Вы находитесь на странице: 1из 11

Int J Adv Manuf Technol (2018) 95:1685–1695

https://doi.org/10.1007/s00170-017-1314-x

ORIGINAL ARTICLE

An assessment of the effect of printing orientation, density,


and filler pattern on the compressive performance of 3D
printed ABS structures by fuse deposition
G. Domínguez-Rodríguez 1 & J. J. Ku-Herrera 2 & A. Hernández-Pérez 3

Received: 30 June 2017 / Accepted: 1 November 2017 / Published online: 13 November 2017
# Springer-Verlag London Ltd., part of Springer Nature 2017

Abstract Acrylonitrile butadiene styrene (ABS) specimens and heavy structures, respectively. A combination of shear
manufactured by fused deposition are tested under uniaxial and local buckling failure appeared in honeycomb structures
compression in order to judge the effectiveness of printing printed transversely with relative densities around 20–40%.
orientation, density, and filler patterns in terms of stiffness
and strength per printing time. The compressive properties Keywords 3D printing . Filling patterns . Compression .
of the 3D printed materials along the three orthogonal direc- Ultimate compressive strength . Failure mechanism . Rapid
tions are studied on cylindrical specimens filled with honey- prototyping
comb and rectangular patterns. In order to achieve different
densities, five filler percentages (0, 20, 30, 40, and 100%) are
employed for each type of structure. Specimens filled with 1 Introduction
honeycomb patterns are stiffer and stronger than those with
rectangular patterns only when they are oriented along the The recent advances in rapid prototyping or additive
applied load. However, structures with rectangular patterns manufacturing technology have provided valuable tools to
only require roughly half of printing time of those filled hon- accelerate the design and manufacturing process. Designing
eycomb cells, which yields effective rectangular structures complex geometrical parts using conventional prototyping
with high elastic properties per printing time. Stress–strain methods such as mold casting, drilling, and milling can be
curves reveal that compressive strength and stiffness increase very expensive or may take weeks to months depending
with respect to the structure density. Patterns printed along the on the complexity of the design [1–4], while using rapid
loading direction present higher strength and stiffness than on prototyping, the part is quickly built in one single step via
the other orthogonal orientations. Local buckling and com- a layered manufacturing process [5]. Rapid prototyping
pressive failure mechanisms are identified for light weight allows not only to evaluate the form, fit, function, and
ergonomics of prototypes [2] but also to produce custom
products at relatively low prices compared to traditional
* A. Hernández-Pérez mass production schemes [6]. Rapid prototyping has
hepadrian@gmail.com quickly gained popularity among engineers and designers,
given its versatility, and it has been explored for a wide
1
Centro de Investigación en Corrosión, Universidad Autónoma de range of applications, such as biomedical for dental and
Campeche, Avenida Héroe de Nacozari No. 480, 24079 San
prosthetics [7], medicine [8], electrochemical [9], elec-
Francisco de Campeche, Campeche, Mexico
2
tronics [10], strain sensing [11, 12], biochemistry [13],
Departamento de Síntesis de Polímeros, CONACYT-Centro de
and automotive [6] and, in general, for mechanical re-
Investigación en Química Aplicada, Blvd. Enrique Reyna
Hermosillo No. 140, San José de los Cerritos, placement of parts [3]. Some rapid prototyping technolo-
25294 Saltillo, Coahuila, Mexico gies available to date are laminated object manufacturing,
3
Departamento de Ingeniería Mecánica, Universidad de Guanajuato stereolithography, solid ground curing, selective laser
Campus Irapuato-Salamanca, Carretera Salamanca-Valle de Santiago sintering, direct shell production casting, and fused deposition
km 3.5 +1.8, 36885 Salamanca, Guanajuato, Mexico modeling (FDM) [2]. Each of these additive manufacturing
1686 Int J Adv Manuf Technol (2018) 95:1685–1695

technologies has its own advantages and limitations including certain printing parameters [4, 14–20, 25, 26], other parame-
material types, built volume, manufacturing speed, part qual- ters such as filling geometrical pattern, pattern orientation, and
ity, and dimensional accuracy [3]. Among all rapid filling content have not been systematically investigated. In
prototyping techniques, FDM is, by far, the most popular order to get insight on this topic, reduce the printing time and
and affordable version of 3D printing and it represents the vast materials, improve their mechanical properties, and provide
majority of consumers nowadays. However, one of the main guidelines to designers, vendors, and users, this work investi-
drawbacks of FDM is its high anisotropy and weak interlayer gates the role of the geometrical filling pattern, filling orien-
bonding which induce failure of the printed parts due to buck- tation, and filling content on the mechanical properties of
ling, delamination, bending, and shear [4]. Therefore, in order ABS parts subjected to uniaxial compression up to failure.
to capitalize the best of the FDM 3D printing technology, it is In addition, the cost-to-benefit ratio of the 3D printed struc-
necessary to study the printing parameters that can affect the tures is judged simultaneously, taking into account the stiff-
mechanical properties under different loading scenarios. Some ness, strength, and density.
printing parameters that have been investigated involve raster
orientation (orientation of filling pattern) [4, 14–16], raster
width (width of raster pattern) [14, 17], raster gap (space be- 2 Materials and methods
tween adjacent rasters) [14, 16, 17], printing orientation (part
build orientation) [18, 19], contour layers [20, 21], layer thick- 2.1 Material and specimen configuration
ness [14, 17], and printing temperature [22]. For unidirection-
al printed parts, it has been shown that elastic properties are All specimens were manufactured by using a FlashForge
practically identical regardless of the layer orientation [23]; Dreamer 3D printer with an extruder diameter of 0.4 mm
however, their tensile [4, 23] and compressive [18] strengths and an ABS filament with a 1.75 mm diameter. All specimens
are highly depended on the layer orientation. For instance, Es- were manufactured with such filament extruded at a tempera-
Said et al. [4] observed that the tensile strength of specimens ture of 240 °C, a layer thickness (raster gap) of 0.2 mm, two
with layers oriented parallel to the loading direction (0° spec- contour layers, and a printing speed of 80 mm/s. The temper-
imens) was twice of that of the specimens with layers oriented ature at the printer bed was set to 100 °C. Specimens with five
perpendicular (90° specimens) to the loading direction and different filling contents were analyzed, two of them for com-
was three times higher than that of the specimens with 45° parative purposes (0 and 100%) and the others are the subjects
layers. Zhang et al. [24] examined the tensile, creep, and fa- of study (20, 30, and 40%). These filling percentages were
tigue responses of acrylonitrile butadiene styrene (ABS) with chosen herein since they are the typically employed ones for
0°, 45°, and 90° printing orientations. They observed that 0° manufacturers to keep a cost-efficient but stable and stiff
oriented structures have the highest Young’s modulus and structure [27]. For illustrative purposes, Fig. 1 shows sche-
ultimate strength whereas the 90° oriented structures present matics of the filling patterns and material coordinate system.
the highest creep resistance. Under compression, Lee et al. As seen in Fig. 1, two rectangular configurations and one
[18] found that the compressive strength of 0° ABS specimens honeycomb configuration were chosen, since they are the
is ~ 12% higher than that of the 90° specimens. Domingo- most commonly used patterns for 3D printing [27]. These
Espin et al. [19] determined the nine independent elastic prop- filler shapes are also made from the same ABS filament used
erties that define the stiffness matrix of an orthotropic material to build the outer surface of the printed parts. The first rect-
on polycarbonate samples. Regarding the elastic deformation, angular filling pattern (labeled as Ra) corresponds to a
they observed that printed specimens can be considered iso- 0/90 cross-ply configuration (Fig. 1a) while the second
tropic; however, beyond yielding, the building orientation one (labeled as Rb, Fig. 1b) corresponds to the + 45/−
plays an important role on their plastic properties [19]. On 45 configuration; the honeycomb pattern is labeled as
the other hand, Croccolo et al. [20] investigated the effect of Hc, and it is shown in Fig. 1c.
the number of contour layers on the tensile properties of ABS Three loading directions are considered, one for each material
specimens, finding that increasing the number of contour coordinate axis (1-, 2-, and 3-directions). Figure 2 illustrates the
layers from 1 to 10, an improvements of ~ 17% in the elastic specimens loaded in compression, indicating their internal filling
modulus and ~ 24% in the ultimate strength are achieved. In a patterns. The 0/90 cross-ply (Ra) filling pattern (Fig. 1a) pro-
step forward, Casavola et al. [15] evaluated the effect of stack- duces a rectangular in-plane geometry, and the material can be
ing sequence and raster orientation on the mechanical proper- considered transversely isotropic (1–2-isotropic plane); i.e., their
ties of ABS and poly(lactic acid) 3D printed parts using clas- elastic properties on the 1- and 2-directions are equal but different
sical laminated plate theory. Using this theory, they were able to the transverse direction (3-direction). The + 45/− 45 specimens
to predict the effective elastic moduli of 3D printed parts with (Rb, Fig. 1b) also form a rectangular filling pattern with the exact
an error of 5–7% compared to the experimental results. Even properties to the 0/90 specimens in the 3-direction, but with
though there are some scientific works addressing the effect of different effective in-plane (1–2-directions) elastic properties.
Int J Adv Manuf Technol (2018) 95:1685–1695 1687

Fig. 1 Schematics of the filling


patterns and material coordinates.
a 0/90 (Ra). b + 45/− 45 (Rb).
c Honeycomb (Hc)

Finally, the honeycomb filling pattern is orthotropic with nine 2.2 Compressive test setup
independent elastic constants required to define their elastic be-
havior. In order to characterize the compressive elastic properties In order to analyze the effect of density, printing orientation,
of the specimen with all pattern configurations, six loading cases and filler pattern on the mechanical properties of 3D printed
are needed, which are represented by the two vertically and four specimens, axial compression tests were conducted following
horizontally printed cylindrical specimens shown in Fig. 2. All the procedure described in the ASTM D695 standard [28].
six cylindrical specimens are 25 mm tall with a radius of Quasi-static uniaxial compression tests were conducted by
12.5 mm, as recommended in the ASTM D695 standard [28]. using an Instron universal testing machine (model 8872)
Table 1 lists all printing configurations (22 sets) for the compres- equipped with a load cell of 25 kN under a 1.3 mm/min
sive testing, their nomenclature, and density (ρ). The left-hand cross-head speed with displacement control. For all tests, a
words in the specimen nomenclature represent the filler shape, 5 N pre-load was applied to the end of specimen cross-section
the numbers at the middle represent the orientation, and the to ensure full contact. Force and cross-head displacement were
right-hand numbers indicate the filler content percentage. recorded until the final collapse of the structure in order to
In this table, filling nomenclature is disclosed in the first identify the failure modes, the crushing loads, and the effect
column (from left to right) while the following columns of densification of the printed structure. In order to character-
describe the printing orientation, filling content, labels, ize the compressive mechanical response of the 3D printed
and densities, respectively. Labels of shell structures (0% specimens, Young’s modulus (E) or stiffness, compressive
filling content) are identified by the initial letter “N” stand- strength (σu), specific compressive stiffness (E∗ = E/ρ), and
ing for none. For each type of configuration, eight repli- specific compressive strength (σ*u =σu/ρ) were obtained from
cates were printed, making a total of 176 specimens the σ–ε curves. E was calculated as the slope of the stress–
manufactured and analyzed. In order to analyze the cell strain (σ–ε) curves in the linear elastic region. Finally,
shape after compression, these specimens were cut trans- crushing loads were identified as peak loads which yield some
versely with a diamond saw disk and then polished with degree of failure, such as local buckling, crack growth, and
sand paper of 3000 grain size. shattering failure.

Fig. 2 Schematics of the


specimen cross-section subjected
to compressive loadings. a 3-
direction for the 0/90 or + 45/− 45
specimen (Ra-3). b 1 (or 2)-
direction for the 0/90 specimen
(Ra-12). c 1 (or 2)-direction for
the + 45/− 45 specimen (Rb-12).
d 3-direction for the honeycomb
specimen (Hc-3). e 1-direction for
the honeycomb specimen (Hc-1).
f 2-direction for the honeycomb
specimen (Hc-2)
1688 Int J Adv Manuf Technol (2018) 95:1685–1695

Table 1 List of specimen


printing configurations for the Filling pattern Direction Filling content (%) Label Density (ρ, kg/m3)
compressive testing
None (N) 1=2 0 N-12-0% 263 ± 3
3 0 N-3-0% 282 ± 3
Rectangular 0/90 (Ra) 1=2 20 Ra-12-20% 455 ± 7
1=2 30 Ra-12-30% 540 ± 6
1=2 40 Ra-12-40% 628 ± 12
1=2 100 Ra-12-100% 993 ± 15
3 20 Ra-3-20% 480 ± 5
3 30 Ra-3-30% 567 ± 5
3 40 Ra-3-40% 644 ± 10
3 100 Ra-3-100% 997 ± 4
Rectangular + 45/− 45 (Rb) 1=2 20 Rb-12-20% 495 ± 10
1=2 30 Rb-12-30% 579 ± 16
1=2 40 Rb-12-40% 662 ± 20
Honeycomb (Hc) 1 20 Hc-1-20% 516 ± 11
1 30 Hc-1-30% 626 ± 8
1 40 Hc-1-40% 724 ± 15
2 20 Hc-2-20% 536 ± 20
2 30 Hc-2-30% 662 ± 30
2 40 Hc-2-40% 728 ± 18
3 20 Hc-3-20% 531 ± 21
3 30 Hc-3-30% 665 ± 11
3 40 Hc-3-40% 743 ± 6

Ra-12-100% = Rb-12-100%; Ra-3-100% = Rb-3-100%

3 Results and discussion 3D printed structures increase. This behavior has also been
observed by Jin et al. [29] in the compressive response of
3.1 Stress–strain curves aluminum hexagonal honeycombs. Ra-3-100% exhibits a
large bilinear σ–ε response whereas N-3-0% presents a short
Compressive strength and stiffness of all ABS printed speci- linear elastic σ–ε behavior until the structure suddenly col-
mens with different densities, filler contents, and orientations lapses after the peak stress. The picture of Ra-3-100% at
were measured from their compressive σ–ε curves. For all the the right-hand side of the σ–ε curves in Fig. 3a confirms
σ–ε curves, σu is considered as the first peak stress and it is that it failed by compression. For all Hc-3 specimens,
indicated with a red circle. Figure 3 shows the representative white marks on their surfaces are observed (see pictures),
σ–ε curves of specimens filled with a honeycomb pattern with suggesting that failure is due to a combination of local
different contents (20, 30, and 40%), loaded along the 3- plastic deformation and buckling. The cross-sectional pic-
direction (Fig. 3a), 1-direction (Fig. 3b), and 2-direction tures of Hc-3 samples show a plastic collapse at the cell
(Fig. 3c). For comparison, the σ–ε curves of solid samples wall corners which creates plastic hinges due to the plastic
(100% of the filler content) and shell samples (0% of the filler moment reached at these corners [30]. A cross-sectional
content) are also plotted. picture of Hc-3-40% shows a radial crack which nucleated
Figure 3a shows that Hc-3-20% has a bilinear σ–ε behav- through three cell walls because the thin wall did not bear
ior. The first linear trend shows a 623 MPa slope from the bending moments at corners. Plastic deformation seems
0 ≤ ε ≤ 4.4% until the first peak stress of 18.8 MPa is reached. to be the dominant failure mechanism for specimens with
Beyond this point, the structure is able to bear more compres- high filling content (Ra-3-100% and Hc-3-40%) while lo-
sive load; then, a second linear σ–ε response is observed until calized buckling is for specimens with low filling content
it reaches a second peak stress of 21 MPa at ε = 20.5%. Hc-3- (N-3-0% and Hc-3-20%). Local buckling arises on N-3-0%
30% and Hc-3-40% specimens exhibit similar σ–ε trends to because it is a thin shell structure with a high aspect ratio
that of Hc-3-20% with the exception that σu increases as the (thickness/length ratio), and hence, it is prone to bending.
filler content increases. Hence, as the in-plane size of honey- Pictures of N-3-0% indicate that its outer surface is bulged
comb cells decreases (higher density), the elastic properties of near the top end as a consequence of local buckling.
Int J Adv Manuf Technol (2018) 95:1685–1695 1689

Fig. 3 Stress–strain curves and


photographs of the representative
Hc samples. a Hc-3, b Hc-1,
and c Hc-2 structures with
different filling percentages.
Scale bars in millimeters

In general, Hc-1 specimens exhibit a small linear elastic that of Hc-1-20%; however, σu and E are enhanced for denser
region until they reach their peak stresses; then, their load structures. Pictures of Hc-1-30% and Hc-1-40% reveal a shear
bearing capability slowly decreases with a slightly jagged failure mechanism at large deformations, occurring at − 45°
σ–ε trend up to the final collapse (see Fig. 3b). Hc-1-20% (see pictures on Fig. 3b), which suggests that its failure mode
shows a small linear elastic σ–ε trend up to ε = 4.4% when is a combination of shear and local buckling. Their cross-sec-
the first peak stress is reached at 12.6 MPa. Further compres- tions show sliding between neighbor cell walls, which also
sive load yields a slightly jagged σ–ε response until the struc- indicates contribution of shearing. White marks on the sur-
ture collapses. For Hc-1-20%, failure is attributed to the local- faces of Hc-1 specimens are only appreciated locally, where
ized buckling of honeycomb walls (inside the structure) ori- shear or buckling failure occurred. N-12-0% shows a small
ented parallel to the compressive loading (see its cross- linear elastic σ–ε behavior followed by a sudden collapse after
sectional picture). Shear and compressive loads also trigger reaching the peak stress. For this specimen, local buckling is
the propagation of cracks along the longitudinal printed direc- identified as its failure mechanism (see pictures in Fig. 3b).
tion. Hc-1-30% and Hc-1-40% exhibit similar σ–ε curves than The local buckling was so severe that a flaw propagates
1690 Int J Adv Manuf Technol (2018) 95:1685–1695

catastrophically along the printing lines (3-direction), typical and the propagation of cracks along the longitudinal axis.
of a brittle fracture. Ra-12-100% displays a stiff bilinear σ–ε Rb-12 specimens show σ–ε curves (Fig. 4c) similar to
response with much higher strength and stiffness than those of those from Ra-12 samples (Fig. 4b) with the exception that
N-12-0% and Hc-1 specimens. As observed in the pictures, Ra-12 has a sharper decrease in the load bearing capacity
compressive plastic deformation is the failure mechanism of once σu is reached. Regarding the filling content, Rb-12-
Ra-12-100%. 40% shows the effect of densification on the smooth linear
Hc-2 specimens (Fig. 3c) show a linear elastic behavior for decaying trend of the σ–ε response beyond the maximum
low strains (ε ≤ 3.0%) similar to that of Hc-1, although Hc-2 stress, which contrasts with the sharper stress drop for Rb-
structures retain higher load bearing capacity once the peak 12-20% and Rb-12-30% at the same strain levels. Local
stress is reached than Hc-1 ones, as seen in Fig. 3b. Hc-2-20% buckling is also identified as the failure mechanism for
and Hc-2-30% exhibit a wavy decreasing σ–ε behavior once Rb-12 structures according to the photographs of Fig. 4c,
σu is reached until failure. In contrast, the load bearing capac- which induced wrinkles and debonding between cells and
ity of Hc-2-40% decreases slightly once σu is reached and then the outer printed surface.
continues increasing up to the final collapse. At the outer In order to directly compare the overall σ–ε behavior
surface, wrinkles and longitudinal cracks are observed in the of all types of structures, Fig. 5 shows specific stress σ*
photographs of Fig. 3c for all Hc-2 structures due to local (kPa*m3/kg) versus ε of all printed specimens with a relative
buckling failure. Cross-sectional photographs indicate that density of 20%. Stress values in kPa for each sample type can
cell walls reached the yield point of the ABS (see the white be obtained by multiplying σ* with the corresponding density
marks within the cells) and local buckling occurred longitudi- listed in Table 1. These parameters (σ* and 20% of relative
nally (it is not appreciated at the in-plane but observed in the density) were chosen in order to disregard the effect of small
outer surface). For Hc-2 samples, the local buckling was not differences on weight between the printed structures (see
severe, allowing the structure to keep pace of the applied load Table 1). As seen in Fig. 5, Hc-3-20% exhibits the highest
due to densification and compression beyond σu. σ*u and E* because its hexagonal extruded pattern is aligned
The representative σ–ε curves of Ra specimens are shown with respect to the applied load (3-direction). The other struc-
in Fig. 4. Even though baseline samples (viz., N-12-0%, N-3- tures display similar peak σ* values among them (they vary
0%, Ra-12-100%, and Ra-3-100%) are discussed in Fig. 3, only between 20 and 25 kPa*m3/kg) although their E* has
they are plotted again in Fig. 4 for comparative purposes. numerical differences (see Fig. 6 for details). It is also ob-
According to Fig. 4a, Ra-3 structures display a proportional served that Hc-1-20%, Hc-2-20%, Ra-12-20%, and Rb-12-
linear elastic σ–ε response for small deformations (ε ≤ 4.41%) 20% are not capable of carrying compressive loads beyond
until the first peak stress is reached. A second linear trend in their σ*u because of their jagged-like σ*–ε behavior. Therefore,
the σ–ε response beyond a yielding region is identified, which Hc-3 and Ra-3 are recommended for compression under large
increases monotonically until it exhibits a second peak stress, deformation regimes due to their linear strain hardening σ*–ε
then σ decreases up to the final collapse of the specimen. For response.
these specimens, σu (first peak stress) improves as the struc-
ture density increases. Photographs of failed Ra specimens
(right-hand side in Fig. 4a) point it out that a combination of 3.2 Specific stiffness and strength
compressive plastic deformation and local buckling are the
dominant failure mechanisms. Cross-sectional photographs In order to quantify the structural effectiveness of the ABS
indicate that local buckling promotes the bending of cell walls specimens (i.e., structural performance considering their
at its corners and, as the cell size becomes smaller, this behav- weigh), specific stiffness and strength are discussed in this
ior decreases and yielding becomes the dominant failure section. To obtain stiffness and strength values in MPa or
mechanism. kPa of a particular structure, E* or σ*u should be multiplied
In Fig. 4b, it can be seen that the σ–ε curves of Ra-12 by the corresponding density listed in Table 1. Figure 6 com-
specimens show a small linear elastic behavior followed by pares the σ*u values for Hc, Ra, and Rb filling patterns at
a steep-down σ–ε response after σu. It is also observed that the different contents (0, 20, 30, 40, and 100%). The histogram
stiffness and strength enhance with respect to density. The shown in Fig. 6 is plotted with three x-axis levels, being the
sudden σ–ε drop is due to the collapse of the inner structure first one (bottom to top) to group filling patterns while the
(filler shape) due to local buckling, as evidenced in pictures of second and third ones for printing orientation and filling con-
Fig. 4b. At the cross-sections, it can be observed that cells tent, respectively. Regardless of the filler shape, all specimens
bend along the longitudinal direction and neighbor cells push loaded along the 3-orientation show the highest strength with
each other as the bending increases, giving rise to the radial respect to those loaded along their 1- or 2-orientation; this
patterns shown. The local buckling is also responsible for response is a consequence of the walls of the printed filler
the σ–ε fluctuating behavior on large deformation regimes shapes being aligned with respect to the compressive load.
Int J Adv Manuf Technol (2018) 95:1685–1695 1691

Fig. 4 Representative stress–


strain curves and photographs of
fractured Ra (a Ra-3, b Ra-12)
and Rb (c Rb-12) specimens.
Scale bars in millimeters

As expected, solid structures display the highest strength filling pattern, it can be seen that filling content does not affect
while shell geometries (0% filling content) present the lowest E*. For shell specimens (N-12-0% and N-3-0%), 3-direction
strength. Small differences in σ*u are found among specimens shows higher stiffness than 1- or 2-direction, whereas for solid
loaded in their 1- or 2-direction, independent of the cell pattern specimens, it is the opposite way; i.e., 1- and/or 2-orientation
employed. The highest σ*u was exhibited by Hc-3 specimens. yields higher E*. This happens to shell specimens because the
Thus, these structures are recommended for low-weight appli- applied load is carried by the thin wall and the stiffness only
cations while solid structures (Ra-3-100% and Ra-12-100%) depends on the structural stability among raster lines of the wall.
are recommended for compressive applications where weight On the other hand, for solid specimens, their dense filling patterns
is not crucial. (it yields smaller voids) promote a very stable structure, which
Figure 7 compares E* for all ABS specimens tested. Shell are unlikely to bend or buckle. For Ra specimens, higher E* is
and solid structures display dissimilar E* values with respect to achieved when specimens are printed along the 1 (1 = 2)-direc-
their orientation. By grouping the specimens based on their tion, while for Hc configurations, higher E* values are achieved
1692 Int J Adv Manuf Technol (2018) 95:1685–1695

45
the required printing time increases with respect to filler con-
Hc-3-20%
40 tent, from 6.1–7.6 min for hollow structures (0% filler) to
35 Ra-3-20%
17.1–17.6 min for solid ones (100% filler), although Hc-3
reached 19.3 min at 40%. In general, honeycomb patterns
30
(kPa*m3/kg)

require more printing time than rectangular patterns. By com-


Hc-2-20% paring Ra-3 and Hc-3, it is observed that Hc-3 specimens
25
*

require 32, 58, and 82% more printing time than Ra-3 ones
20
for 20, 30, and 40% of the filler content, respectively. These
15 differences may be due to the honeycomb structure taking
10 extra steps in the printing process as a consequence of its
Hc-1-20% hexagonal array. For most cases, regardless of filler percent-
Ra-12-20%
5 Rb-12-20% age, printing time increases according to the following orien-
0 tation sequence: Ra-3, Rb-12, Ra-12, Hc-2, Hc-1, and Hc-3;
0 5 10 15 20 25 30 35 40
i.e., Ra-3 specimens take the lowest printing time whereas Hc-
(%)
3 specimens the highest (Fig. 8). Printing orientation has a
Fig. 5 Representative specific stress–strain curves for Hc-1, Hc-2, Hc-3, small influence on the printing time of rectangular patterns,
Ra-12, Ra-3, and Rb-12 specimens with 20% filling content
whereas for honeycomb patterns, printing time is practically
the same regardless the orientation. For rectangular patterns,
in the 3-direction. If solid specimens are excluded, Hc-3 struc- the 3-orientation yields the shortest printing time whereas the
tures present the highest E* values. As suggested in Fig. 7, ap- 1–2-orientations yield the longest. Results of Fig. 8 evidence
plications that demand stiff structures should employ the Hc-3 that for rectangular patterns (Ra-12, Ra-3, and Rb-12), print-
configuration as a first option followed by Ra-12 as a second ing time slightly increases with respect to the filler percentage.
choice. No significant differences on the E* values of Ra-3-, Rb- For instance, the required printing time for Ra-3 slightly in-
12, Hc-1, and Hc-2 specimens were observed. creases from 9.0 min at 20% to 10.6 min at 40%. On the other
hand, honeycomb samples show an important variation of
3.3 Mechanical properties as functions of the printing time printing time with respect to filler percentage. For example,
time for Hc-1 increases from 11.9 to 18.6 min for an increase
Since printing time represents one of the dominant factors on of filler content from 20 to 40%. Therefore, dense structures
the production cost of 3D printed structures, comparative can be achieved by using rectangular patterns in a short period
analyses of the cost associated to printing time and cost-to- of time.
benefit ratio regarding the strength and stiffness as a function Figure 9 compares the E values of structures with 20, 30,
of the printing time were conducted. Figure 8 shows the and 40% of the filler content as a function of their printing
time required to print a single specimen for all printing con- time. Data for structures with 0 and 100% of filler content
figurations; here, the structures are sorted from short to long are not plotted in Figs. 9 and 10 in order to make systematic
printing time for each filler percentage. It is observed that comparisons between honeycomb and rectangular patterns.

Fig. 6 Comparison of specific 60


compressive strength (σ*u ) for
48
ABS printed specimens 50
46 45 44
41 41
40 38
(kPa·m3/kg)

34 32
32
30 28 28
25 26 24
22 24 24 24 24
22 23
u
*

20

10

0
Filler % 0 20 30 40 100 20 30 40 100 20 30 40 20 30 40 20 30 40 20 30 40
Orientation 1=2 3 1=2 3 1=2 1 2 3
Pattern N Ra Rb Hc
Int J Adv Manuf Technol (2018) 95:1685–1695 1693

Fig. 7 Comparison of specific 1.6


stiffness for all ABS printed

9
1.2
1.4

6
structures

2
1.2

1.2
8
6
1.1
1.1
4
1.1
1.2

4
76

2
1.0
77
1.0
0.9
E* (MPa· m3/kg)

0.9

75
1.0

54

42
0.8
27
15

3
0.8

0.8

56
77

53
0.8

52

0.8
0.8

34

4
0.7
0.7

0.7
0.7

0.7
0.7
0.8

0.6

0.4

0.2

0.0
Filler % 0 20 30 40 100 20 30 40 100 20 30 40 20 30 40 20 30 40 20 30 40
Orientation 1=2 3 1=2 3 1=2 1 2 3
Pattern N Ra Rb Hc

As discussed in Section 3.1, stiffness of the specimens in- the other hand, E varies linearly with respect to printing
creases as a function of filling content; however, this also time. However, Ra-12 specimens are stiffer than Rb-12 ones
implies that stiffer structures require longer printing time as but shorter printing times are required. Honeycomb struc-
suggested in Fig. 9. In general, honeycomb structures are tures present a linear behavior of E with respect to their
stiffer than the rectangular ones but their required printing printing time. Hc-3 raises from 626 to 908 MPa for an
time is much longer. Thus, honeycomb structures are more increase of printing time from 11.9 to 19.3 min; i.e., impor-
mass-efficient but their production is less time-efficient. tant enhancements in stiffness are obtained but large print-
Both rectangular and honeycomb patterns show that the 3- ing times are needed. On the other hand, Young’s modulus
orientation yields stiffer structures than the 1–2-orientations of Hc-1 and Hc-2 samples are very similar at 11.9 and
for similar printing times. For example, Hc-3-20% has 15.3 min; a difference of 86 MPa is found between Hc-1
626 MPa whereas Hc-1-20% and Hc-2-20% have 426 and and Hc-2 at 18.6 min.
405 MPa, respectively, for the same 11.9 min period. Ra-3 The compressive strength of specimens with 20, 30,
structures show a nonlinear increase in E with respect to and 40% of the filler content was also plotted against
printing time, raising from 445 to 652 MPa while printing printing time per sample, as it is shown in Fig. 10. In
time only increases from 9 to 10.6 min; i.e., they become this figure, it is evident that σu has a similar trend to that
stiffer in short periods of printing. For Ra-12 and Rb-12, on of E with respect to printing time (Fig. 9). Structures with

Fig. 8 Printing time per sample


of all tested structures
1694 Int J Adv Manuf Technol (2018) 95:1685–1695

Ra-3 40 %
900
Ra-12
printing time, and filler shape on their stiffness, strength, and
Rb-12 failure mechanisms. Analysis of the stress–strain (σ–ε) curves
800 Hc-3 30 %
Hc-1 suggests that stiffness and ultimate strength are rendered
Hc-2
700
40 %
higher accordingly by increasing the filling content.
20 %
E (MPa)

Specimens with patterns printed along the longitudinal direc-


600 30 %
40 % tion show a bilinear σ–ε response whereas those transversely
40 %
500
20 % 30 % printed show a short linear elastic response for small deforma-
30 %
tions and a nonlinear (jagged-like) behavior beyond the peak
400
20 %
20 % stress. Specimens printed along the longitudinal direction (3-
300 orientation) are stiffer and stronger than those transversely
7.5 10.0 12.5 15.0 17.5 20.0
Printing time per specimen (min)
printed (1–2-orientations). Specimens filled with honeycomb
patterns exhibit better elastic properties compared to those
Fig. 9 Compressive Young’s modulus as a function of the printing time
per sample of representative samples with a filler percentage of 20, 30, filled with rectangular patterns when they are printed in the
and 40% longitudinal direction, while no significant differences in their
elastic properties are observed in the transversal direction.
Local buckling failure occurred in lightweight (low filling
content) structures, promoting crack propagation along the
3-orientation bear higher compressive loads than those longitudinal direction. Compressive failure was identified for
with 1- and 2-orientations for similar printing times. relatively dense structures (30, 40, and 100%) and those with
Honeycomb structures carry more compressive loads than longitudinal extruded patterns. A combination of shear and
rectangular ones although their printing time is longer. local buckling failure appeared in honeycomb structures trans-
Structures with patterns along the 3-direction show a non- versely printed with relative densities around 20–40%.
linear increase of σu with respect to printing time, but Structures with rectangular patterns employ much shorter
patterns along the 1- and 2-directions present a linear printing time than structures with honeycomb cells for the
response. According to results in Figs. 8, 9, and 10, it same filling content. Therefore, the compressive stiffness
is suggested to employ Ra-3 structures in large-scale pro- and strength normalized with the printing time of structures
duction systems because they can reach high E and σu with rectangular cells were higher than those from honeycomb
values at relatively low printing time. On the other hand, patterns, indicating that rectangular filler shape is more suit-
Hc-3 structures are recommended for the production of able for fast mass production systems. Results suggest that 3D
stronger and stiffer parts despite of the longer printing printed structures filled with honeycomb patterns in the lon-
time required. gitudinal direction have the most effective strength/mass ratio,
and thus, they are recommended for low-weight applications.
Solid structures (rectangular or honeycomb with 100% of rel-
4 Conclusions ative density) exhibit the highest strength being recommended
for applications where weight is not limited.
3D printed ABS structures were tested under compressive
loads in order to analyze the role of printing direction, density, Acknowledgments The authors acknowledge Diego Gomez Marquez
and Miguel Nuñez Cardenas (Universidad de Guanajuato) for the com-
pression tests.
35
40 %

30 % Funding information The partial support from the “Laboratorio


30
Nacional en Innovación y Desarrollo de Materiales Ligeros para la
40 %
Industria Automotriz (LANI-Auto)” through CONACYT grant no.
25 280425 is greatly appreciated. A. Hernández-Pérez acknowledges the
(MPa)

30 % 20 %
PRODEP program (UGTO-PTC-539) for the economic support.
20
40 % 40 %
u

20 % References
15 30 % Ra-3
30 %
Ra-12
20 % Rb-12
10 Hc-3 1. Lin AC, Liang S-R (2002) Rapid prototyping through scanned
20 %
Hc-1 point data. Int J Prod Res 40(2):293–310
Hc-2
5 2. Kamrani AK, Nasr EA (2009) Rapid prototyping, engineering de-
8 10 12 14 16 18 20 sign and rapid prototyping. Springer, Boston, pp 339–354
Printing time per specimen (min)
3. Conner BP, Manogharan GP, Martof AN, Rodomsky LM,
Fig. 10 Compressive strength as a function of the printing time per Rodomsky CM, Jordan DC, Limperos JW (2014) Making sense
sample of representative samples with a filler percentage of 20, 30, and of 3-D printing: creating a map of additive manufacturing products
40% and services. Additive Manufacturing:1, 64–4, 76
Int J Adv Manuf Technol (2018) 95:1685–1695 1695

4. Es-Said OS, Foyos J, Noorani R, Mendelson M, Marloth R, 18. Lee CS, Kim SG, Kim HJ, Ahn SH (2007) Measurement of aniso-
Pregger BA (2000) Effect of layer orientation on mechanical prop- tropic compressive strength of rapid prototyping parts. J Mater
erties of rapid prototyped samples. Mater Manuf Process 15(1): Process Technol:187, 627–188, 630
107–122 19. Domingo-Espin M, Puigoriol-Forcada JM, Garcia-Granada A-A,
5. Um D (2016) Rapid prototyping, solid modeling and applications. Llumà J, Borros S, Reyes G (2015) Mechanical property character-
Springer, Cham, pp 191–221 ization and simulation of fused deposition modeling polycarbonate
6. Savastano M, Amendola C, D’Ascenzo F, Massaroni E (2016) 3-D parts. Mater Des 83:670–677
printing in the spare parts supply chain: an explorative study in the 20. Croccolo D, De Agostinis M, Olmi G (2013) Experimental charac-
automotive industry. In: Caporarello L, Cesaroni F, Giesecke R, terization and analytical modelling of the mechanical behaviour of
Missikoff M (eds) Digitally supported innovation. Springer, fused deposition processed parts made of ABS-M30. Comput
Cham, pp 153–170 Mater Sci 79:506–518
7. Bose S, Vahabzadeh S, Bandyopadhyay A (2013) Bone tissue en- 21. Song Y, Li Y, Song W, Yee K, Lee KY, Tagarielli VL (2017)
gineering using 3D printing. Mater Today 16(12):496–504 Measurements of the mechanical response of unidirectional 3D-
8. Chya-Yan L, Murat G (2017) Current and emerging applications of printed PLA. Mater Des 123:154–164
3D printing in medicine. Biofabrication 9(2): 024102 22. Yang C, Tian X, Li D, Cao Y, Zhao F, Shi C (2017) Influence of
9. Ambrosi A, Pumera M (2016) 3D-printing technologies for elec- thermal processing conditions in 3D printing on the crystallinity
trochemical applications. Chem Soc Rev 45(10):2740–2755 and mechanical properties of PEEK material. J Mater Process
10. Espalin D, Muse DW, MacDonald E, Wicker RB (2014) 3D print- Technol 248:1–7
ing multifunctionality: structures with electronics. Int J Adv Manuf 23. D.P. Cole, J.C. Riddick, H.M. Iftekhar Jaim, K.E. Strawhecker,
Technol 72(5–8):963–978 N.E. Zander, Interfacial mechanical behavior of 3D printed ABS,
11. Kumbay Yildiz S, Mutlu R, Alici G (2016) Fabrication and charac- Journal of Applied Polymer Science 133(30) (2016) n/a-n/a
terisation of highly stretchable elastomeric strain sensors for pros-
24. Zhang H, Cai L, Golub M, Zhang Y, Yang X, Schlarman K, Zhang J
thetic hand applications. Sensors Actuators A Phys 247:514–521
(2017) Tensile, creep, and fatigue behaviors of 3D-printed acrylo-
12. Muth JT, Vogt DM, Truby RL, Mengüç Y, Kolesky DB, Wood RJ,
nitrile butadiene styrene. J Mater Eng Perform
Lewis JA (2014) Embedded 3D printing of strain sensors within
25. Pilipović A, Raos P, Šercer M (2009) Experimental analysis of
highly stretchable elastomers. Adv Mater 26(36):6307–6312
properties of materials for rapid prototyping. Int J Adv Manuf
13. Gross BC, Erkal JL, Lockwood SY, Chen C, Spence DM (2014)
Evaluation of 3D printing and its potential impact on biotechnology Technol 40(1):105–115
and the chemical sciences. Anal Chem 86(7):3240–3253 26. Mueller J, Shea K, Daraio C (2015) Mechanical properties of parts
14. Lee BH, Abdullah J, Khan ZA (2005) Optimization of rapid fabricated with inkjet 3D printing through efficient experimental
prototyping parameters for production of flexible ABS object. J design. Mater Des 86:902–912
Mater Process Technol 169(1):54–61 27. Salinas R (2014) 3D printing with RepRap cookbook. Packt,
15. Casavola C, Cazzato A, Moramarco V, Pappalettere C (2016) Birmingham
Orthotropic mechanical properties of fused deposition modelling 28. ASTM D2344, standard test method for short-beam strength of
parts described by classical laminate theory. Mater Des 90:453–458 polymer matrix composite materials and their laminates, ASTM
16. Dawoud M, Taha I, Ebeid SJ (2016) Mechanical behaviour of ABS: International, West Conshohocken, 2015
an experimental study using FDM and injection moulding tech- 29. Jin T, Zhou Z, Wang Z, Wu G, Shu X (2015) Experimental study on
niques. J Manuf Process 21:39–45 the effects of specimen in-plane size on the mechanical behavior of
17. Sood AK, Ohdar RK, Mahapatra SS (2010) Parametric appraisal of aluminum hexagonal honeycombs. Mater Sci Eng A 635:23–35
mechanical property of fused deposition modelling processed parts. 30. Gibson LJ, Ashby MF (1999) Cellular solids: structure and proper-
Mater Des 31(1):287–295 ties, 2nd edn. Cambridge University Press, Cambridge

Вам также может понравиться