You are on page 1of 11

Materials and Design 146 (2018) 249–259

Contents lists available at ScienceDirect

Materials and Design

journal homepage: www.elsevier.com/locate/matdes

Performance of biocompatible PEEK processed by fused deposition


additive manufacturing
M.F. Arif a, S. Kumar a,⁎, K.M. Varadarajan b,c, W.J. Cantwell d
a
Department of Mechanical and Materials Engineering, Khalifa University of Science and Technology, Masdar Institute, Masdar City, P.O. Box 54224, Abu Dhabi, United Arab Emirates
b
Department of Orthopaedic Surgery, Harris Orthopaedics Laboratory, Massachusetts General Hospital, 55 Fruit St, Boston, USA
c
Department of Orthopaedic Surgery, Harvard Medical School, A-111, 25 Shattuck Street, Boston, USA
d
Department of Aerospace Engineering, Khalifa University of Science and Technology, P.O. Box 127788, Abu Dhabi, United Arab Emirates

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Horizontally 3D printed PEEK with ras-


ter angles of 0o and 90o exhibits tensile
flexural and fracture properties compa-
rable to those of molded PEEK
• Specimens built vertically are prone to
delamination, exhibiting poorer me-
chanical performance due to high ther-
mal gradient in the build-direction
• Stick-slip fracture and lower Poisson’s
ratio are observed for specimens built
vertically, due to the presence of interfa-
cial voids.
• Minimizing thermal gradients across
beads is the key to producing parts
with excellent macroscopic properties

a r t i c l e i n f o a b s t r a c t

Article history: In this study, the tensile, flexural and fracture behavior of PEEK processed by fused filament fabrication (FFF) is
Received 13 December 2017 reported. Three different configurations, viz., specimens built horizontally with a raster angle of 0° (H-0°) and
Received in revised form 4 March 2018 90° (H-90°), and vertically with a raster angle of 90° (V-90°) are examined. The best performing specimen in
Accepted 5 March 2018
terms of its tensile, flexural and fracture toughness properties is H-0°, followed by H-90° and V-90°. The H-0°
Available online 7 March 2018
and H-90° specimens exhibit 85% and 75% of tensile and flexural strengths of molded bulk PEEK, respectively.
Keywords:
The fracture toughnesses of the H-0° and H-90° specimens are 78% and 70% of molded bulk PEEK, respectively.
Additive manufacturing However, V-90° specimens show lower tensile, flexural and fracture toughness properties compared to those
Fused filament fabrication (FFF) of H-0°, H-90° and molded PEEK. The fracture surface and microtomography analyses indicate that the degree
Biocompatible PEEK of interfacial bonding between beads during layer-by-layer buildup, is affected by the thermal gradient across
Fracture toughness the beads. The PEEK specimen configurations examined here have different thermal gradient in the build direc-
Interfacial bonding tions and such variations manifest themselves in their macroscopic mechanical behavior. The findings of this
Fused deposition modeling (FDM) study provide guidelines for FFF of PEEK to enable its realization in applications such as orthopedic implants.
© 2018 Elsevier Ltd. All rights reserved.

⁎ Corresponding author.
E-mail address: s.kumar@eng.oxon.org (S. Kumar).

https://doi.org/10.1016/j.matdes.2018.03.015
0264-1275/© 2018 Elsevier Ltd. All rights reserved.
250 M.F. Arif et al. / Materials and Design 146 (2018) 249–259

1. Introduction support the overhang portions, unlike FFF which requires supporting
parts. However, the inherent cost of SLS is high, owing to relatively ex-
Compared to established technologies, additive manufacturing pensive capital costs and high power laser source.
(AM) offers the possibility of cost-effective automation of the fabrica- Recently, FFF has been promoted as a cost-effective alternative to SLS
tion process, it enables a digital inventory, and provides greater flexibil- for the fabrication of high-performance polymers, such as PEEK.
ity to locally design the material architecture in three-dimensions [1–4]. Valentan et al. [25] developed FFF systems to fabricate two grades of
Fused filament fabrication (FFF), also known as fused deposition model- PEEK-Optima®. The strength of the FFF samples in their study was ap-
ing (FDM) is the most commonly used AM technique for thermoplastic proximately half of the tensile strength of molded PEEK (~100 MPa).
polymers. In the FFF process, a thermoplastic filament is fed into a Vaezi and Yang [26] developed the FFF system to fabricate PEEK-
heated nozzle, melted or liquefied, and subsequently extruded and de- Optima® samples. A nozzle temperature of 400–430 °C, an ambient
posited onto a build plate. A gantry moves the nozzle in the horizontal temperature of 80 °C and a build plate temperature up to 130 °C were
x − y plane as the molten material is deposited. The build plate then identified as optimal parameters. The FFF process with 100% infill den-
moves vertically (in the z axis) after completing the deposition in the sity results in a sample porosity of 14% and a tensile strength of
x − y plane. The deposited layers solidify and bond/weld with adjacent 75 MPa. Wu et al. [27] showed that the warping deformation of FFF-
layers, forming the desired 3D geometry. FFF has been used in the AM of PEEK samples reduces with increasing chamber and nozzle tempera-
numerous thermoplastics. The most widely used polymer for FFF is ac- ture. Wu et al. [28] evaluated the influence of layer thickness (200,
rylonitrile butadiene styrene (ABS) [5,6] and polylactic acid (PLA) [7]. 300 and 400 μm) and raster angle (0°/90°, 30°/−60° and 45°/−45°)
Moreover, high impact polystyrene [8], Nylon [8], Ultem® [9], polycar- on the mechanical properties of FFF–PEEK samples. A maximum tensile
bonate [8], poly(vinyl alcohol), polypropylene [10] and the mixtures strength of 56.6 MPa was achieved for samples with a 300 μm layer
of any two thermoplastics [11] or reinforced thermoplastic composites thickness and a 0°/90° raster angle. Cicala et al. evaluated the tensile
[12,13] are suitable for FFF. Nanocomposite feedstocks, such as properties of FFF-PEEK samples with raster angle of 0°/90°, 30°/60°
graphene [14,15] or CNT [16] reinforced filaments, are currently and 45°/−45° and found an average tensile strength of 69.04 MPa and
attracting interest, due to their superior thermal, electrical and mechan- a tensile modulus of 3.53 GPa for all tested samples [29]. Yang et al.
ical properties. [30] studied the effect of ambient temperature, nozzle temperature
Processing parameters, material type and geometrical features were and post heat-treatments on crystallinity and tensile properties of
found to influence the final properties of the products made by FFF. For FFF–PEEK with a 0° raster angle. In general, a higher ambient tempera-
optimal properties of components or devices manufactured by FFF, tem- ture, higher nozzle temperature and a higher post treatment tempera-
perature control is critical. The distribution and variation of temperature ture and time resulted in higher crystallinity, strength and stiffness.
and residual stresses in structures or parts produced by FFF must be op- The maximum tensile strength and modulus were found to be 84 MPa
timally controlled so as to ensure sufficient internal stress relaxation and 4 GPa, respectively.
and a good bonding quality between the adjacent polymer beads. For The aforementioned studies mostly focused on the development of
example, warping and interlayer delamination are frequently observed new FFF machines to accommodate the high processing temperature
due to the residual stresses accumulated during the layer-by-layer build of PEEK, and to evaluate the effect of process parameters on the final
up. These stresses originate from the thermal gradient generated during properties of the samples. However, for the latter, the analysis of the
the solidification process that influence the layer's volume contraction, mechanical performance of the fabricated parts is presented with little
warping and bonding/diffusion quality between beads [17,18]. This detail, especially on how the spatial thermal gradients during the layer
has shown to pose important challenges particularly for processing of by layer build-up of the sample, can influence the properties. In the cur-
semicrystalline polymers as a high degree of crystallinity and high rent study, the tensile and flexural properties of samples built horizon-
shrinkage coefficient can negatively affect the dimensional stability of tally (x − y plane) with raster or bead angles of 0° and 90°, and vertically
the part [19]. By using a heated build plate and by controlling the tem- with a bead angle of 90° with respect to loading direction are evaluated.
perature of the 3D printer chamber, thermal gradients and their effects, These three configurations are associated with different thermal histo-
such as warping can be minimized [20,21]. ries during FFF-PEEK part build up. Digital image correlation (DIC) anal-
In recent years, there has been a growing interest in the fabrication ysis is used to measure the axial and lateral strains in the gauge length
of high-performance polymers such as polyphenylsulfone, polyether zone of dogbone specimens. Moreover, the fracture behavior of FFF–
ether ketone (PEEK), polyether ketone ketone (PEKK), etc. via AM due PEEK samples is also investigated. The fracture study is important,
to their outstanding mechanical properties and chemical stability. since formation of micro-cracks, particularly in the diffusion/bonding
PEEK is an expensive thermoplastic having excellent mechanical and zones between beads, could lead to premature failure of fabricated
chemical resistance properties even at temperatures up to 240 °C. parts during their service life. To date, the fracture behavior and inter-
PEEK also exhibits excellent hydrolysis resistance and provides fire, layer cohesion studies of FFF fabricated samples are limited to ABS
smoke, and toxicity performance. PEEK is one of the few polymers con- [31,32] and PLA [33,34] through single edge notch bending (SENB) or
sidered for metal replacement in high temperature applications. PEEK double cantilever beam (DCB) testing configurations. In this work, the
has been used in automotive, aerospace, oil & gas and space applica- fracture behavior of FFF–PEEK specimens is evaluated for the first time
tions. This study focuses on FFF of PEEK for bioimplants in order to effec- using compact tension tests for the same fabrication configurations as
tively utilize its properties such as biocompatibility, excellent fatigue those used in tensile and flexural testing.
and wear resistance for such applications [22]. PEEK is typically fabri-
cated by an injection molding or extrusion process. AM of high perfor- 2. Experimental methods
mance PEEK in particular, holds tremendous promise for biomedical
applications, such as for design of novel knee replacement implants The configurations used to fabricate the specimens by FFF are
that more closely mimic native physiology. Effort has been made to ad- discussed, followed by tensile, flexural and fracture toughness testing,
ditively manufacture PEEK via the selective laser sintering (SLS) method including optical strain mapping with DIC.
[23]. Oxford Performance Materials, Inc. has developed OsteoFab® tech-
nology for 3D printing of facial and cranial implants for bone replace- 2.1. FFF configurations
ment using PEKK powders via SLS process [24]. SLS uses a high power
laser to selectively sinter powders, which causes them to coalesce to- FFF specimens were fabricated using an Indmatec HPP 155 device
gether due to interfacial and free surface diffusion. SLS can be used to (Apium Additive Technologies GmbH). This FFF device is able to fabri-
fabricate very complex geometries because it uses the same powder to cate samples with a maximum build volume of 135 × 145 × 148 mm3.
M.F. Arif et al. / Materials and Design 146 (2018) 249–259 251

Filaments with 1.75 mm diameter made of Victrex® PEEK 450G were brim. Brim is an outline to create a fabricated large ring that surrounds
used. The filament exhibited glass transition and melting temperatures the part and is attached to it. This holds the edges of the FFF fabricated
of 153 °C and 340 °C, respectively, based on the 2nd heating cycle data part and thus helps to improve bed adhesion and prevent warping.
from a differential scanning calorimetry (DSC) investigation with a After fabrication, the chamber was cooled to room temperature by nat-
heating rate of 10 °C/min. These transition temperatures closely corre- ural convection and the parts were then removed from the build plate.
spond to those found in the Victrex® PEEK 450G database [35]. The fil- The brim structure surrounding the specimen was detached by hand.
ament was fed into a 0.4 mm diameter nozzle by a feeding pressure Three FFF configurations for the tensile, flexural and compact tension
mechanism via a driver motor and a counter-rotating set of grooved specimens used in this work are depicted in Fig. 1.
gears. The Simplify3D software package was used to determine the slic- One specimen for each of the configurations was produced per fab-
ing sequence and define the FFF process. The FFF process parameters rication cycle, except for V-90°, tensile and flexural specimens. Four
used in this study were as follows: and three V-90° tensile and flexural specimens were fabricated respec-
tively in a single fabrication sequence. The molten polymer in the V-90°
• Nozzle movement speed: 800 mm/min; first layer: 300 mm/min dogbone and flexural samples is deposited onto a small cross-sectional
• Nozzle temperature: 410 °C; first layer: 390 °C area, as seen in Fig. 1b and thus the nozzle travel time in the x − y plane
• Bed temperature: 100 °C. is very short prior to moving in the z direction. This can induce a fabri-
• Layer height: 0.1 mm; first layer: 0.18 mm cation failure, since the molten polymer is deposited onto a layer that
• Extrusion width: 0.48 mm is still in a molten state. To avoid this, we increased the number of spec-
• Infill pattern: Rectilinear imens so that the fabrication cycle could be done successfully, while en-
• Infill density: 100%. suring acceptable degree of adhesion between the beads normal to the
loading direction (z-direction). A reduced number of specimens re-
sulted in fabrication failure while increasing the number of specimens
The nozzle movement speed, nozzle temperature and layer height might increase the thermal gradient between beads in the vertical di-
for the first layer were set differently in order to enhance adhesion be- rection, ΔT Δz , due to the longer travel time in x − y plane, leading to a
tween the sample and the glass build plate. Adhesion was further en- weak bonding. The length of the clamping zone of V-90° sample was re-
hanced by applying a glue on the plate surface and by the addition of duced to reduce the fabrication time, as seen in Fig. 2. H-90° samples

a) q b)

r
p s

c) d) e)
p p
r

q q
s

Horizontally fabricated in Horizontally fabricated in Vertically fabricated in direction,


plane, 0o beads orientation w.r.t plane, 90o beads orientation w.r.t 90o beads orientation w.r.t loading
loading direction loading direction direction
(H-0o) (H-90o) (V-90o)

Fig. 1. AM of PEEK by FFF: FFF configurations of tensile and compact tension specimens. The specimens were fabricated either a) horizontally or b) vertically. The deposition pattern of
specimens fabricated horizontally with either c) 0° beads orientation or d) 90° beads orientation, and e) specimens fabricated vertically with 90° beads orientation. Sub-figures p and q
in c) and d) illustrate the magnified view of the corresponding regions p and q in a). Similarly, r and s in e) illustrate the magnified zone of corresponding regions r and s in b).
Flexural specimen configuration is similar to that of tensile specimen.
252 M.F. Arif et al. / Materials and Design 146 (2018) 249–259

V-90o

H-90o

H-0o

Fig. 2. FFF-PEEK tensile specimens.

were fabricated without outside perimeter/outline in order to avoid 2.3. Flexural testing
beads with opposing orientations. In H-0° sample, the polymer melt
was initially deposited in the x − y plane following the outline contour Three point flexural tests were performed according to ISO 178 at a
of the dogbone sample and was continued until completely printing the constant crosshead speed of 2 mm/min at ambient temperature (~20
entire gauge length zone, leaving the gap only at the central areas of the °C) on a Zwick-Roell Z005 universal testing machine (UTM) with a
clamping zone. These clamping zones were then fabricated with beads 2.5 kN load cell. The H-0° specimens slipped from the supports after
of 45°/−45° orientation to ensure sufficient strength in the transition the maximum load had been reached but didn't exhibit final failure.
zone and thus avoiding sample failure in this zone. Therefore, testing for H-0° was stopped when the load had reduced by
20% from the maximum. Each of the specimen configurations was
tested three times to ensure repeatability.
2.2. Tensile testing
2.4. Fracture tests
Tensile tests were performed according to ISO 527 at a constant
crosshead speed of 1 mm/min at ambient temperature (~20 °C) on a Fracture tests were conducted to evaluate the Mode I fracture tough-
Zwick-Roell Z005 universal testing machine (UTM) with a 2.5 kN load ness of the 3D printed PEEK. The compact tension fracture properties of
cell. The load cell accuracy was b±0.25% for a load N 10 N and b±1% FFF-PEEK specimens were evaluated following ASTM D5045-14 stan-
for a load range of 2.5 to 10 N, and crosshead travel resolution was dard for measuring the plane strain fracture toughness of polymers.
0.041 μm. DIC technique was used for full-field measurement of strain Compact tension samples were fabricated in the configuration shown
on the specimens. Random black–white speckle patterns of acrylic in Fig. 3a, where W = 13–15 mm. After fabricating rectangular blocks,
paint were sprayed on the specimen using an airbrush prior to testing. holes for loading pins (∅ 3 mm) were drilled using a benchtop drilling
Consecutive speckle images were acquired as a function of load using machine. A notch was introduced by a high precision circular saw of
a monochrome 5.0 MP camera for strain evaluation. The average engi- 152.4 mm diameter and 0.508 mm thickness. A sharp notch was subse-
neering strain in the axial and lateral directions over the gauge length quently generated by a slow tapping motion with a fresh razor blade.
zone was evaluated using Vic-2D software. These strain values were These specimens were manufactured without either holes or a notch,
then used to construct the engineering stress–engineering strain to ensure a consistent thermal history. The FFF of the holes and the ini-
curve. The tensile properties of the samples were extracted from the tial cracks would clearly require a change in the deposition patterns of
stress-strain curve. Each of the specimen configurations was tested at the filament beads, leading to a change in specimen's thermal history.
least three times to assess repeatability. The FFF–PEEK tensile Prior to testing, the samples were kept in oven for 12 h at 70 °C to re-
specimens are shown in Fig. 2. move moisture. A crosshead speed of 1 mm/min in a Zwick-Roell

a) b) c)

3‒15 mm

Fig. 3. a) Geometry of the compact tension specimen, b) FFF-PEEK compact tension specimen and c) magnified view of the sharp crack tip region of the specimen prior to testing.
M.F. Arif et al. / Materials and Design 146 (2018) 249–259 253

a) b)
A B C D
A
D (%) (%) (%) (%)
H-0o
B 4.50 3.00 0.40 25.50
4.05 2.70 0.36 22.95
H-90o 3.60 2.40 0.32 20.40
3.15 2.10 0.28 17.85
2.70 1.80 0.24 15.30
2.25 1.50 0.20 12.75
1.80 1.20 0.16 10.20
1.35 0.90 0.12 7.65
0.90 0.60 0.08 5.10
C 0.45 0.30 0.04 2.55
V-90o 0.00 0.00 0.00 0.00

Fig. 4. Tensile properties of the FFF-PEEK tested in different directions corresponding to the different build directions: a) Representative stress-strain curve of H-0°, H-90° and V-90°, insert
shows extended ×-axis of H-0° specimen, b) axial strain contour at point A, B, C and D in a).

Z005 UTM was used. Critical stress intensity factor KIC was determined and tensile strength. The H-90° offered a Young's modulus and tensile
following the method given in ASTM D5045. The adopted scheme as- strength values that were 7% and 12% lower than those of H-0°, respec-
sumes a linear elastic fracture behavior of compact tension specimen tively. The H-90° specimens exhibited properties approaching those of
and therefore restriction on linearity of load-displacement response the H-0° specimen, due to the excellent interfacial bonding between
was imposed. The 5% secant line criterion of the load-displacement beads. The H-90° had a short bead deposition path in the horizontal
curve was employed to ensure validity of the KIC, as described in plane as the nozzle travelled across the width of the specimen, as seen
ASTM D5045 Section 9.1.1. Moreover, the specimen size was appropri- in Fig. 5a. Consequently, the molten polymer diffused and bonded well
ately chosen to ensure a plane strain-state at the crack tip. The size cri- with adjacent beads, since the deposited beads were still in a molten
terion stipulates that the value of 2.5(KIC/σy)2 must be less than the state. The H-0° samples had a longer deposition path in the horizontal
specimen thickness, crack and ligament lengths. In the current study, plane as the nozzle travelled across the length of the specimen (see,
the yield strength (σy) of the molded PEEK was used as a reference Fig. 5b). It is likely that the bead was deposited when the neighboring
due to different thermal histories associated with 3D printed tensile bead had been solidifying, resulting in poorer interfacial bonding be-
and compact tension specimens. Moreover, introducing a sharp crack tween beads compared to that of the H-90°. Even though the nozzle
exactly terminating at the bead-to-bead interface is practically travel speed was constant, the build-rates were different for H-0° and
impossible. H-90° configurations due the different deposition paths. It can be seen
from Fig. 5 that the global in-plane (x − y) build direction for H-90°
sample was in the direction of specimen's length, while the build direc-
3. Results and discussion
tions for H-0° and V-90° samples were in the direction of specimen's
width and thickness, respectively. Therefore, the build-rate in x direc-
3.1. Tensile and flexural properties
tion for H-90° samples, V x ¼ Δx Δt and the build-rate in y direction for H-

The tensile performance of specimens fabricated in three different 0° and V-90° samples, V y ¼ Δy
Δt . The build-rate in z direction for H-0°,
configurations was evaluated. According to ISO 527-1:2012, tensile H-90° and V-90° samples, V z ¼ Δz Δt . As seen in Table 2, the global in-
strength is the stress at which the first local maximum is observed dur- plane build-rate for H-90° sample (62.5 mm/min) was higher than
ing a tensile test. Therefore, the tensile strength of the H-0° sample was that of H-0° sample (4.74 mm/min) and hence lower thermal gradient
evaluated at the yield point while those of the H-90° and V-90° speci- leading to superior interfacial bonding between beads in x − y plane.
mens were evaluated at the failure point, as seen in Fig. 4. The H-0° It should be noted that Vy for H-0° specimen was evaluated without con-
specimen exhibited the highest Young's modulus and tensile strength, sidering the build-speed in the middle portion of the clamping zones
followed by the H-90° and V-90° (Fig. 4 and Table 1). In the H-0° spec- that had alternating 45°/−45° bead orientations.
imen, the tensile loading direction is parallel to the orientation of the fil- The difference in thermal history between H-0° and H-90° speci-
aments or beads and thus this sample showed higher Young's modulus mens can be seen qualitatively from the specimen color (see, Fig. 2).

Table 1
Mean values and standard errors for the tensile and flexural properties of H-0°, H-90° and V-90° specimens. Tensile elongation is evaluated at break. The properties of molded Victrex®
PEEK 450G [35] are presented for comparison.

H-0° H-90° V-90° PEEK 450G

Tensile properties
Young's modulus, Et (GPa) 3.80 ± 0.03 3.54 ± 0.05 3.03 ± 0.01 4.0
Tensile strength, σm (MPa) 82.58 ± 1.03 (at yield) 72.88 ± 1.92 (at break) 9.99 ± 0.94 (at break) 98 (at yield)
Tensile elongation, εb (%) 110.97 ± 5.31 2.91 ± 0.14 0.33 ± 0.03 45
Poisson's ratio, ν 0.44 ± 0.02 0.43 ± 0.01 0.37 ± 0.005 –

Flexural properties
Flexural modulus, Ef (GPa) 3.08 ± 0.22 3.06 ± 0.21 2.54 ± 0.07 3.8
Flexural strength, σf (MPa) 142.0 ± 5.56 124.3 ± 7.98 16.40 ± 2.18 165
254 M.F. Arif et al. / Materials and Design 146 (2018) 249–259

a) deposition path
c)

b)

deposition path
Δy
Fig. 5. Schematic of the thermal gradient and build-rate during FFF-PEEK process in a) H-90°, b) H-0°, and c) V-90° samples. Vx ¼ Δx
Δt is the build-rate in x for H-90° samples. Vy ¼ Δt is the
build-rate in y for H-0° and V-90° samples. Vz ¼ Δz
Δt is the build-rate in z for H-0°, H-90° and V-90° samples.

The gauge length zone of the H-0° specimen with a higher thermal gra- is observed, as seen in Fig. 6c. It can also be seen clearly from the SEM
dient between beads in the x − y plane, and thus weak interfacial bond- images of the H-0° and H-90° specimens where the weak interface
ing, had a brownish color. While the H-90° specimen that had a lower can be observed between the beads in vertical fabrication direction
thermal gradient between beads in the x − y plane, and thus better in- (Fig. 6a and b). It should be noted that, the time to build a single layer
terfacial bonding between beads, had a more pale or whiter color. This in the horizontal plane (x − y) of H-0° and H-90° samples was
color was relatively similar to that of the clamping zone for the H-0° 1.52 min, which was approximately five times higher than that of the
specimen which was purposely fabricated with alternating 45°/−45° V-90° samples (0.28 min), due to larger x − y deposition area compared
bead orientations to have a shorter deposition path, and to avoid failure to that of V-90° samples. Therefore, the H-0° and H-90° samples had
in the transition or clamping zone during tensile testing. Moreover, the lower build-rate (0.07 mm/min) in vertical direction than that of V-
surface finish can be used as an indicator of the degree of diffusion be- 90° samples (0.36 mm/min), as seen in Table 2, and thus larger thermal
tween beads. As seen from Fig. 2, the surface finish of the H-90° speci- gradient in z direction. Therefore, we expect poorer interfacial bonding
men was relatively rough. Since the molten polymer was deposited in the vertical direction for H-0° and H-90° samples. However, the effect
next to a bead that was still in its molten state (due to the short deposi- of thermal gradient in the z direction has less influence on the mechan-
tion path), the beads flowed more easily, forming an overlap region be- ical properties of the H-0° and H-90° samples since the weak interfaces
tween beads. In contrast, the H-0° specimen had a relatively smooth between beads in the z direction do not experience normal traction
surface, since the molten polymer was deposited next to a bead that under tensile loading. A heating zone at the top of the printer's chamber
had started to solidify. However, even though the interfacial bonding has been installed to maintain the top surface temperature of the sam-
in H-90° was better than that in H-0° specimen, the H-0° specimen ex- ple and to regulate the chamber temperature, with temperature control
hibited superior properties, since the macroscopic performance was pri- being a proprietary technology of the manufacturer. Though it ensures
marily due to the strength and strain tolerance of the filament, as that the top surface of the sample remains heated, it cannot fully elimi-
discussed earlier. nate the low interfacial bonding in the z direction as the minimum ther-
The interfacial bonding between beads is critical, especially between mal gradient between beads is restricted by the fact that the preceding
the layers in the vertical direction, particularly for V-90° specimens
where the loading direction is normal to the interface between beads Table 2
in the z direction (see, Fig. 5c). A lower thermal gradient between Build-rate (Vi) in x, y and z directions, and the deposition time (txy) in a single horizontal
beads in the z direction required for promoting interfacial adhesion in plane during the FFF-PEEK process.
vertically-fabricated specimens was restricted by the fact that the pre- Vx Vy Vz txy
ceding layer should be fully or partially solidified so as to lay the (mm/min) (mm/min) (mm/min) (min)
succeeding layer properly. Therefore, four tensile specimens of V-90°
Tensile specimens H-0° – 4.74 0.07 1.52
configurations were fabricated simultaneously to increase the deposi- H-90° 62.5 – 0.07 1.52
tion time in the horizontal plane (x − y), and thus reducing the build- V-90°a – 12.64 0.36 0.28
rate in the vertical (z) direction and avoiding the fabrication failure. Re- Flexural specimens H-0° – 5.68 0.05 1.9
H-90° 42.11 – 0.05 1.9
sults show that the tensile strength and modulus of V-90° samples were
V-90°b – 10.96 0.27 0.33
significantly lower compared to those of H-0° and H-90° samples, as Compact tension H-0° – 27.63 0.16 0.58
seen in Fig. 4 and Table 1. These results indicate incomplete sintering specimens H-90° 24.81 – 0.16 0.58
between layers in the vertical direction due to unavoidable large ther- V-90° – 26.17 0.4 0.25
mal gradient in the vertical fabrication direction (z). This can be seen a
Four specimens in one fabrication batch.
from the SEM image of V-90° sample where brittle-like fracture surface b
Three specimens in one fabrication batch.
M.F. Arif et al. / Materials and Design 146 (2018) 249–259 255

a) b) c)

Fig. 6. SEM fracture surface morphology of a) H-0°, b) H-90° and c) V-90° specimens. The H-0° specimen was evaluated at 110% strain to failure and thus lower in cross sectional area and
larger macroscopic void. Red, yellow and white (arrows and circles) indicate the z-fabrication direction, the raster angle and the load direction, respectively. The circle indicates the
direction normal to the surface.

layer should be fully or partially solidified so as to lay the succeeding tensile fracture surface, as seen in Fig. 6a and b, and also from the SEM
layer. images of fractured surface of compact tension H-0° and H-90° samples
To further confirm the interfacial defects and morphology in the z di- (see, Fig. 11a and b). In fact, some interfacial defects were found in the
rection, micro-computed tomography (μCT) images of the H-0°, H-90° μCT H-0° and H-90° samples. However, the number of defects observed
and V-90° tensile samples under no-load have been acquired via Phoe- was small and the defects were randomly scattered in a few zones of the
nix Nanotom® m system. μCT analysis is important to complement the sample volume. It is interesting to note that our H-0° and H-90° samples
SEM observations since it provides details of 3D microstructure and en- possess qualitatively a very low void fraction, although FFF process
ables visualization of cross-sectional 2D image from the 3D volume, tends to produce large gaps or voids between deposited beads. A low
though it has lower resolution than that of the SEM images. As seen in void fraction was achieved in this study by choosing a 100% infill den-
Fig. 7a and b, the interfacial defects between the beads in the vertical di- sity, a small layer height (0.1 mm) and high nozzle temperature (410
rection in H-0° and H-90° samples cannot be seen as the void dimen- °C) that allowed an optimum viscous flow of the molten polymer. More-
sions were mostly below the μCT voxel size or resolution (3 μm). over, the vertical build-rate for the H-0° and H-90° samples was lower
However, the presence of interfacial defects between beads in the verti- (0.07 mm/min) due to the large fabrication area in the x − y plane
cal fabrication direction was confirmed earlier from the SEM images of and thus the preceding layer had sufficient time to solidify. Therefore,

a) c)

1 mm

d)
voids

b)

1 mm

Fig. 7. μCT images of FFF-PEEK samples at no-load; a) 2D image of H-0° sample, the insert shows voids at cross junctions between beads, b) 2D image of H-90° sample, c) 2D image of V-90°
tensile sample, the insert shows 3D reconstructed image of voids, d) 2D image of V-90° compact tension sample, the insert shows 3D reconstructed image of voids. Red, white and yellow
(arrows or circles) respectively indicate z-fabrication direction, load direction and raster angle. The circle indicates direction normal to the surface.
256 M.F. Arif et al. / Materials and Design 146 (2018) 249–259

Table 3
Mean values and standard errors of the fracture toughness (KIC) of the FFF-PEEK
H-0o samples.

H-90o KIC (MPa·m1/2)

H-0° 5.32 ± 0.17


H-90° 4.76 ± 0.39
V-90° 0.97 ± 0.03
Molded PEEK
Karger-Kocsis and Friedrich [37] 7
Beguelin and Kausch [38] 6.75
Gensler et al. [39] 6.8
Rae et al. [40] 6.87

V-90o Samples loaded below the yield point (for H-0°) and below failure
point (for H-90° and V-90°) showed uniform axial strain along the
gauge length zone, as can be seen in Fig. 4b, with the maximum stan-
dard deviation in longitudinal strains of 0.15%, 0.12% and 0.019%, re-
spectively. This indicates that the samples were macroscopically
Fig. 8. Flexural stress-strain curves for the H-0°, H-90° and V-90° FFF-PEEK specimens.
uniform, even though the samples were built layer-by-layer. The sam-
ple loaded beyond the yield point for H-0° didn't exhibit a uniform
axial strain along the gauge length, with the maximum strain occurring
in the middle portion due to its excessive straining. The Poisson's ratio
the succeeding horizontal layer was deposited onto a surface that had of the H-0° and H-90° samples were approximately the same (see the
already solidified, inducing a stable deposition of the polymer melt mean values and their corresponding standard error, in Table 1).
and thus preventing a void formation at the interface. Concomitantly, Much lower Poisson's ratio was observed in V-90° sample and this
this led to weak interfacial bonding due to high thermal gradient be- could be attributed to two factors, namely, the bead orientation and
tween beads along the vertical direction. Different interfacial defects the proliferation of voids during loading. The tensile loading in V-90°
and morphologies were observed for V-90° tensile samples under no- specimen is normal to numerous weak bead-to-bead interfaces. This re-
load, and those results were further confirmed from the V-90° compact sults in significant void growth during loading and thus exhibits lower
tension samples. Voids at the interfaces between beads were observed Poisson's ratio.
in both tensile and compact tension V-90° samples, as seen in Fig. 7c The flexural properties of the FFF-PEEK were also evaluated. It was
and d. As discussed earlier, the deposition of molten polymer in V-90° found that the flexural properties followed the same trend as those of
configuration occurred on the surface of the preceding horizontal tensile properties, as the H-0° specimen had the highest flexural
layer that was not fully solidified in order to promote an acceptable de- strength and modulus, followed by the H-90° and V-90°, as seen in
gree of interfacial adhesion between the beads in the vertical fabrication Table 1 and Fig. 8. This is due to nearly the same build-rate and deposi-
direction, without inducing a fabrication failure. Consequently, good ad- tion time in x − y plane for tensile and flexural samples, as seen in
hesion between beads was achieved over a certain area of the interface, Table 2. The flexural properties reported here confirms the observation
but voids were created primarily due to undulations of the molten layer of the effect of printing configurations on tensile properties discussed
being laid on a partially solidified preceding layer, and shrinkage of the earlier. Both tensile and flexural strength of H-0°, H-90° and V-90° spec-
polymer. Inserts in Fig. 7c and d show the 3D reconstructed image of V- imens were 15%, 25% and 90% lower than those of molded PEEK 450G,
90° tensile and compact tension samples, respectively, which describe respectively (Table 1). However, the tensile and flexural moduli didn't
more clearly the voids formed at the interface between the beads in exhibit similar levels of reduction compared to the PEEK 450G. The ten-
the vertical fabrication direction. It should be noted that, these voids sile and (flexural) moduli of the H-0°, H-90° and V-90° specimens were
cannot be seen clearly from the SEM images of tensile and compact ten- 5% (19%), 11% (19%) and 24% (33%) lower than those of molded PEEK
sion samples in Figs. 6c and 11c. These images were taken on the 450G, respectively. This is due to the fact that the interface zones of
interfacially debonded surfaces between beads in the vertical fabrica- the layered structure of the FFF-PEEK samples provide strain-tolerance
tion direction. Voids were mostly developed between the undulating to the system (due to their lower compliance) under bending load
beads normal to surface view and thus they cannot be seen clearly. and thus lower the flexural modulus of the samples. The Young's mod-
Some voids were also observed in H-0° samples particularly at the ulus and tensile strength of the H-0° and H-90° specimens were similar
cross-junction between beads, as seen in insert of Fig. 7a. It shows an av- to those reported elsewhere [30] or indeed higher [25,26,28,29] than
erage bead dimension of 0.5 × 0.1 mm2, which corresponds to a layer the previously reported FFF-PEEK samples. The flexural modulus and
height of 0.1 mm and extrusion width of 0.48 mm. strength of the H-0° and H-90° specimens were higher than the

a) b) c)

Fig. 9. Load-displacement curve of FFF PEEK with a) H-0°, b) H-90° and c) V-90° specimens.
M.F. Arif et al. / Materials and Design 146 (2018) 249–259 257

almost the same and thus we expect similar degree of interfacial bond-
ing between beads in x − y plane for both samples. For the ASTM D5045
standard adopted here, it is most likely that the sharp tip of the intro-
duced crack in H-0° sample is terminating well within the bead (but
not exactly at the interface between beads) and thus the fracture tough-
ness value estimated doesn't fully represent the fracture behavior of lay-
ered structure of FFF-PEEK. Such limitation was also encountered for H-
90° and V-90° samples as it was difficult to generate a sharp crack ex-
actly terminating at the interface between beads. Regardless of these
limitations, it can be seen that the trend of KIC was almost the same as
those observed for the tensile and flexural tests. Results indicate that
the bead orientation with respect to the loading direction contributes
the most to the mechanical performance, followed by the interfacial ad-
hesion between beads. Interfacial adhesion is affected by the thermal
gradient dictated by the in-plane and out-of-plane build-rates (see
Table 2).
The crack tip in H-90° and V-90° compact tension specimens exhib-
ited a higher probability of advancement through the bead-bead inter-
face where the interfacial zone between beads acted as the weakest
link, unlike the H-0° specimen wherein the crack tip advanced normal
to the bead orientation (see, Fig. 10). The crack tip in H-0° specimen re-
quired more energy to break the beads transversely, resulting in higher
fracture toughness. This effect can also be seen on the fracture surface,
Fig. 10. Photograph of fractured compact tension samples of a) H-0°, b) H-90° and c) V-90° where the crack path in H-0° specimen sometimes deflected into the
specimens.
weakest link, leading to out-of-plane crack propagation, while the
crack propagation in H-90° and V-90° specimens consistently
previously-reported FFF-PEEK samples [26,28]. For the V-90° sample, al- delaminated along the bead-bead interface, as shown in Figs. 10 and
though it had a relatively high Young's modulus, the tensile strength 11. The difference in performance between H-90° and V-90° specimens
was significantly lower than those of H-0° and H-90° specimens and was related to the different degree of interfacial bonding between beads
molded PEEK. The tensile strength of V-90° falls within the range of associated with the different thermal gradients, as discussed in the pre-
molded LDPE [36]. vious section. After the maximum load had reached, a softening re-
sponse was observed as seen in Fig. 9b due to the deflection of crack
3.2. Fracture toughness into neighboring weak bead-bead interface. The SEM images of the frac-
ture surfaces of compact tension samples shown in Fig. 11a, b and c
The fracture test results showed the influence of the FFF fabrication helped to draw a similar conclusion to those of tensile samples as
configurations and bead orientation on the fracture toughness. The H- weak interfacial bonding can be found between beads in the vertical
0° specimen showed the best performance, followed by H-90° and V- fabrication direction. The stick-slip crack growth (after the maximum
90°, as can be seen in the load-displacement curves (Fig. 9) and KIC load) was observed in V-90° sample under constant applied extension
(Table 3). Based on build-rate data presented in Table 2, the interfacial rate loading conditions due to the presence of voids at the interface be-
bonding between beads in H-0° compact tension sample was better tween beads. These voids act as a crack arrester. Stick-slip fracture in
than that in H-0° tensile and flexural samples. On the other hand, H- this case was characterized by oscillatory crack tip velocities and crack
90° compact tension sample exhibited lower interfacial bonding be- growth jumps resulting in periodic load fluctuations (see, Fig. 9c) [41].
tween beads compared to those of tensile and flexural samples. The The strain contour at maximum load of the three sample configurations
build-rates in z direction (Vz) were almost the same for compact ten- was evaluated. It showed that the strain was a maximum at the crack
sion, tensile and flexural samples and thus exhibited poor interfacial front, as seen in Fig. 12. The highest value of maximum strain was
bonding between beads, as discussed earlier. Interestingly, the build- found in H-0°, followed by H-90° and V-90° specimens. The H-0° speci-
rate in x − y plane for H-0° and H-90° compact tension samples was men had the highest strain tolerance as the crack tip advanced normal

a) In-plane crack
propagation
Out-of-plane crack
propagation
b) In-plane crack
propagation
c)

Fig. 11. SEM images of fracture surfaces of a) H-0°, b) H-90° and c) V-90° specimens. Red, yellow and white (arrows or circles) respectively indicate z-fabrication direction, raster angle and
load direction. Circle indicates direction normal to the surface.
258 M.F. Arif et al. / Materials and Design 146 (2018) 249–259

a) (%)
6.10
b) (%)
4.30 c) (%)
0.47
5.23 3.71 0.41
4.36 3.11 0.34
3.49 2.51 0.28
2.62 1.92 0.21
1.75 1.33 0.15
0.88 0.73 0.09
0.01 0.14 0.02
-0.86 -0.46 -0.04
-1.73 -1.05 -0.11
-2.6 -1,65 -0.17

Fig. 12. DIC contour of strain at maximum load of a) H-0°, b) H-90° and c) V-90° specimens.

to the bead orientation, and thus the highest fracture toughness of the melting temperature is the key to producing parts with excellent mac-
sample. While comparing the current results with those from tests on roscopic material properties and dimensional stability. Thermal gradi-
bulk, molded PEEK using compact tension geometry at room tempera- ent effects remain an important issue in FFF, leading to limitations in
ture, we noticed that the fracture toughnesses of H-0° and H-90° sam- mechanical performance relative to conventionally molded samples.
ples, respectively were 78% and 70% of the molded PEEK fracture Ways such as precise control of the chamber temperature, post heat-
toughness (see, Table 3). The low fracture toughness in the V-90° con- treatment of the FFF fabricated samples [30] and microwave heating
figuration reduced the full potential of PEEK to be applied in high of FFF samples fabricated with coated-CNT filament [43], would help
load-bearing applications but the value was still within the range of to minimize these limitations. The Poisson's ratio calculation can be
brittle polymers such as acrylic, epoxy and polystyrene [42]. used as a comparative indicator of the specimen's quality. For example
the V-90° specimen had much lower Poisson's ratio compared to the
4. Summary H-0° and H-90° specimens which was mostly due to the opening of
the voids during loading. The current study serves as an important
The tensile and flexural properties of PEEK fabricated by FFF have first step in the development of comprehensive guideline for the FFF
been studied. The fracture behavior of FFF-PEEK has been reported for of high temperature and high performance thermoplastics (e.g. PEEK)
the first time. Three specimen configurations; H-0°, H-90° and V-90° to enable high-end engineering applications such as for biomedical
have been tested to failure in conjunction with optical strain mapping. implants.
The configuration that exhibited a superior performance, in terms of
tensile strength and modulus, flexural strength and modulus, and frac-
ture toughness, was H-0°, followed by H-90° and V-90°. The H-0° and Acknowledgements
H-90° showed 85% and 75% attainable tensile and flexural strengths of
the bulk molded PEEK respectively. The fracture toughnesses of the H- Authors would like to thank to Abu Dhabi Education Council (ADEC)
0° and H-90° specimens were 78% and 70% of the molded PEEK fracture for providing the research grant (EX2016-000006) through “the ADEC
toughness, respectively. However, the V-90° sample exhibited low ten- Award for Research Excellence (A2RE) 2015”. SK would like to thank
sile, flexural and fracture toughness properties compared to those of H- Professor Brian Wardle, Massachusetts Institute of Technology for his
0°, H-90° and bulk PEEK. The performance of these three configurations comments on the MS.
represents different thermal gradients between beads during layer by
layer buildup in FFF. References
It was found that different fabrication path can promote a different
[1] S. Kumar, B.L. Wardle, M.F. Arif, Strength and performance enhancement of bonded
degree of interfacial adhesion between beads which in turn can have a joints by spatial tailoring of adhesive compliance via 3D printing, ACS Appl. Mater.
significant influence on the tensile and fracture toughness properties. Interfaces 9 (2017) 884–891, https://doi.org/10.1021/acsami.6b13038.
This information is directly relevant to the fabrication of biomedical im- [2] S. Kumar, B.L. Wardle, M.F. Arif, J. Ubaid, Stress reduction of 3D printed compliance-
tailored multilayers, Adv. Eng. Mater. (2017)https://doi.org/10.1002/adem.
plants with sufficient mechanical strength for long-term in-vivo appli- 201700883 (1700883).
cations. Short polymer deposition path in x − y plane promoted the [3] J. Liljenhjerte, S. Kumar, Pull-out performance of 3D printed composites with em-
best bead-to-bead interfacial adhesion since the diffusion is faster bedded fins on the fiber, MRS Proc. 1800 (2015)https://doi.org/10.1557/opl.2015.
645 (mrss15-2135597).
when the beads were in molten condition. However, it had drawback [4] J. Liljenhjerte, P. Upadhyaya, S. Kumar, Hyperelastic strain measurements and con-
in the final surface roughness due to the formation of bead-to-bead stitutive parameters identification of 3D printed soft polymers by image processing,
overlap zone. Therefore, short deposition path could be considered for Addit. Manuf. 11 (2016) 40–48, https://doi.org/10.1016/j.addma.2016.03.005.
[5] D.P. Cole, J.C. Riddick, H.M. Iftekhar Jaim, K.E. Strawhecker, N.E. Zander, Interfacial
the inner part of the sample while the long deposition path could be
mechanical behavior of 3D printed ABS, J. Appl. Polym. Sci. 133 (2016) 1–12,
employed for the outer surface to obtain better surface quality. The https://doi.org/10.1002/app.43671.
short deposition path need not necessarily be implemented with a 90° [6] R. Gautam, S. Idapalapati, S. Feih, Printing and characterisation of Kagome lattice
bead orientation as alternate bead orientations such as 45°/−45° and structures by fused deposition modelling, Mater. Des. 137 (2018) 266–275,
https://doi.org/10.1016/j.matdes.2017.10.022.
60°/−60° also exhibit almost the same thermal gradient in the horizon- [7] A. Tsouknidas, M. Pantazopoulos, I. Katsoulis, D. Fasnakis, S. Maropoulos, N.
tal layer as that of the 90° bead orientation. Michailidis, Impact absorption capacity of 3D-printed components fabricated by
Diffusion lines (interfaces at which the bonding is weak) were found fused deposition modelling, Mater. Des. 102 (2016) 41–44, https://doi.org/10.
1016/j.matdes.2016.03.154.
at the interface between the beads in the vertical fabrication direction in [8] N.G. Tanikella, B. Wittbrodt, J.M. Pearce, Tensile strength of commercial polymer
all specimen configurations due to high thermal gradient across beads. materials for fused filament fabrication 3D printing, Addit. Manuf. 15 (2017)
This had a dramatic effect when the load was applied normal to this https://doi.org/10.1016/j.addma.2017.03.005.
[9] R.J. Zaldivar, D.B. Witkin, T. McLouth, D.N. Patel, K. Schmitt, J.P. Nokes, Influence of
weakest zone, as exemplified by the tensile, flexural and fracture tough- processing and orientation print effects on the mechanical and thermal behavior
ness properties of V-90° samples. Results suggest that minimizing ther- of 3D-Printed ULTEM® 9085 Material, Addit. Manuf. 13 (2017) 71–80, https://doi.
mal gradients across beads during FFF of thermoplastics with a high org/10.1016/j.addma.2016.11.007.
M.F. Arif et al. / Materials and Design 146 (2018) 249–259 259

[10] S. Hertle, M. Drexler, D. Drummer, Additive manufacturing of poly(propylene) by [27] W.Z. Wu, P. Geng, J. Zhao, Y. Zhang, D.W. Rosen, H.B. Zhang, Manufacture and ther-
means of melt extrusion, Macromol. Mater. Eng. 301 (2016) 1482–1493, https:// mal deformation analysis of semicrystalline polymer polyether ether ketone by 3D
doi.org/10.1002/mame.201600259. printing, Mater. Res. Innov. 18 (2014) S5-12–16, https://doi.org/10.1179/
[11] A.R. Torrado, C.M. Shemelya, J.D. English, Y. Lin, R.B. Wicker, D.A. Roberson, Charac- 1432891714Z.000000000898.
terizing the effect of additives to ABS on the mechanical property anisotropy of [28] W. Wu, P. Geng, G. Li, D. Zhao, H. Zhang, J. Zhao, Influence of layer thickness and ras-
specimens fabricated by material extrusion 3D printing, Addit. Manuf. 6 (2015) ter angle on the mechanical properties of 3D-printed PEEK and a comparative me-
16–29, https://doi.org/10.1016/j.addma.2015.02.001. chanical study between PEEK and ABS, Materials (Basel) 8 (2015) 5834–5846,
[12] Z. Weng, J. Wang, T. Senthil, L. Wu, Mechanical and thermal properties of ABS/mont- https://doi.org/10.3390/ma8095271.
morillonite nanocomposites for fused deposition modeling 3D printing, Mater. Des. [29] G. Cicala, A. Latteri, B. Del Curto, A. Lo Russo, G. Recca, S. Farè, Engineering thermo-
102 (2016) 276–283, https://doi.org/10.1016/j.matdes.2016.04.045. plastics for additive manufacturing: a critical perspective with experimental evi-
[13] F. Ning, W. Cong, Z. Hu, K. Huang, Additive manufacturing of thermoplastic matrix dence to support functional applications, J. Appl. Biomater. Funct. Mater. 15
composites using fused deposition modeling: a comparison of two reinforcements, (2017) 0, https://doi.org/10.5301/jabfm.5000343.
J. Compos. Mater. (2017)https://doi.org/10.1177/0021998317692659 [30] C. Yang, X. Tian, D. Li, Y. Cao, F. Zhao, C. Shi, Influence of thermal processing condi-
(2199831769265). tions in 3D printing on the crystallinity and mechanical properties of PEEK material,
[14] K. Prashantha, F. Roger, Multifunctional properties of 3D printed poly(lactic acid)/ J. Mater. Process. Technol. 248 (2017) 1–7, https://doi.org/10.1016/j.jmatprotec.
graphene nanocomposites by fused deposition modeling, J. Macromol. Sci. A 54 2017.04.027.
(2017) 24–29, https://doi.org/10.1080/10601325.2017.1250311. [31] K.R. Hart, E.D. Wetzel, Fracture behavior of additively manufactured acrylonitrile bu-
[15] S. Dul, L. Fambri, A. Pegoretti, Fused deposition modelling with ABS–graphene nano- tadiene styrene (ABS) materials, Eng. Fract. Mech. 177 (2017) 1–13, https://doi.org/
composites, Compos. A: Appl. Sci. Manuf. 85 (2016) 181–191, https://doi.org/10. 10.1016/j.engfracmech.2017.03.028.
1016/j.compositesa.2016.03.013. [32] N. Aliheidari, R. Tripuraneni, A. Ameli, S. Nadimpalli, Fracture resistance measure-
[16] W.W. Yu, J. Zhang, J.R. Wu, X.Z. Wang, Y.H. Deng, Incorporation of graphitic nano- ment of fused deposition modeling 3D printed polymers, Polym. Test. 60 (2017)
filler and poly(lactic acid) in fused deposition modeling, J. Appl. Polym. Sci. 134 94–101, https://doi.org/10.1016/j.polymertesting.2017.03.016.
(2017) 1–11, https://doi.org/10.1002/app.44703. [33] Y. Song, Y. Li, W. Song, K. Yee, K.Y. Lee, V.L. Tagarielli, Measurements of the mechan-
[17] A. Kantaros, D. Karalekas, Fiber Bragg grating based investigation of residual strains ical response of unidirectional 3D-printed PLA, Mater. Des. 123 (2017) 154–164,
in ABS parts fabricated by fused deposition modeling process, Mater. Des. 50 (2013) https://doi.org/10.1016/j.matdes.2017.03.051.
44–50, https://doi.org/10.1016/j.matdes.2013.02.067. [34] M. Spoerk, F. Arbeiter, H. Cajner, J. Sapkota, C. Holzer, Parametric optimization of
[18] C. Kousiatza, D. Karalekas, In-situ monitoring of strain and temperature distributions intra- and inter-layer strengths in parts produced by extrusion-based additive
during fused deposition modeling process, Mater. Des. 97 (2016) 400–406, https:// manufacturing of poly(lactic acid), J. Appl. Polym. Sci. (2017), 45401. https://doi.
doi.org/10.1016/j.matdes.2016.02.099. org/10.1002/app.45401.
[19] M. Spoerk, J. Sapkota, G. Weingrill, T. Fischinger, F. Arbeiter, C. Holzer, Shrinkage and [35] https://www.victrex.com/en/datasheets (accessed November 20, 2017).
warpage optimization of expanded-perlite-filled polypropylene composites in [36] J.E. Mark (Ed.), Polymer Data Handbook, Oxford University Press, 1999.
extrusion-based additive manufacturing, Macromol. Mater. Eng. (2017)https://doi. [37] J. Karger-Kocsis, K. Friedrich, Temperature and strain-rate effects on the fracture
org/10.1002/mame.201700143 (1700143). toughness of poly(ether ether ketone) and its short glass-fibre reinforced compos-
[20] B.N. Turner, R. Strong, S.A. Gold, A review of melt extrusion additive manufacturing ite, Polymer (Guildf.) 27 (1986) 1753–1760, https://doi.org/10.1016/0032-3861
processes: I. Process design and modeling, Rapid Prototyp. J. 20 (2014) 192–204, (86)90272-7.
https://doi.org/10.1108/RPJ-01-2013-0012. [38] P. Beguelin, H.H. Kausch, The effect of the loading rate on the fracture toughness of
[21] S.F. Costa, F.M. Duarte, J.A. Covas, Thermal conditions affecting heat transfer in FDM/ Poly(methyl methacrylate), Polyacetal, Polyetheretherketone and modified PVC, J.
FFE: a contribution towards the numerical modelling of the process, Virtual Phys. Mater. Sci. 29 (1994) 91–98, https://doi.org/10.1007/BF00356577.
Prototyp. 10 (2015) 35–46, https://doi.org/10.1080/17452759.2014.984042. [39] R. Gensler, P. Béguelin, C.J.G. Plummer, H.-H. Kausch, H. Münstedt, Tensile behaviour
[22] S.M. Kurtz, J.N. Devine, PEEK biomaterials in trauma, orthopedic, and spinal im- and fracture toughness of poly(ether ether ketone)/poly(ether imide) blends,
plants, Biomaterials 28 (2007) 4845–4869, https://doi.org/10.1016/j.biomaterials. Polym. Bull. 37 (1996) 111–118.
2007.07.013. [40] P.J. Rae, E.N. Brown, E.B. Orler, The mechanical properties of poly(ether-ether-ke-
[23] S. Berretta, K.E. Evans, O.R. Ghita, Predicting processing parameters in high temper- tone) (PEEK) with emphasis on the large compressive strain response, Polymer
ature laser sintering (HT-LS) from powder properties, Mater. Des. 105 (2016) (Guildf.) 48 (2007) 598–615, https://doi.org/10.1016/j.polymer.2006.11.032.
301–314, https://doi.org/10.1016/j.matdes.2016.04.097. [41] T.W. Webb, E.C. Aifantis, Stick-slip instabilities in fracture, Comput. Mech. '95,
[24] OsteoFab® Implants, http://oxfordpm.com/cmf-orthopedics/osteofab-implants, Springer Berlin Heidelberg, Berlin, Heidelberg 1995, pp. 1353–1358, https://doi.
Accessed date: 1 January 2018. org/10.1007/978-3-642-79654-8_219.
[25] B. Valentan, Ž. Kadivnik, T. Brajlih, A. Anderson, I. Drstvenšek, Processing poly(ether [42] M.F. Ashby, Materials Selection in Mechanical Design, 4th ed. Elsevier, 2011.
etherketone) on a 3d printer for thermoplastic modelling, Mater. Technol. 47 (2013) [43] C.B. Sweeney, B.A. Lackey, M.J. Pospisil, T.C. Achee, V.K. Hicks, A.G. Moran, et al.,
715–721. Welding of 3D-printed carbon nanotube–polymer composites by locally induced
[26] M. Vaezi, S. Yang, Extrusion-based additive manufacturing of PEEK for biomedical microwave heating, Sci. Adv. 3 (2017), e1700262. https://doi.org/10.1126/sciadv.
applications, Virtual Phys. Prototyp. 10 (2015) 123–135, https://doi.org/10.1080/ 1700262.
17452759.2015.1097053.

You might also like