You are on page 1of 7

journal of the mechanical behavior of biomedical materials 119 (2021) 104544

Contents lists available at ScienceDirect

Journal of the Mechanical Behavior of Biomedical Materials


journal homepage: http://www.elsevier.com/locate/jmbbm

3D printing of dental restorations: Mechanical properties of thermoplastic


polymer materials
Lisa Marie Schönhoff a, Felicitas Mayinger a, *, Marlis Eichberger a, Elena Reznikova b,
Bogna Stawarczyk a
a
Department of Prosthetic Dentistry, University Hospital, LMU Munich, Goethestrasse 70, 80336, Munich, Germany
b
Apium Additive Technologies GmbH, Siemensallee 84, 76187, Karlsruhe, Germany

A R T I C L E I N F O A B S T R A C T

Keywords: In the seminal field of 3D printing of dental restorations, the time and cost saving manufacturing of removable
Thermoplastic polymer materials and fixed dental prostheses from thermoplastic polymer materials employing fused filament fabrication (FFF) is
Polyphenylene sulfone gaining momentum. As of today, the additive manufacturing of the established semi-crystalline poly­
Polyetheretherketone
etheretherketone (PEEK) requires extensive post-processing and lacks precision. In this context, the amorphous
3D printing
polyphenylene sulfone (PPSU) may provide a higher predictability and reliability of the results. The aim of this
Mechanical properties
study was to investigate the mechanical properties of PPSU and PEEK processed by FFF (PPSU1-3D (PPSU Radel)
and PPSU2-3D (Ultrason P 3010 NAT)) or extrusion (PPSU1-EX (Radel R-5000 NT) and PEEK-CG (PEEK Juvora)).
Three-point flexural strength, two-body wear, and Martens hardness (HM) and indentation modulus (EIT) were
tested after aging. One-way ANOVA, the Kruskal-Wallis and the Pearson’s and Spearman’s correlation tests were
computed (α = 0.05).
PPSU1-3D and PPSU2-3D showed lower flexural strength values than PPSU1-EX and PEEK-CG. PPSU1-3D
showed the highest, and PEEK-CG and PPSU1-EX the lowest height loss. The highest HM and EIT results were
observed for PEEK-CG and the lowest for PPSU1-3D. Correlations were observed between all parameters except
for the application height.
In conclusion, the manufacturing process affected the flexural strength of PPSU, with 3D printed specimens
presenting lower values than specimens cut from prefabricated molded material. This finding indicates that the
3D printing parameters employed for the additive manufacturing of PPSU specimens in the present investigation
require further optimization. For 3D printed specimens, the quality of the filament showed an impact on the
mechanical properties, underlining the importance of adhering to high quality standards during filament
fabrication. Extruded PPSU led to comparable results with PEEK for flexural strength and two-body wear,
indicating this novel dental restorative material to be a suitable alternative to the established PEEK for the
manufacturing of both removable and fixed dental prostheses.

1. Introduction prepared by milling from prefabricated blanks (Miyazaki et al., 2009),


the manufacturing process of 3D-printing is gaining importance as a
The ever-increasing demand for highly esthetic prosthetic restora­ potential material and time saving option (Kessler et al., 2020; Ngo
tions (Samorodnitzky-Naveh et al., 2007) has invigorated research ac­ et al., 2018). As of today, orthodontic splints and tooth models are
tivities in the field of tooth-shaded materials for both removable and increasingly manufactured from low-viscosity resins employing an ad­
fixed dental prostheses. The success of new material compositions is, ditive work-flow. Unreacted monomers on the surface of the print
however, not only measured by the clinical long-term performance of structures, do, however, require extensive post-processing and can limit
the fabricated restorations, but the ease and reliability of their the application of this material group in the oral cavity due to their high
manufacturing process. residual monomer content (Kurzmann et al., 2017). Against this back­
While silicate and zirconia ceramics are nowadays routinely drop, high-performance thermoplastics present a favorable alternative,

* Corresponding author.
E-mail address: felicitas.mayinger@med.uni-muenchen.de (F. Mayinger).

https://doi.org/10.1016/j.jmbbm.2021.104544
Received 25 January 2021; Received in revised form 14 April 2021; Accepted 16 April 2021
Available online 21 April 2021
1751-6161/© 2021 Elsevier Ltd. All rights reserved.
L.M. Schönhoff et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104544

as these inert materials possess a high biocompatibility (Wenz et al., To evaluate the mechanical properties of the thermoplastic polymer
1990). For fused filament fabrication (FFF), a thermoplastic filament is PPSU in comparison with the established PEEK, three-point flexural
melted in a nozzle, which moves to create a 3D-structure on a fixed strength, two-body wear and Martens parameters were tested initially
building platform layer by layer. In this context, the melting point of the and after artificial aging. The following null hypotheses were investi­
3D print material plays a key role. The high-performance polymer pol­ gated: (1) neither the material nor (2) aging show an influence on the
yetheretherketone (PEEK), which has been successfully employed in the flexural strength, (3) the material shows no influence on the two-body
additive manufacturing of implants and inlays (Prechtel et al., 2020b; wear, and (4) neither the material nor (5) the aging level show an in­
Sharma et al., 2020), has a melting temperature of 343 ◦ C (Zanjanijam fluence on the Martens parameters (HM and EIT).
et al., 2020), which has to be reached in the extrusion process of FFF. As
a lower melting temperature renders the printing process more practi­ 2. Material and methods
cable and cheaper, the thermoplastic polyphenylene sulfone (PPSU)
might represent a suitable alternative. In contrast to PEEK, which un­ In total, 368 specimens were fabricated from two different PPSU
dergoes a phase transformation during cooling and occurs in a materials (PPSU1, n = 184; PPSU2, n = 92) and PEEK (n = 92) (Fig. 1,
semi-crystalline state after complete solidification, with the Table 1).
crystalline-amorphous phase fraction being highly dependent on the The following properties were tested, and compared: i. flexural
challenging adjustment of the cooling parameters, as rapid cooling strength (N = 240, n = 20), ii. two-body wear (N = 48, n = 12), and iii.
prevents the formation of crystalline structures, PPSU only exists in an Martens parameters, namely Martens hardness (HM) and indentation
amorphous solid state (Dorf et al., 2018). This could entail a higher modulus (EIT) (N = 80, n = 20).
consistency on the part of printed PPSU restorations compared to PEEK,
where the crystallinity exerts a strong impact on material properties,
2.1. Specimen preparation
with a high percentage of crystalline phase increasing a restoration’s
mechanical properties, while an increase in amorphous phase content
Specimens with dimensions of 10 × 10 × 4 mm (cubic geometry, n =
can improve its optical properties, such as translucency (Wang et al.,
128) for two-body wear and Martens parameters measurements and 2 ×
2019; Yang et al., 2017; Zanjanijam et al., 2020). As PPSU possesses a
3 × 15 mm (bar geometry, n = 80) for flexural strength testing were
tensile strength of 55.0–76.9 MPa (in comparison: PEEK: 90.0–100.0
prepared using different manufacturing techniques.
MPa) and a modulus of elasticity of 2.14 GPa (in comparison: PEEK:
3D-printed specimens (PPSU1-3D and PPSU2-3D, n = 184) were
3.5–20 GPa), as well as a high resistance to hydrolysis and different
printed (P115, Apium Additive Technologies) from the two different
sterilization methods (Dizon et al., 2018; Evans et al., 2016; Lee et al.,
filaments PPSU1 and PPSU2 with a 0.4 mm nozzle, tempered to 410 ◦ C,
2017), this material meets the requirements for a potentially successful
in a controlled ambient temperature resulting from the 110 ◦ C printing
use in the additive manufacturing of dental restorations. The indication
bed temperature and the 150 ◦ C adaptive heating system (Apium Ada­
spectrum of PPSU could include single and multi-unit fixed dental
pative Heating System; Apium Additive Technologies) at a speed of 400
prostheses, clasps and frameworks for removable dental prostheses and
mm/min and layer thickness of 0.15 mm. Layers (XY-plane: 10 × 10
implant structures. A recent publication has reported promising results
mm) were printed with a Z building height of 4 mm. The force applied
for the tensile bond strength between PPSU and a veneering composite
during testing was perpendicular to the layer orientation.
resin, hereby providing proof of principle for the feasible esthetic
Extruded specimens (PPSU1-EX and PEEK-CG, n = 184) were cut
veneering of this material group (Schönhoff et al., 2021).
from molded material with a diamond saw (Secotom-50, Struers) under

Fig. 1. Study design.

2
L.M. Schönhoff et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104544

Table 1
Abbreviations, materials, manufacturers, compositions, geometry, color and lot numbers used.
Abbreviation Material Manufacturer Composition Geometry Color Lot No.

PPSU1-EX Radel R-5000 NT Solvay PPSU 170 × 20 × 4 mm plate Translucent n.a.


PPSU1-3D (Fil-A-Gehr PPSU) PPSU Radel Gehr PPSU filament n.a.
PPSU2-3D Ultrason P 3010 NAT BASF PPSU n.a.
PEEK-CG PEEK Juvora Juvora PEEK 98 × 22 mm circular blank Opaque (grey) WO000042IDML

water-cooling by using a diamond disc (M1D13, Struers). The edges of with σ: Flexural strength (MPa), P: Fracture load (N), l: the test span
the bar specimens were broken by pulling them over SiC-Foil P1200 (mm), w: Specimen width (mm), b: Specimen thickness (mm).
(Struers).
Cubic specimens were embedded in acrylic embedding resin (Scan­ 2.3. Two-body wear
DiQuick A, and B, ScanDia), and polished up to grain P2000 (SiC-Foil, Ø
200 mm, Struers) under water cooling with an automatic grinding and Embedded cubic specimens (n = 48) were subjected to 400,000
polishing device (Tegramin 20, Struers), and cleaned in distilled water chewing- and 2000 thermocycles between 5 ◦ C, and 55 ◦ C (duration of
for 5 min using an ultrasonic unit (Sonorex, Bandelin). 60 s for each thermocycle) in a chewing simulator (Chewing Simulator
CS-4, SD Mechatronik) with a vertical load of 50 N, and a sliding
2.2. Flexural strength movement of 0.7 mm employing steel antagonists. This aging procedure
is supposed to correlate with in vivo conditions after 3 years (Leinfelder
The definitive dimensions of the bar specimens (n = 300) were and Suzuki, 1999). Afterwards, specimens were cleaned in an ultrasonic
determined with a digital micrometer (Mitutoyo IP65, Mitutoyo Cor­ unit for 5 min (Sonorex, Bandelin) with distilled water. For two-body
poration) with an accuracy of ±0.05 mm. Specimens were tested wear analysis, the surface of each specimen was digitized with a trian­
initially (n = 20), as well as after 5000 (n = 20), and 10,000 thermo­ gulation sensor (Las 20, SD Mechatronik) after chewing simulation. The
cycles (TC) between 5 ◦ C and 55 ◦ C with 20 s dwelling time. Aging with dataset was loaded in a software for digital image correlation (GOM
5000 and 10,000 TC was aimed to mimic clinical conditions after Correlate, GOM) and a plane was created to simulate the surface before
approximately 6 months or 1 year in vivo (Gale and Darvell, 1999). chewing simulation, which was superimposed with the scanning data
Three-point flexural strength was tested in a universal testing machine using a 3D matching procedure. The loss (removal) and application of
(Zwick/Roell 1445 RetroLine, ZwickRoell) with a crosshead speed of 1 material was displayed graphically and computed numerically (Fig. 2).
mm/min. Flexural strength was calculated using the following formula:
3Pl
σ=
2wb2

Fig. 2. Graphic display of the height loss for a PPSU1-EX specimen.

3
L.M. Schönhoff et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104544

2.4. Martens parameters < 0.001), the parameter height was displayed and discussed to allow a
greater comparability to the current literature. The material showed an
On cubic embedded specimens (N = 80), HM (N/mm2), and EIT (kN/ influence on the height loss (η2p = 0.824, p < 0.001). PPSU1-3D showed
mm2) were assessed (n = 20) using a universal hardness testing machine the highest height loss (p < 0.001), while PEEK-CG and PPSU1-EX
(ZHU 0.2/Z2.5, ZwickRoell) by pressing a Vickers diamond indenter (α showed the lowest height loss (p < 0.001). No differences in the
= 136◦ ) with a maximum load of 9.807 N for 20 s vertically into the height application were found between materials (p > 0.790).
specimen’s surface. As 11 out of 15 HM groups and all EIT groups showed a violation of
HM, and EIT were calculated, and force-displacement diagrams the assumption of normality (Table 4), non-parametric tests were
recorded (testX-pert V12.3 Master, ZwickRoell). All specimens were performed.
tested longitudinally after aging by thermocycling between 5 ◦ C, and The material showed an influence on HM and EIT (η2p>0.938, p <
55 ◦ C with 20 s dwelling time after 5000 TC, 10,000 TC, 10,000 TC plus 0.001). The highest HM and EIT values were found for PEEK-CG (p <
36 days dry storage, and 20,000 TC. These aging conditions were aimed 0.001) and the lowest for PPSU1-3D (p < 0.001). For both HM and EIT,
to mimic clinical conditions after approximately 6 months, 1 year, 1 year no differences were found between aging levels (p > 0.098).
and 1 month, and 2 years in vivo (Gale and Darvell, 1999). Correlations were observed between all parameters (R > 0.617, p <
0.001) except for the application height (R > − 0.013–0.057, p > 0.351).
2.5. Statistical analysis
4. Discussion
The statistical evaluation of the results was carried out with a
descriptive analysis, followed by the Kolmogorov-Smirnov test to verify The aim of this study was to evaluate the mechanical properties of
compliance with the normal distribution. Correlation coefficients were the high-performance polymer PPSU in comparison with the established
determined using Pearson’s and Spearman’s correlation. One-way PEEK. As the choice of material showed an influence on flexural
ANOVA, Scheffé-post-hoc and Kruskal-Wallis test were performed to strength, two-body wear and Martens parameters, the null hypotheses
analyze significant differences between the tested groups (IBM Statistics (1), (3) and (4) were rejected. The second null hypothesis was partially
SPSS 26.0, International Business Machines Corporation) (α = 0.05). rejected since differences in flexural strength were observed for PPSU1-
EX and PPSU1-3D with respect to the different aging levels. For the
3. Results Martens parameters, aging showed no influence on either HM or EIT. The
fifth null hypothesis was thus accepted.
For flexural strength, 3 out of 12 tested groups showed a violation of Since PPSU1-3D and PPSU1-EX specimens were manufactured from
the assumption of normality (Table 2). the same core material, but PPSU1-EX showed superior flexural strength
Thus, non-parametric tests were performed. The material showed the results, the manufacturing process seems to play an essential role. While
highest influence on flexural strength (partial eta-squared ((η2p) = 0.770, PPSU1-EX specimens were cut from a prefabricated PPSU bar, that had
p < 0.001) followed by the interaction of material and aging level (η2p = been produced from PPSU granulate by injection molding, PPSU1-3D
0.191, p < 0.001) and the aging level (η2p = 0.079, p < 0.001). Regarding specimens were processed in an FFF printer from a filament, that had
the different materials, no differences between PPSU1-3D and PPSU2- been extruded from PPSU granulate in a preceding step. For PPSU1-3D,
3D (p > 0.052) and between PPSU1-EX and PEEK-CG (p > 0.097) one of the processing steps seems to have weakened the material. Prior
were observed initially and after 10,000 TC, as PPSU1-3D and PPSU2-3D to filament extrusion, the material must be carefully dried and exact
showed lower flexural strength values than PPSU1-EX and PEEK-CG (p temperatures have been maintained in the extrusion process itself to
< 0.001). After 5000 TC, PPSU1-EX and PEEK-CG showed the highest ensure high quality standards. In the printing process, additional factors
flexural strength values, while the lowest values were observed for to the necessary renewed pre-drying of the filament have been observed
PPSU2-3D (p < 0.001) (Fig. 3). to show an impact on the mechanical properties of the printed objects,
For PPSU1-EX, initial flexural strength results were the highest, fol­ namely the printing parameters (printing speed, degree of infill and
lowed by values at 5000 TC and 10,000 TC (p < 0.008). In PPSU1-3D, no printing strategy) and the employed printing temperatures (temperature
differences were observed between initial flexural strength results and of the nozzle, printing bed and printing chamber) (Li and Lou, 2020;
values after 5000 TC (p = 0.116), and between 5000 TC and 10,000 TC Prechtel et al., 2020a; Wang et al., 2019). While additive manufacturing
(p = 0.222), while values at 10,000 TC were higher than initial flexural promises a time efficient and material saving manufacturing process
strength results (p = 0.009). For the materials PPSU2-3D and PEEK-CG, (Kessler et al., 2020), it is yet to be determined in how far individually
no differences in flexural strength were observed regarding the different printed layers fuse together to form a homogenous structure that
aging levels (p > 0.252). matches material properties obtained by injection molding (Wickra­
For two-body wear, none of the removal or application groups masinghe et al., 2020). The benefit of this standardized process, that
showed a violation of the assumption of normality (Table 3). employs pressure and heat to ensure consistent material compositions,
Thus, parametric tests were performed. As the Pearson coefficient has been described for denture bases made from poly­
indicated a correlation between volume and height values (R = 0.966, p methylmethacrylate (Keenan et al., 2003). It has yet to be determined
whether the lack of this manufacturing step is causal for the observed
Table 2
impaired mechanical properties, or whether the printing process itself
Descriptive statistics (Min/median/max) of flexural strength values (N/mm2) of that requires an additional heating of the filament during additive
the different materials after aging. manufacturing shows a negative effect. Future studies are needed to
investigate whether a reduction of the printing layer thickness below
PPSU1-EX PPSU1-3D PPSU2-3D PEEK-CG
0.15 mm and an adaption of the printing temperature may improve
Initial 149.2/158.1/ 49.5/83.0/ 55.3/78.8/ 84.7/139.1/ PPSU properties to match those of injection molded counterparts, a feat
169.4Bc 94.1*Aa 99.0Aa 222.9Ba
5000 TC 127.0/148.4/ 80.3/84.5/ 61.0/72.4/ 97.2/141.4/
recently reported for 3D printed PEEK (Li and Lou, 2020).
162.2Cb 92.24Bab 87.0Aa 194.7Ca After artificial aging with 5000 TC, PSSU1-3D showed higher flex­
10,000 97.4/112.8/ 80.0/87.6/ 62.2/74.6/ 106.6/130.1/ ural strength results than PPSU2-3D. For HM, EIT and two-body wear,
TC 157.9*Ba 92.03Ab 88.5*Aa 198.0Ba the performance of the two materials was, however, reversed, with
* indicates deviations from the normal distribution. PPSU2-3D showing higher Martens parameters and a higher resistance
ABC
indicate differences between materials within one aging level. to wear than PPSU1-3D. According to the manufacturers’ specifications,
abc
indicate differences between aging levels within one material. both PPSU1-3D and PPSU2-3D filaments are manufactured from unfilled

4
L.M. Schönhoff et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104544

Fig. 3. Box plot of the flexural strength values of the different materials after different aging procedures.

Table 3
Descriptive statistics (Mean ± SD; 95% CI) of height loss and application (mm) of the different materials after 400,000 chewing cycles.
PPSU1-EX PPSU1-3D PPSU2-3D PEEK-CG

Mean ± SD 95% CI Mean ± SD 95% CI Mean ± SD 95% CI Mean ± SD 95% CI

Height loss − 0.285 ± [− 0.318; − 0.717 ± [− 0.786; − 0.389 ± [− 0.446; − 0.218 ± [− 0.291;
0.050AB − 0.253] 0.106C − 0.649] 0.088B − 0.332] 0.113A − 0.146]
Height 0.050 ± 0.023A [0.034; 0.065] 0.054 ± 0.027A [0.035; 0.072] 0.043 ± 0.031A [0.023; 0.063] 0.048 ± 0.023A [0.032; 0.063]
application

CI, confidence interval; SD, standard deviation.


ABC
indicate differences between materials.

homogeneity. In the extrusion process, the employed pressure influences


Table 4
the continuity of the filament diameter and the surface quality. The
Descriptive statistics (Min/median/max) for HM (N/mm2) and EIT (kN/mm2) of
pressure is in turn controlled by the extrusion speed in conjunction with
the different materials after aging.
the die diameter. The filament quality thus exerts an important influence
PPSU1-EX PPSU1-3D PPSU2- PEEK-CG on the print result (Wang et al., 2020). Noteworthy, PPSU2-3D was the
3D
only material besides the PEEK control group where no differences in
Initial HM 115/121/ 102/111/ 90/113/ 204/207/ flexural strength were observed for the different aging levels. This un­
124* 113* 149* 209
derlines the strong performance of this additively manufactured group
EIT 2.9/2.9/ 2.5/2.6/ 2.2/2.6/ 4.6/4.8/
3.1* 2.7* 3.6* 4.9* already seen for the Martens parameters and two-body wear, just as
5000 TC HM 106/115/ 99/109/ 85/108/ 193/206/ PPSU1-3D was affected by aging, with its flexural strength values at 10,
118* 111* 145* 209* 000 TC being higher than initial results. This might be caused by an
EIT 2.7/2.8/ 2.5/2.6/ 2.2/2.6/ 4.4/4.8/ ongoing post-polymerization process induced by the high temperature
3.0* 2.6* 3.6* 4.8*
10,000 TC HM 98/116/ 97/108/ 84/109/ 197/204/
water bath employed for thermocycling. As the constant performance of
119* 110* 146* 208 PEEK during aging is seen as one of its strongest features (Liebermann
EIT 2.6/2.8/ 2.5/2.6/ 2.3/2.6/ 4.6/4.7/ et al., 2016; Micovic et al., 2020; Taufall et al., 2016), follow-up studies
3.3* 2.6* 3.6* 4.8* are warranted to examine this promising novel material composition in
10,000 TC + 36 HM 108/116/ 97/108/ 65/112/ 197/206/
the field of dentistry with regard to its resistance to aging more closely.
days dry storage 127* 114 148* 210*
EIT 2.7/2.9/ 2.4/2.6/ 2.2/2.7/ 4.4/4.8/ The puzzling observation that the two additively manufactured mate­
3.4* 2.7* 3.6* 4.9* rials differed in their performance regarding flexural strength on the one
20,000 TC HM 99/116/ 88/105/ 99/110/ 199/203/ side, and Martens parameters and wear resistance on the other side, will
126* 109* 146* 205 have to be investigated further.
EIT 2.7/2.8/ 2.3/2.6/ 2.4/2.6/ 4.6/4.8/
3.6* 2.6* 3.6* 4.9*
As PSSU1-EX showed comparable properties to the tried-and-tested
control group (Schwitalla et al., 2015; Wimmer et al., 2016) for both
* indicates deviations from the normal distribution. flexural strength and height loss during two-body wear, this is a prom­
ising finding for the future application of PPSU materials. At the same
PPSU. As the two core materials were processed with the same printing time, PPSU1-EX was the only material that showed a negative impact of
parameters, a possible difference can therefore only be found in the artificial aging. Although a hydrophobic material, PPSU may be weak­
fabrication of the filament. During this process, contaminations (e.g., ened by water absorption during the thermocycling process. As the
dental wax) can be incorporated into the filament and impair its flexural strength of the additively manufactured group PPSU2-3D was,

5
L.M. Schönhoff et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104544

however, not affected by aging and none of the tested groups showed an Projektträger des BMWi).
impact of aging on the surface properties HM and EIT, this property
needs to be examined in follow-up studies. Ethical approval
Interestingly, correlations were observed between all parameters
except for the height application. Furthermore, no differences in the This article does not contain any studies with human participants or
height application were found between materials. Although this animals performed by any of the authors.
parameter may provide valuable information about a thermoplastic’s
mechanical performance, as this material group has been observed to Informed consent
undergo plastic deformation during force application (El-Qoubaa and
Othman, 2015), this parameter may be of negligible importance for the For this type of study, formal consent was not required.
tested materials.
One limitation of the present investigation is the potentially uneven
CRediT authorship contribution statement
distribution of flaws in the test specimens distorting the observed three-
point flexural strength values. Future investigations should thus include
Lisa Marie Schönhoff: Investigation, Formal analysis, Writing –
the computation and analysis of Weibull characteristic strength and
original draft. Felicitas Mayinger: Formal analysis, Writing – original
Weibull modulus. Furthermore, the present study only investigated one
draft, Writing – review & editing. Marlis Eichberger: Investigation.
printing and testing modality. As previous investigations have reported
Elena Reznikova: Investigation. Bogna Stawarczyk: Conceptualiza­
a vertical printing strategy to yield lower results than observed for
tion, Methodology, Formal analysis, Writing – review & editing,
specimens manufactured in a horizontal printing direction, with a
Supervision.
delamination taking place between layers during force application, a
horizontal printing strategy and perpendicular application of force was
used in the present study (Kessler et al., 2021; Prechtel et al., 2020a; Declaration of competing interest
Unkovskiy et al., 2018). While this strategy should be employed in a
clinical setting to ensure the best outcome for the patient, with a hori­ The authors declare that they have no known competing financial
zontal printing and perpendicular testing set-up providing the highest interests or personal relationships that could have appeared to influence
reliability (Kessler et al., 2021), it could embellish the obtained in vitro the work reported in this paper.
results, as both vertical and horizontal forces occur during masticatory
function. Future studies should thus focus on test set-ups, where the
Acknowledgements
employed force is applied in the same direction as the horizontal print
layers.
The authors would like to thank Gehr GmbH and BASF SE for
Based on the present findings, the choice of material and
providing material for this study, as well as Apium Additive Technolo­
manufacturing process seems to play an important role, which is
gies GmbH for providing material and printing specimens.
underlined by the parameter material showing the highest influence on
both flexural strength and Martens parameters. While this study in­
References
dicates a potentially successful long-term clinical use of PPSU in the field
of both removable and fixed dental prostheses with regard to its me­ Dizon, J.R.C., Espera, A.H., Chen, Q., Advincula, R.C., 2018. Mechanical characterization
chanical properties, it has yet to be determined whether an optimization of 3D-printed polymers. Additive Manufacturing 20, 44–67.
in the production of the printing filament and the additive Dorf, T., Perkowska, K., Janiszewska, M., Ferrer, I., Ciurana, J., 2018. Effect of the main
process parameters on the mechanical strength of polyphenylsulfone (PPSU) in
manufacturing process itself will allow comparable values to conven­
ultrasonic micro-moulding process. Ultrason. Sonochem. 46–58.
tionally fabricated specimens. El-Qoubaa, Z., Othman, R., 2015. Tensile behavior of polyetheretherketone over a wide
range of strain rates. International Journal of Polymer Science, 275937, 2015.
Evans, N.T., Irvin, C.W., Safranski, D.L., Gall, K., 2016. Impact of surface porosity and
5. Conclusions
topography on the mechanical behavior of high strength biomedical polymers.
J Mech Behav Biomed Mater 59, 459–473.
Based on the findings of this in vitro study, the following conclusions Gale, M.S., Darvell, B.W., 1999. Thermal cycling procedures for laboratory testing of
were drawn: dental restorations. J. Dent. 27, 89–99.
Keenan, P.L., Radford, D.R., Clark, R.K., 2003. Dimensional change in complete dentures
fabricated by injection molding and microwave processing. J. Prosthet. Dent 89,
1. The manufacturing process affected the flexural strength of the 37–44.
thermoplastic polymer PPSU, with 3D printed specimens presenting Kessler, A., Hickel, R., Ilie, N., 2021. In vitro investigation of the influence of printing
direction on the flexural strength, flexural modulus and fractographic analysis of 3D-
lower values than specimens cut from prefabricated molded mate­ printed temporary materials. Dent. Mater. J. https://doi.org/10.4012/dmj.2020-147
rial. This finding indicates that the 3D printing parameters employed online ahead of print.
for the additive manufacturing of PPSU specimens in the present Kessler, A., Hickel, R., Reymus, M., 2020. 3D printing in dentistry-state of the art. Operat.
Dent. 45, 30–40.
investigation require further optimization. Kurzmann, C., Janjic, K., Shokoohi-Tabrizi, H., Edelmayer, M., Pensch, M., Moritz, A.,
2. The quality of the filament processed in 3D printing showed an Agis, H., 2017. Evaluation of resins for stereolithographic 3D-printed surgical
impact on the mechanical properties of PPSU, underlining the guides: the response of L929 cells and human gingival fibroblasts. BioMed Res. Int.
4057612.
importance of adhering to high quality standards during filament Lee, J.-Y., An, J., Chua, C.K., 2017. Fundamentals and applications of 3D printing for
fabrication. novel materials. Applied Materials Today 7, 120–133.
3. The flexural strength and height loss during two-body wear of Leinfelder, K.F., Suzuki, S., 1999. In vitro wear device for determining posterior
composite wear. J. Am. Dent. Assoc. 130, 1347–1353.
extruded PPSU was comparable to PEEK. PPSU, a novel material for
Li, Y., Lou, Y., 2020. Tensile and bending strength improvements in PEEK parts using
dental restorations, could thus represent a suitable alternative to the fused deposition modelling 3D printing considering multi-factor coupling. Polymers
established PEEK for the manufacturing of both removable and fixed 12, 2497.
dental prostheses. Liebermann, A., Wimmer, T., Schmidlin, P.R., Scherer, H., Loffler, P., Roos, M.,
Stawarczyk, B., 2016. Physicomechanical characterization of polyetheretherketone
and current esthetic dental CAD/CAM polymers after aging in different storage
Funding media. J. Prosthet. Dent 115, 321–328 e322.
Micovic, D., Mayinger, F., Bauer, S., Roos, M., Eichberger, M., Stawarczyk, B., 2020. Is
the high-performance thermoplastic polyetheretherketone indicated as a clasp
This work was supported by research program ZF4052006AW8 (AiF material for removable dental prostheses? Clin. Oral Invest. https://doi.org/
Projekt GmbH, Berlin, Germany, ZIM-Kooperationsprojekte, 10.1007/s00784-020-03603-y online ahead of print.

6
L.M. Schönhoff et al. Journal of the Mechanical Behavior of Biomedical Materials 119 (2021) 104544

Miyazaki, T., Hotta, Y., Kunii, J., Kuriyama, S., Tamaki, Y., 2009. A review of dental Unkovskiy, A., Bui, P.H., Schille, C., Geis-Gerstorfer, J., Huettig, F., Spintzyk, S., 2018.
CAD/CAM: current status and future perspectives from 20 years of experience. Dent. Objects build orientation, positioning, and curing influence dimensional accuracy
Mater. J. 28, 44–56. and flexural properties of stereolithographically printed resin. Dent Mater 34,
Ngo, T.D., Kashani, A., Imbalzano, G., Nguyen, K.T.Q., Hui, D., 2018. Additive e324–e333.
manufacturing (3D printing): a review of materials, methods, applications and Wang, P., Zou, B., Xiao, H., Ding, S., Huang, C., 2019. Effects of printing parameters of
challenges. Compos. B Eng. 143, 172–196. fused deposition modeling on mechanical properties, surface quality, and
Prechtel, A., Reymus, M., Edelhoff, D., Hickel, R., Stawarczyk, B., 2020a. Comparison of microstructure of PEEK. J. Mater. Process. Technol. 271, 62–74.
various 3D printed and milled PAEK materials: effect of printing direction and Wang, Y., Muller, W.D., Rumjahn, A., Schwitalla, A., 2020. Parameters influencing the
artificial aging on Martens parameters. Dent Mater 36, 197–209. outcome of additive manufacturing of tiny medical devices based on PEEK. Materials
Prechtel, A., Stawarczyk, B., Hickel, R., Edelhoff, D., Reymus, M., 2020b. Fracture load of 13, 466.
3D printed PEEK inlays compared with milled ones, direct resin composite fillings, Wenz, L.M., Merritt, K., Brown, S.A., Moet, A., Steffee, A.D., 1990. In vitro
and sound teeth. Clin. Oral Invest. 3457–3466. biocompatibility of polyetheretherketone and polysulfone composites. J. Biomed.
Samorodnitzky-Naveh, G.R., Geiger, S.B., Levin, L., 2007. Patients’ satisfaction with Mater. Res. 24, 207–215.
dental esthetics. J. Am. Dent. Assoc. 138, 805–808. Wickramasinghe, S., Do, T., Tran, P., 2020. FDM-based 3D printing of polymer and
Schönhoff, L.M., Mayinger, F., Eichberger, M., Lösch, A., Reznikova, E., Stawarczyk, B., associated composite: a review on mechanical properties, defects and treatments.
2021. Three-dimensionally printed and milled polyphenylene sulfone materials in Polymers 12, 1529.
dentistry: tensile bond strength to veneering composite resin and surface properties Wimmer, T., Huffmann, A.M., Eichberger, M., Schmidlin, P.R., Stawarczyk, B., 2016.
after different pretreatments. J. Prosthet. Dent. https://doi.org/10.1016/j. Two-body wear rate of PEEK, CAD/CAM resin composite and PMMA: effect of
prosdent.2020.12.042 online ahead of print. specimen geometries, antagonist materials and test set-up configuration. Dent Mater
Schwitalla, A.D., Spintig, T., Kallage, I., Muller, W.D., 2015. Flexural behavior of PEEK 32, e127–136.
materials for dental application. Dent Mater 31, 1377–1384. Yang, C., Tian, X., Li, D., Cao, Y., Zhao, F., Shi, C., 2017. Influence of thermal processing
Sharma, N., Aghlmandi, S., Cao, S., Kunz, C., Honigmann, P., Thieringer, F.M., 2020. conditions in 3D printing on the crystallinity and mechanical properties of PEEK
Quality characteristics and clinical relevance of in-house 3D-printed customized material. J. Mater. Process. Technol. 248, 1–7.
polyetheretherketone (PEEK) implants for craniofacial reconstruction. J. Clin. Med. Zanjanijam, A.R., Major, I., Lyons, J.G., Lafont, U., Devine, D.M., 2020. Fused filament
9, 2818. fabrication of PEEK: a review of process-structure-property relationships. Polymers
Taufall, S., Eichberger, M., Schmidlin, P.R., Stawarczyk, B., 2016. Fracture load and 12, 1665.
failure types of different veneered polyetheretherketone fixed dental prostheses.
Clin. Oral Invest. 20, 2493–2500.

You might also like