linear relationship in the whole range. However, they are found to
show a fairly good linearity at the primary and middle
crystallization stage but only to present a deviation at the
final crystallization stage, which may be ascribed to the
secondary crystallization at that stage. Therefore, the
Avrami model seemed to successfully fit the crystallization data of POM domain for the relative degree of crystallinity value up to 90 %. In this case, the Avrami
parameters n and k were obtained from the linear regression of the Avrami plots in the relative crystallinity range
of 590%, because the determination coefficients of linear
fits in this regime were greater than 0.92.
The isothermal crystallization kinetic parameters of
pure POM and its LFT composites were calculated from
their Avrami plots and summarized in Table 1. It is
known that the value of n strongly depends on both the
mechanism of the nucleation and the crystalline growth,
DOI 10.1002/pc

and that ideally n would be an integer. Pure POM has an


average value of n around 1.8, indicating that the homogeneous nucleation acts in a predominant way, and meanwhile, the crystallization may correspond to the circular
lamellar. This probably corresponds to a threedimensional growth with a combination of thermal and
athermal nucleation, resulting in the fractional values of n
observed. Nevertheless, the n value of POM domain in its
LFT composites varies in a broad range of 1.342.19 at
the given crystallization temperatures. This fact is a hint
that the crystalline growth of POM domain acts in different manners from the two-dimensional lamellar crystal,
nucleating thermally with diffusion-controlled rate to the
three-dimensional truncated spherulites as a result of heterogeneous nucleation [42]. Moreover, the transformation
of crystalline structures may occur at different stages
throughout the whole crystallization process; therefore,
the n value is also dependant on crystallization conditions
like crystallization temperature in most cases. The notable
variety of n indicates a concurrent action of various
nucleation mechanisms and the growth geometry of
spherulites during isothermal crystallization process. This
also reflects very complicated modes of two- and threedimensional (lamellar or spherulitic) crystal growth as
well as the one-dimensional growth of fibrils along with
fibers, nucleating thermally with diffusion-controlled rate
crystals from a three-dimensional truncated growth of
spherulites. The athermal nucleation implies that there is
no contribution from nucleation rate to the activation
energy [43, 44]. It is also interesting to note that the LFT
composite with 40 wt% of glass fiber showed a trend of
reduction in the n value. This may be due to the extension confinement during the growth of POM spherulites
resulting from the high density of fibers and spherulites.
In fact, the bulk crystallization rate is hardly compared
with one another only based on the crystallization rate
constant k. Therefore, the crystallization half-time (t0.5),
defined as the time required for a sample to reach an Xt
of 50%, or its reciprocal (1/t0.5) has been used to evaluate
the crystallization rate of diverse polymers [30], and it is
calculated by the following formula:


ln 2
t0:5 5
k

1=n
(7)

The advantage for using experimental t0.5 over the


model-predicted value is that the experimental value is
independent of the model validity range and related
assumptions [45]. The values of t0.5 and 1/t0.5 for pure
POM and its LFT composites are also summarized in
Table 1. The effects of crystallization temperature and
fiber content on overall crystallization rates are clearly
distinguished on the basis of the data listed in Table 1.
The POM-based LFT composites show slightly lower values of t0.5 (corresponding to higher values of 1/t0.5) at the
given crystallization temperatures compared to pure
POM. This result confirms that the crystallization of
POLYMER COMPOSITES2014 11

FIG. 9. Polarized optical micrographs of the crystalline morphology for (a) pure POM and its composite
containing (b) 20 wt% and (c, d) 30 wt% of glass fiber. [Color figure can be viewed in the online issue,
which is available at wileyonlinelibrary.com.]

POM domain could be accelerated due to the heterogeneous nucleating effect of glass fiber. However, it is also
observed that the values of t0.5 are almost not reduced
with the increase of fiber content, indicating that the
space hindrance from long fibers may prevent the crystalline growth and thus extends the crystallization time. This
factor evidently dominated the isothermal crystallization
rate for the LFT composite with 40 wt% of glass fiber,
resulting in an increase of t0.5 accordingly. In addition, it
is noteworthy that, for both pure POM and its LFT composites, the lower the crystallization temperature, the less
the t0.5 value. Such a trend indicates that the overall isothermal crystallization rate decreases with increasing the
crystallization temperature due to the low supercooling in
the range of given temperatures. Despite the heterogeneous nucleation of glass fiber, it is more difficult to form
the crystal nuclei at higher temperatures, especially as
close to the melting point, due to the high mobility of
polymeric chains. Furthermore, at higher temperatures,
only a small part of fibers could act as nuclei for crystallization, but the rest of them generated a retardation effect
on the movement of polymeric chains. Consequently, the
induction time became longer, and the crystallization process was postponed accordingly.
12 POLYMER COMPOSITES2014

Spherulitic Morphology
POM is well known for its high crystallinity and can
usually form well-defined spherulites when crystallizing
from the melt. These spherulites were reported to consist
of folded lamellar crystals and extended chain crystals
[46]. Figure 9 shows the crystalline morphologies of pure
POM and its LFT composites observed under a polarized
optical microscope. The polarized optical micrograph
confirms the reported result for pure POM as shown by
Fig. 9a, where the spherulitic structures based on parallelly packed-lamellar crystals with distinct boundary
could be clearly distinguished. The formation of these
large and partially interconnected spherulites is ascribed
to few of nucleation centers among the POM molecules.
Differing from the short glass fiber reinforced composites,
the residual fibers in the LFT composites have much
greater aspect ratios (ratio of fiber length to diameter).
This may reduce the heterogeneous nucleation sites but
makes it possible to induce the transcrystallinity of POM
domain. Figure 9bd shows the morphology of the POMbased LFT composites formed during isothermal crystallization, where a narrow zone of transcrystallinity along
the fibers could be clearly distinguished in addition to
bulk nucleated spherulites. The formation of
DOI 10.1002/pc

FIG. 10. TGA thermograms of pure POM and its LFT composites.
Inset: Derivate TG curve of pure POM and its LFT composites. [Color
figure can be viewed in the online issue, which is available at wileyonlinelibrary.com.]

transcrystallinity may be a heterogeneous process occurring on impurities collected on the long fiber surfaces.
Furthermore, the observed transcrystallinity appears at the
surfaces of long fibers without any interruption by the
bulk-nucleated spherulites, and the transcrystallization of
POM domain is also a process of growth with spatial constraints [47]. It is no doubt that the formation of transcrystallinity during the molding of POM-based LFT
composites facilitates the interfacial adhesion between the
matrix and fibers and thus enhances the reinforcement
effect.

Thermal Stability
The effect of long glass fiber on the thermal stability
of POM-based composites has been investigated by TGA,
and the obtained thermograms are shown in Fig. 10.
According to the weight-loss profiles, pure POM and its
LFT composites exhibited a typical one-step thermal degradation behavior under a nitrogen atmosphere. Such a
pyrolysis may be attributed to the chain scission of the
POM backbone as the prevailing decomposition reaction.
Accordingly, pure POM was found to encounter an initial
decomposition at 320.4 C to generate a weight loss of 3
wt%, and then, a rapid decomposition took place with a
maximum weight loss at 418.5 C due to the random scission of CAO bonds in the molecular chains. A thermally
induced chain rupture led to complete unzipping of the
damaged chain and thus almost left no residual char. On
the other hand, it seems that the incorporation of long
glass fiber did not notably affect the thermal degradation
behavior of POM domain. However, it is noteworthy that
the characteristic temperatures at the maximum weightloss rate were considerably improved with increasing the
fiber content as shown by derivate TG thermograms in
DOI 10.1002/pc

the inset of Fig. 10. It was reported that the incorporation


of nonflammable inorganic materials in polymeric matrix
could reduce the heat release rate, which plays a key role
in retarding the decomposition temperature. Moreover,
the glass fiber at a high loading level can effectively act
as physical barriers to hinder the transport of volatile
decomposed products out of POM-based composites during thermal decomposition. Therefore, the POM-based
LFT composites present an increasing improvement in
thermal stability with increasing the fiber content. In
addition, the char yields of the composites are found to
slightly exceed the weight percentage of glass fiber in
composites as a result of the enhancement of carbonization toward the POM matrix by long glass fiber. Similarly, the glass fiber also has a good barrier effect on the
thermal degradation process, resulting in the retardation
of weight loss of thermal degradation products as well as
the thermal insulation of POM matrix. Consequently,
much higher char yields were achieved at the end of the
thermal decomposition for the composites.
CONCLUSIONS
The long glass fiber reinforced POM composites were
prepared through a thermoplastic pultrusion on the
custom-designed pultrusion equipment. The resultant
composites achieved much higher tensile and flexural
strength than the short glass fiber reinforced ones at the
same fiber loadings. Such a promising reinforcement
effect is attributed to the feature that the residual fiber
length in the injection-molded LFT products is greatly
superior to that in short fiber reinforced ones. This takes
full advantage of the strength of the reinforcing fiber
itself. The POM-based LFT composites also gained a significant improvement in impact toughness through the
energy dissipation by fiber fracture as a result of long
fiber effect. The thermal stability of the POM-based composites was also improved in the presence of long glass
fiber. The morphologies indicate that the fiber fracture
and fiber pull-out concurred on the tensile and impact
fracture surfaces, and the former preceded the latter. The
SEM investigation also confirms a good interfacial adhesion between the matrix and fibers, thus leading to subsidiary enhancement in mechanical performance. The
isothermal kinetic investigations indicates that the crystallinity of POM domain was enhanced as a result of the
heterogeneous nucleation of glass fiber, but the crystallization rate did not show a distinct improvement due to
the interspace restriction toward crystalline growth caused
by long glass fiber. These kinetic parameters provide
information on the processing conditions of POM-based
LFT composites for the injection and compression molding. Moreover, the transcrystallinity formed during isothermal process was advantageous for enhancing the
interfacial adhesion between the matrix and long fibers,
thus improving the reinforcement effect of long glass
fiber. It is prospective that the development of POMPOLYMER COMPOSITES2014 13

based LFT composites will provide opportunities for the


expanded applications of thermoplastic composites with
POM matrix.
REFERENCES
1. R.F. Gibson, Compos. Struct., 92, 2793 (2010).
2. A.P. Mouritz, A.P. Leong, and I. Herszberg, Compos. Part
A Appl. Sci. Manuf., 28, 979 (1997).
3. M. Bannister, Compos. Part A Appl. Sci. Manuf., 32, 901 (2001).
4. M. Schemme, Reinf. Plast., 52, 32 (2008).
5. A. Goel, K.K. Chawla, U.K. Vaidya, N. Chawla, and M.
Koopman, Mater. Charact., 60, 534 (2009).
6. V.S. Chevali and G.M. Janowski, Compos. Part A Appl. Sci.
Manuf., 41, 1253 (2010).
7. A. Hassan, R. Yahya, A.H. Yahaya, A.R.M. Tahir, and P.R.
Hornsby, J. Reinf. Plast. Compos., 23, 969 (2004).
8. E. Lafranche, P. Krawczak, J.P. Ciolczyk, and J. Maugey,
Polym. Adv. Technol., 24, 114 (2005).
9. T. Hartness, G. Husman, J. Koenig, and J. Dyksterhouse,
Compos. Part A Appl. Sci. Manuf., 32, 1155 (2001).
10. R. Marissen, L.T. van der Drift, and J. Sterk, Compos. Sci.
Technol., 60, 2029 (2000).
11. Y. Du, T. Wu, N. Yan, M.T. Kortschot, and R. Farnood,
Compos. Part B Eng., 56, 717 (2014).
12. K. Han, Z. Liu, and M. Yu, Macromol. Mater. Eng., 290,
688 (2005).
13. S. Kalia, B.S. Kaith, and I. Kaur, Polym. Eng. Sci., 49,
1253 (2009).
14. T. Bayerl and P. Mitschang, J. Appl. Polym. Sci., 131,
39716 (2014).
15. A.R. Melro, P.P. Camanho, and S.T. Pinho, Compos. Sci.
Technol., 68, 2092 (2008).
16. J. Karger-Kocsis, T. Harmia, and T. Czigany, Compos. Sci.
Technol., 54, 287 (1995).
17. E. Lafranche, P. Krawczak, J.P. Ciolczyk, and J. Maugey,
Expr. Polym. Lett., 1, 456 (2007).
18. S.D. Bartus and U.K. Vaidya, Compos. Struct., 67, 263 (2005).
19. J.L. Thomason, Compos. Part A Appl. Sci. Manuf., 36, 995
(2005).
20. J.L. Thomason, Compos. Part A Appl. Sci. Manuf., 39, 1732
(2008).
21. D. Teixeira, M. Giovanela, L.B. Gonella, and J.S. Crespo,
Mater. Design, 47, 287 (2013).

14 POLYMER COMPOSITES2014

22. A. Carlsson and B.T. Astrom, Compos. Part A Appl. Sci.


Manuf., 29A, 585 (1998).
23. D.H. Kim, W.I. Lee, and K. Friedrich, Compos. Sci. Technol., 61, 1065 (2001).
24. C.H. Chen and C.H.M. Ma, Compos. Part A Appl. Sci.
Manuf., 28A, 65 (1997).
25. L. Cilleruelo, E. Lafranche, P. Krawczak, P. Pardo, and P.
Lucas, Expr. Polym. Lett., 6, 706 (2012).
26. K. Ramani, H. Borgaonkar, and C. Hoyle, Compos. Manuf.,
6, 35 (1995).
27. A.H. Miller, N. Dodds, J.M. Hale, and A.G. Gibson, Compos. Part A Appl. Sci. Manuf., 29A, 773 (1998).
28. A. Luisier, P.E. Bourban, and J.A.E. Manson, Compos. Part
A Appl. Sci. Manuf., 34, 583 (2003).
29. Y. Li, Z. Tao, Z. Chen, J. Hui, L. Li, and A. Zhang, Polymer, 52, 2059 (2011).
30. K. Pielichowska, E. Dryzek, Z. Olejniczak, E. Pamua, and
J. Pagacz, Polym. Adv. Technol., 24, 318 (2013).
31. A. Durmus, A. Kasgoz, N. Ercan, D. Akn, and S. Sanli,
Polymer, 53, 5347 (2012).
32. A. Kelly and W. Tyson, J. Mech. Phys. Solids, 13, 329
(1965).
33. J.L. Thomason, Compos. Part A Appl. Sci. Manuf., 33, 1283
(2002).
34. J.K. Kim and Y.W. Mai, Compos. Sci. Technol., 41, 333
(1991).
35. A.H. Cottrell, Proc. R. Soc. Lond. A, 282, 2 (1964).
36. V. Mirjalili and P. Hubert, Compos. Sci. Technol., 70, 1537
(2010).
37. J.L. Thomason, Compos. Part A Appl. Sci. Manuf., 28, 277
(1997).
38. K.J. Pielichowska, J. Appl. Polym. Sci., 123, 2234 (2012).
39. Z. Qiu, S. Zhu, and W.J. Yang, J. Nanosci. Nanotechnol., 9,
4961 (2009).
40. M.J. Avrami, Chem. Phys., 7, 1103 (1939).
41. M.J. Avrami, Chem. Phys., 8, 212 (1940).
42. W.B. Xu and P.S. He, Polym. Eng. Sci., 41, 1903 (2001).
43. M. Raimo, Acta Mater., 56, 4217 (2008).
44. J. Li, C. Zhou, G. Wang, Y. Tao, Q. Liu, and Y. Li, Polym.
Test., 21, 583 (2002).
45. T. Chatterjee, A.T. Lorenzo, and R. Krishnamoorti, Polymer, 52, 4938 (2011).
46. H. Hama and K. Tashiro, Polymer, 44, 2159 (2003).
47. C. Wang and C.R. Liu, Polymer, 40, 289 (1999).

DOI 10.1002/pc

You might also like