You are on page 1of 13

Materials and Design 140 (2018) 209–221

Contents lists available at ScienceDirect

Materials and Design

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

Screw extrusion-based additive manufacturing of PEEK


Jian-Wei Tseng a, Chao-Yuan Liu a, Yi-Kuang Yen a, Johannes Belkner a, Tobias Bremicker a,
Bernard Haochih Liu c, Ta-Ju Sun a, An-Bang Wang a,b,⁎
a
Institute of Applied Mechanics, National Taiwan University, Taipei 106, Taiwan
b
Institute of Medical Device and Imaging, National Taiwan University, Taipei 106, Taiwan
c
Department of Materials Science and Engineering, National Cheng Kung University, Tainan 701, Taiwan

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

• A new screw extrusion-based 3D print-


ing system was designed to solve the
problems of conventional filament feed-
ing ones.
• High viscosity and melting temperature
material e.g., polyether-ether-ketone
(PEEK), has been printed with good
quality.
• Stable flow ability (b3% variation) with
high printing speed (up to 370 mm/s)
and reproducible quality was demon-
strated.
• 96% of the mechanical strength of the
bulk material (PEEK) in additive
manufacturing printing was achieved.
• SEM and XRD results revealed the ther-
mal condition was the major factor af-
fecting mechanical strength.
• The post annealing process had no signif-
icant effect on the material properties.

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

Article history: Polyether-ether-ketone (PEEK) has high mechanical strength, thermal performance, and biocompatibility, and is
Received 31 August 2017 widely used in biomedical and chemical engineering applications. However, to date there are no guidelines for
Received in revised form 15 November 2017 the building of a 3D printer for highly viscous materials, e.g., PEEK. In this paper, a screw extrusion method
Accepted 16 November 2017
was developed to overcome the existing problems of the filament-feeding method. Excellent flow stability
Available online 23 November 2017
(b 3% variation) and high printing speed (up to 370 mm/s) for PEEK printing were achieved. Highly reproducible
Keywords:
mechanical tests of the printing products were demonstrated with 96% of the bulk material strength for the first
Screw extrusion time. Furthermore, an exchangeable printing head was built to cover both line- and plane-printing needs to
3D printing system widen its applications and improve printing surface quality (up to 0.945 nm in Ra). All printed material had a
Polyether-ether-ketone more brittle character in comparison with the bulk material and the post annealing process was found to have
Printing strength and quality no significant effect on the mechanical strength. Additionally, porous artificial intervertebral cages with control-
lable size and distribution were manufactured to demonstrate potential applications.
© 2017 Elsevier Ltd. All rights reserved.

⁎ Corresponding author at: Institute of Applied Mechanics, National Taiwan University, Taipei 106, Taiwan.
E-mail address: abwang@iam.ntu.edu.tw (A.-B. Wang).

https://doi.org/10.1016/j.matdes.2017.11.032
0264-1275/© 2017 Elsevier Ltd. All rights reserved.
210 J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221

1. Introduction printing quality due to the manufacturing imperfectness of the pushrod


mechanism, especially for extruding the highly viscous PEEK material
For the applications in the biomedical field, polyether-ether-ketone for fast printing.
(PEEK), a semi-crystalline thermoplastic, is a well-known alternative In addition to printing PEEK material, different methods of generat-
to metal implants because of its excellent biocompatibility and compa- ing 3D PEEK composite for various biomedical applications have been
rable modulus to human cortical bones, which can potentially reduce developed. For instance, Vaezi et al. [20] used first FDM to print a porous
the stress shielding effect after implantation [1–3]. Moreover, it is hydroxyapatite (HA) scaffold and then compressed melted PEEK into
radio-transparent, which can provide better evaluation of postoperative the HA scaffold to form a PEEK/HA composite to enhance the bioactivity
recovery situations and has been widely used in clinical applications of PEEK material. Similarly, in an attempt to develop smart materials,
such as for intervertebral cage fusion and craniomaxillofacial recon- Yang et al. [21] added carbon fiber (CF) to the melted PEEK in the print-
struction [4–6]. Although PEEK possesses several bio-favorable proper- ing process, forming PEEK/CF and PLA/CF composites, to bend the
ties, it still suffers from inadequate osseointegration [7], rendering resulting material by the electrocaloric effect for biomimetic robotic ap-
porosity expression in artificial implants necessary to enhance the bio- plication, e.g., artificial muscle. Based on this trend of composite print-
interaction inside the human body to more closely mimic natural tissue. ing, it is likely that more PEEK composites will be developed
Meeting this requirement would largely increase the manufacturing combining new 3D printing technology for future applications.
complexity, especially in forming pore interconnectivity and would in- In the following Experimental section, the main facilities and test
troduce contamination issues [8,9]. procedures are described. The design of the screw extrusion mechanism
Recently, particulate leaching and selective laser sintering (SLS) by directly using raw pellets, system stability & operation test for the
have been identified as promising methods to overcome the line- and plane-printing head, tensile and compression tests, determi-
manufacturing bottlenecks due to their more physical mechanism and nation of crystallinity under various operation conditions with some
flexible platform compared to currently available fabrication processes printed PEEK examples (e.g., an artificial intervertebral cage with pore
[10,11]. However, these techniques still suffer from limited capability size of 0.4 mm and 50% porosity and a high aspect ratio of hollowed
of designing microstructure for particulate leaching and expensive ap- Taipei 101) are described in the Results and Discussion section.
paratus requirements for the SLS method [12]. An alternative, which
balances the economic concerns, is fused deposition modeling (FDM), 2. Experimental descriptions
which has been recognized as the most efficient method in additive
manufacturing (AM) to construct products for tissue engineering 2.1. PEEK extrusion system
(e.g., scaffolds) due to its wide material compatibility, low-cost appara-
tus, and great flexibility in terms of structural design [13]. Commercial The schematic diagram of the newly designed additive manufactur-
FDM incorporates a metal (e.g., copper) nozzle in which the melted fil- ing system for highly viscous material (here PEEK) is shown in Fig. 1(a).
ament is extruded to construct layer-by-layer structures by the relative This system includes three main subsystems, i.e., the extruder, thermal
motion between the nozzle and substrate. control, and traversing subsystems. The extruder subsystem was firmly
The filament feeding mechanism is the most commonly used meth- installed on a rigid frame fixed on a heavy granite plate (L × W × H:
od in FDM additive manufacturing, but it is mainly restricted to the ex- 600 mm × 600 mm × 50 mm). To eliminate the need for pre-
trusion of low viscosity materials such as poly lactic acid (PLA, 10 Pa·s) generation of the filament (commonly 1.75 or 3 mm in diameter)
and acrylonitrile butadiene styrene (ABS, 140 Pa·s) to avoid buckling used in conventional filament-based 3D printers and to maintain the
and back-flow induced flow instability. The gap between the filament printing quality without interference by the filament cross-section
and feeding tubes is the main cause for these problems, especially dur- change, a user-friendly pellet feeding design was used to satisfy eco-
ing filament feeding in the fast printing process [14]. Once the melted nomic concerns for practical use of the system. Two versions of the ex-
material flows back to the lower temperature region, it increases in re- trusion systems have been previously tested in the development of the
sistance and begins to solidify, which inhibits feeding efficiency due to 3D PEEK printing process, the first incorporating the pushrod mecha-
the buckling effect of the feeding filament and lowers the printing pre- nism [19] and the other by screw extrusion mechanism. The latter
cision and can even cause clogging. method is used in this paper and its detailed design is described in the
Despite these challenges, Valentan et al. [15] first demonstrated the next section. A microbalance (Nevada Weighing LLC) was used for the
feasibility of using FDM for the highly viscous PEEK by a custom-made extruded flow rate measurement.
machine in 2013, although it suffered from quality incompleteness The thermal control subsystem includes a thermal control panel,
(e.g., warpage, delamination and bubbles) and non-reproducibility heat plate (installed on the XY traversing unit), and heaters (installed
(e.g., due to non-uniform filament) of the printed products. In 2015 either within or around the extruder). Simulations for heat up time
Vaezi & Yang [16] also used filaments to print PEEK specimens with dif- and thermal distribution in the extruder design were conducted by
ferent porosities (14% and 31%) for mechanical test comparison with the circuit analogue method using the free open-source software-suite
the bulk material. The printed specimens had reduced mechanical Scilab Xcos [22]. For easy control of the extruder operation temperature,
strength because of air gaps between the infill pattern and entrapped three thermal modules were designed to separately control the temper-
micro-bubbles inside the filaments. Similarly, Wu et al. [17] and Yang ature settings in the feeding, compression, and metering zones
et al. [18] investigated the mechanical strength of the FDM products (Fig. 1(b)) using proportional–integral–derivative (PID) controllers.
and showed the importance of printing orientation with respect to the The typical temperature settings were 180, 340, and 380 °C for the feed-
load direction and the necessity of thermal management during print- ing, compression, and metering zones, respectively. Electric heaters
ing process to improve the mechanical properties of the printed were installed in the middle of each functional zone and covered by
specimens. an insulator and stainless-steel box for easy operation and maintenance.
In contrast to the filament feeding mechanism, Vaezi & Yang [16] The substrate was installed on the traversing subsystem with a
also chose powder as a raw material to avoid the use of filaments and stroke of 100 mm in three directions. Two different substrates, quartz
used a pushrod (or syringe) to provide higher pressures to overcome glass and stainless-steel, were tested and their influence on the process
the back flow issue, and successfully extruded melted PEEK through and final products were compared. The latter was mainly used in the
the printing nozzle. A similar design appeared in a previous study by measurements described herein, if not specially mentioned. To control
the present author [19] except a PEEK pellet was used as the raw mate- the thermal boundary condition of the substrate, the stainless-steel
rial. However, the pushrod mechanism has the inherent drawbacks of a base plate (60 × 120 mm2) was fixed at 280 ± 2.5 °C by a PID controller
batch process incapable of continuous printing, and suffers from lower (CAHO, model H481).
J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221 211

Fig. 1. The additive manufacturing system for the PEEK material. (a) Schematic diagram of the present screw extrusion printing system and (b) configuration of screw sections, and
(c) notations for screw geometry.

The X-Y-Z traversing unit is comprised of an X-Y stage (OPM In- 90G pellets had a median diameter size of 2.6 mm and height of
strument Inc., M-JTH200-100-C7, with repeatability of ± 1 μm and 2.8 mm with a melt viscosity of 90 Pa·s. The PEEK 450G had a higher vis-
positioning accuracy of ± 50 μm) and a Z stage (OPM Instrument cosity of 350 Pa·s with a median diameter size of 2.1 mm and height of
Inc., M-MT90-100-C7, with repeatability of ± 1.25 μm and position- 3.3 mm, and was used to evaluate the flow rate stability. All raw mate-
ing accuracy of ± 50 μm). To control the traverse position with re- rials were baked to ensure dehumidification at 180 °C for 4 h before the
spect to the extruder head, the G-code commands of moving printing process to avoid void formation.
segments in 3 axes from a commercial slicing program (Reptier-
host) for the printing path were interpreted by the LabView program
and synchronized by a stepper motion control card (National Instru- 2.3. Mechanical tests
ments, PCI-7344).
Additionally, a light, compact, and exchangeable coater, namely a One of the challenges in additive manufacturing is achieving good
slot printing head, which has been previously used for bio-material mechanical strength of the printed parts compared to the bulk material,
coating [23,24], was modified and adapted for use in the present system. especially for medical implants, in which sufficient strength must be en-
The slot printing head was designed as a fishtail-type structure with an sured for practical usage. Based on the ASTM standards, the mechanical
upstream spread angle of 60°. This coater was designed to ensure a uni- strength of the printed material was assessed in tensile (ASTM D638-
form flow velocity across the slit width and was constructed with struc- 02a, Type V) and compressive (ASTM D695-02a) tests in this study.
tural steel via the dual phase doping process in a typical screw Moreover, the samples used in the tensile tests were printed with a 0°
manufacturing process. The slit of coater is 0.3 mm in depth and raster angle (defined as the loading direction with respect to the print-
20 mm in width at the coater exit. The total interior volume of the coater ing path direction in the additive manufacturing process), 100% infill
is approximately 160 μL. Two plane heater strips were built in the coater rate and 3.2 mm in height. The compressive test samples were printed
for heating to 390 °C. An atomic force microscope (AFM, MultiMode 8- by the route of concentric circles to form cylinder stubs with the size
HR, Bruker Corporation, USA) was used to measure the printed surface of 12.7 mm in diameter and 25.4 mm in height. Due to the different
profile. loading range, two material test machines, a MTS Criterion 42.503
(with a full scale of 2 kN) and a MTS Landmark 370.02 (with a full
2.2. Adhesion test and printing material scale of 25 kN) were used for the tensile and compressive tests in an
air-conditioned room (temperature of 25 ± 2 °C), respectively. The
Since good and uniform adhesion strength between the semi- strain rate was set as 3.8 × 10−5 s− 1 for the tensile test, and 8.5
melted materials and the substrate/previous layer relies on a well- × 10− 4 s−1 for the compressive test. To test the annealing effect on
controlled thermal environment in the printing process, a pair of infra- the mechanical performance, the printed specimens were annealed fol-
red (IR) heaters were mounted around the nozzle head. This enabled lowing the Victrex® protocol for PEEK injection molding. The samples
the control of the melting material properties (e.g., surface tension) of were first baked at 150 °C for 3 h, heated at 10 °C/h to 200 °C where it
each newly extruded layer to prevent delamination and/or warpage remained for 4 h, cooled at a rate of 10 °C/h to 140 °C, and then cooled
caused by the residual thermal stress. The substrate temperature (Ts) to room temperature by switching off the power of the furnace.
was controlled to have less than a 1.3% variation within the test area After material mechanical test, the fracture surface of samples was
of 2 × 2 cm2 over the range of 200 to 340 °C. For the adhesion test, a measured by scanning electron microscopy (SEM, NOVA NanoSEM
pushrod-force-sensor system was designed for the adhesion force mea- 450, FEI, USA) to analyze differences in the microstructure. Further-
surement (Fig. 4) and a pushrod was programmed to move at a constant more, to calculate the polymer crystallinity of the samples, X-ray diffrac-
speed of 6.25 mm/s to push the adherent drop after it is laid on the sub- tion (XRD) and differential scanning calorimetry (DSC) were used. An
strate with a push delay time (Δt) of 3, 4, and 10 s, and the force was X-ray diffractometer (Bruker D2 phaser) was used to scan within the
measured by a three-axes balance (Bertec, model 3191111). range of 2θ = 0–60°. DSC (Perkin Elmer - DSC4000) was performed
In this study, two kinds of raw PEEK pellet material from Victrex® by first holding the sample temperature at 50 °C for 5 min, then repeat-
(PEEK 90G and PEEK 450G), were used for assessment of additive ing 3 cycles of sample heating (50 to 400 °C) and cooling (400 to 50 °C)
manufacturing by using both line- and plane-printing heads. The PEEK with same heating and cooling rates of 10 °C/min.
212 J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221

3. Results and discussion alloys (e.g., 18NiCr5 steel) were chosen for the screw material to ensure
lower friction between the PEEK and screw compared with that be-
3.1. Mechanics design of the screw extruder tween the PEEK and barrel for better extrusion efficiency.
Another important aspect is to ensure the flight, the thinnest part of
To further improve the feeding mechanism of the 3D printer for the screw, has sufficient strength to withstand the strongest friction and
highly viscous materials with high printing speed, flexibility, and better shear forces during the extrusion process. For a simplified but conserva-
printing quality, the screw extrusion mechanism was used in this study. tive estimation, a Newtonian fluid assumption was applied for estimat-
This concept is based on a conventional plastic extrusion machine, ing the upper bound of the screw strength under the most harsh
which offers high extrusion pressure and good and consistent quality extrusion conditions with high shear rate and viscosity, although the
of the products. Based on the mechanical principles of a screw, this tech- melted PEEK has a shear-thinning nature. There are two kinds of forces
nique can provide high pressure to inhibit backflow, purge the at work on the screw extrusion, the pressure and shear force. The force
entrapped air, and precisely control the extrusion flow volume by the on the flight can be estimated using the following equation:
screw spin rate. In fact, similar ideas have also been adopted by Gibson
et al. [25] and Tarik et al. [26] who used screw extrusion printing for fab- F ¼ ΔPAN þ μAN γ: ð2Þ
ricating polycaprolactone (PCL) material with much lower viscosity
where ΔP is the required extrusion pressure, AN is the active surface of
(1 Pa·s) and melt temperature (60 °C). No additional descriptions of
screw parallel to the screw axis in the metering zone (276 mm2), μ is the
the screw extrusion design for 3D printers could be found in the litera-
viscosity of the printing material (here using 350 Pa·s for PEEK 450G),
ture. Therefore, a brief description of the design process of the screw
and γ is the shear rate (using 200 rpm, i.e., 7536 s− 1). Based on the
extrusion-based 3D printer will be given here for the printing of highly
Hagen-Poiseuille equation, ΔP can be estimated by:
viscous material, i.e., PEEK.
As shown in Fig. 1(b), the screw configuration can be divided into 3 8μΔLQ N
sections according to their functions, i.e., material feeding, compression, ΔP ¼ ð3Þ
πr4
and metering zones. The nomenclature of the screw apparatus can be
found in Fig. 1(c). As previously mentioned, the design of the screw ex- where ΔL is the channel length, QN is the flow rate, and r is the nozzle
trusion system used in this study is based on a conventional screw ex- radius. In general, the flow rate is set to 60 mg/min for screw extrusion,
trusion machine, but with some modification for fitting the practical which is much higher than that used in this study and the correspond-
machining constraints. In the screw design of a PEEK extrusion machine, ing value of ΔP for the upper bound is 1.2 MPa. Therefore, the maximum
two critical parameters should be first determined, the aspect ratio L/D shear stress is estimated to be 3.84 MPa (τmax = F/AS, where AS is the
(where L is the total length and D is screw exradius) and the compres- active surface of screw perpendicular to the screw axis in the metering
sion ratio ε (defined as the ratio of feeding to metering zone depths). zone (276 mm2)) which is far lower than τallowed (850 MPa), ensuring
Generally, to ensure sufficient time for melting PEEK and exhausting the safety of the screw design.
entrapped air, the optimal aspect ratio L/D is between 18:1–24:1, and
the compression ratio ε is between 2:1–3:1 for a stable flow at the noz- 3.2. Thermal design of the screw extruder
zle outlet based on the Victrex PEEK material processing guide [27]. In
the present design, the aspect ratio was chosen to be 20:1 and the To prevent PEEK degradation, the residence time of the melted ma-
length of the feeding, compression, and metering zones were set as 9, terial in the extruder was designed to be shorter than 30 min. The sim-
5, and 6 times the screw exradius D, respectively. To minimize the sys- ulation of the whole screw extruder using Scilab was conducted and
tem size, D was chosen to be 14 mm, which is the minimum screw shown to be quite promising for optimizing the trial-and-error process
exradius that can be fabricated. The other specifications of the screw in the system design. The simulation result for part of nozzle section is
are the fixed pitch (t = 14 mm = D), clearance to barrel (λ = 100 shown as a representative example in Fig. 2. Fig. 2(a) is the real model
μm), flight width (s = 1 mm), and helix angle (ψ = 17.65°), as listed of the nozzle section, including the nozzle and its adapter, used in this
in Table 1. The core diameter Dc in the compression section was study. Since the thermal transfer is more complicated in reality, the
linearly increased from 8 (Dc, F) to 12 (Dc, M) mm, which represents block diagram of the heat transfer model was built by Scilab and
the fixed Dc in the feeding and metering section. The compression shown in Fig. 2(b) with its associated symbols. Numbers (1–7) in
ratio is defined as the volumetric compression ratio and can be Fig. 2(b) represent the heat transfer conditions of their related positions
described by: in Fig. 2(a), e.g., 1: the thermal resistance between the nozzle and
adapter, 2: the conductive resistance and thermal capacity of the nozzle,
h F ðt−sÞðD−h F Þ 3: the convective and radiative transfer to the ambient surroundings
ε¼ ; ð1Þ
hM ðt−sÞðD−hM Þ (20 °C), 4: the convective and radiative transfer to the melted PEEK,
5: the conductive resistance and thermal capacity of the nozzle adapter,
where hF is the feeding zone depth and hM is the metering zone depth. 6: the convective and radiative transfer to the ambient surroundings
By setting hF = 3 mm and hM = 1 mm, the designed value of ε becomes (20 °C), and 7: the convective and radiative transfer to the melted
2.56, which is within the suggested range of 2:1–3:1 [28]. Furthermore, PEEK. Fig. 2(c) shows the steady simulation results of temperature
since PEEK adheres well to metal, dual-phase nickel and chromium-rich distribution around the extruder by using different arrangements of
heating cartridge(s) under the same heating power (50 W). The blue
Table 1 point shows the position of heater. The type III arrangement with sym-
Detailed dimensions of the screw. metrical heating installation at four corners has the most uniform tem-
perature distribution with a 3.1 °C maximal temperature difference. The
type I arrangement has the worst temperature distribution with up to a
25 °C temperature difference, which may cause degradation of the ex-
truding material. Therefore, the type III arrangement, where each corner
has one heater, was adopted to achieve temperature homogeneity in
the present design and 16 heating cartridges (50 W) were installed
around the extruder tube together with an auxiliary heating cartridge
(20 W) at the nozzle adapter. By using the overall heating power
(820 W), the instantaneous temperature of the extruder wall and
J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221 213

Fig. 2. Thermal modeling and heating performance. (a) Real model of the nozzle section, (b) electrical-circuit analogue modeling, (c) temperature distribution. The biggest temperature
difference for Type I: 25 °C, Type II: 6.2 °C, Type III: 3.1 °C (d) Heat up time of the screw extruder estimation.

melted PEEK could be obtained and are shown as solid-lines and dotted- variations. The dotted- and solid-lines indicate the fitting curves of the
lines in Fig. 2(d), respectively. It is worth noting that the initial condi- experimental data for the upward and downward operation of screw
tions for the melted PEEK were determined under the steady setting spin speed, i.e., the flow rate. As shown in Fig. 3, a highly linear Q – S re-
temperature of the extruder wall. The heat up time is defined as the lationship for Dn = 400 μm was found without hysteresis for both PEEK
time when all temperatures at the A–D monitoring locations 90G and 450G. Using least–squares curve fitting, the linear equations of
(i.e., different functional zones) of the extruder have reached their tar- Q = 0.1236S − 0.0185 and Q = 0.133S − 0.0251 were obtained with
get temperature after overshoot. The simulation values of the heat up the root-mean-square error of 0.0042 and 0.0021 for PEEK 90G and
time were 23 and 21 min for the extruder wall and melted PEEK, respec- 450G, respectively. The extrusion flow rate was insensitive to the extru-
tively. The former is quite consistent with the experimental data of sion viscosity. Similarly, no significant hysteresis was found for the
24 min and within the design limit of 30 min that was set to avoid ma- smaller nozzle diameter of Dn = 300 μm, and all measured data are
terial degradation. within the band of 3% deviation from the mean value. A linear logarith-
mic equation of Q = 0.8804 log S − 1.7257 obtained using least–
3.3. Flow rate stability test squares curve fitting for the operation design with a root-mean-
square error of 0.1174. This result indicates that a solid base of reliable
The screw extrusion system was designed and developed to prevent extrusion flow rate for the 3D printing can be obtained only by the con-
back flow and achieve a stable flow rate during the high speed additive trol of screw spin speed.
manufacturing process. The relationship between PEEK flow rate, Q, and
screw spin speed, S, was calibrated for two types of nozzles (0.3 and 3.4. Choice of substrate temperature
0.4 mm in diameter) used in this study and shown in Fig. 3. The flow
rate was measured by the extruded weight within a 1 min period. Five Substrate temperature control during the printing process is vital to
extruded PEEK samples were measured for each data point. The symbol ensure good adhesion without delamination. The removing force of the
represents the average value and the error bar shows its related PEEK drop was measured by the pushrod-force-sensor system, shown
214 J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221

the printed material. It is important to note that with longer the delay,
the peak of the μs-Ts curve moves to the right, i.e., toward the higher
substrate temperature. This is because more adhesion area forms due
to the longer solidification time. For Ts near the melting temperature
(343 °C), a low adhesion force of melted PEEK is observed for all cases.
Similar phenomena are observed for the quartz glass substrate with
lower thermal conductivity (data not shown).
Considering the printing process is not isothermal and it takes time
to build the structures, Fig. 4(b) offers a good reference for the macro-
mechanic interactions between the printed material and substrate for
determining the suitable substrate temperature. Based on Fig. 4(b),
one can easily find the optimal operation window for the substrate tem-
perature as well as the best adhesion conditions (i.e., highest μs). It is
therefore reasonable to choose the lower range of the substrate temper-
ature (to save energy), with higher adhesion coefficients in the longer
time curve (Δt = 10 s) to determine the optimized substrate tempera-
ture. In this study, unless otherwise mentioned, the substrate tempera-
ture was maintained at 280 °C.
Fig. 3. The flow rate calibration curves for different nozzle sizes. The dotted- and solid-line
indicates the fitting curves of the flow rate by upward and downward measurements,
respectively, to specify the severity of hysteresis. Error bars represent the standard 3.5. Plane printing test
deviations of five independent experiments.

To smoothen the printing surface and to minimize the iterative ef-


in Fig. 4(a), and normalized by the drop weight to obtain the equivalent forts by integration of printing lines for the regular plane or block for-
coefficient of adhesion friction (μs). Fig. 4(b) shows the measured μs mation, the melted material was uniformly extruded from a slot
data related to substrate temperature for three delay times, Δt, of 3, 4, channel (of the slot printing head) to print a plane film, as shown in
and 10 s. In general, the μs value increases to a peak of approximately Fig. 5(a), instead of from the commonly used point nozzle to form a
3 and then decreases again with further increasing Ts. Among the line film. The plane-printing process is basically similar as the conven-
three experimental curves, it is worth noting that the increasing curve tional line-printing except the former should overcome more difficul-
shifts toward higher Ts for Δt = 10 s. This is likely caused by the crack ties in layers binding due to the much wider printing film, which may
at the edge perimeter of the cone due to large temperature gradient cause delamination or bubbles entrapment in interlayers. This means
for Ts ≤ 250 °C. After enough propagation time, the edge crack grows to- that more cares have to be taken to ensure good binding between layers
ward the middle, e.g., for Δt = 10 s, and the contact area between the during the nozzle design and printing processes. In Fig. 5(a), a pair of IR
cone and substrate is reduced so that the pushrod can easily remove heaters (as shown in green bars) were used not only to control the
the cone. For higher Ts values, the viscosity of the printed material be- suitable thermal environment around the printing head but also to pro-
comes lower for better wetting and the time to achieve full solidification long the enough wetting time for the printed film on the substrate dur-
is longer when Δt is constant. The real contact area between the printed ing the solidification process. In Fig. 5(b), the AFM-scanned surface
material and substrate can increase due to the reduced volume of the profile of a printed five-layered (n = 5) PEEK film with dimensions of
trapped air during printing and solidification processes, increasing the 40 × 20 × 1.5 mm (L × W × H) is compared with the single-layered
adhesion force with higher surface temperature. (n = 1) printing films of thicknesses ranging from 123 to 313 μm. All
After reaching the μs peak, the adhesion force decreases for profiles show variation in their surface height within a similar range
increasing substrate temperature due to incomplete solidification of of ± 4 nm. To get a detailed overall view for comparison, the profiles

Fig. 4. (a) Schematic diagram of the pushrod mechanism for adhesion force measurement and (b) the measured coefficient of adhesion friction for different substrate temperatures.
J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221 215

of the thinnest single-layered and the thickest five-layered printing manufactured PEEK can possibly be mitigated with proper control of
films are shown in Fig. 5(c) and (d) with surface roughness (Ra) of nozzle temperature and annealing treatment. Further investigations
0.945 and 3.46 nm, respectively. The nanometer level surface roughness on the process parameters should be conducted as a separate study.
demonstrates the smoothness of the surface resulting from the highly The comparison of tensile strength in samples with different thermal
stable flowrate of the system. Moreover, this integrated printing system histories is shown in Fig. 6(c). The best ultimate tensile strength (Us) of
comprising of line- and plane-printing process is highly suitable for fast the printed parts is obtained at 390 °C for cases 3 and 4 with the value of
printing and to meet the manufacturing needs of high quality and uni- 94.0 ± 3.88 MPa, which is almost the same (96%) as that of the bulk
formity (e.g., for microfluidic channels) [29,30]. specimen in case 5 (98.0 ± 1.99 MPa). The data obtained in this study
was also compared with literature data in Table 2 and shows that the
3.6. Mechanical properties assessment – tensile specimens test printing material that can achieve high ultimate tensile strength compa-
rable to that of the bulk material. The present data in both cases at 390
The dimensions and the photos of the test specimens produced by °C (cases 3 and 4) have significantly higher Us-value than reported pre-
line- and plane- printing are shown in Fig. 6(a). To analyze the influence viously [16,17]. It is worth noting that cases 3 and 4 show much better
of thermal conditions on the line- and plane-printing in the printing property, defined as the ratio of ultimate tensile strength by
manufacturing processes, specimens were prepared by line-printing printing with respect to that of the bulk specimen, at 96% compared to
(denoted as cases 1–3) and plane-printing (denoted as case 4) and 66% in the work by Vaezi & Yang [16]. A similar comparison can be
were tested for comparison with the bulk material (denoted as case made with respect to the porosity, P, of the samples, which is calculated
5). The thermal treatment conditions for cases 1–4 were: 370 °C with- Vi −ðmρÞ
by P ¼ Vi  100%, where Vi, m, and ρ are the nominal volume, mass
out annealing for case 1, 370 °C with annealing for case 2, 390 °C with-
out annealing for cases 3 and case 4, respectively. Fig. 6(b) shows the and, density of the printed specimen, respectively. Here, the value of
results of the stress-stain relationship for the PEEK specimens of cases ρ(1.3 g/cm3) is provided by Victrex plc. The P value of 14% in [16] is sig-
1–5. All strains referred to in this study are engineering strains. For re- nificantly higher than that obtained in the present work (2.6%) by line-
peatability analysis, six samples were tested independently for each printing, so the Us-value reduces correspondingly.
case and the error bars show the variations of the measured data of The analysis of fractured surface by tensile tests was performed for
the ultimate tensile strength tests in Fig. 6(c). It is important to note cases 1–5 by SEM, and shown in Fig. 7(a–e). The inset in each panel
that, in Fig. 6(b), the measured data of all cases overlap excellently in shows the five-times magnified image of the local area. From the SEM
the elastic region and have almost the same average value of tensile images in Fig. 7(a) and (b), it is clear that inferior layer-to-layer bond-
elastic modulus (obtained by the slopes in case 1–4) of 3.93 GPa, ing, as shown by the arrow, occurs for the lower operation temperature
which is only 3.2% different from that of the bulk material (case 5). All of 370 °C in cases 1 and 2. The cross sections of cases 3 and 4 at 390 °C
printed specimens have much lower strains (0.018–0.029) than that operation temperature revealed a solid-like cross section (Fig. 7(c–d))
of the bulk material (about 0.25) and the zoom-in view in the red instead of the layered morphology as in cases 1 and 2, suggesting a
framed region is shown in the upper-right panel of Fig. 6(b). The results higher strength than the parts printed at a lower nozzle temperature.
reveal that the different thermal conditions and line- and plane-printing In case 5, Fig. 7(e) shows the clear necking of the cross-section on the
in the manufacturing processes do not affect the elastic moduli of PEEK right-hand side and the larger area of burnished surface due to the
but all printing specimens exhibit brittle behavior. The brittleness of the slow crack propagation for a ductile material.

Fig. 5. The laminated thin films by plane-printing additive manufacturing: (a) schematic diagram of the plane-printing process; (b) the AFM-measured surface profiles of printed single (n = 1)
and five layered (n = 5) PEEK films and the photo of the latter (inset); (c) AFM surface morphology of the single layered film with a surface roughness of 0.945 nm in Ra; (d) AFM surface
morphology of the five layered (n = 5) film with a surface roughness of 3.46 nm in Ra by plane-printing.
216 J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221

Fig. 6. (a) The (i) dimensions and the printed test specimens by (ii) line- and (iii) plane- printing before the tension test; comparisons of (b) tensile stress and strain curves and (c) the
ultimate tensile strengths of the materials prepared using different manufacturing processes.

3.7. Compressive specimen test variations in the measured data in Fig. 8(c). Again, the modulus of elas-
ticity was calculated from the slope of the initial linear section and the
The dimensions of the compressive tests and the photos of the spec- compressive yield strength was determined from the stress-strain dia-
imen prepared by line printing and machined from bulk material are gram by the same linear slope at a 0.2% strain. Similar to the results of
shown in Fig. 8(a). Fig. 8(b) shows the result of the compressive test tension test (in Fig. 6(b)), Fig. 8(b) shows excellent overlap of the data
for four different manufacturing processes in cases 1–3 (by line- points for all printed samples, although the averaged compressive elas-
printing) and case 5 (machined from the bulk material). Six samples tic moduli obtained by the slopes of cases 1–3 (2.55 ± 0.22 GPa) is
were tested independently for each group and the error bars show the slightly lower (20.5%) than that in the bulk material (case 5). The reason

Table 2
The comparison of mechanical properties of the AM-PEEK specimens made by different manufacturing processes.
J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221 217

for the major difference in the compressive stress-strain curve between


the bulk specimen and the printed specimens is likely caused by radial
deformation, which is perpendicular to the printing direction and con-
tributes to the weakened tensile strength of the printed samples.
These deformations also provoke the delamination of the printed layers
during the compression process and the printed specimens often failed
in the manner of a big burst of either part (as shown in the sketched se-
quence A for cases 1 and 3 in the inset of Fig. 8(b)) or complete breakup
(as shown the sketched sequence B for case 2 in the inset of Fig. 8(b))
after accumulating enough compressive energy. The whole breakup of
the specimen explains the sudden fall to zero for the green stress
curve in Fig. 8(b), but this behavior can also be interpreted as a two-
stage bursts from A and then B due to the unavoidable structural hetero-
geneity (e.g., manufacturing defects).
The average of compressive yield strength in cases 1–3 is 98.5 ±
2.79 MPa, which is similar to that (96% of) the bulk material
(103 MPa), as shown in Table 2. The ultimate compressive strength
values are 135, 130, and 140 MPa for cases 1, 2, and 3, respectively.
However, the bulk-PEEK was determined to be more ductile and ho-
mogeneous than the printed samples since its stress-strain curve
continuously moved upward with a strain up to 40% without rup-
ture, as shown in Fig. 8(a)–(iv), up until the loading limit. The com-
parison of the compressive strength for samples with different
thermal histories is shown in Fig. 8(c). The result exhibits that the
mechanical strength of printed specimens is not influenced by the
annealing process.
When quantitatively comparing the mechanical properties of the
AM-PEEK specimens and those from the literature in Table 2, all values
of the compressive ultimate strength for the samples prepared herein
are 2.2 times higher than those reported by Wu. et al. [17]. It is worth
mentioning that the bulk material processed by injection molding
with annealing from Vaezi & Yang [16] has the same value of yield
strength (Ys), but much lower elastic modulus (E) compared to the
present data. Combining the previous data and the result obtained
here, the annealing process could likely be eliminated to save energy
in the manufacturing process. Furthermore, the porosity values listed
in Table 2 were 8.2 ± 2.0% for cases 1 and 2, and 2.6 ± 1.0%, and 1.9
± 0.5% for cases 3 and 4, respectively. By assuming the porosity, P, to
be zero for the bulk specimens, a clear trend of increasing porosity
emerges for decreasing Us and Ys in the compressive and tensile tests
presented herein.
The porosity can be further examined from the measurement of frac-
tured surfaces by SEM, as shown in Fig. 9(a–c). The inset in each photo
shows the five-times magnified image of the local area. The PEEK spec-
imens printed at 370 °C (the lower temperature, cases 1 and 2 corre-
sponding to SEM images in Fig. 9(a) and 9(b), respectively) have
regular “holes” and clear welding lines (as indicated by the arrow) be-
tween the infill filaments. However, the SEM image in Fig. 9(c) shows
no layered morphology for specimens printed at the higher tempera-
ture of 390 °C (case 3). This coincides with the higher P value at 370
°C compared with that at 390 °C, creating more voids at the microscale
to allow multiple crack propagations in the shear mode along the
welding line direction, resulting in a lower strength than the sample
printed at 390 °C. This further indicates that void formation during the
printing process can be greatly restrained by the screw feeding mecha-
nism, resulting in higher printing stability. This can also be achieved by
increasing the liquidity of the printing material by increasing the pro-
cessing temperature to 390 °C.

Fig. 7. SEM images of the fractured cross-section after tension tests of the specimens
manufactured at (a) 370 °C without annealing, (b) 370 °C with annealing, 390 °C
without annealing by (c) line-printing and (d) plane-printing, and (e) the bulk material.
The inset in each photo shows the five times magnification of the area indicated by the
red rectangular frame.
218 J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221

Fig. 8. (a) The (i) dimensions, (ii) printed test specimens and test specimens manufactured from bulk PEEK (iii) before and (iv) after the compressive test; comparisons of (b) compressive
stress and strain curves and (c) the ultimate compressive strengths of samples produced by different manufacturing processes.

3.8. Crystallinity of the PEEK specimen mechanical strengths of cases 1 and 2 mentioned previously (only
a 3–4% difference in mechanical properties). Meanwhile, despite
The crystallinity of both the annealed and unannealed AM-PEEK the crystallinity in the group of specimens printed at 390 °C being
specimens was measured after the tensile and compressive tests. As comparable to those at 370 °C, their ultimate tensile strength is
a reference, the crystallinity of the bulk specimens after the me- higher by 34%. This further indicates that the temperature of the
chanical tests was also measured. The percentages of crystallinity nozzle head during the printing process is a more critical factor af-
in the specimens are summarized in Table 3. The results show that fecting the mechanical properties of the material than its crystallin-
the crystallinity in the 370 °C annealed group is 39.5% after the ten- ity. In fact, as a semi-crystalline polymer, PEEK possess a maximum
sile test and 36.2% after the compressive test, which are both slight- crystallinity between 35% and 39% [31], and once the crystallinity
ly higher than that at 370 °C without annealing (38.8% in the tensile approaches this level, the post processing heat treatment could
test and 35.0% in compressive test). The average of all the not further increase this value. Based on this approach, the screw
unannealed groups in this study is 39.7 ± 4.1%, which is much extrusion PEEK printer was shown to provide a high degree of crys-
higher than the bulk material (20.8%). In fact, unlike the random ori- tallinity and is compatible with pellets, gives high printing preci-
entation of the polymer chain structure in the bulk specimen, the sion, and rapid printing speed, making it suitable for industrial
melted PEEK polymer is subjected to a shearing strain by the extru- applications.
sion process and the configuration of the polymer adopts a more or-
dered state, increasing the crystallinity of the long-chain polymers. 3.9. Additive manufactured PEEK structure
The annealing process for the 370 °C printed specimens contributes
slightly to the increased crystallinity (1.7% and 3.3% in the tensile Fig. 10 shows structures printed by the screw extrusion printing sys-
and compressive tests, respectively). This corresponds with the tem including: (a) an artificial intervertebral cage (50% porosity); (b) a

Table 3
Crystallinity (%) of the AM-PEEK specimens produced from different manufacturing processes.
J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221 219

Fig. 10. 3D PEEK structures: (a) intervertebral cage; (b) corrosion protective sleeve;
(c) bio-scaffold; and (d) Taipei 101, an empty structure without a supporting frame.

supporting frames for building floating structures (e.g., inverted cone


and hollowed pyramid), commonly used in conventional additive
manufacturing processes, can potentially be eliminated due to the high-
ly viscous nature of PEEK with suitable printing design. The hollowed
Taipei 101 was printed without any supporting frame during the print-
ing process because of the viscous polymer, naturally preventing struc-
tural collapse. This is a significant feature that distinguishes PEEK from
conventional FDM, which require an additional nozzle to fabricate a
water-soluble or break-away supporting structure (BASS) [32,33] for
constructing some 3D materials.
In short, the novel screw extrusion PEEK printer has demonstrated
the ability to perform line and plane printing without the need for fila-
ments, and offers a reliable manufacturing platform producing printed
parts with excellent mechanical properties that are applicable in general
CAD/bio-laboratories.

4. Conclusion

To improve the reliability of the printing properties of the fused de-


position modeling method, a new screw extrusion additive manufactur-
Fig. 9. SEM images of the fractured cross-section after compression tests of the samples ing system with a pellet-feeding mechanism was successfully
prepared at (a) 370 °C without annealing, (b) 370 °C with annealing, (c) 390 °C without developed and described in this paper for the first time. This system
annealing. The inset in each photo shows the five times magnification of the area was systematically tested using PEEK biomedical material under differ-
indicated by the red rectangular frame.
ent operating conditions and allowed the melted PEEK to be efficiently
and homogeneously layered onto a heating substrate by suitable control
of the temperature and extrusion rate. The commonly used annealing
process did not improve the mechanical properties of the material
after printing and its removal is recommended for the related additive
corrosion protective sleeve; (c) a bio-scaffold structure, and (d) a brac- manufacturing processes. Nozzle temperatures in the range of
ing frame-free hollowed structure (Taipei 101). The achieved printing 370–390 °C and a heated plate up to 280 °C were ideal conditions to
resolution was 125 μm. In fact, the artificial implants, such as the inter- maintain an extrusion rate of 1.5 g/min within a 3% variation to meet in-
vertebral cages with interconnected channels of different pore sizes and dustrial demands. In addition, maintenance of the extrusion tempera-
changeable porosity distributions, can be easily realized by using the ture at 390 °C proved to significantly inhibit void formation/residual
flexibility of the 3D printing process. stresses during the printing process, and to greatly improve the me-
A summary of the mechanical properties of cortical bone and related chanical properties of the printed materials which approached (96%)
biomaterials are listed in Table 4. The mechanical strengths that lie those of the bulk material. Moreover, this novel screw extrusion print-
within the range of cortical bone in the first row are marked grey. It is ing system further realized the line- and plane-printing additive
obvious that PEEK has good biocompatibility and extraordinary me- manufacturing through an exchangeable printing head design, which
chanical properties that can be used to mimic bone structure to prevent improved surface roughness (up to 0.945 nm Ra for plane-printing)
stress sheltering. Furthermore, the 30 vol% carbon fiber (CF30)/PEEK and printing speed (up to 370 mm/s for line-printing). This highly reli-
and glass fiber (GF30)/PEEK composites have better mechanical able, user-friendly, and easily adaptable screw extrusion additive
strength and have the potential to be used in 3D printing. In addition manufacturing method is expected to be used in laboratories to perform
to the potential of PEEK for use in medical applications, the use of 3D printing in the future.
220 J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221

Table 4
Summary of the mechanical properties of cortical bone and biomaterials.

Acknowledgement [15] B. Valentan, Z. Kadivnik, T. Brajlih, A. Anderson, I. Drstvensek, Processing poly(ether


etherketone) on a 3d printer for thermoplastic modelling, Mater. Tehnol. 47 (6)
(2013) 715–721.
The authors appreciate the financial support of Ministry of Science [16] M. Vaezi, S. Yang, Extrusion-based additive manufacturing of PEEK for biomedical
and Technology, Taiwan (MOST 104-2221-E-002-138-MY3 and 104- applications, Virtual. Phys. Prototyp. 10 (3) (2015) 123–135, https://doi.org/10.
1080/17452759.2015.1097053.
2218-E-002-009). [17] W.Z. Wu, P. Geng, G.W. Li, D. Zhao, H.B. Zhang, J. Zhao, Influence of layer thickness
and raster angle on the mechanical properties of 3D-printed PEEK and a compara-
tive mechanical study between PEEK and ABS, Materials 8 (9) (2015) 5834–5846,
References https://doi.org/10.3390/ma8095271.
[18] C. Yang, X. Tian, D. Li, Y. Cao, F. Zhao, C. Shi, Influence of thermal processing conditions in
[1] S.M. Kurtz, J.N. Devine, PEEK biomaterials in trauma, orthopedic, and spinal im- 3D printing on the crystallinity and mechanical properties of PEEK material, J. Mater.
plants, Biomaterials 28 (32) (2007) 4845–4869, https://doi.org/10.1016/j. Process. Technol. 248 (2017) 1–7, https://doi.org/10.1016/j.jmatprotec.2017.04.027.
biomaterials.2007.07.013. [19] C.Y. Liu, Y.W. Hsieh, T.J. Sun, J.W. Zeng, A.B. Wang, N.T. Lee, S.W. Chau, W.C. Wei, B.H.
[2] J. Anguiano-Sanchez, O. Martinez-Romero, H.R. Siller, J.A. Diaz-Elizondo, E. Flores- Liu, R.C. Luo, Design and test of additive manufacturing for coating thermoplastic
Villalba, C.A. Rodriguez, Influence of PEEK coating on hip implant stress shielding: PEEK material, IEEE International Conference on Industrial Technology, Taipei,
a finite element analysis, Comput. Math. Methods Med. 2016 (2016)https://doi. 2016https://doi.org/10.1109/ICIT.2016.7474919.
org/10.1155/2016/6183679. [20] M. Vaezi, S. Yang, A novel bioactive PEEK/HA composite with controlled 3D inter-
[3] M. Roskies, J.O. Jordan, D.D. Fang, M.N. Abdallah, M.P. Hier, A. Mlynarek, F. connected HA network, Int. J. Bioprint. 1 (1) (2015)https://doi.org/10.18063/IJB.
Tamimi, S.D. Tran, Improving PEEK bioactivity for craniofacial reconstruction 2015.01.004.
using a 3D printed scaffold embedded with mesenchymal stem cells, J. [21] C. Yang, B. Wang, D. Li, X. Tian, Modelling and characterisation for the responsive per-
Biomater. Appl. 31 (1) (2016) 132–139, https://doi.org/10.1177/ formance of CF/PLA and CF/PEEK smart materials fabricated by 4D printing, Virtual.
0885328216638636. Phys. Prototyp. 12 (1) (2017) 69–76, https://doi.org/10.1080/17452759.2016.1265992.
[4] J. Parthasarathy, 3D modeling, custom implants and its future perspectives in cra- [22] J.W. Tseng, On the Development of a Novel Pellet-based and Screw-typed 3D Printer
niofacial surgery, Ann. Maxillofac. Surg. 4 (1) (2014) 9–18, https://doi.org/10. for PEEK Material and Its Applications(Master Thesis) Institute of Applied Mechan-
4103/2231-0746.133065. ics, National Taiwan University, 2017 1–132, https://doi.org/10.6342/
[5] S. Najeeb, M.S. Zafar, Z. Khurshid, F. Siddiqui, Applications of polyetheretherketone NTU201703171.
(PEEK) in oral implantology and prosthodontics, J. Prosthodont. Res. 60 (1) (2016) [23] Y.K. Yen, Y.W. Jiang, S.C. Chang, A.B. Wang, Western blotting by thin-film direct coating,
12–19, https://doi.org/10.1016/j.jpor.2015.10.001. Anal. Chem. 86 (10) (2014) 5164–5170, https://doi.org/10.1021/ac5010162.
[6] I.V. Panayotov, V. Orti, F. Cuisinier, J. Yachouh, Polyetheretherketone (PEEK) for [24] C.Y. Liu, D.C. Lu, Y.W. Jiang, Y.K. Yen, S.C. Chang, A.B. Wang, Easy and fast western
medical applications, J. Mater. Sci. Mater. Med. 27 (7) (2016)https://doi.org/10. blotting by thin-film direct coating with suction, Anal. Chem. 88 (12) (2016)
1007/S10856-016-5731-4. 6349–6356, https://doi.org/10.1021/acs.analchem.6b00699.
[7] M.Z. Zhao, M.R. An, Q.S. Wang, X.C. Liu, W.J. Lai, X.Y. Zhao, S.C. Wei, J.G. Ji, [25] I. Gibson, M.M. Savalani, C.X. Lam, R. Olkowski, A.K. Ekaputra, K.C. Tan, D.W.
Quantitative proteomic analysis of human osteoblast-like MG-63 cells in response Hutmacher, Towards a medium/high load-bearing scaffold fabrication system,
to bioinert implant material titanium and polyetheretherketone, J. Proteome 75 Tsinghua Sci. Technol. 14 (2009) 13–19, https://doi.org/10.1016/S1007-
(12) (2012) 3560–3573, https://doi.org/10.1016/j.jprot.2012.03.033. 0214(09)70060-4.
[8] H.M. Aydin, A.J. El Haj, E. Piskin, Y. Yang, Improving pore interconnectivity in poly- [26] M. Tarik Arafat, I. Gibson, X. Li, State of the art and future direction of additive
meric scaffolds for tissue engineering, J. Tissue Eng. Regen. Med. 3 (6) (2009) manufactured scaffolds-based bone tissue engineering, Rapid Prototyp. J. 20 (1)
470–476, https://doi.org/10.1002/term.187. (2014) 13–26, https://doi.org/10.1108/RPJ-03-2012-0023.
[9] L.O. Dandy, G. Oliveux, J. Wood, M.J. Jenkins, G.A. Leeke, Accelerated degradation of [27] Victrex plc, (Home Page) https://www.victrex.com/en/victrex-peek, Accessed date:
Polyetheretherketone (PEEK) composite materials for recycling applications, Polym. 31 August 2017.
Degrad. Stab. 112 (2015) 52–62, https://doi.org/10.1016/j.polymdegradstab.2014. [28] T. Womer, Basic screw geometry “things your screw designer never told you about
12.012. screws!!”(983), Antec Conference Proceedings 2002, pp. 74–79.
[10] M. Schmidt, D. Pohle, T. Rechtenwald, Selective laser sintering of PEEK, CIRP [29] S. Waheed, J.M. Cabot, N.P. Macdonald, T. Lewis, R.M. Guijt, B. Paull, M.C. Breadmore,
Ann. Manuf. Technol. 56 (1) (2007) 205–208, https://doi.org/10.1016/j.cirp. 3D printed microfluidic devices: enablers and barriers, Lab Chip 16 (11) (2016)
2007.05.097. 1993–2013, https://doi.org/10.1039/C6LC00284F.
[11] A.R. Siddiq, A.R. Kennedy, Porous poly-ether ether ketone (PEEK) manufactured by a [30] B.C. Gross, J.L. Erkal, S.Y. Lockwood, C. Chen, D.M. Spence, Evaluation of 3D printing
novel powder route using near-spherical salt bead porogens: characterisation and and its potential impact on biotechnology and the chemical sciences, Anal. Chem. 86
mechanical properties, Mater. Sci. Eng. C 47 (2015) 180–188, https://doi.org/10. (7) (2014) 3240–3253, https://doi.org/10.1021/ac403397r.
1016/j.msec.2014.11.044. [31] A.A. Mehmetalkan, J.N. Hay, The crystallinity of peek composites, Polymer 34 (16)
[12] S. Singh, V.S. Sharma, A. Sachdeva, Progress in selective laser sintering using metallic (1993) 3529–3531, https://doi.org/10.1016/0032-3861(93)90487-U.
powders: a review, Mater. Sci. Tech. Lond. 32 (8) (2016) 760–772, https://doi.org/ [32] R. Liska, F. Schwager, C. Maier, R. Cano-Vives, J. Stampfl, Water-soluble photopoly-
10.1179/1743284715Y.0000000136. mers for rapid prototyping of cellular materials, J. Appl. Polym. Sci. 97 (6) (2005)
[13] J. Korpela, A. Kokkari, H. Korhonen, M. Malin, T. Narhi, J. Seppala, Biodegradable and 2286–2298, https://doi.org/10.1002/app.22025.
bioactive porous scaffold structures prepared using fused deposition modeling, J [33] S. McMains, Layered manufacturing technologies, Commun. ACM 48 (6) (2005)
Biomed Mater Res B Appl Biomater 101B (4) (2013) 610–619, https://doi.org/10. 50–56, https://doi.org/10.1145/1064830.1064858.
1002/jbm.b.32863. [34] Q. Chen, C. Zhu, G.A. Thouas, Progress and challenges in biomaterials used for bone
[14] S. Qnagoruwa, S. Bose, A. Bandyopadhyay, Fused Deposition of Ceramics (FDC) and tissue engineering: bioactive glasses and elastomeric composites, Prog. Biomater. 1
Composites, Solid FreeForm Fabrication, Texas, 2001 224–231. (1) (2012) 1–22, https://doi.org/10.1186/2194-0517-1-2.
J.-W. Tseng et al. / Materials and Design 140 (2018) 209–221 221

[35] T. Kokubo, H.M. Kim, M. Kawashita, Novel bioactive materials with different me- [41] J.P. Davim, R. Cardoso, Effect of the reinforcement (carbon or glass fibres) on friction
chanical properties, Biomaterials 24 (13) (2003) 2161–2175, https://doi.org/10. and wear behaviour of the PEEK against steel surface at long dry sliding, Wear 266
1016/S0142-9612(03)00044-9. (7) (2009) 795–799, https://doi.org/10.1016/j.wear.2008.11.003.
[36] M. Amaral, M.A. Lopes, R.F. Silva, J.D. Santos, Densification route and mechanical [42] D. Garcia-Gonzalez, M. Rodriguez-Millan, A. Rusinek, A. Arias, Investigation of me-
properties of Si3N4-bioglass biocomposites, Biomaterials 23 (3) (2002) 857–862, chanical impact behavior of short carbon-fiber-reinforced PEEK composites,
https://doi.org/10.1016/S0142-9612(01)00194-6. Compos. Struct. 133 (2015) 1116–1126, https://doi.org/10.1016/j.compstruct.
[37] A. Ravaglioli, A. Krajewski, Bioceramics: Materials Properties Applications, Springer 2015.08.028.
Science & Business Media, 2012. [43] I. Hanafi, F.M. Cabrera, F. Dimane, J.T. Manzanares, Application of particle swarm op-
[38] R. Domingo, M. García, M.R. Gómez, Determination of energy during the dry drilling timization for optimizing the process parameters in turning of peek cf30 compos-
of PEEK GF30 considering the effect of torque, Procedia Eng. 63 (2013) 687–693, ites, Proc. Technol. 22 (2016) 195–202, https://doi.org/10.1016/j.protcy.2016.01.
https://doi.org/10.1016/j.proeng.2013.08.195. 044.
[39] F. Mata, V. Gaitonde, S. Karnik, J.P. Davim, Influence of cutting conditions on machin- [44] K.S. Katti, Biomaterials in total joint replacement, Colloids Surf. B: Biointerfaces 39
ability aspects of PEEK, PEEK CF 30 and PEEK GF 30 composites using PCD tools, J. (3) (2004) 133–142, https://doi.org/10.1016/j.colsurfb.2003.12.002.
Mater. Process. Technol. 209 (4) (2009) 1980–1987, https://doi.org/10.1016/j. [45] M. Niinomi, Mechanical properties of biomedical titanium alloys, Mater. Sci. Eng. A
jmatprotec.2008.04.060. 243 (1–2) (1998) 231–236, https://doi.org/10.1016/S0921-5093(97)00806-X.
[40] A. Avanzini, G. Donzella, A. Mazzù, C. Petrogalli, Wear and rolling contact fatigue of
PEEK and PEEK composites, Tribol. Int. 57 (2013) 22–30, https://doi.org/10.1016/j.
triboint.2012.07.007.

You might also like