You are on page 1of 12

Research Collection

Journal Article

Pultrusion of large thermoplastic composite profiles up to Ø 40


mm from glass-fibre/PET commingled yarns

Author(s):
Volk, Maximilian; Wong, Joanna; Arreguin, Shelly; Ermanni, Paolo

Publication Date:
2021-12-15

Permanent Link:
https://doi.org/10.3929/ethz-b-000513801

Originally published in:


Composites Part B: Engineering 227, http://doi.org/10.1016/j.compositesb.2021.109339

Rights / License:
Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International

This page was generated automatically upon download from the ETH Zurich Research Collection. For more
information please consult the Terms of use.

ETH Library
Composites Part B 227 (2021) 109339

Contents lists available at ScienceDirect

Composites Part B
journal homepage: www.elsevier.com/locate/composites

Pultrusion of large thermoplastic composite profiles up to Ø 40 mm from


glass-fibre/PET commingled yarns
Maximilian Volk a ,∗, Joanna Wong b , Shelly Arreguin a , Paolo Ermanni a
a
Laboratory of Composite Materials and Adaptive Structures, Department of Mechanical and Process Engineering, ETH Zürich, Leonhardstrasse
21, CH 8092, Zürich, Switzerland
b
Department of Mechanical and Manufacturing Engineering, Schulich School of Engineering, University of Calgary, Canada

ARTICLE INFO ABSTRACT

Keywords: Pultrusion is a rapid and cost-effective manufacturing technology for continuous fibre reinforced thermoplastic
Pultrusion composite profiles. As the cross-sections of pultruded profiles grow to meet increasing performance require-
Thermoplastic resin ments, manufacturing challenges concerning heat transfer are encountered. In this study, a two-dimensional
Glass fibres
finite element model was used to simulate the heat transfer and fluid flow physics of the pultrusion process
Thermal analysis
for increasing diameters from Ø 5–Ø 40 mm. To facilitate the experimental validation, a novel batch-wise
pultrusion concept is introduced in which the impregnation process is observed in-situ using a transparent die.
The pultrusion studies, conducted on glass-fibre/amorphous polyethylene terephthalate (GF/PET) commingled
yarns, show that – with proper design – pultrusion is able to deliver consistent, high quality (void content <
2%) profiles up to at least Ø 40 mm.

1. Introduction In state-of-the-art thermoset composite pultrusion, the exothermic


curing reaction of the monomer is the primary cause restricting the
Large cross-section composite profiles are critical lightweight load- profile cross-section and manufacturing speed [8]. If not carefully
bearing structures found in wind turbine blade root reinforcements, controlled, the temperature gradients during curing result in stresses
high voltage insulator cores, and beams for construction applications that can lead to the formation of cracks on the surface of the pro-
[1]. Historically, these large profiles have been pultruded from rein- file [9], delamination, warpage [10], or thermal degradation [11].
forcement fibres and low viscosity thermosetting polymers [2]. How- Thermoplastic composite processing does not require an exothermic
ever, growing health and safety concerns about the anhydride harden- curing reaction. Therefore, this study wants to investigate whether
ers used for this application have led to legislative pressure in Europe manufacturing strategies for these materials may be able to circumvent
to eliminate these chemicals from manufacturing processes [3]. At the the limitations imposed by thermosetting materials by applying the
same time, increasing environmental awareness has brought criticism pultrusion process to large cross-section thermoplastic composites.
of the limited end-of-life options for thermoset composites. Thermoplas- Even though thermoplastic polymers only undergo a physical phase
tic composites do not require chemical curing and can be processed change during solidification, heat extraction may still be a challenge as
without volatile organic compounds (VOCs). Additionally, they can be heat can only be removed through the surface of the profile. Moreover,
reshaped, welded and recycled and possess a virtually infinite shelf the processing temperatures of thermoplastics are significantly higher
life [4]. Despite these clear advantages, manufacturing processes for with 240–280 ◦ C used for the PET polymer in this study compared
thermoplastic composites remain underdeveloped. Notably, pultrusion to 175 ◦ C for GF-epoxy profiles [21]. For small thermoplastic profiles,
of thermoplastic composite profiles has been not been investigated for characterized by large surface to volume ratios, heat extraction can be
profiles exceeding 4.5 mm in thickness [5] or Ø 8 mm diameter [6], performed quickly, enabling pultrusion speeds up to 10 m/min [19].
compared to the Ø 130 mm diameters [7] commercially available in However, heat extraction and solidification can become a bottleneck
thermoset composite rods. A selection of the recent literature on the in the manufacturing process for large thermoplastic profiles with low
profile dimensions and intermediate materials used for thermoplastic surface to volume ratios due to the low thermal conductivity of polymer
composite pultrusion is presented in Table 1. materials.

∗ Corresponding author.
E-mail address: mvolk@ethz.ch (M. Volk).
URL: http://structures.ethz.ch (M. Volk).

https://doi.org/10.1016/j.compositesb.2021.109339
Received 25 July 2021; Received in revised form 13 September 2021; Accepted 18 September 2021
Available online 24 September 2021
1359-8368/© 2021 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
M. Volk et al. Composites Part B 227 (2021) 109339

Table 1
Overview about pultruded thermoplastic profile cross-sections.
Author Dimensions Material
Asensio (2020) [12] 20 × 2 mm PET recycled/GF towpreg
Lapointe (2019) [13] Ø 4.8 mm PEEK/CF commingled
Tomic (2018) [6] Ø 8 mm PET/GF commingled
Novo (2016) [14] 20 × 2 mm PP/GF towpreg
Nunes (2013) [15] 24 × 4 U-shaped PP/GF towpreg
Babeau (2015) [16] 50 × 4 mm Low viscosity PA6/GF melt impregnation
Linganiso (2014) [17] 30 × 3 mm PLA/flax commingled
Kamble (2008) [5] 22.75 × 4.5 mm PP/GF commingled
Wiedmer (2006) [18] Ø 4.9 mm, 10 × 3.5 mm PA12/GF commingled
Miller (1998) [19] Ø 2 mm, 20 × 2 mm PP/GF commingled
Jürss (1995) [20] 3 × 1.5 mm PP/GF commingled

Thermoplastic matrix systems also present challenges for impreg- Table 2


Ø 10 mm die geometry parameters, all measurements in mm.
nation and consolidation due to their high viscosity at processing
temperature. Typical melt viscosities of thermoplastics are in the order Øentrance Øexit ltaper lstraight lcooling lgap

of 500–5000 Pa s compared to the 0.03–1 Pa s typical of uncured Ø 25 Ø 10 114 20 90 5


thermosets [22]. To overcome the processing issues associated with the
high melt viscosities of thermoplastics, different materials have been
developed for pultrusion processes. Reactive thermoplastics facilitate a fibre volume content of 55%, was used in this study. PET was
impregnation due to their low viscosities. However, similarly to their selected for its favourable shrinkage and good impregnation behaviour,
thermosetting counterparts, reactive thermoplastics also exhibit disad- as well as its competitive price point. A study comparing the results of
vantages such as exothermic curing reactions, furthermore they are pultrusion on commingled yarns containing other thermoplastic matrix
only available for a limited range of materials i.e. PA12 [23], PA6 systems was previously published [29]. Before use, the commingled
or PMMA [24] and require precisely controlled processing conditions yarns were dried at 40 ◦ C for 6 h. The dried yarns were heat-set under
such as e.g. nitrogen atmosphere. Direct melt impregnation of the tension inside the preheating oven at 150 ◦ C for 15 min to allow the
fibres can be achieved using specially formulated low-viscosity ther- highly oriented polymer chains to relax after the fibre spinning process
moplastics. However, good results have only been achieved for profiles and to prevent them from contracting during preheating.
with short impregnation lengths and large width to height ratios [25].
Alternatively, intermediate material such as tapes [26], towpregs [27] 2.2. Material characterization
or commingled yarns [28] can be used to reduce impregnation lengths.
While tapes provide excellent impregnation results and towpregs allow 2.2.1. Thermal analysis
a homogeneous polymer distribution among the fibres, their high cost Thermogravimetric analysis (TGA) using a Perkin Elmer Pyris 1
is a limiting factor. Commingled yarns were selected for this study as TGA was run from 20 ◦ C to 800 ◦ C at heating rates of 10 ◦ C/min in
they are available in a wide selection of polymers and at a reasonable air atmosphere to determine the thermal degradation behaviour of the
cost while still showing a sufficiently good impregnation behaviour. polymer. Differential scanning calorimetry (DSC) on a Mettler Toledo
This study investigates the feasibility of protruding high quality DSC 1 running from 20 ◦ C to 300 ◦ C at heating rates of 10 ◦ C/min in
large cross-section solid thermoplastic rods with diameters larger than nitrogen atmosphere was used to determine glass transition tempera-
Ø 8 mm with void contents below 2% from commingled yarns. The tures. The DSC measurements showed a fully amorphous polymer with
research focus lies on the thermal management of the pultrusion pro- a glass transition temperature of 80 ◦ C. The DSC and TGA curves are
cess as it changes with increasing profile diameter. A multiphysics provided in the Appendix A. The thermal conductivity of the compos-
finite element model (FEM) was implemented in Comsol 5.4 to simulate ite materials at room temperature was measured in longitudinal and
the heat transfer and fluid flow physics during the pultrusion process transverse direction to the fibres using a custom built device according
for a Ø 10 mm profile. The model was validated experimentally and to Ref. [30], with the values provided in Table 3 . The emissivity of
then used to design pultrusion dies for Ø 5 mm, Ø 20 mm, and the material of 0.94 for a wavelength ranging from 8 μ m to 14 μ
Ø 40 mm profiles. Pultrusion experiments were conducted using a m has been determined by fitting the temperature measurement of a
novel batch pultrusion system, to minimize the cost of setting up a composite rod section using a calibrated thermocouple thermometer to
continuous pultrusion line, while still enabling the study of steady- the measurements of a pyrometer.
state processing conditions. A transparent die was used to evaluate
this concept in terms of resin backflow, void evacuation, and fibre 2.2.2. Rheology
alignment. The numerical model guided the design of the pultrusion Samples of neat polymer melt were extracted from the commingled
dies and processing parameters and enabled the manufacture of high yarn by hot pressing under vacuum in a Fontijne TP 400 press at
quality profiles with void contents < 2% for diameters up to Ø 40 mm, 270 ◦ C. The extracted polymer was used to characterize its rheological
the largest thermoplastic composite rods to be pultruded to date. These behaviour using an Anton Paar MCR302 plate rheometer operating
results indicate that large cross-section thermoplastic profiles may be in frequency sweep at 270 ◦ C from 100 to 1 rad/s and temperature
manufactured by pultrusion. However, the speed advantage often asso- ramp from 270 ◦ C to 150 ◦ C. The shear viscosity was determined from
ciated with thermoplastic composite processing is significantly reduced the oscillatory measurements through the Cox–Merz rule [31]. The
with increasing diameters due to the limited cooling capabilities. viscosity curves are found in the Appendix A.

2. Material and methods 2.2.3. Microscopy & void content measurement


Commingled yarns and pultruded rods were embedded in epoxy
2.1. Feedstock resin and cross-sectional surfaces were obtained by polishing on a
Struers Abramin machine using, at the finest, a 3 μ m diamond suspen-
Compofil-PET-70-R-2690N from Jushi (China), a commingled yarn sion. Imaging was done with a Keyence VHX 6000 digital microscope
material based on polyethylene terephthalate (PET) and E-glass with with a defined brightness threshold under coaxial light to distinguish

2
M. Volk et al. Composites Part B 227 (2021) 109339

Fig. 1. Schematic representation of the pultrusion die assembly.

Table 3
Material properties of E-glass, PET and of the composite with 55% fibre volume fraction.
Material property glass fibre PET composite
Density (20–300 ◦ C) [g/cm3 ] 2.55–2.53 [32] 1.33–1.14 [33] 2.01–1.92
Heat capacity (20–300 ◦ C) [J/(gK)] 0.8-1.1 [34] 1.14–2.1 [35] 1.03–1.34
Viscosity [Pa s] 100–270 ◦ C – 8228–44 –
Glass transition temperature Tg [◦ C] – 80 ◦ C –
Thermal conductivity literature (20 ◦ C) [W/(mK)] 0.72 (longitudinal)
1.05 [32] 0.29 [36]
0.55 (transverse)

Thermal conductivity experimental (20 C) [W/(mK)] 0.7 (longitudinal)
– –
0.6 (transverse)
Emissivity – – 0.94

voids from the surrounding material. Void content was averaged over and processing temperatures. First, a 3D model of the die assembly
5 different measurements evenly spaced out over the length of the was used to determine the placement of the heating cartridges and the
pultruded rod, without considering the first and last sections of the rod. distribution of cooling channels to achieve a homogeneous temperature
distribution of the die surfaces in contact with the commingled yarns.
2.3. Numerical simulation This model was then simplified into a 2D axis-symmetric model with
predefined temperature boundary conditions at the die/pultrudate in-
2.3.1. Pultrusion die geometry terfaces. The fluid flow model follows a similar approach developed for
The geometry of the heating die consists of a tapered and straight tapes by Åström, who treated the material as a fibre filled fluid inside
section, while the cross-section of the cooling die’s cavity remains the tapered section where backflow can occur, and as a solid in the con-
constant over its whole length. A schematic of the die geometry is stant cross-sections of the die where plug flow prevails [37]. To avoid
dividing the pultrudate into different physical phases, i.e. melt and
represented in Fig. 1. In the tapered region of the die, the same ratio
solid, the approach from Haffner [38] of implementing phase change
between diameter change to length change of 1:10 was used for all
through a temperature-dependent viscosity is used. The polymer flow
rod diameters to assure comparability, which corresponds to a taper
is therefore modelled through the superposition of two submodels. One
angle of 5.71◦ . This ratio was chosen due to its favourable compromise
submodel considers the backflow of the polymer through a porous
between impregnation performance, relatively short length, and ease of
media of fibres, while the second submodel accounts for the motion of
manufacture using tapered reamers. The length of the straight section
the pultrudate through the die, assuming Stokes flow. The Brinkman
of the heating and cooling dies was increased linearly with the outlet
equation is used as the governing equation of flow through porous
diameter, effectively increasing the length by a factor of two when media, due to the large variation of fibre volume content through-
doubling the diameter. This allows the total surface areas available out the heating die. The transversely isotropic permeability tensor is
for heating and cooling to scale linearly with the pultrudate volume determined using Gebart’s model [39].
to ensure sufficient heating and cooling power as the rod diameter A more detailed modelling approach for describing the pultru-
increases. The geometrical parameters of the initial Ø 10 mm pultrusion sion of commingled yarns has been developed by Kim et al. [40]. In
die tested are summarized in Table 2. contrast to Åström’s model, Kim’s model considers a dual scale flow
behaviour in which polymer matrix first flows along the dry glass
2.3.2. Model formulation fibres agglomerations before impregnating them. Kim’s approach was
A 2D axis-symmetric coupled heat transfer-fluid flow finite element not applied here because micrographs indicated that the dual-scale flow
model was developed in Comsol 5.4 to simulate the temperature distri- effects were negligible after the first 25% of the die, as the glass fibre
bution inside a Ø 10 mm pultrusion die for different pultrusion speeds agglomerations are completely impregnated with polymer.

3
M. Volk et al. Composites Part B 227 (2021) 109339

Table 4
Preheating temperature for a heating die temperature of 260 ◦ C.
Diameter/speed Ø 5 mm Ø 10 mm Ø 20 mm Ø 40 mm
◦ ◦ ◦
25 mm/min 160 C 180 C 200 C 240 ◦ C
◦ ◦ ◦
50 mm/min 180 C 240 C 245 C 250 ◦ C
75 mm/min 220 ◦C 245 ◦C 248 ◦C –
100 mm/min 230 ◦C 250 ◦C 253 ◦C –
◦ ◦ ◦
150 mm/min 240 C 253 C 255 C –

The fluid flow model is directly coupled to the heat transfer model,
which is described through the 2D steady-state heat transfer equa-
tion [10]. Inside the heating die, only conduction is considered while
additional free convection and radiation terms are added once the
material exits the die. The convection heat transfer coefficient is deter-
mined for round cylinders with the corresponding Nusselt and Prandtl
Fig. 2. Micrograph of unconsolidated commingled yarn, with polymer fibres
values for ambient air at 20 ◦ C according to Kind [41], while the emis-
highlighted in violet and surrounded by a dashed line.
sivity value is determined experimentally as described in Section 2.2.
The resulting non-linear multiphysics problem is solved through a seg-
regated solver approach, where the problem is subdivided into several
2.3.7. Heat capacity
steps, solved separately, and joined at the end to compose the final
The temperature dependent heat capacity of the composite is com-
solution. A mesh convergence study was conducted to optimize the
puted based on literature values of amorphous PET [35] and E-glass
mesh refinement and element type for all sub-models while maintaining
fibre [34] using the rule of mixture. A linear regression function is fitted
the robustness of the model.
to the results:

2.3.3. Boundary conditions 𝑐𝑝 (𝑇 ) = 1.11−3 𝑇 + 1.010, [◦ C, J∕(g K)] (3)


In contrast to the majority of existing modelling approaches, a slip
2.3.8. Thermal conductivity
boundary condition is applied over the whole die length as the move-
Instead of assuming an isotropic conductivity as found in state
ment of the pultrudate is determined by the fibres. This assumption
of the art thermoplastic pultrusion modelling, both longitudinal and
is supported by the fact that no visible polymer-rich region on the transverse thermal conductivity are considered in this work to increase
surface of the pultrudate was observed. The reason for this is likely the accuracy of the temperature prediction for large diameters. The
the comparably high fibre volume fractions used, which were between longitudinal thermal conductivity is determined through the rule of
55% and 65%. The inlet boundary condition is set to atmospheric mixtures, while the transverse conductivity is computed using Chamis’
pressure to account for potential backflow, while the outlet boundary equations [43], based on literature values for the glass fibre [32] and
condition is set to the pultrusion speed. A preheating temperature value PET [36]. The experimental validation of the conductivity at room tem-
according to Table 4 has been applied to the material entering the perature given in Table 3 showed good agreement with the computed
heating die from the preheating oven. A constant temperature of 260 ◦ C values with a 8% deviation for the transverse and 3% for the longi-
has been applied to the heating die surface and 30 ◦ C was applied tudinal conductivity that lay within the measurement accuracy. The
to the cooling die surfaces for all simulations and experiments. No conductivity is assumed to remain constant over the temperature range
temperature conditions were applied to the cooling die liner. as no reliable temperature dependent material data for the amorphous
PET grade was available.
2.3.4. Material models
2.3.9. Permeability
Table 3 summarizes the numerical values of the material properties
The permeability is assumed to be transversely isotropic to the
used for the FEM-simulation for a fibre volume content of 55%. All tem-
pulling direction. The longitudinal and transverse permeabilities were
perature dependent property curves are provided in the Appendix A. computed using the Gebart model [39] with a hexagonal fibre arrange-
ment as suggested by Raper [44]. In Fig. 2 the glass fibres appear white
2.3.5. Density with an average glass fibre diameter of 12.6 μm while the thermoplastic
The temperature dependent density of the glass fibre and PET fibres are highlighted in blue, with an average fibre diameter of 18.6
material are derived from literature values [32,42]. The temperature μm. The computed permeability values over the length of the die
dependent density 𝜌(𝑇 ) of the composite is then computed for the depend on the die radius and are provided in the Appendix A. The
variable fibre volume fractions using the rule of mixture to which the resulting permeability values are consistent with those that have been
following regression function is then fitted: experimentally determined for glass fibre yarns in other pultrusion
processes [16,45].
𝜌(𝑇 ) = −3.41 × 10−4 𝑇 + 2.02 [◦ C, g∕cm3 ] (1)
2.3.10. Viscosity
2.3.6. Degradation The temperature-dependent viscosity of the PET is implemented
through a power regression function based on the rheometry measure-
To predict the maximum exposure time the polymer could re-
ments described in Section 2.2.2. The resulting equation is given as:
main at a defined temperature with minimal degradation, a 4th order
polynomial regression function has been fitted to the TGA degrada-
tion measurements, assuming an acceptable weight loss of 0.5%. The 𝜇(𝑇 ) = 1016 𝑇 −5.959 , [Pa s] (4)
maximum exposure time 𝑡(𝑇 ) as a function of temperature is given as:
No shear thinning effect is taken into account as its influence on
𝑡(𝑇 ) = 8.03 × 103 − 104.50 𝑇 + 0.51 𝑇 2 − 1.08 × 10−3 𝑇 3 the viscosity of the polymer was found to be insignificant compared to
the influence of temperature [38] at the pultrusion speeds used in this
+ 8.44 × 10−7 𝑇 4 [◦ C, min] (2) study.

4
M. Volk et al. Composites Part B 227 (2021) 109339

Fig. 3. Pultrusion process for sample manufacturing.

2.4. Experimental cooling die [46]. This liner is easily replaced in case of damage. The
temperature gradient between the heating and cooling temperatures
2.4.1. Batch pultrusion approach was found to be sensitive to the thickness of the liner at the transition
The batch pultrusion approach allows the space- and cost-efficient as well as to the distance between the heating and cooling dies.
pultrusion of large cross-section profiles of limited lengths. Conven- Prior to being drawn into the heating die, the commingled yarns
tional continuous pultrusion requires a bobbin creel and a yarn guid- were preheated in a custom built convection oven. Historically, pre-
ance system for each yarn that is fed into the pultrusion die. For heating set-ups were conductive systems using metal tubes wrapped
the pultrusion of a Ø 40 mm rod, 850 bobbins would be required with heater tape [47], or heated metal plates [48]. Later forced con-
assuming the heaviest available yarn weight of 2690 tex. Batch pul- vection systems using hot air blowing, such as the one used here,
trusion circumvents the space requirements and associated equipment were developed to increase the efficiency and homogeneity of the
and material cost by winding material from a single bobbin into one preheating [26,27]. Other preheating set-ups that have been proposed
thick loop to achieve the required throughput for the pultrusion die. for pultrusion include infrared [49] and microwave radiation [17] or
As illustrated in Fig. 3, the loop is cut on its trailing end after winding hybrid systems in which infrared or microwaves are combined with
and pulled through the die by a steel pulling wire attached to a load forced convection. To achieve the desired temperature distribution
cell and pulling mechanism consisting of a sled on linear rails that is inside the pre-heating oven, a custom forced convection preheating
powered by an electric winch capable of up to 100 kN pulling force concept was developed using a longitudinal airflow of adjustable speed.
and speeds up to 1000 mm/min. The pulling wire can be reused for the The temperature of the preheater was set according to the desired pro-
next experiment after removing the pultruded rod. The pultrusion speed cessing temperature, pultrusion speed, profile diameter, and the time
and pulling forces are recorded at a 2 Hz frequency using a magnetic and temperature dependent degradation behaviour determined through
position sensor with an accuracy of ±15 μm and S-type strain gauge the TGA analysis. It was chosen to be as low as possible to minimize
load cell with an accuracy of ±2 N, respectively. The length of the loop, degradation while also being high enough to enable a homogeneous
and therefore of the pultruded rod, is set to 1500 mm. temperature distribution to be reached at the outlet of the heating die.
Generally, thermoset pultrusion is thought to take several metres A maximum temperature difference of ±5 ◦ C between the centre of
before steady-state pultrusion is reached. This is due to the high pultru- the rod and the surface at the exit of the heating die was allowed. All
sion speeds used in industry and the exothermic reaction of thermosets processing parameters were set according to the results of the numer-
which requires time to develop. In thermoplastic pultrusion, steady- ical simulations. Table 4 shows the preheating temperature required
state conditions can be obtained once the temperature profile is static for the pultrusion of different diameter rods with a die temperature
and the matrix flow has stabilized. Pulling force and temperature of 260 ◦ C at different pultrusion speeds. An intermediate conduction
measurements were used to verify steady-state conditions. High heating heating element set to the preheating temperature was placed between
and cooling capacities were used so that the contact surfaces were the preheating oven and the inlet of the heating die to prevent the
maintained at a constant temperature, not influenced by the transient commingled yarns from cooling down through the transition.
material. Furthermore, shear strength testing and void content analysis All heating modules (preheater, intermediate heater, heating die)
over the length of the pultruded samples confirmed no measurable were heated by electrical resistance and regulated by PID (proportional–
influence of the position within the rod. This combined analysis showed integral–derivative) controllers. The cooling die was water-cooled with
that the pultrusion speed was the most important parameter in deter- the temperature regulated through a PID-controlled heater and fan-
mining the pulling length needed to achieve steady-state conditions. At supported radiator. Temperatures were monitored by strategically placed
the highest speed studied here of 200 mm/min, steady-state conditions thermocouples as well as pyrometers for contactless temperature ac-
were reached after a rod section with a length equivalent to the length quisition. The temperature distribution and evolution predicted by the
of the heating die was pultruded. To analyse only material made under model were validated by inserting a K-type thermocouple into the
steady-state conditions, the first and last section of the rod with the commingled yarn bundle and recording the temperature during the
same length as the heating die were omitted from the analysis. Batch pultrusion process [50]. After the experiment, the exact position of
pultrusion lengths up to 3000 mm were investigated throughout the the thermocouple inside the rod was determined and compared to the
study. A length of 1500 mm was determined to be ideal to reach steady- equivalent location in the simulation.
state conditions interesting for the study of continuous pultrusion while
providing sufficient material for characterization. 2.4.2. Transparent die
The heating and cooling dies were manufactured to the geometries A major concern with the batch pultrusion approach when com-
determined to be most effective by the numerical simulations. The pared with continuous pultrusion is the absence of fibre pretensing and
heating die is machined out of steel, while the cooling die consists fibre guidance. Fibre pretensing and guidance ensure that the fibres are
of an aluminium body with a steel liner. The steel liner protrudes straight and aligned along the pulling direction. Without these systems,
into the heating die and is made from a precision ground tube as yarns may entangle or buckle resulting in poor mechanical properties
represented in green in Fig. 1. This continuous die approach prevents in the pultruded structure. To investigate whether the lack of preten-
the recurring problem of deconsolidation of the material upon exit from sion and fibre guidance had detrimental effects on fibre alignment, a
the heating die, and the shearing and recompression upon entry into the transparent observation die was developed as shown in Fig. 4. The

5
M. Volk et al. Composites Part B 227 (2021) 109339

3. Results and discussion

3.1. Coupled thermal-fluid flow model

Representative temperature distributions predicted by the coupled


heat transfer and fluid flow FEM model are shown in Fig. 5. Specifically,
these are the simulation results for a Ø 10 mm profile pultruded with a
preheating temperature of 250 ◦ C, heating die temperature of 260 ◦ C,
and cooling die temperature of 30 ◦ C run at two different pultrusion
speeds of 25 and 100 mm/min. Because the material is thoroughly
preheated prior to entering the heating die, it exhibits a homogeneous
temperature distribution throughout its transit through the heating die.
Upon entering the cooling die, the surface of the pultrudate solidifies
first, leading to a bell curve-shaped phase transition profile that extends
in length with increasing pultrusion speed.
The importance of preheating is illustrated in Fig. 6 which show-
cases the effect of preheating on the temperature distributions for a
Fig. 4. Transparent heating die for optical in-situ observation of the pultrusion process.
Ø 10 mm pultrudate simulated using different preheating temperatures
of 250 ◦ C and 230 ◦ C at a pultrusion speed of 100 mm/min. These
simulations show that a preheating temperature of 250 ◦ C leads to a
transparent set-up consists of a heating die without cooling section homogeneous temperature distribution within the allowed 5 ◦ C toler-
where half of it was replaced by a 15 mm thick glass plate that is ance, while 230 ◦ C leads to inhomogeneous and incomplete heating
sealed with high-temperature silicone. The transparent die had a taper at this relatively high pultrusion speed. This is due to a combination
angle of 18.43◦ , an exit diameter of Ø 10 mm, an entrance diameter of the low thermal conductivity of the polymer material and the low
of Ø 30 mm, a taper length of 60 mm, and a straight section length residence time inside the heating die at high speeds.
of 30 mm. Although the same heating strategy was employed for the While preheating can be used to ensure homogeneous heat intro-
transparent die, the temperature profile deviated up to 3 ◦ C from a fully duction, heat extraction during the solidification of the pultrudate in
axisymmetric profile of the full die due to the constant heat flux from the cooling die is more critical as it only occurs through the surface
in contact with the cooling die. Fig. 7 shows the simulated tempera-
the heated glass plate into the pultruded material. The transparent die
ture evolution as a function of position from the inlet of the heating
enabled the phenomena of resin backflow and void evacuation to be
die for Ø 10 mm rods pultruded at a heating die temperature of
observed during pultrusion. Although both void evacuation and resin 260 ◦ C, cooling die temperature of 30 ◦ C, and pulling speeds of 50,
backflow mechanisms are included in several pultrusion models [40, 100, and 200 mm/min. Temperatures are reported for the centre and
51], to the author’s knowledge these behaviours have not been directly 1 mm below the surface of the rod. Higher pulling speeds decrease
experimentally observed until now. the residence time of the material in each die section leading to lower

Fig. 5. Temperature distribution during pultrusion of a Ø 10 mm profile based on a preheating temperature of 250 ◦ C, heating die temperature of 260 ◦ C, cooling die temperature
of 30 ◦ C, and pultrusion speeds of (a) 25 and (b) 100 mm/min.

Fig. 6. Temperature distribution of Ø 10 mm pultrudate simulated using preheating temperatures of (a) 250 ◦ C and (b) 230 ◦ C with a heating die temperature of 260 ◦ C, a
cooling die temperature of 30 ◦ C and a pulling speed of 100 mm/min.

6
M. Volk et al. Composites Part B 227 (2021) 109339

scale bars indicating the actual dimensions. This comparison indicates


that the cooling efficiency of the process does not scale linearly with
increasing pultrudate diameter with high-temperature regions propa-
gating far into the cooling die for larger diameters. The main reason for
this behaviour is the ratio between the volume of the pultrudate and
the available cooling surface which decreases with larger diameters.
The volume of the pultrudate scales with radius squared, whereas the
available cooling surface scales only linearly with radius. Fig. 8(b)
provides a plot of the time required to cool a composite rod from a
core temperature of 260 ◦ C to 80 ◦ C using a cooling die temperature of
30 ◦ C. While the cooling time for a Ø 5 mm pultrudate is around 9 s, it
would take one hour for a Ø 100 mm rod to reach the same temperature
because the thermal conductivity of the composite material becomes
the limiting factor. Even though this issue can be mitigated to a certain
extent by increasing the length of the cooling die, this would be at the
expense of increasing friction and therefore pultrusion forces.
The results of the fluid-flow simulations for a Ø 10 mm rod pul-
truded at a speed of 50 mm/min are shown in Fig. 9. Due to the
axisymmetric nature of the flow, only half of the die is represented.
Fig. 7. Temperature evolution as a function of die position taken at the centre and The flow field shown in Fig. 9 a represents the superimposed flow
edge of a Ø 10 mm pultruded rod for speeds of 50, 100, and 200 mm/min.
generated by the motion of the fibre bed in pultrusion direction and
the matrix backflow in the opposite direction. At the entrance of the
tapered section, the speed of the total flow is relatively low and reverses
temperatures in the heating die, higher temperatures in the cooling its direction after the dashed line towards the die walls. This can be
die, and larger temperature gradients between the surface and the explained through the increased backflow at those locations, which is
centre of the pultrudate. At high speeds, the centre of the rod is not caused by higher temperatures through the heat input which in turn
given sufficient time inside the cooling die to come to equilibrium leads to a decrease of the matrix viscosity. This backflow is highlighted
with the surface temperature. The residual heat in the core causes the in Fig. 9 b and represents the movement of the matrix relative to the
outer surface of the rod to be reheated through conduction as can be
fibre bed. It decreases towards the end of the tapered section with
observed for pultrusion speeds of 100 and 200 mm/min, and potentially
decreasing pressure gradient until no backflow is present in the straight
lead to deconsolidation if surface temperatures are raised above the
section and the total flow therefore equals the pultrusion speed.
solidification temperature of the thermoplastic matrix.
For any fixed pulling speed, heat extraction inside the cooling die
becomes more critical with increasing pultrudate diameter. Fig. 8(a) 3.2. Experimental validation of finite element models
depicts the temperature distributions for Ø 5 mm, Ø 10 mm, Ø 20 mm,
and Ø 40 mm rods inside the cooling die at the same heating and To experimentally validate the finite element models, the custom-
cooling die temperatures as well as the same pulling speed. For ease built batch pultrusion setup shown in Fig. 10(a) was used. Examples of
of comparison, the images have been scaled to the same size with the the profiles obtained are shown in Fig. 10(b).

Fig. 8. Temperature distributions inside the cooling dies for pultrudate diameters of Ø 5–Ø 40 mm (a). Pulling speeds were fixed at 100 mm/min, and preheating and heating
temperatures were set to 245 ◦ C and 260 ◦ C, respectively. Time required to cool the core of a thermoplastic composite rod from 260 ◦ C to 80 ◦ C assuming a cooling die temperature
of 30 ◦ C (b).

Fig. 9. Total matrix flow through pultrusion die (a), matrix backflow relative to the fibre bed at 50 mm/min and Ø 10 mm pultrusion diameter (b).

7
M. Volk et al. Composites Part B 227 (2021) 109339

Fig. 10. Photographs of Ø 20 mm pultrusion setup during pultrusion of Ø 20 mm rod (a), Ø 40 and Ø 20 mm profiles made from amorphous PET, semi-crystalline PET (b).

Fig. 11. Tapered section of pultrusion die throughout pultrusion experiment with 50 mm/min pulling speed.

The metal pulling wire from Fig. 3 can be seen in Fig. 11(a) Attempts were made to track the resin backflow visually using
highlighted in red pulling in a bundle of commingled yarns through differently sized iron oxide and graphite particles as well as Astrazon
the transparent heating die. The images (a), (b) and (c) in Fig. 11 show BR200 colour dye to determine a flow distribution inside the die.
the development of a resin backflow front, highlighted in blue. The Unfortunately, the solid particles were quickly trapped by the fibres and
polymer volume displaced by the metal pulling wire creates the initial carried along as can be observed in Fig. 11(a), while the coloured dye
backflow. The shear forces created by pulling the yarns through the remained up to ten times longer in the tapered section than the glass
straight section of the die rapidly straighten any initially misaligned fibres as depicted in Fig. 11(e) and (f). While these observations did
fibres. Void extraction can be observed to occur towards the middle of not provide a measurable velocity field, it did confirm the presence of
the tapered die, with bubbles being pushed towards the die inlet. For resin backflow as found in the fluid flow simulation of Fig. 9. The yarn
speeds greater than 100 mm/min some of the bubbles were observed to bundles were observed to increase slightly in diameter upon entry into
be trapped by the fibres, rapidly decreasing in size and likely collapsing the tapered section. As no pretension is applied to the yarns, matrix
due to high matrix pressures at the transition between the tapered can fill the volume between yarns through backflow and capillary
and straight section of the heating die. The fibre straightening and pressure and increase the bundles’ diameter, potentially improving
void evacuation mechanisms observed through the transparent die impregnation.
suggest that the batch pultrusion process introduced here is not signif- The simulated temperature distributions were experimentally vali-
icantly disadvantaged compared to continuous pultrusion systems with dated for different temperature combinations, pultrusion speeds, and
extensive tensioning and fibre guidance systems. A video where GF- diameters using the batch pultrusion system. The results from the
PP commingled yarns are pultruded is accessible through a hyperlink numerical and experimental investigations are compared in Fig. 12
https://youtu.be/qKukzqm0J-I. which shows the temperature evolution at the centre of the selected

8
M. Volk et al. Composites Part B 227 (2021) 109339

Fig. 12. Validation of modelling results for Ø 10 mm at 50 and 100 mm/min at Fig. 14. Void content of different diameter pultruded rods for different pultrusion
preheating, heating and cooling temperature of 245 ◦ C, 260 ◦ C and 40 ◦ C and Ø 40 mm speeds.
at 25 mm/min at preheating, heating and cooling temperature of 220 ◦ C, 240 ◦ C and
30 ◦ C.

might be the main cause of void formation and not incomplete impreg-
nation. The fibre distribution is not completely homogeneous over the
rods during the pultrusion process. Note that the x-position has been
cross-section, indicating that the mingling quality plays an important
normalized to account for the differences in die length between the
role for the final consolidation quality.
Ø 10 and Ø 40 mm die. The temperature evolution over the heating
The results of the void content analysis over the whole diameter
and cooling die length shows good agreement between simulation and
and speed range investigated are summarized in Fig. 14. For all rod
experiment for all the rod dimensions tested even though the model was
diameters, the void content was below 1%, apart from the Ø 20 mm,
initially developed for Ø 10 mm rods. A slight mismatch between the
25 mm/min experiment, where the material degraded due to prolonged
simulated and measured temperatures was consistently found towards
preheating. Experimental results for the Ø 40 mm rod could only be
the end of the pultrusion die. This effect can be explained by shrinking performed for low speeds as space constraints limited the maximum
effects of the rod, not considered in the model, where the pultrudate size of the die. However, there is nothing to indicate that the high
detaches from the wall, decreasing the thermal transfer coefficient and quality of impregnation could not be maintained at higher pulling
causing higher temperatures than predicted. speeds for this profile size. To investigate the speed limits of this
pultrusion set-up, higher speed pultrusion experiments were performed
3.3. Material quality of pultruded profiles for Ø 10 mm profile with the results depicted in Fig. 15. To be able
to achieve pulling speeds above 200 mm/min while still applying
Designing the pultrusion dies and selecting the processing parame- reasonable pulling forces, the volume of material fed through the dies
ters according to the developed finite element models resulted in very was reduced to the extent that a void content of 2% would result from
good material quality. Fig. 13 shows a representative micrograph taken processing the material within the parameter window the dies were
from a Ø 40 mm rod at different magnifications, showing very low void designed for. Under these conditions, void content was observed to
content and good consolidation. The voids were located mainly inside increase linearly with speed until 400 mm/min after which the void
resin-rich areas towards the centre of the rod, indicating that shrinkage content drastically increased. This rapid increase in void content is

Fig. 13. Micrographs of Ø 40 mm pultruded rod at different resolutions.

9
M. Volk et al. Composites Part B 227 (2021) 109339

Acknowledgements

We thank Massimo Brivio and Achim Kupferschmitt for their contri-


bution to the model development and Damiano Keller, Nadine Hertäg,
and Felix Fischer for the experimental validation during their student
thesis at CMASLab. This work has been supported by the Swiss In-
novation Agency (Innosuisse) through the Swiss Competence Center
for Energy Research (SCCER) Efficient Technologies and Systems for
Mobility.

Appendix A. Supplementary data

Supplementary material related to this article can be found online


at https://doi.org/10.1016/j.compositesb.2021.109339.

References

[1] Baran I. Analysis of pultrusion process for thick glass/polyester composites:


transverse shear stress formations. Adv Manuf: Polymer Comp Sci 2016;2(3–
Fig. 15. Void content for lower throughput Ø 10 mm rods at speeds exceeding the 4):124–32. http://dx.doi.org/10.1080/20550340.2016.1269037, URL https://
heating and cooling capacity of the die set-up. www.tandfonline.com/doi/full/10.1080/20550340.2016.1269037.
[2] Starr T. Pultrusion for engineers. In: The evaluation of research by sciento-
metric indicators. CRC Press; 2000, p. 303, URL http://linkinghub.elsevier.com/
retrieve/pii/B9781843345725500159.
caused by the limited die length which prevents complete impregnation [3] European Parliament and Council. Regulation (EC) no 1907/2006 concerning the
inside the heating die and incomplete cooling inside the cooling die, registration, evaluation, authorisation and restriction of chemicals (REACH). 55,
(726):2006, p. 1–65.
leading to deconsolidation of the material at the die exit.
[4] Biron M. Thermoplastics and thermoplastic composites. William Andrew; 2018.
[5] Kamble V, Vaidya U. Optimization of thermoplastic pultrusion process using
4. Conclusion & outlook commingled fibers. 2008.
[6] Tomić N, Vuksanović M, Međo B, Rakin M, Trifunović D, Stojanović D,
We report for the first time the pultrusion of a solid cross-section Uskoković P, Jančić Heinemann R, Radojević V. Optimizing the thermal gradient
thermoplastic composite rod with a diameter up to Ø 40 mm and void and the pulling speed in a thermoplastic pultrusion process of PET/E glass
fibers using finite element method. Metall Mater Eng 2018;24(2):103–12. http:
contents below 2%. This study shows that with careful die design and
//dx.doi.org/10.30544/367.
selection of the processing parameters based on accurate FEM models, [7] Taporel. Electrical e-glass insulator rods . 2020, http://www.taporel.com/
high quality thermoplastic composite profiles can be obtained from products/pultruded-rod/e-glass-rod-2/.
commingled yarns. [8] Baran I, Tutum C, Nielsen M, Hattel J. Process induced residual stresses
The advantages of thermoplastic composites with regards to cycle and distortions in pultrusion. Composites B 2013;51:148–61. http://dx.
doi.org/10.1016/J.COMPOSITESB.2013.03.031, URL https://www.sciencedirect.
time is limited in pultrusion processes by the need for effective cooling.
com/science/article/pii/S1359836813001339.
As the profile diameter increases so does the volume to surface ratio. [9] Safonov A, Gusev M, Saratov A, Konstantinov A, Sergeichev I, Konev S, Gusev S,
As a result, the speed advantage of processing thermoplastic profiles Akhatov I. Modeling of cracking during pultrusion of large-size profiles. Compos
decreases significantly with increasing diameter, with the time required Struct 2020;235:111801. http://dx.doi.org/10.1016/j.compstruct.2019.111801.
for cooling increasing quadratically with diameter. Even though this [10] Baran I. Pultrusion: State of the art process models. Smithers Rapra; 2015,
p. 248, URL https://research.utwente.nl/en/publications/pultrusion-state-of-the-
issue can be mitigated to a certain extent by increasing the length of
art-process-models.
the cooling die, this would be at the expense of increasing friction and [11] Oh J, Lee D. Cure cycle for thick glass / epoxy composite laminates. J
pultrusion forces. Compos Mater 2002;36(01):19–45. http://dx.doi.org/10.1106/00219980202330,
Experimental validation of the FEM models was made possible using URL http://journals.sagepub.com/doi/10.1177/0021998302036001300.
a novel batch pultrusion process that did not require an expensive creel [12] Asensio M, Esfandiari P, Núñez K, Silva JF, Marques A, Merino JC, Pastor JM.
and fibre guiding system, but still enabled the study of pultrusion under Processing of pre-impregnated thermoplastic towpreg reinforced by continuous
glass fibre and recycled PET by pultrusion. Composites B 2020;200:108365.
steady-state conditions. The fibre straightening, and void evacuation http://dx.doi.org/10.1016/j.compositesb.2020.108365.
mechanisms observed through the transparent die indicated that the [13] Lapointe F, Laberge Lebel L. Fiber damage and impregnation during multi-
developed batch pultrusion process is valid for approximating the die vacuum assisted pultrusion of carbon/PEEK hybrid yarns. Polym Compos
material quality obtainable in continuous pultrusion. 2019;40(S2):E1015–28. http://dx.doi.org/10.1002/pc.24788.
In conclusion, thermoplastic pultrusion is demonstrated to be a [14] Novo P, Silva J, Nunes J, Marques A. Pultrusion of fibre reinforced thermoplastic
pre-impregnated materials. Composites B 2016;89:328–39. http://dx.doi.org/10.
promising approach for large profiles provided thermal management
1016/j.compositesb.2015.12.026.
is taking into account the interaction of the pultrusion die and pul- [15] Nunes J, Silva J, Novo P. Processing thermoplastic matrix towpregs by pul-
trudate depending on the processing parameters. Further investigation trusion. Adv Polymer Technol 2013;32(S1):E306–12. http://dx.doi.org/10.1002/
of pulling forces created by friction and bulk compaction of interme- adv.21279, URL http://doi.wiley.com/10.1002/adv.21279.
diate materials during pultrusion of large cross-sections profiles are [16] Babeau A. Modélisation de la pultrusion de composites thermoplastiques et
propriétés induites. 2015.
suggested as topics of future study.
[17] Linganiso L, Bezerra R, Bhat S, John M, Braeuning R, Anandjiwala R. Pul-
trusion of flax/poly(lactic acid) commingled yarns and nonwoven fabrics. J
CRediT authorship contribution statement Thermoplast Compos Mater 2014;27(11):1553–72. http://dx.doi.org/10.1177/
0892705713486137.
Maximilian Volk: Conceptualization, Methodology, Software, Data [18] Wiedmer S, Manolesos M. An experimental study of the pultrusion of carbon
fiber–polyamide 12 Yarn. J Thermoplast Compos Mater 2006;19. http://dx.doi.
curation, Formal analysis, Writing – original draft. Joanna Wong: Con-
org/10.1177/0892705706055448.
ceptualization, Writing – review & editing, Funding acquisition. Shelly [19] Miller A, Dodds N, Hale J, Gibson A. High speed pultrusion of thermoplastic
Arreguin: Writing – review & editing. Paolo Ermanni: Supervision, matrix composites. Composites A 1998;29(7):773–82. http://dx.doi.org/10.1016/
Conceptualization, Project administration, Reviewing. S1359-835X(98)00006-2.

10
M. Volk et al. Composites Part B 227 (2021) 109339

[20] Michaeli W, Jürss D. Thermoplastic pull-braiding: pultrusion of profiles with [35] Gaur U, Lau S, Wunderlich B, Wunderlich B. Heat capacity and other thermo-
braided fibre lay-up and thermoplastic matrix system (PP). Composites Part A dynamic properties of linear macromolecules. VIII. Polyesters and polyamides. J
1996;2:3–7. Phys Chem Ref Data 1983;12(1):65–89. http://dx.doi.org/10.1063/1.555678.
[21] Sandberg M, Yuksel O, Baran I, Hattel J, Spangenberg J. Numerical and [36] SD3d materials LTD. PETG technical datasheet. 2020, https://sd3d.com/wp-
experimental analysis of resin-flow, heat-transfer, and cure in a resin-injection content/uploads/2017/06/MaterialTDS-PETG_01.pdf.
pultrusion process. Composites A 2021;143:106231. [37] Åström B, Pipes R. A modeling approach to thermoplastic pultrusion. I: Formu-
[22] Minchenkov K, Vedernikov A, Safonov A, Akhatov I. Thermoplastic pultrusion: lation of models. Polym Compos 1993;14(3):173–83. http://dx.doi.org/10.1002/
A review. Polymers 2021;13(2):180. http://dx.doi.org/10.3390/polym13020180, pc.750140302, URL http://doi.wiley.com/10.1002/pc.750140302.
URL https://www.mdpi.com/2073-4360/13/2/180. [38] Haffner S, Friedrich K, Hogg P, Busfield J. Finite-element-assisted modelling of
[23] Luisier A, Bourban P, Månson J. Reaction injection pultrusion of PA12 compos- a thermoplastic pultrusion process for powder-impregnated yarn. Compos Sci
ites: process and modelling. Composites A 2003;34(7):583–95. http://dx.doi.org/ Technol 1998;58(8):1371–80.
10.1016/S1359-835X(03)00101-5, URL http://www.sciencedirect.com/science/ [39] Gebart B. Permeability of unidirectional reinforcements for RTM. J Compos Mater
article/pii/S1359835X03001015. 1992;26(8):1100–33.
[24] Ma C, Yn M, Chen C, Chiang C. Processing and properties of pultruded [40] Kim D-H, Lee W, Friedrich K. A model for a thermoplastic pultrusion process
thermoplastic composites (I). Composites Manufacturing 1990;1(3):191–6. http: using commingled yarns. Compos Sci Technol 2001;61(8):1065–77. http://dx.
//dx.doi.org/10.1016/0956-7143(90)90167-U. doi.org/10.1016/S0266-3538(00)00234-7, URL http://www.sciencedirect.com/
[25] Babeau A, Comas-Cardona S, Binetruy C, Orange G. Composites : Part A modeling science/article/pii/S0266353800002347.
of heat transfer and unsaturated flow in woven fiber reinforcements during [41] Kind M, Martin H. VDI-Wärmeatlas. Springer; 2013, page 759, 765.
direct injection-pultrusion process of thermoplastic composites. Composites Part [42] Kolb H, Izard E. Dilatometric studies of high polymers. I. Second-order transi-
A 2015;77:310–8. http://dx.doi.org/10.1016/j.compositesa.2015.04.017. tion temperature. J Appl Phys 1949;20(6):564–71. http://dx.doi.org/10.1063/1.
[26] Devlin B, Williams M, Quinn J, Gibson A. Pultrusion of unidirectional composites 1698426.
with thermoplastic matrices. Composites Manufacturing 1991;2. [43] Chamis C. Simplified composite micromechanics equations for hygral, thermal
[27] Ramani K, Borgaonkar H, Hoyle C. Experiments on compression moulding and mechanical properties. 1983.
and pultrusion of thermoplastic powder impregnated towpregs. Composites [44] Raper K, Roux J, Vaughan J, Lackey E. Permeability impact on the pressure rise
Manufacturing 1995;6(1):35–43. in a pultrusion die. J Thermophys Heat Transfer 1999;13(1):91–9.
[28] Larock J, Hahn H, Evans D. Pultrusion processes for thermoplastic compos- [45] Sandberg M, Kabachi A, Volk M, Bo Salling F, Ermanni P, Hattel JH, Span-
ites. J Thermoplast Compos Mater 1989;2(3):216–29. http://dx.doi.org/10.1177/ genberg J. Permeability and compaction behaviour of air-texturised glass fibre
089270578900200304. rovings: A characterisation study. J Compos Mater 2020;54(27):4241–52.
[29] Volk M, Arreguin S, Ermanni P, Wong J, Bär C, Schmuck F. Pultruded thermo- [46] Bechtold G. Pultrusion von geflochtenen und axial verstärkten Thermoplast-
plastic composites for high voltage insulator applications. IEEE Trans Dielectr Halbzeugen und deren zerstörungsfreie Porengehaltbestimmung (Ph.D. thesis),
Electr Insul 2020;27(4):1280–7. IVW; 2000.
[30] Zhao S, Manic M, Ruiz-Gonzalez F, Koebel M. Aerogels. In: The sol-gel hand- [47] Vaughan J, Dillard T, Seal E. A characterization of the important parameters for
book. John Wiley & Sons, Ltd; 2015, p. 519–74. http://dx.doi.org/10.1002/ graphite/peek pultrusion. J Thermoplast Compos Mater 1990;3(2):131–49.
9783527670819.ch17, URL https://onlinelibrary.wiley.com/doi/abs/10.1002/ [48] Astroem T, Larsson P, Pipes R, et al. Development of a facility for pultrusion of
9783527670819.ch17. thermoplastic-matrix composites. Compos Manufacturing 1991;2(2):114–23.
[31] Cox W, Merz E. Correlation of dynamic and steady flow viscosities. J Polym Sci [49] Carlsson A, Åström B. Experimental investigation of pultrusion of glass fibre
1958;28(118):619–22. reinforced polypropylene composites. Composites A 1998;29(5–6):585–93.
[32] Barbero E. Introduction to composite materials design. CRC Press; 2017. [50] Åström B. Development and application of a process model for thermoplastic
[33] Lin J, Shenogin S, Nazarenko S. Oxygen solubility and specific volume of rigid pultrusion. Compos Manufacturing 1992;3(3):189–97.
amorphous fraction in semicrystalline poly (ethylene terephthalate). Polymer [51] Hepola P, Advani S, Pipes R. A process model to describe matrix flow and
2002;43(17):4733–43. heat transfer in thermoplastic pultrusion. In: 49th annual conference, composite
[34] Inaba S, Oda S, Morinaga K. Heat capacity of oxide glasses measured by AC institute, society of the plastics industry. 58, (8):Elsevier; 1994, p. 1371–80.
calorimetry. J Non-Crystalline Solids 2002;306(1):42–9. http://dx.doi.org/10.
1016/S0022-3093(02)01055-4.

11

You might also like