You are on page 1of 36

Viscosity characterization and flow simulation and

visualization of polytetrafluoroethylene paste extrusion using


a green and biofriendly lubricant
George A. Schmidt, II,1,2 Yu-Jyun Lin,1,2 Yiyang Xu,1,2 Dongfang Wang,1,2,3,4 Galip
Yilmaz,1,2 and Lih-Sheng Turng1,2
Author information Copyright and License information PMC Disclaimer

Associated Data
Supplementary Materials

Go to:

Abstract

Polytetrafluoroethylene (PTFE) and expanded PTFE (ePTFE) are ideal


for various applications. Because PTFE does not flow, even when heated
above its melting point, PTFE components are fabricated using a process
called paste extrusion. This process entails blending PTFE powder
particles with a lubricant to form PTFE paste, which is subsequently
preformed, extruded, expanded (in the case of ePTFE), and sintered. In
this study, ethanol was proposed as an alternative green lubricant for
PTFE processing. Not only is ethanol benign and biofriendly, it provides
excellent wettability and processing benefits. Using ethanol as a
lubricant, the shear viscosity of PTFE paste and its flow behavior during
paste extrusion were investigated. Frequency sweeps using a parallel-
plate rheometer were performed on PTFE paste samples and various
grits of sandpaper were used to reduce wall slip of PTFE paste. A
viscosity model was generated and a multiphysics software was used to
simulate PTFE paste extrusion. The simulated extrusion pressure was
compared to experimental data of actual paste extrusion. Flow
visualization experiments using colored PTFE layers were conducted to
reveal the flow profile of the PTFE paste. The morphology of the
expanded ePTFE tubes was examined using scanning electron
microscopy and the effect of expansion ratio on ePTFE morphology was
quantified.

Keywords: expanded polytetrafluoroethylene (ePTFE), flow


visualization, green lubricant, polytetrafluoroethylene (PTFE), viscosity
characterization
Go to:

1 |. INTRODUCTION

Polytetrafluoroethylene (PTFE), also known as Teflon®, is a widely used


synthetic fluoropolymer that has several superior material properties,
including extremely low surface energy, low coefficient of friction,
excellent chemical resistance and inertness, hydrophobicity, ultraviolet
resistance, high electrical insulation, and good biocompatibility.[1] These
unique properties make it a popular material choice for the chemical,
coating, composites, defense, medical industries, and
telecommunications, among others.[2–5] PTFE also has the unique ability
to fibrillate when expanded at a high strain rate, which produces a
unique porous structure.[6] This expanded form of the material, called
ePTFE, exhibit drastically different mechanical properties compared to
the original material; particularly, higher compliance and lower tensile
strength.[1] Since ePTFE is a soft material with numerous fibrillated
regions, attractive characteristics like strong flexibility, air permeability,
and porosity can be obtained. The highly multifunctional nature and soft,
inert, and flexible characteristics of ePTFE make it ideal for various
applications, particularly in filtration, implants, and medical devices
applications.[7,8]

One application of interest is in fabricating small diameter (<6 mm)


vascular grafts using ePTFE. Cardiovascular diseases (CVDs) are a
leading cause of death and morbidity across the world, accounting for
17.3 million deaths per year (2016) and representing approximately
31.5% of all deaths.[9,10] Correlated with the increasing number of deaths
is the rising economic burden to treat CVDs. The estimated global cost of
$863 billion per year (2010) is expected to increase over $1 trillion per
year by 2030.[11] As the population ages and grows, it is anticipated that
the need for treatments and healthcare for CVDs will increase
substantially. A common procedure to prevent heart attacks caused by
occluded blood vessels is a coronary bypass surgery, performed about
400,000 times a year in the United States.[12] For synthetic grafts used in
bypass surgery, mimicking the mechanical properties of the native tissue
is of utmost importance, as the elasticity of the tube greatly affects the
hemodynamics within the tube.[13] Unfortunately, current synthetic grafts
face issues of patency at small diameters and the inability to facilitate
tissue growth around and within them.[1,12,13] Some of the significant
factors that cause these grafts to fail are differences in compliance
between the graft and native tissue, hemodynamic factors related to
difference in flow conditions through the graft, and lack of sufficient
endothelialization.[12]

ePTFE, which has been successfully used for large diameter (>6 mm)
artificial blood vessels, offers a possible solution to these problems. By
adjusting the expansion ratio of PTFE tubing, ePTFE structures of
varying compliance can be fabricated to mimic the mechanical
properties of the native tissue to be replaced. In addition, the porous
structure of surface-modified ePTFE can allow for endothelial tissue
growth, as can additional biomolecules such as angiogenic growth
factors and/or anti-thrombogenic drugs added to the graft.[14–17] Thus, the
fabrication of ePTFE vascular grafts is of significant interest.

Because PTFE is not melt-processible, conventional extrusion is not


possible. Instead, the raw powder-like PTFE material must be mixed
with a lubricant, compressed to make a preform, and processed through
a paste extruder. In order to create the porous structure in ePTFE, a
paste extruded solid PTFE profile will undergo an expansion stage prior
to sintering, the latter of which will remove all residual lubricant and
secure the final structure through amorphic locking. To quantify the
porous morphology on both surfaces of the porous ePTFE, the internodal
distance, which characterizes the islands (nodes) and fibrils distribution
after expansion, was proposed. It is well known that PTFE powder
particles tends to fibrillate under shear, due to molecular entanglement
at the particle surfaces that leads to unwinding of the fibrils from the
crystallite structures.[1] The fibrillation among PTFE powder particles
increases the tensile strength of the extrudate and the elastic nature of
the overall flow within the extruder.[18] The unique viscoelastic
properties of the solid-like PTFE paste preform and its tendency to
fibrillate under shear make it a highly interesting candidate for
rheological study.
Go to:

2 |. PRIOR WORK, STATE OF THE ART, AND MOTIVATION


2.1 |. An alternative green lubricant for PTFE processing

Due to the special fibrillation properties of PTFE powder particles,


lubrication is required for the PTFE extrusion process in order to
conduct a smooth and uniform extrusion.[1,19] Because of PTFE’s
extremely low surface energy with high hydrophobic characteristics,
hydrocarbon lubricants are used in PTFE processing. Since the lubricant
must be removed after extrusion, lubricants with high volatility are
chosen. Two of the most common commercial lubricants are Isopar G
and Naphtha.[1] However, those hydrocarbon solutions are highly
flammable and potentially toxic, which not only leads to environmental
pollution issues, but also becomes a potential hazard to patients if there
is any lubricant residue left in PTFE/ePTFE biomedical products. Due to
the unique processing characteristics of PTFE and the proprietary nature
in making various biomedical products, the literature on lubricants in
PTFE processing is scarce, and there is no report focused on potential
substitution for traditional lubricants. Therefore, developing a new type
of lubricant for PTFE processing that is both environmentally friendly
and non-toxic is extremely important and can have huge potential in
current and future PTFE/ePTFE processing.

Recently, we have discovered that ethanol can serve as an effective


lubricating agent for the processing of PTFE. Ethanol, which is benign
and has a low boiling point, enables PTFE to be processed at room or
body temperature, thereby, leading to energy savings and cost reduction.
Furthermore, since ethanol is miscible with water in any proportions,
many water-soluble biomolecules can be fully dissolved in the
ethanol/water mixture. This provides a very effective way of introducing
biofunctionality to the ePTFE vascular grafts. In this study, ethanol was
compared with a number of potential lubricant candidates in terms of
wettability and contact angle on ePTFE. In addition, an optimal mixing
ratio of PTFE powders and ethanol was determined and used in the
fabrication of ePTFE grafts that includes preforming, extrusion, and
expansion. The morphology, expansion ratio, and internodal distance of
the resulting ePTFE grafts were carefully examined and statistical
analysis was performed. Finally, flow visualization during the paste
extrusion was conducted to understand the ethanol-lubricated PTFE
paste movement inside the barrel. The recorded extrusion pressure and
observed velocity profile were also compared with simulated results.

2.2 |. Viscosity characterization and simulation of PTFE for paste


extrusion

Despite its unique flow properties, there are only a few studies involving
PTFE paste extrusion or simulation of PTFE paste flow. An experimental
paste preforming study as well as extrusion studies using three different
PTFE fine powders through both capillary and annular dies with high
reduction ratio (RR) was reported.[20,21] The effects of extrusion die
design, resin molecular structure, as well as lubricant properties, and
concentration on the properties of PTFE paste processing and extrudates
were studied.[22,23] Modeling of incompressible PTFE paste extrusion was
documented in the literature using the commercial finite element
method code FEMLAB.[18] This simulation accounted for the viscoelastic
effects within paste flow and used a “structural parameter” to
characterize the degree of fibrillation. A shear-thinning viscosity model
was created from data obtained using a parallel plate rheometer with
sandpaper glued to each plate. The reported power-law index, n, for the
PTFE paste in this study was 0.5.[18] The model was supplemented with
elasticity data obtained using an extensional rheometer. The variables of
length-to-diameter (L/D) ratio, RR, and entrance angle and their effect
on the extrusion pressure were shown. For each factor, different
setpoints for that factor were varied while the rest remained constant in
order to show the relationship between that factor and the resulting
extrusion pressure.[18] The study concluded that the paste could be
modeled as a shear-thinning fluid until the onset of fibrillation, in which
case a hyperelastic Mooney–Rivlin model was implemented.
Subsequently, a model incorporating compressibility and wall slip was
created to develop further understanding of the flow behavior.[24]

Both studies implemented a steady-state flow simulation with an inlet at


the top of the barrel, where the ram would be located. However, due to
the nature of the ram extruder, each PTFE preform must be loaded into
the barrel and extruded separately, making the extrusion a transient
batch process. Although the flow does reach steady state and long
lengths of tubing can be extruded from a small preform, the startup
effects and transient development of the flow and the transient extrusion
process were not considered in those studies. Therefore, little is known
about the flow behavior and required extrusion force over time. In
addition, extruded tubing from the beginning of the process (before
steady state is reached) is often neglected. Understanding of these
transient flow effects through simulation and flow visualization may
shed light on the properties of the extrudate, resulting in less material
waste and better extrudate quality.
Go to:

3 |. EXPERIMENT AND SIMULATION


3.1 |. Materials

Commercially available fine PTFE powder particles (Daikin) sieved to


produce a diameter range from 250 μm to 2000 μm were used in this
study. The ethanol used in the experiment was provided by Decon’s Pure
Ethanol. Ethanol was chosen as the lubricant of choice in this study due
to its boiling point of 70°C, solubility in water, low viscosity, and non-
toxic nature. The material characterized and simulated in this study was
a paste consisting of fine PTFE powder particles mixed with 18% pure
ethanol by weight as a lubricating agent. Traditional lubricants such as
Isopar G and Naphtha as well as other potential lubricant candidates
such as avocado oil, glycerol, isopropyl alcohol, methanol, mineral oil,
and deionized (DI) water were purchased from common commercial
vendors and used as received.

3.2 |. Wettability and contact angle test for initial screening of


lubricants

Initial screening of various potential lubricant candidates was conducted


by visual wettability and absorption observation of 10 microliters (μl) of
selected lubricant droplet on a PTFE membrane surface for 90 s. These
potential lubricants include ethanol, Naphtha, glycerol, avocado oil,
mineral oil, and DI water. Based on the initial screening results and the
promising performance of ethanol, droplet contact angle was conducted
using an optical water contact angle device (Powereach JC2000D1).
About 20 μl of methanol, ethanol, and isopropyl alcohol from the alcohol
family, traditional lubricants (i.e., Naphtha, Isopar G) as positive control,
mineral oil, and DI water (as negative control) were dropped on the
same PTFE membrane and the contact angle was measured immediately
after the droplet application.

3.3 |. Determination of suitable PTFE/ethanol mixing ratio

After the conclusion that ethanol can serve as an effective lubricant for
the processing of PTFE, the optimal mixing ratio between PTFE and
ethanol needed to be determined. In the case of traditional lubricants
(Isopar G and Naphtha), around 15%−25% by weight of the total
compound was needed.[25] In this study, ethanol at a weight percentage of
15%, 18%, 22%, 25%, 30%, and 40% of total weight of the mixture,
respectively, was added to 10 g of PTFE powder particles. The
PTFE/lubricant mixture of each sample was then mixed by a roller mixer
for 30 min under room temperature. After mixing, the appearance and
morphology of the mixture were observed, and the optimal ethanol
percentage was determined based on the homogeneity of the mixture
and the feasibility of being extruded. A water-based dye was added into
the ethanol for easier determination of the homogeneity upon mixing
and lubrication.

3.4 |. Preforming and paste extrusion of PTFE

To create 100 g of PTFE paste, 82 g of PTFE powder particles and 18 g of


ethanol were added to a jar and blended using a bottle roller for 30 min.
The resulting mixture was then preformed within a cylindrical barrel of
50 mm in diameter using a pressure of 2 MPa. After preforming, the cone
and die were attached to the extruder. The cone used in this study had a
RR of 90. The barrel was then loaded with the preform and an Instron
5967 Universal Testing Machine was used in compression mode to move
the ram downward, which forced material through the cone and die. In
this study, the ram velocity was kept constant at 5 mm/min and the
required pressure as a function of time was recorded. Due to the
convergent geometry of the paste extruder, the shear rate distribution
inside the extruder was not uniform. The extruded tube had an inner
diameter of 6 mm and a uniform wall thickness of 1 mm.

3.5 |. PTFE expansion and sintering

After the paste extrusion, PTFE extrudates were expanded to different


expansion ratios. To accomplish this, an extruded PTFE tube was cut into
samples with lengths of 2 cm, and the ends of each sample were wrapped
with PTFE tape to prevent edge cracking during expansion. Each sample
was pre-heated under 200 °C for 3 min to ensure consistency of the
material temperature. After 3 min, the extrudate was expanded
unidirectionally with the Instron 5967 Universal Testing Machine at an
expansion rate of 1000 mm/min at 200°C inside an environmental
chamber. In the experiment, 100%, 200%, 300%, 400%, 500%, 600%,
and 700% expansion ratios were achieved by incrementally increasing
the maximum expansion length of the PTFE sample. In order to further
understand the relationship between the expansion ratio and ePTFE’s
morphology, a statistical analysis was performed on the internodal
distance based on different expansion ratios of ePTFE. The internodal
distance differences between external surface and internal surface of
ePTFE were compared and analyzed. After the expansion process, the
resulting ePTFE tubes were sintered in the oven for 12 h at 360°C. In
addition, the evaporation rate of ethanol as a lubricant of PTFE after
paste extrusion was also measured as described below.

3.6 |. Fourier-transform infrared spectroscopy and scanning electron


microscopy

To confirm complete removal of ethanol through evaporation within the


ePTFE tubing, the sliced PTFE extrudates were dried under room
temperature for 24 h. Then, Fourier-transform infrared spectroscopy
(FTIR) was used to characterize the infrared absorption peaks of as-
extruded PTFE tubes and dried PTFE samples in the range of 4000–400
cm−1, with a resolution of 4 cm−1 (Tensor 27, Bruker, Germany). The
surface morphology and fibrillated microstructure of the ePTFE samples
on the internal and external surfaces of the ePTFE tube at different
expansion ratios were examined with a scanning electron microscope
(NeoScope JEOL JCM-5000, Japan).

3.7 |. Flow visualization of PTFE during paste extrusion

A flow visualization experiment was conducted to observe the


movement of the PTFE paste within the extruder. A multilayered
preform was created using multiple layers of thin PTFE performs at 5
mm thickness. Each layer was formed separately as a single preform and
the interface was painted by a blue, water-resistance laboratory marker
to serve as a flow tracker. The colored thin preform layers were stacked
and loaded into the extruder and extruded with a ram velocity of 5
mm/min. Halfway through the paste extrusion, the remaining paste
inside the barrel was removed and cut open, and the cross-section of the
paste was observed and examined.

3.8 |. Viscosity characterization


Samples were created to fit into a parallel plate rheometer and mimic the
properties of the preform. The decision to use the parallel plate
rheometer was based on its ubiquity in polymer engineering labs and the
suitability for highly viscous materials like PTFE paste. To prepare the
viscosity characterization samples, PTFE paste was compression molded
into disks with a diameter of 44 mm and a height of 2.54 mm. Using the
known dimensions of the compression mold cavity, the mass of paste
required to achieve a density of 1.69 g/cm3 was calculated and loaded
into the mold. After compression molding, the samples were trimmed to
25.4 mm in diameter and characterized in a TA AR 2000 EX rheometer.

Due to the extremely low friction of the PTFE paste, additional


precautions had to be taken in attempt to record accurate rheological
data. Three parallel plate geometries were used for viscosity
characterization, each intended to reduce slip during experiments. The
geometries consisted of three different grits of adhesive sandpaper disks
glued to a 25.4 mm diameter steel plate (180, 120, and 60 grit,
respectively). The sandpaper was applied to both the bottom and top
plate of the rheometer. The viscosity of each sample was characterized
using an oscillatory frequency sweep using a strain amplitude of 1% to
further minimize the effect of wall slip while ensuring linear
viscoelasticity. The Cox-Merz rule was applied as an assumption.

3.9 |. Transient flow simulation

COMSOL Multiphysics 5.3a was used to simulate paste extrusion. In this


study, the Laminar Flow, Moving Mesh, and Level Set physics modules
were used, as well as the Stationary, Phase Initialization, and Time-
Dependent study modules. The Laminar Two-Phase Flow, Level Set
Multiphysics module was used to couple the equations of momentum for
the fluid flow and the level set equation.

The paste extruder was modeled using a 2D axisymmetric coordinate


system. Using this coordinate system drastically reduced the simulation
time (as compared to 3D) by assuming θ-symmetry (radial flow
hypothesis). The model was created using the dimensions of the paste
extruder used to acquire experimental data.

To further reduce simulation time and increase accuracy only in


locations of interest, the model was meshed in the following way. A
mapped mesh was applied to the barrel region, which was allowed to
deform as the ram moved downward. The cone and die regions were
meshed using a free triangular mesh. The entire mesh of the geometry is
shown below in Figure 1.

FIGURE 1
Mesh of the 2D axis-symmetric model of the paste extruder. (A) Plot showing both the
mapped mesh in the barrel region and the free triangular mesh in the cone and die regions.
(B) Magnified view of the free triangular mesh in the cone and die regions [Color figure can
be viewed at wileyonlinelibrary.com]

The laminar flow module was solved using the incompressible


formulations of the continuity equation and momentum equation, shown
below. Gravity was neglected due to the dominance of viscous terms in
polymer flows.

∇⋅u=0.
(1)
ρ∂u∂t+ρ(u⋅∇)u=∇⋅[−pI+μ(∇u+(∇u)T)].
(2)

In the above equations, u is the velocity vector, ρ is the density, p is the


pressure, and μ is the viscosity. The viscosity model used was the
Carreau model as shown below.

μ=μinf+(μ0−μinf)[1+(λγ˙)2]n−12.
(3)

Carreau model constants were obtained from the oscillatory shear


rheometer tests by fitting the model curve to the data obtained from the
frequency sweep. TA Rheology Advantage Data Analysis software was
used to generate both of the Carreau model and power-law model
constants. However, the simulation failed to converge when using the
power-law viscosity model, likely due to the near infinite viscosity at low
shear rates based on the power-law equation.

In order to model the paste leaving the barrel and entering the empty
cone, the cone had to be simulated as initially filled with air. Because of
this, an additional equation was necessary to track the fluid–air interface.
The level set function was chosen based on its compatibility with the
existing physics modules and its ability to predict any bubble or droplet
formation (if any). The level set function adds an additional partial
differential equation to the solver, shown below.

∂ϕ∂t+u⋅∇ϕ=γ∇⋅(ϵls∇ϕ−ϕ(1−ϕ)∇ϕ|∇ϕ|).
(4)

In the above equation, ϕ is a scalar field called the level set variable, γ is a
reinitialization parameter, and ϵls controls the interface thickness. The
level set variable has value ranging from 0 to 1, namely, ϕ is equal to zero
in areas of the domain containing solely fluid 1 (air) and equal to one in
areas of the domain containing solely fluid 2 (PTFE paste). ϕ smoothly
transitions from zero to one across the fluid interface. The motion of the
interface is tracked via the left-hand side of Equation (4); the other terms
provide numerical stability.[26]

Due to the extremely low friction of PTFE and the ethanol lubricant, as
well as the works by Patil et al., Mitsoulis et al., and Hatzikiriakos et al.,
the slip wall condition was used for all walls in the simulation
model.[18,24,27] The mesh in the cone was set as fixed, since the ram will
never cross into the cone region. The wall representing the ram was
given a prescribed normal mesh velocity of −5 mm/min to match the
ram speed used in the experimental data. A smooth step function was
used to ramp up from 0 to this velocity. The mesh in the barrel freely
deformed based on the movement of the ram.
Go to:

4 |. RESULTS AND DISCUSSION


4.1 |. Wettability and contact angle tests

Figures 2(A–C) illustrate the absorbability and wettability properties of


various potential lubricant candidates on the PTFE surface at three time
instants: immediately after the lubricant droplet contacted the PTFE, 60
s after the contact, and 90 s after the contact. It can be seen that ethanol
and Naphtha had similar level of absorbability with fairly large wetting
areas on the PTFE. This method also showed that ethanol and Naphtha
could be absorbed by PTFE over time. In contrast, avocado oil and
mineral oil showed smaller contact areas and partial absorption. Finally,
for glycerol or DI water, neither wetting nor absorption was observed,
suggesting poor lubricating performance

FIGURE 2
Wettability and absorption tests for initial screening of potential lubricant candidates:
Ethanol, naphtha, glycerol, avocado oil, mineral oil, and DI water. (A) Immediately after
droplet application; (B) 1 min (60 s) after contact; (C) one and a half minutes (90 s) after
contact. (D) Contact angle tests for methanol, ethanol, isopropyl alcohol, naphtha, Isopar G,
mineral oil, and DI water. (E-F) evaporation of various lubricant candidates 1 and 3 min
after contact [Color figure can be viewed at wileyonlinelibrary.com]

The contact angle results for the alcohol family (i.e., methanol, ethanol,
and isopropyl alcohol) and traditional lubricants (i.e., Naphtha, and
Isopar G) as well as mineral oil and DI water (as negative controls) on
the PTFE are shown in Figure 2(D). As the plot indicates, ethanol had the
smallest contact angle. Finally, a simple evaporation test with ethanol,
methanol, and isopropyl alcohol as well as glycerol showed that, after
contacting with the PTFE surface for 3 min, ethanol had the fastest
evaporation rate of all three alcohols, with the evaporation rate of
methanol being the second fastest and isopropyl alcohol being the
slowest (cf. Figures 2[E–F]).

4.2 |. PTFE/ethanol mixing ratio

All PTFE/lubricant blends with different ethanol contents exhibited a


uniform mixture coloring, suggesting that ethanol was able to coat PTFE
powder particles evenly regardless of the amount of ethanol. However,
formation of large agglomerates and excessive powder particles on the
surface of the mixing container were found in the mixtures with over
30% ethanol. When the weight ratio of ethanol was reduced to between
15% and 25%, the large agglomerates could be broken up into smaller
ones by increasing the mixing time and gently tapping the container.
Accordingly, three batches of PTFE/ethanol mixtures with 15%, 18%,
and 25% ethanol were employed and further evaluated in the PTFE
preforming and paste extrusion experiments.

4.3 |. PTFE preforming and paste extrusion

The purpose of the paste preforming is to remove air from the


PTFE/ethanol mixture by providing a compaction force to the paste.
Under normal conditions, PTFE paste after preforming should be
compacted to around one-third of its initial heights.[1] The preforming
process helps the paste to density and maximizes paste material quantity
while ensuring a smooth paste extrusion and reducing defects.
After the preforming process, paste extrusion was conducted with
PTFE/ethanol mixtures with 15%, 18%, and 25% of ethanol to produce
PTFE extrudate in a tubular shape using the paste extruder. The
extrusion result of PTFE paste with 25% ethanol lubricant exhibited
defects such as weak structural integrity and large cracks. Excessive
lubricant will reduce the contact area among the PTFE powder particles,
thereby hampering the extrusion effectiveness and integrity of the
extrudate. The extrusion result of 18% ethanol lubricant paste showed a
clean and flawless structure with good structural integrity. No cracks or
fractures were observed. For PTFE paste with 15% ethanol lubricant, no
extrudate was collected due to the high-pressure build-up in the
extruder barrel. The maximum extrusion pressure of the machine was
reached shortly after the paste reached the conical section of the
extruder barrel. The significant reduction of cross-sectional area and
lower amount of lubricant in the paste required a pressure that exceeded
the current Instron machine pressure limit. Thus, the 18% ethanol
lubricant content was used for the rest of the extrusion experiments and
viscosity characterization.

4.4 |. PTFE expansion and ePTFE structure

With sufficient expansion rate and processing temperature close to but


below PTFE’s own melting temperature, PTFE can be expanded to form
ePTFE with strong a porous structure and high flexibility. Figure 3 shows
SEM image of the typical ePTFE structures and the definition of
internodal distance. After expansion, the structure of ePTFE is formed by
numerous small lumps of unexpanded PTFE domains called islands
(nodes) that are connected with fibrillated ePTFE nanofibers. The
internodal distance, L, is determined as the length of the ePTFE fiber
sandwiched between two islands. The fibers not only provide high
flexibility and softness, but also suitable condition for cell attachment
and nutrition transport in biomedical applications.w PTFE’s own mel
FIGURE 3
SEM image showing a typical ePTFE structure expanded at 400% with islands and
fibrillated ePTFE nanofibers. The internodal distance, L, is the length of the ePTFE fiber
between two islands. ePTFE, expanded polytetrafluoroethylene; SEM, scanning electron
microscopy [Color figure can be viewed at wileyonlinelibrary.com]

Figure 4 shows the SEM images of ePTFE with different expansion


ratios. Figure 4(A) presents SEM images of the ePTFE tube’s external
surface under different expansion ratios and Figure 4(B) presents SEM
images of the ePTFE tube’s internal surface under different expansion
ratios. Due to the conical paste extruder design and a large RR,
PTFE/ethanol paste travels at different speeds along the straight internal
surface and the conical external surface, resulting in different
morphologies on the internal surface and the external surface of the
ePTFE tubes.

FIGURE 4
SEM images of ePTFE on the (a) external surface and (B) internal surface of ePTFE under
different expansion ratios. Scale bars are all 100 μm. The arrows indicate the expansion
direction for all expansion ratios. SEM, scanning electron microscopy [Color figure can be
viewed at wileyonlinelibrary.com]
As it can be seen, the internodal distances increased as the expansion
ratio of the PTFE increased. This phenomenon was observed for both the
external and the internal surfaces of the ePTFE. The images in Figure
4 also show that the length of the ePTFE fibers gradually increased while
maintaining properly aligned and connecting fibers from 100% to 600%
expansion ratio. However, at 700% expansion ratio, a large amount of
disconnected ePTFE fibers were found on the surface of the ePTFE. The
disconnection of the ePTFE fibers indicated that the ePTFE tube started
to lose its structural integrity when the expansion ratio exceeded 600%.

4.5 |. Average internodal distance and distribution and fiber density


of ePTFE

The average internodal distance of each expansion ratio was calculated


from the SEM images of the samples (cf. Figure 3). The average fiber
length was calculated from the sum of all ePTFE fiber lengths divided by
the total number of the fibers. For example, the average of the three
fibers in Figure 3 was calculated as (L1 + L2 + L3)/3. In order to capture
the most representative fiber length average, fibers with lengths shorter
than 30% of the majority fibers were excluded from the averaging
calculation. Examples of those short ePTFE fibers are indicated by red
circles in Figure 3.

The average internodal distances on the external and internal surfaces of


each ePTFE sample under different expansion ratios were plotted
in Figure 5. The internodal distance increases with the expansion ratio,
and there is a difference between the external and internal intermodal
distance under the same expansion ratio. Based on the calculated results,
the 100% expansion ratio created an internodal distance ranging from 5
to 10 μm. As expansion ratio increased to 600%, an average intermodal
distance of 355 μm was observed on the external surface and 220 μm on
the internal surface. The fact that the ePTFE tube has a larger internodal
distance on the external surface compared to the interior surface was
caused by the force distribution in the conical section and different paste
velocity and movement during the paste extrusion process. The faster
paste movement on the internal surface resulted in a higher compaction
between PTFE powder particles and hence a shorter internodal distance
after expansion. This will be discussed in detail in the flow visualization
of PTFE paste extrusion section below.

FIGURE 5
Averaged internodal distance on the external and internal surfaces of ePTFE tubes under
different expansion ratios. ePTFE, expanded polytetrafluoroethylene [Color figure can be
viewed at wileyonlinelibrary.com]

Detailed statistical analysis on distribution of fiber lengths measured


using the SEM images was performed in order to investigate the
characteristics of island and fibril morphology of ePTFE after expansion.
The statistical results of internodal distance distribution on the external
and internal surfaces under different expansion ratios are shown
in Figure 6. From the statistical results, fiber length data from all the
expansion ratios exhibited a normal distribution. This phenomenon was
observed for both ePTFE external and internal surfaces. Differences in
the range of internodal distance were also observed between the
external and internal surfaces. By comparing Figures 6(A) and 6(B), the
range of internodal distance on the external surface was wider compared
to that on the internal surface, a phenomenon that was observed for all
samples under different expansion ratios. This is due to the different
compression forces, shear, and velocity the PTFE paste on the external
and internal surfaces experienced. The largest difference was observed
in the smallest expansion ratio at 100% and the range difference
gradually decreased when the expansion ratio increased. By monitoring
the force required during the expansion process, a relatively high
expansion force was always required to initialize the expansion process.

FIGURE 6
Statistical analysis of ePTFE internodal distance distribution. (A) ePTFE internodal distance
distribution on the external surfaces. (B) ePTFE internodal distance distribution on the
internal surfaces. ePTFE, expanded polytetrafluoroethylene [Color figure can be viewed
at wileyonlinelibrary.com]

Due to the unique island-fibrils morphology of ePTFE, a new quantity


called fiber density, D, defined as D = (L1 + L2 + L3) / L (cf. Figure 3), was
used to quantify the extent of fibrous regions on the ePTFE surface under
different expansion ratios. Figure 7 depicts the fiber density of ePTFE on
the external surface and internal surface under different expansion
ratios. Note that the ePTFE fiber density showed a rapid initial increase
from 100% to 200%, and then the increase lessened as the expansion
ratio increased. The fiber density reached its maximum value at the
expansion ratio of 500%, beyond which it appeared to drop slightly as
the expansion ratio reached 600%. This coincided with the onset of fiber
breakage and disconnection beyond an expansion ratio of 600%.

FIGURE 7
Fiber density of ePTFE on the external and internal surfaces under different expansion
ratios. ePTFE, expanded polytetrafluoroethylene [Color figure can be viewed
at wileyonlinelibrary.com]
4.6 |. Flow visualization of PTFE during paste extrusion

Figure 8 shows three different flow visualization samples, with layers of


blue ink used as flow trackers, prior to paste extrusion (Figure 8(A)) and
removed from the paste extruder during the paste extrusion stage
(Figures 8[B–C]). In particular, Figure 8(B) shows the deformed PTFE
paste right before the extrudate reached the extrusion die,
whereas Figure 8(C) shows the other sample after the paste started to
exit the extrusion die. A clear flow pattern of PTFE paste movement can
be observed based on the deformed flow tracker shape. It clearly
demonstrated that the PTFE paste moved relatively faster along the
inner surface (facing the mandrel rod), and slower along the outer
surface on the surface of the barrel. It is also evident that the ethanol-
lubricated PTFE paste exhibited slip along both the inner and outer
surfaces.
FIGURE 8
Ethanol-lubricated PTFE paste visualization result (A) prior to paste extrusion, (B)
deformed and sectioned PTFE paste right before it reached the extrusion die, and (C)
deformed and sectioned PTFE paste after it reached the extrusion die. PTFE,
polytetrafluoroethylene [Color figure can be viewed at wileyonlinelibrary.com]

4.7 |. Complete evaporation of ethanol from the extruded PTFE paste


FTIR was used to check the complete removal of ethanol through
evaporation 24 h after the paste extrusion. Before evaporation of
ethanol, characteristic peaks of ethanol showed up in the corresponding
FTIR curve (cf. Figure S1 in Supporting Information), including –OH at
3425.6 cm−1, C-H at 2974.5 cm−1, and C-O at 1055.8 cm−1. With proper
ventilation and enough wait time, all of these peaks disappeared from
the sample, which demonstrated ethanol could be easily removed
without compromising the processibility of ePTFE, demonstrating its
utility as an ideal lubricating agent for green fabrication of PTFE and
ePTFE.

4.8 |. Viscosity characterization and modeling

Using the Cox–Merz transformation, complex viscosity versus angular


frequency data was converted to viscosity versus shear rate data. Each
sample was tested using values of normal force ranging from 10 N to 40
N. The TA AR 2000 EX rheometer allowed for control of normal force
applied to the top plate within a certain tolerance range during the
frequency sweep. This was done by measuring the normal force with a
load cell and varying the gap between plates to keep the normal force
within the desired range. The normal forces were applied in ascending
order. It was hypothesized that higher normal forces would result in
higher apparent viscosities and less evidence of slip. The shear viscosity
results of compressed PTFE samples from the frequency sweeps are
shown in Figure 9 for the three sandpaper grits.
FIGURE 9
Shear rate versus viscosity data for the (a) 180-grit, (B) 120-grit, and (C) 60-grit sandpaper
disk. A strain amplitude of 1% was used. Dotted line is the power-law regression fit and
solid line in (C) is the Carreau model fit for 60 grit, 40 N normal force data (n = 3) [Color
figure can be viewed at wileyonlinelibrary.com]

As shown in Figure 9, each set of data exhibited a distinct drop in


viscosity from the first to the second datapoint. This was likely due to the
onset of slip that occurred at the second shear rate measured (~0.1 1/s).
Each dataset also showed deviation from expected power-law behavior
at higher shear rates, which indicated that increased slip was occurring.
This was consistent with expectations, as higher velocities would
increase the chance of the top plate losing traction with the PTFE paste
sample. Because of these commonalities, the data from each experiment
were evaluated using the following criteria: the percent drop from the
first datapoint to the second datapoint, and the R2 value when fitted to a
power-law regression. A lower percent drop indicated less slip. A higher
R2 value indicated a more accurate fit to power-law behavior and thus
less evidence of slip. For each geometry, the 40 N dataset exhibited the
least amount of deviation from power-law behavior; thus, it was chosen
to characterize the effectiveness of each grit.

Table 1 summarizes the evaluation criteria for the sandpaper


experiments. Because the 60 grit, 40 N dataset showed the lowest
percent drop and highest R2 value, it was chosen to create the viscosity
model. It should be noted that although the data in Figures 9(A), 9(B),
and 9 (C) look very similar, viscosity models created from different grits
and different normal forces other than the chosen 60 grit, 40 N normal
force produced noticeably different simulated extrusion pressure results.
Equation (3) and Table 2 show the Carreau model and the model
parameters derived from the data. The Carreau model fitted to the
experimental data is shown in Figure 9(C).

TABLE 1
Percent drop and R2 value for each grit of sandpaper using 40 N normal force
Sample (grit) Percent drop(%) R2 value
180 82.24 0.9828
120 68.26 0.9610
60 55.92 0.9939
Open in a separate window

TABLE 2
Carreau model parameters for the viscosity data obtained using the 60-grit, 40 N normal
force conditions on three samples

Carreau model parameter Value


Zero-Rate Viscosity (Pa-s) 9.771 × 107
Infinite-Rate Viscosity (Pa-s) 2278
Relaxation Time (s) 1952
Power Index 0.3508
Open in a separate window
Transient flow simulation

COMSOL Multiphysics was used to model the paste extrusion for a


simulated time of 20 min. Figure 10 shows the pressure distribution at
four time instants whereas Figure 11 shows a zoomed-in view of the
velocity profile through the cone and die at steady-state and comparison
with visualized paste movement. The flow profile within the cone and die
resembles a plug flow. There is a deviation from plug flow behavior at
the intersection of the cone and die, likely due to entrance effects caused
by the sharp corner in the geometry. However, the experimental
observation showed earlier acceleration of paste along the inner surface
of the extruder.
FIGURE 10
Simulated pressure distribution at various time instants. (A) t = 0 min. (B) t = 4 min. (C) t =
8 min. (D) t = 12 min [Color figure can be viewed at wileyonlinelibrary.com]
FIGURE 11
Zoomed-in view of the simulated velocity profile in the cone and die at steady-state and
comparison with visualized paste movement [Color figure can be viewed
at wileyonlinelibrary.com]

Simulated extrusion pressure over time was compared to experimental


extrusion pressure measured by an Instron machine. Figure 12 shows
the experimental pressure as compared to the simulation.
FIGURE 12
Experimental versus simulated extrusion pressure over time

As shown, the simulation under-predicted the steady-state pressure by


about 6 MPa. The rise behavior of each curve was also noticeably
different: the simulated pressure rose much more abruptly to its steady-
state value, whereas the experimental pressure rose much more slowly
and reached steady-state 3 to 4 min later. This was explained by the
assumption of incompressibility in the model. In reality, the PTFE paste
was slightly compressible: as the ram forced the paste into the cone and
die, it compressed slightly and thus allowed the pressure buildup to rise
slowly. Once the PTFE paste was fully compressed and the flow through
the die entered steady-state, the extrusion pressure reached a steady
value. In the simulation, however, the paste was modeled as an
incompressible fluid. This caused the pressure to increase sharply as
soon as the flow area was reduced within the cone and die. Another
reason for the discrepancy was due to the fact that the viscosity model
used in the simulation was based on a normal force of 40 N whereas the
paste extrusion was conducted at a much higher pressure range, leading
to under-prediction of the viscosity value used in the simulation.

Because of the slip wall assumption, the simulated velocity profile was
uniform across the r-direction, similar to a plug flow. The flow
visualization showed evidence of plug flow in wider regions of the cone
similar to the predicted velocity profile; however, in smaller regions of
the cone, a higher velocity was observed on the inner (mandrel) surface
of the extruder. This could be attributed to the convergent conical
geometry and the amount of slip on the walls and the geometry of the
cone. Overall, the flow visualization experiments provided valuable data
to validate the simulation and revealed areas that caused discrepancy,
such as compressibility and viscoelasticity of the paste and the necessity
of a more accurate viscosity measurement that reflects the real paste
extrusion operating conditions. Future viscosity modeling of PTFE paste
should also take into account the increasing fibrillation between PTFE
powder particles during paste extrusion that contributed to the increase
in extensional forces and hence the pressure drop.
Go to:

5 |. CONCLUSIONS

In this study, a new ethanol lubricant that is non-toxic and has a high
evaporation rate and good wettability to PTFE was evaluated and used
for PTFE/ePTFE fabrication and showed great performance. The
ethanol-lubricated PTFE paste was extruded and expanded under
different expansion ratios from 100% to 700%. The SEM images of the
ePTFE showed the unique island-fibril morphology and the averaged
fiber length in terms of internodal distance and fiber density increased
with the expansion ratio until it exceeded an expansion ratio of 600%
when the fibers started to break. The FTIR results confirmed complete
evaporation of ethanol after the paste extrusion.

Flow visualization on the paste extrusion was performed in order to fully


understand the flow profile and PTFE paste movement inside the
extrusion barrel. Colored layers as flow trackers were introduced on the
PTFE paste preform, and the paste movement was visualized by the
deformation of the flow trackers. A complete record of the paste
movement was revealed from the cross-section of the PTFE paste
removed from the extruder. The PTFE paste was found to move faster
along the inner surface of the extruder facing the mandrel rod and
slower on the outer surface. The speed difference of the PTFE paste was
caused by the converging conical section, which, in turn, resulted in
different expanded morphology between the internal and external
surfaces of the ePTFE tubes.

The transient flow simulation provided useful information on predicting


the order of magnitude of extrusion pressure. However, further work is
needed to attain accurate viscosity measurement and slip conditions for
better predictions of the velocity profile within the entire flow domain.
In addition, the model can be updated to properly account for the effect
of fibrillation on the paste flow. Once these limitations are addressed,
this model will be a powerful tool in predicting various transient flow
properties of different pastes and different cone and die geometries
without investing significant time and capital. Knowledge of the flow
properties within the extruder will shed light on the mechanical
structure of the extrudate and help greatly in the design process of small-
diameter vascular grafts.
Go to:

Supplementary Material
Supplementary Material
Click here to view.(267K, docx)
Go to:

ACKNOWLEDGMENTS
The authors would like to acknowledge the support and research facility
of the Wisconsin Institute for Discovery (WID) at the University of
Wisconsin–Madison as well as the support of the NHLBI of the National
Institutes of Health (grant number U01HL134655) and the Kuo K. and
Cindy F. Wang Professorship.

Funding information

National Institutes of Health, Grant/Award Number: U01HL134655;


Wisconsin Institute for Discovery
Go to:

Footnotes

SUPPORTING INFORMATION

Additional supporting information may be found online in the Supporting


Information section at the end of this article.

Go to:

REFERENCES
[1] Ebnesajjad S, Expanded PTFE Applications Handbook: Technology, Manufacturing
and Applications, William Andrew; 2017. [Google Scholar]
[2] Ying L, Fu Z, Wu K, Wu C, Zhu T, Xie Y, Wang G, Coating 2019. 9, 59. [Google
Scholar]
[3] Park BH, Kim SB, Jo YM, Lee MH, Aerosol Air Qual. 2012, 12(5), 1030. [Google
Scholar]
[4] Huang A, Kharbas H, Ellingham T, Mi HY, Turng LS, Peng XF, Polym Eng
Sci 2017, 57(5), 570. [Google Scholar]
[5] Zhou Y, He W, Wang N, Xu D, Chen X, He M, Guo J, Polym Eng Sci 2019, 59(8),
1593. [Google Scholar]
[6] Kitamura T, Kurumada K-I, Tanigaki M, Ohshima M, Kanazawa S-I, Polym Eng
Sci 1999, 39(11), 2256. [Google Scholar]
[7] Shadfar S, Farag A, Jarchow AM, Shockley WW, JAMA Otolaryngol. Head Neck
Surg. 2015, 141(8), 710. [PubMed] [Google Scholar]
[8] Aturaliya R, Wang D, Xu Y, Lin YJ, Li Q, Turng LS, ACS Appl. Mater.
Interfaces 2020, 12(34), 38241. [PubMed] [Google Scholar]
[9] Townsend N, Wilson L, Bhatnagar P, Wickramasinghe K, Rayner M, Nichols
M, European Heart J 2016, 37(42), 3232. [PubMed] [Google Scholar]
[10] World Health Organization, Cardiovascular diseases (CVD) Fact
sheet 2017. http://www.who.int/mediacentre/factsheets/fs317/en/.
[11] Benjamin EJ et al., Circulation 2019, 139(n10), e56. [PubMed] [Google Scholar]
[12] Pashneh-Tala S, MacNeil S, Claeyssens F, Present Future Tissue Eng Part B:
Rev 2016, 22, 1. [PMC free article] [PubMed] [Google Scholar]
[13] Humphrey JD, O’Rourke SL, An Introduction to Biomechanics, Springer, New
York: 2015. [Google Scholar]
[14] Mi H-Y, Jing X, Thomsom JA, Turng L-S, J Mater Chem 2018, 6(21), 3475. [PMC
free article] [PubMed] [Google Scholar]
[15] Wang D, Xu YY, Wang L, Wang XF, Ren C, Zhang B, Li Q, Thomson JA, Turng
LS, ACS Appl. Mater. Interfaces 2020, 12(26), 29844. 10.1021/acsami.0c07868.
[PubMed] [CrossRef] [Google Scholar]
[16] Wang D, Xu YY, Lin YJ, Yilmaz G, Zhang J, Schmidt G, Li Q, Thomson JA, Turng
LS, Biomacromolecules 2020, 21, 3807. 10.1021/acs.biomac.0c00897. [PubMed]
[CrossRef] [Google Scholar]
[17] Yan S, Xu YY, Lin YJ, Zhang Z, Zhang X, Yilmaz G, Li Q, Turng LS, Appl. Surf. Sci.
2020, 511, 145565. 10.1016/j.apsusc.2020.145565. [CrossRef] [Google Scholar]
[18] Patil PD, Ochoa I, Feng JJ, Hatzikiriakos SG, Non-Newtonian Fluid Mech J
2008, 153, 25. [Google Scholar]
[19] Luntz JF, Jaffe JA, Robb LE, Ind. Eng. Chem 1952, 44(8), 1805. [Google Scholar]
[20] Ariawan AB, Ebnesajjad S, Hatzikiriakos SG, Powder Technol. 2001, 121,
249. [Google Scholar]
[21] Patil PD, Ochoa I, Stamboulides C, Hatzikiriakos SG, Polastri F, Kapeliouchko
V, Appl Polym Sci 2008, 108, 1055. [Google Scholar]
[22] Ariawan AB, Ebnesajjad S, Hatzikiriakos SG, Polym Eng Sci 2002, 42(6),
1247. [Google Scholar]
[23] Ochoa I, Hatzikiriakos SG, Powder Technol. 2005, 153, 108. [Google Scholar]
[24] Mitsoulis E, Hatzikiriakos SG, J. Non-Newtonian Fluid Mech 2009, 157,
26. [Google Scholar]
[25] Ebnesajjad S, Fluoroplastics: Non-Melt Melt Processible Fluoropolymers, 2nd ed.,
Vol. 1, Elsevier, Amsterdam: 2014. [Google Scholar]
[26] COMSOL, COMSOL Multiphysics, The Level Set
Method 2019. https://www.comsol.com/forum/thread/attachment/37361/The-
level-set-methodfrom-MEMS-Module-5198.pdf.
[27] Hatzikiriakos SG, Ariawan AB, Ebnesajjad S, The Canadian J Chem Eng 2002, 80,
1153. [Google Scholar]

You might also like