You are on page 1of 10

Optics & Laser Technology 89 (2017) 3140

Contents lists available at ScienceDirect

Optics & Laser Technology


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

A Round Robin study for Selective Laser Sintering of polyamide 12:


Microstructural origin of the mechanical properties

crossmark

Thomas Stichela,b, , Thomas Fricka, Tobias Laumera,b, Felix Tennerc, Tino Hausotteb,d,
Marion Merkleinb,e, Michael Schmidta,b,c
a

Bayerisches Laserzentrum GmbH, 91052 Erlangen, Germany


CRC Collaborative Research Center 814 Additive Manufacturing, Germany
c
Friedrich-Alexander-University Erlangen-Nrnberg (FAU), Institute of Photonic Technologies, 91052 Erlangen, Germany
d
Friedrich-Alexander-University Erlangen-Nrnberg (FAU), Institute of Manufacturing Metrology, 91052 Erlangen, Germany
e
Friedrich-Alexander-University Erlangen-Nrnberg (FAU), Institute of Manufacturing Technology, 91058 Erlangen, Germany
b

A R T I C L E I N F O

A BS T RAC T

Keywords:
Additive manufacturing
Polymers
Polyamide-12
Selective Laser Sintering
Round Robin

The mechanical and microstructural investigation of polymer parts (polyamide 12) fabricated by Selective Laser
Sintering as part of a Round Robin initiative is presented. The paper focuses on the microstructural analysis of
the Round Robin samples and their evaluation regarding their eect on mechanical properties with respect to
each other. Therefore optical microscopy on microtomed samples, X-ray computed tomography and Dierential
Scanning Calorimetry is used to determine the morphology of residual particle cores and of internal pores.
The mechanical tensile testing revealed a high variability of the ductility of the samples among the used
machines and a distinctive anisotropic mechanical response. Especially the quite brittle characteristic along the
building direction has shown to be still a crucial challenge for the process. However, one machine delivered
samples with outstanding ductility with total elongation values of about 21% along the building direction and of
about 32% planar to the layer. This result was back traced to a distinctive pore and residual particle morphology
which is characterized by low pore concentration, the absence of coplanar pore or residual particle
arrangements and the highest degree of particle melting measured. Furthermore, the analysis depicts that
both features, pores and residual particles, contribute to the mechanical properties signicantly and that they
are not necessarily linked since they can vary independently in a certain range depending on the machine
conguration.

1. Introduction
Selective Laser Sintering (SLS) is one of the most widespread
Additive Manufacturing processes that emerged in the last decades
[1,2]. It allows the fabrication of three-dimensional polymer components with complex shapes according to CAD data directly out of
powdery material without the need of tooling and setup [3]. Therefore
parts are built up layer by layer. Three steps are repeated for each layer:
First, the platform is lowered by the thickness of one layer. Second,
powder is deposited on the building platform using a moving roller or
blade device and preheated to a temperature close to the material
melting point. And third, a laser beam scans the powder layer with
arbitrary trajectories in order to melt specic powder layer areas. These
steps are applied alternately until the part is completed. Before the
parts can be extracted from the machine, homogeneous cooling has to
take place whereas the viscose melted polymer crystallizes with

minimal distortions.
The main advantage of SLS over other Rapid Prototyping or
Additive Manufacturing technologies using plastics is the ability to
produce parts which feature material properties that are very close to
the material properties obtained by injection molding [4]. However,
this is only achieved when a robust process and material routine is
established leading to high quality and repeatable results. In many
cases, SLS machine users observe minor reproducibility regarding to
mechanical and geometrical properties [5], limiting the industrial
application and cost-eciency of SLS.
Especially the mechanical properties along the build direction have
shown to be a challenge for the process. Ductility and strength are
generally lower along the build direction than perpendicular to it and
they are also very sensitive to varying process conditions and parameters. This can be explained by the stacking-layer-process which
leads easily to insucient interlayer connection, e. g., if energy input or

Corresponding author at: Bayerisches Laserzentrum GmbH, 91052 Erlangen, Germany.


E-mail address: t.stichel@blz.org (T. Stichel).

http://dx.doi.org/10.1016/j.optlastec.2016.09.042
Received 12 July 2016; Received in revised form 23 August 2016; Accepted 26 September 2016
0030-3992/ 2016 Elsevier Ltd. All rights reserved.

Optics & Laser Technology 89 (2017) 3140

T. Stichel et al.

the local process temperatures are too low for the complete melting of
the layer section. Insucient interlayer connection is related to
distinctive microstructural features such as pore formation or residual
particles (or cores) [6,7] which can occur with varying frequency and
morphology (size, shape, orientation).
The occurrence of dierent morphologies of porosity inside polymer parts built by SLS is well known [68]. While it is aected by many
material and process parameters [8], an evidentiary understanding of
its morphology and impact on the mechanical properties is still
missing. Very often the interlayer porosity is blamed for inferior
mechanical properties along the building direction [6]. Besides pores,
residual particle and particle core arrangements due to partial melting
were also identied to aect the mechanical response of SLS processed
samples. It was stated that an increase of the degree of particle melt
(DPM) leads to an improvement of the mechanical properties and
coplanar aligned residual particles and cores promote mechanical
weakness along the building direction [6,9]. This is explained by the
many short binding necks creating high stiness which result in short
elongation [10].
Nevertheless, a quantication of the inuence of both eects
(porosity and residual particles) on the mechanical properties in
comparison has not been reached yet. Typically applied research
methodology base on the variation of few parameters which leads to
complementary changes in both features and thus do not allow the
estimation of their respective signicance for the mechanical properties. This can be exemplarily seen in [8] where the variation of the laser
beam energy leads to congruent change of the crystal weight fraction
and the porosity with strong impact on the elongation at break but with
no clue of which eect is more important. Thus experimental methodologies should be applied which include the broad variation of
parameters and dierent machines resulting in more versatile microstructural morphologies.
In order to measure the inuence of pore formation and residual
particle (cores) on mechanical properties with respect to each other, a
comprehensive microstructural analysis is performed upon tensile test
samples built during a Round Robin initiative based on six dierent
machines and parameter sets. The internal pore morphology of the
samples is measured by X-ray computed tomography while the residual
particles (and particle cores) are investigated using microscopy on thin
sections and Dierential Scanning Calorimetry (DSC). The results are
used to trace back conspicuous features of the tensile testing outcome
to the source in order to value the signicance of pore and residual
particle morphology.

Fig. 1. CAD design of a testing sample set consisting of three dierently orientated
tensile bars: Horizontal orientation corresponds to in-plane direction, vertical orientation to build direction.

medium.
3. Characterization methods
In Fig. 3 an overview of the performed characterization methods is
given. The mechanical properties which resemble the functionality of
the samples are determined via tensile testing according to ISO 3167.
Microstructural features which are represented by the residual particle
(core) arrangements and the porosity can help to explain the appearance of functionality (mechanical) issues. While arrangement of
residual particles and cores are veried by optical analysis of thin
sections, the general amount of residual particle is resembled by the
degree of particle melt which is determined using Dierential Scanning
Calorimetry. In order to analysis the pore morphology of the samples
comprehensively, X-ray computed tomography is used which delivers
information about porosity, pore density, pore arrangement, pore
orientation and pore size distribution. All respective determination
methods are commented in detail in the following sections.

2. Round Robin methodology and tensile test samples

3.1. Mechanical testing

The Round Robin includes six dierent SLS machines and users
(M1-M6) which produced a range of tensile samples (ISO 3167, Type
A) with dierent build orientations (see Fig. 1). For each machine
dierent parameter sets (laser power, scanning speed or layer thickness, etc.) were used, which were specied by the applicants to be the
optimum for the respective machine. The samples were made from the
standard material polyamide 12 Duraform PA (3D Systems) or PA2200
(EOS) which base on the same basic powder VESTOSINT produced by
Evonik. An overview of the production parameters are displayed in
Table 1.
Six sample sets were produced in machine M1, M3, M4 and M6,
three sets in machine M2 and one set in machine M5. The nominal and
measured linear dimensions of the samples as well as the classication
details according to ISO 2768-1 are displayed in Fig. 2. The samples of
M6 could not be measured due to strong deformations and displacements and are beyond of any fabrication tolerances. The samples
fabricated with machine M5 cannot be classied by the norm since the
sample length is exceeding the dened deviation of 2.5 mm for the
very coarse category by far. Instead, the samples of the machines M1,
M3 and M4 can be categorized as coarse, the samples of M2 as

Tensile tests are carried out using the Zwick/Roell 100 AllroundLine machine with a test speed of 5 mm/min. The values that are
evaluated and compared are the ultimate tensile strength (UTS) and
the total elongation (TE) which is also known as elongation at break.
The yield point is dened for plastics as the rst point on the stressstrain curve at which an increase in strain occurs without an increase in
stress [11].
3.2. Thin section analysis
Optical analysis was performed on thin sections with a thickness of
50 m prepared by a microtome. Polarized light was used in order to
distinguish between non-molten, partly molten and fully molten
particles in the sample inside. The preparation of the section was
realized in that way that they display the layer stack.
Already mechanically tested tensile test samples were picked and
prepared by the microtome. The thin sections were extracted from the
furthest clamping areas of the tensile bars where no plastic deformation aecting the samples microstructure is introduced upon tensile
and clamping force. This was veried by controlling the geometry of the
32

Optics & Laser Technology 89 (2017) 3140

T. Stichel et al.

Table 1
Parameter protocol for the different machines M1M6.
Machine

M1

M2

M3

M4

M5

M6

Machine manufacturer
Sample count (H, D, V)
Laser power (W)
Beam diameter (FWHM) (m)
Hatch distance (m)
Hatch strategy
Layer thickness (m)
Scanning speed (mm/s)
Recoating speed (mm/s)
Powder material
Powder mixing rate (used-new)
Processing temperature (C)
Total build time (hh: mm)
Cooling time in machine (h)
Finishing routine

3D Systems
6, 6, 6
42
464
279.4
X and Y alternately
101.6
10000
254
Duraform PA
5050
182
12:35
1+4 (at room temp.)
bead blast

EOS GmbH
3, 3, 3
21
430
250
X and Y alternately
100
2500
n. a.
PA 2200
5050
173
12:30
~10
bead blast+water wash

EOS GmbH
6, 6, 6
n. a.
600
n. a.
n. a.
120
n. a.
120
PA 2200
5050
160
10:20; 27:37
10
sand blast+water wash

EOS GmbH
6, 6, 6
38
500
310
X and Y alternately
120
6000
200
Duraform PA
5050
165.5
16:31
48
sand blast

DTM Corporation
1, 1, 1
7
400
150
X and Y alternately
100
1256
120
PA 2200
5050
172
n. a.
n. a.
sand blast

EOS GmbH
6, 6, 6
30
600
600
X and Y alternately
150
300
50
PA 2200
5050
178
n. a.
2
compressed air

clamping area before and after the tensile testing, whereas no


permanent inuence could be detected.
3.3. Dierential Scanning Calorimetry
DSC measurements were performed using TA Instruments equipment and software. Sample masses were 10 0.5 mg and samples were
heated from ~20 C to 200 C at a rate of 10 C per minute. Samples
were held at 200 C for two minutes, and then cooled at 10 C per
minute. Then a second heating/cooling cycle was undertaken subsequently. For each charge a powder probe as well as a part probe was
measured. An example measurement is shown in Fig. 4. The evaluation
was carried out according to Majewski et al. [9]. The percentage
crystallinity is dened as the relation between the heat of melting
and the known heat of melting for a 100% crystalline specimen. The
latter value was taken to be 209.3 J/g, as reported by Gogolewski et al.
[12]. The degree of particle melt (DPM) resembles the relation between
both melted and crystallized regions, and non-melted particle cores
within the part [9]. It is calculated with formula [9]:

Fig. 3. Overview over the characterization methods used.

.
While the total crystallinity of the sample can be extracted from the
rst heating cycle of the DSC measurement of the sample, the crystallinity of melted and recrystallized material is derived from the second
heating cycle of the same sample. Instead the core (or powder)
crystallinity is determined by means of the DSC measurement using
pure powder probes of the powders used with the respective machines.

DPM (%)
Total crystallinity of sample Core (or powder) crystallinity
=
Crystallinity of melted and crystallized material Core crystallinity
(1)

Fig. 2. Boxplot of the sample thickness (a), width (b) and length (c) measured using tactile method and compared to nominal value (dotted line) and general tolerance values according
to ISO 2768-1. The measured samples were built with dierent orientation (D: diagonal; H: horizontal; V: vertical) with dierent machines (M1-M6) and parameters.

33

Optics & Laser Technology 89 (2017) 3140

T. Stichel et al.

4. Results and interpretation


At rst, the tensile testing results of the Round Robin samples will
be presented with respect to the state of the art, before the analysis of
the microstructure reveals the origin of conspicuous features. This
analysis includes the discussion of the residual particle (cores) and the
porosity regarding their inuence on the mechanical properties of SLS
built samples with respect to each other.
The following measurements of the mechanical properties, the
porosity, the pore density, the degree of particle melt and the crystallinity are resembled by boxplots which give the interquartile range
(IQR) of the respective measured values. The top line, the middle line
and the bottom line indicate the value that 75%, 50% (median) or 25%
of data are equal or less than this value, respectively. The upper and
lower spread bars (whiskers) show the lowest value still within 1.5 IQR
of the lower quartile, and the highest value still within 1.5 IQR of the
upper quartile.

Fig. 4. Thermogram of a DSC measurement of a sintered sample built with machine M3.

3.4. X-ray computed tomography

4.1. Tensile testing results

The porosity, shapes, dimensions and positions of pores were


investigated by X-ray computed tomography [13]. The samples were
scanned with the Werth TomoCheck 200 3D system with a CT sensor.
The detector possesses 10241024 pixels at a pixel size of 50 m. With
a geometrical magnication of ve voxel sizes of 10 m were achieved.
The measured X-ray data set was analyzed using a defect detection
function modeling pores with a minimum pore size of 8 voxels (see
Fig. 5). These detected defection volumes were used as basis for the
analysis of the sample porosity which is resembled by the relation
between defect volume and overall volume of the measured sample.
Since the determination of overall volume considers the outer sample
topography, the estimated sample porosity can be referred as closed
porosity. In order to analyze the pore size distribution and the
respective contribution of them to porosity, the detected defect
volumes were simply recalculated to a pore size value which is equal
to the side length of assumed cubic pores.
For the porosity determination of the sample, already mechanically
tested tensile test samples were picked and prepared. Therefore,
samples of a size of 5520 mm3 were extracted from the clamping
areas of the tensile bars where no tensile force aected the samples
microstructure. For each machine charge three samples built with
dierent orientations were prepared and measured.

Fig. 6 displays the UTS and the TE of the tensile testing results of
the samples of the machines M1 to M6. Generally, the measured UTS
and TE values show that they depend on build orientation. So the
values decrease with the build angle increasing from 0 (H) and 45 (D)
to 90 (V). Moreover, scatter is also quite large for the D and V builds.
Only the H builds are close to the specication of the powder supplier
which are 45 MPa for PA2200 (EOS) and 43 MPa for DuraForm PA
(3D Systems) for the UTS and 20% for PA2200 (EOS) and 14% for
DuraForm PA (3D Systems) for the TE. Instead, the V builds show
mostly brittle fracture behavior resulting in TE values below 10%.
These observed anisotropic characteristics which are a consequence of
the stacking-layer-process are well known in literature [6,9].
Exceptions are the samples of machine M2 which achieved best
values by far for the UTS and the TE. Despite to the other machines,
higher UTS and higher TE for the V builds than for the D builds were
measured. The V builds feature a TE of around 20% which is as high as
the TE values of the H builds of the other machines. Thus, they are the
only V builds which yield the desired ductile fracture mode which
indicates a fracture that occurs after the onset of strain hardening [6].
Exceptionally distinctive necking with a high TE value of ca. 32% were
measured for the H builds of M2 proving a very high ductility
compared to the other samples. The TE values (TE for H and V builds)
are also clearly higher than the ones reported by the VDI upon a Round
Robin initiative using PA 12 powder with laser sintering [5]. However,
ductility of injection molded PA12 samples is still higher [6].
The outstanding properties of the samples built by machine M2
show that even along the building direction ductile mechanical properties can be reached. This must be the result of distinctive microstructural dierences compared to the samples built in the other machines.
Thus, in the following sections, a comprehensive microstructural
analysis is performed in order to trace back the mechanical performance to distinctive microstructural features.
4.2. Inuence of the pore morphology on the mechanical properties
A detailed analysis of the morphologies of porosity of the dierent
samples measured by X-ray computed tomography is given and
correlated to the mechanical testing results. The overall degree of
porosity of the samples, the pore concentration as well as the pore size
distributions are determined. The generated volume data sets of the
pores are visually reviewed in order to identify conspicuous porosity
arrangements.
4.2.1. Validity of the porosity measurement
In order to estimate the validity of the measurements, an example
of a pore size distribution is compared with its respective contribution

Fig. 5. Visualized CT data and detected pores (Sample size: ca. 1584 mm3).

34

Optics & Laser Technology 89 (2017) 3140

T. Stichel et al.

Fig. 6. Ultimate tensile strength (UTS) (a) and total elongation (TE) (b) of samples built with dierent orientation (D: diagonal; H: horizontal; V: vertical) in dierent machines (M1
M6). N is the number of samples.

Fig. 7. Exemplary measurement: relative frequency of the dierent pore sizes (pore occurrence relative to pore size) (a) and its contribution to the overall porosity (b). Cross section
image of a tensile test sample showing pores of dierent sizes.

to the overall porosity, see Fig. 7. Fig. 7a shows that there is a high
count of small pores, allowing the conclusion that pores with sizes
below the metrological structural resolution are also frequent.

However, Fig. 7b indicates that their contribution to the overall amount


of porosity can probably be neglected as the clear maximum of the
porosity contribution is located between 100 m and 200 m. With

Fig. 8. Boxplots of overall porosity and pore concentration of samples built with the machines M1-M6, analyzed using x-ray computed tomography.

35

Optics & Laser Technology 89 (2017) 3140

T. Stichel et al.

the formula [14]

pore sizes approaching the metrological structural resolution, the


porosity contribution is close to 0%. Nevertheless, the pore concentration can be expected to be underestimated, since cross section images,
see Fig. 7c, show indeed the presence of small pores below the
detection limit of the CT analysis.

2
1 K
ac= Ic
Yf

(2)

using a plane-strain fracture toughness KIc for nylon of 2.8 MPa m1/2
[14], and fracture stress of 40 MPa as well as a geometrical shape factor
Y of 1. Thus, a critical crack size ac of about 3.5 mm is estimated which
is clearly larger than the initial pore sizes measured for all samples.
Thus mechanical behavior is predominated by sub-critical crack growth
until the critical crack size is reached, leading nally to fast fracture.
The strong anisotropy of the mechanics (best TE and UTS values for
H built specimen) and the sensitivity of the TE values along the
building direction suggest that pores support directly an anisotropic
mechanical response. Hence, elongated pore shapes or coplanar
alignments have to be the reason which will be discussed in the
following section.
However, the correlation between mechanics and pore concentration is not valid for the H built samples which can be noticed when
comparing the TE and UTS of M1 with M2. They dier strongly despite
similar pore concentration distributions are observed. This shows that
pore morphology is not the only relevant factor inuencing the
mechanics of samples built by SLS (see Section 4.3).

4.2.2. Overall porosity and pore concentration


Fig. 8 displays the overall porosities and the pore concentration
(number of pores per cubic millimeter) of the samples of the dierent
machines. For the measurement three samples of each machine charge
(N=3) were prepared. The results show porosity values between 2.4%
and 3.9% which are in very good agreement with the values measured
by Dupin et al. with X-ray computed tomography for samples made
from Duraform PA by SLS [8]. Generally the highest porosity values are
observed with M1 and M4, the lowest values with M2 and M5.
Comparing the porosity with the mechanical testing results, a direct
relation between the overall porosity values and the mechanical
properties cannot be derived, which can be especially seen with the
results of M2 and M5 which both possess best porosity values but
dierent mechanical properties. Fig. 8b depicts the pore concentration
being ranged between 45 and 100 pores per mm3. It is remarkable that
the dierences of the pore concentrations between the machine charges
are not analogous to the porosity behavior. This shows that dierent
pore size distributions exist. The lowest pore concentrations were
measured for the samples built with the machines M1 and M2, which
also provided the best mechanical properties (see Section 4.1). Thus we
could suspect that the pore concentration is a superior indicator than
the overall porosity value for tracing back the origin of the mechanical
properties. However, an analysis of the pore size distribution is needed
for a more profound understanding of this observation.

4.2.4. Pore shapes and arrangements


Pore shapes and arrangements of samples built with the dierent
machines are shown on rendered images from the CT data sets in
Fig. 10. It can be recognized that shape as well as arrangement vary
depending on the machine used. The medium and large pores of M1
M4 possess quite complex, stretched shapes, while the corresponding
pores of M5 and M6 seem to be more round. The stretched pores which
consist obviously of small interconnected pores seem to be predominantly orientated along the layer direction.
Moreover, the pores of most samples are also aligned along the
layer orientation (coplanar). The pore morphology of M2 seems to be
an exception with a mostly arbitrary pore arrangement. M1, M4, M5,
M6 feature a strong coplanar pore morphology.
Despite to the coplanar alignment of the pores indicating the
anisotropic character of the SLS process, a distinctive interlayer
porosity cannot be seen as conrmed as it was often stated in literature
[6,8]. The aligned pores are too large with a size of roughly a whole
layer and, moreover, show large interspaces. However, since the
resolution limit of the procedure hinders the detection of pores smaller
than 20 m, the interlayer porosity might not be detected. According to
literature distinctive interlayer porosity would come along with poor
interlayer connection. This would induce delamination with TE values
near 1% upon tensile load along the build direction [6]. Since all
sample included in the Round Robin exceed this value it is likely that
no distinctive interlayer porosity exist. This is conrmed by additionally performed cross section analysis which revealed indeed the
presence of pores smaller than 20 m but no distinguishable interlayer
porosity.
The detailed mechanism for the coplanar pore arrangement is
unclear so far. It is imaginable that the pores align themselves to the
residual particle formations (see Section 4.3) during the melt state. So,
pores could rise like bubbles in the melt till they a hindered by aligned
residual particle material and end up in aligned pores after the cooling
cycle.
Consequently, the anisotropic mechanical response could be the
direct result of distinct pore shapes or arrangements since pores are
analogous to small pre-existing cracks. Thus, coplanar pore orientation
or arrangement could facilitate brittle fracture upon perpendicular
tensile load. The more coplanar the pores are the more directional and
fast will be the crack propagation. But if the pores are less ordered, the
fracture may originate in specic layers, but will probably travel
through pores in neighboring layers. However, in order to estimate

4.2.3. Pore size distribution


In order to clarify the inuence of porosity, the pore concentration
is plotted in dependence on the pore size in Fig. 9. The distributions
imply a maximum at the smallest pore sizes measurable and sometimes
a side maximum between 100 m and 150 m. The graphs depict clear
dierences of the pore size distribution between the measured samples.
Comparing the graphs of M1 and M4 in Fig. 9, which both provide
the highest porosity values but dierent mechanics, clearly dierent
pore size distribution can be recognized. While the samples built with
M1 possess a noticeable concentration of very large pores ( > 150 m)
which obviously contribute a lot to the porosity value, the samples built
with M4 feature a higher concentration of small (2040 m) and of
medium sized pores (40100 m). This leads to the clearly higher total
pore concentration of M4 compared to M1 (see Fig. 8b).
Moreover, the graphs show that the samples of M2 and M5 which
feature both a low porosity values possess a quite similar pore
concentration distribution for pore sizes between 60 m and 150 m.
However, the sample of M5 exhibits additionally a very high amount of
small pores (2040 m) which do not contribute much to the porosity
value but lead to the clearly higher total pore concentration displayed
in Fig. 8b.
When correlating the mechanical testing results (UTS and TE) of
Fig. 6 to the morphology observed, it is apparent that lower pore
concentrations which imply lower amount of small and medium sized
pores come along with superior mechanical properties for the V built
samples. This is given for the samples of M1 and M2. In case of
machine M2 we can depict that the measured D built sample possess
signicant higher pore concentration than the V and H built specimens.
This would explain that the D built samples of M2 feature lower
ductility compared to the V built samples (see Fig. 6).
Pores can be seen as small pre-existing cracks. However, the large
pores of the samples built with machine M1 do not result in exceptional
weak mechanics. Instead, second best values were achieved for the V
builds. This can be understood by calculating the critical crack size ac
for crack growth initiation using Linear Elastic Fracture Mechanics and
36

Optics & Laser Technology 89 (2017) 3140

T. Stichel et al.

Fig. 9. Pore concentration in dependency on the pore size of the samples built with dierent machines M1M6.

ductility in perpendicular direction is a logical consequence. Thus,


the lack of aligned non-molten particle (cores) and the supposedly
lower amount of residual particles (and cores) could explain the
superior mechanical properties of the samples of M2 observed in
Section 4.1. The amount of residual particles is estimated in the
following section.

the signicance of this eect, the inuence of residual particles and


particle cores will be investigated in the next section.
4.3. Inuence of residual particle (core) arrangements on the
mechanical properties
Fig. 11 shows microtomed thin sections of dierent specimen built
in vertical direction. They allow the identication of non-molten
material regions which can be particle cores or even complete particles.
The presence of non-molten material, surrounded by melted areas
inside SLS processed parts was identied already in previous research
[15,16]. The images in Fig. 11 show that each sample features partial
melting, whereby particles receive insucient energy to melt completely. Either inside the sample or adherent to the outside of the part
completely non-molten particles can be detected.
In the samples of the machines M1, M3, M4, M5 and M6 nonmolten particles aligned along the layer extent exist. Instead, the
sample of M2 lacks of aligned non-molten particles. Besides scattered
non-molten particles inside the part and the adherent particles at the
edge, the microstructure of the sample of M2 seems to consist mainly
of partly molten particles or completely molten areas and is homogeneous even at the border areas of the sample.
The presence of the non-molten particles is suggested to aect the
mechanical properties of the resulting samples [15,16]. Since short
binding necks between such particles occur which create high stiness
with little elongation, they support brittle macroscopic mechanical
behavior. If they are aligned along the layer orientation, as it is
observed for the samples of most machines in Fig. 11, the poor

4.4. Inuence of the degree of particle melt and crystallinity on the


mechanical properties
Since in SLS molten semi-crystalline materials continue to maintain
at a high process temperature until the whole manufacturing process is
nished, crystallization characteristics of a low degree of cooling rate
can be expected, resulting in a higher degree of crystallinity, formation
of crystal phase and coarse microstructure in SLS parts. On the
contrary, in injection molding (IM), fully molten semi-crystalline
materials are rapidly cooled down to room temperature after injection,
and thus show a higher cooling rate and rapid crystallization, which
gives rise to a low degree of crystallinity, absence of phase and a ner
microstructure in IM parts basically crystallized in form [17].
Moreover, it is known that for materials processed by IM an
increase in the crystalline content leads to an increase in tensile
strength and stiness, but at the detriment of the TE, as it is the
amorphous regions of a polymer that provide its ability to yield without
breaking [9,18]. In that case single phase materials (completely melted
and recrystallized: DPM =100%) are meant which percentage crystallinity is adjusted by the subsequent cooling rate. For materials
processed by SLS, however, double phase materials can be expected
37

Optics & Laser Technology 89 (2017) 3140

T. Stichel et al.

Fig. 10. Pore morphology of samples of the dierent machine charges visualized using CT data with a depth of eld of approximately 3 mm. Samples were visualized with vertical layer
orientation perpendicular to the image plane.

containing melted and recrystallized material as well as non-melted


particle cores and thus a DPM smaller than 100%. It is known that

DPM inuences the percentage crystallinity of the processed samples,


but also leads to versatile mechanical properties. Thus, it was observed

Fig. 11. Thin section images of samples built with vertical (V) orientation in the dierent machines M1-M6 allow the identication of non-molten particles.

38

Optics & Laser Technology 89 (2017) 3140

T. Stichel et al.

Fig. 12. Boxplots of percentage crystallinity and degree of particle melt (DPM) of samples of the machines M1-M6 estimated out of DSC charts.

5. Discussion

that the TE is increasing with decreasing crystallinity with a signicant


leap when the DPM is completed [9]. Moreover, the UTS also decreases
with increasing crystallinity until the point when DPM is completed,
where the direction of the trend is reversed [9].
According to Section 3.3, Dierential Scanning Calorimetry (DSC)
was used to obtain the percentage crystallinity of the pure powder
probe as well as the crystallinity and the degree of particle melt (DPM)
of laser sintered probes extracted from the built specimens.
The measurement of the native powders conrms that the respective powders agree concerning crystal phase. For all powder probes
single endotherm peaks were measured at a temperature of 185.6
0.8 C. This indicates the common and thermodynamically stable form crystal structure since the metastable -form crystals melt with
clearly lower temperatures resulting usually in recrystallization during
DSC an thus in a double endotherm peak in the thermogram [19]. By
estimation of the heat of melting, a percentage crystallinity of 52.5
0.7% for the PA 12 native powders is calculated.
The results of Fig. 12 show the percentage crystallinity and the
DPM measured for the dierent laser sintered samples. Generally,
percentage crystallinity values between 27% and 37% were received.
While M1, M3, and M4 provide roughly the same crystallinities at 30%,
the smallest value was measured for M2 with around 27% and the
largest for M5 and M6 around 34%. The DPM values lie between 60%
and 95%, whereas the highest value was achieved by M2. The smallest
values were measured for M5 and M6 with values below 70%.
For comparing the values with the tensile testing results of Section
4.1, the H builds and V builds are considered separately.
When focusing on the H builds, the inuence of possible inadequate
interlayer connection caused by distinct pore or residual particle
arrangement in layer orientation is excluded. M2 clearly provides the
lowest crystallinity and highest DPM. Since we know that both trends
should come along with higher UTS and TE, the tensile test results of
the specimen of M2 agree with that expectation. Especially the
outstanding TE value of ca. 32% for the H builds of machine M2
accounts for the theory. Instead, clearly lower TE values of 1520%
were measured for the other machines.
When considering the V builds, the inuence of the DPM on the
mechanics can easily depicted. The samples of M5 and M6 with the
lowest DPM result also in the lowest UTS and TE. However when
comparing the UTS and TE values of M1, M3 and M4 which all feature
a similar DPM and crystallinity, it is noticeable that M1 provides by far
the highest UTS and TE in this group. This shows that porosity
arrangements indeed play a signicant role for the mechanical properties along the building direction and are not completely outperformed
by the morphology of the residual particles and cores.

Results show that residual non-molten or partly molten particle


arrangements (indicating the layer structure), which is the result of
incomplete melting process upon energy input, are a key characteristic
which strongly inuences the mechanics along the building direction.
This is valid for most samples, resulting in total elongation values of 1
11% along the building direction. However, machine M2 delivered
samples without coplanar residual particle core arrangements, which is
obviously the main reason for the remarkable ductility with a total
elongation of around 20% and the highest ultimate strength values
along the building direction. Since the samples of M2 involve also the
highest degree of melting, resulting also in exceptional high total
elongation values of over 30% in the plane, it can be stated that
achieving a high melt eciency without residual non-molten or partly
molten particle arrangements is most crucial for mechanically advanced laser melting results.
In comparison to this nding, the observed dierent pore morphologies seem to be of minor signicance for the mechanical properties.
Nevertheless, its inuence can be depicted by comparing the characterization results of samples built with machine M1 to the ones of M3
and M4. While all these samples possess the same degree of particle
melt and a similar residual non-molten or partly molten particle
characteristic, the mechanical properties along the building direction
of M1 are the best for this group with total elongations of around 11%.
This is obviously related to the pore concentration which is the lowest
for the samples of M1 among all. A second proof of the signicance of
the pore morphology can be derived from the results of M2. While an
isotropic microstructure without residual particle arrangements was
observed, an anisotropic mechanical response is still present. This can
be related to the anisotropic pore morphology resembled by elongated
pores in coplanar direction, which could facilitate brittle sample
fracture upon perpendicular tensile load since then the critical crack
size is reached more easily.
Results hint also that both features (porosity and residual particles
and cores) are not necessarily linked since they can vary independently
in a certain range depending on the machine conguration. This can be
especially seen with the samples of M1 which feature pronounced
residual particle arrangements on the one hand but relatively low pore
concentration on the other hand. Thus, porosity and residual particle
morphology does not always change congruently with parameter
variation as it was observed, for example, by Dupin et al. [8].
6. Conclusions
The comprehensive microstructural analysis (X-ray computed
tomography, optical microscopy of thin sections, Laser Scanning
Microscopy and Dierential Scanning Calorimetry) of tensile test
39

Optics & Laser Technology 89 (2017) 3140

T. Stichel et al.

Ann. 40 (1991) 603614.


[2] G.N. Levy, R. Schindel, J.-P. Kruth, Rapid manufacturing and rapid tooling with
layer manufacturing technologies: state of the art and future perspectives, CIRP
Ann. 52 (2003) 89609.
[3] N. Hopkinson, R.J.M. Hague, P.M. Dickens, Rapid-Manufacturing An Industrial
Revolution for the Digital Age, Wiley, 2005 (ISBN 0470016132).
[4] R.D. Goodridge, C.J. Tuck, R.J.M. Hague, Laser sintering of polyamides and other
polymers, Prog. Mater. Sci. 57 (2012) 229267.
[5] VDI 3405 Blatt 1: Additive Fertigungsverfahren, Rapid Manufacturing LaserSintern von Kunststobauteilen Gteberwachung.
[6] D.K.Leigh, A comparison of polyamide 11 mechanical properties between laser
sintering and traditional molding, in: SFF Proceedings of International Solid
Freeform Fabrication Symposium 23, Austin, Texas, 2012, pp. 574605.
[7] D.L. Bourell, T.J. Watt, J. Trevor, D.K. Leigh, B. Fulcher, Performance limitations
in polymer laser sintering, Phys. Procedia 56 (2014) 147156.
[8] S. Dupin, O. Lame, C. Barrs, J.-Y. Charmeau, Microstructural origin of physical
and mechanical properties of polyamide 12 processed by laser sintering, Eur.
Polym. J. 48 (2012) 16111621. http://dx.doi.org/10.1016/j.eurpolymj.2012.06.007.
[9] C.MajewskiH.ZarringhalamN.Hopkinson, Eect of the degree of particle melt on
mechanical properties in selective laser-sintered Nylon-12 parts, in: Proceedings of
the Institution of Mechanical Engineers Part B Journal of Engineering
Manufacture, 222, 2008, pp. 10551064, doi:10.1243/09544054JEM1122.
[10] J.-P.KruthG.LevyR.SchindelT.CraeghsE.Yasa, Consolidation of Polymer Powders
by Selective Laser Sintering, in: PMI 2008 Proceedings of 3rd International
Conference, Ghent, Belgium, Sep1719, 2008, pp. 116.
[11] ASTM D63802a: Standard Test Method for Tensile Properties of Plastics.
[12] S. Gogolewski, K. Czerntawska, M. Gastorek, Eect of annealing on thermal
properties and crystalline structure of polyamides, Colloid Polym. Sci. 258 (1980)
11301136.
[13] J.C. Elliott, P. Anderson, G. Davis, S.D. Dover, S.R. Stock, T.M. Breunig,
A. Guvenilir, S.D. Antolovich, Application of X-ray microtomography in materials
science illustrated by a study of a continuous ber metal matrix composite, J. Xray
Sci. Technol. 2 (1990) 249258 (ISSN 0895-3996).
[14] M.A. Meyers, K.K. Chawla, Mechanical Behavior of Materials, 2nd edition,
Cambridge University Press, 2009 (ISBN 978-0-521-86675-0).
[15] E.MoeskopsN.KampermanB.van de VorstR.Knoppers, Creep Behaviour of
Polyamide in Selective Laser Sintering, in: SFF Proceedings of International Solid
Freeform Fabrication Symposium 15, Austin, Texaspp, 2004, pp. 6067.
[16] Y. Shi, Z. Li, H. Sun, S. Huang, F. Zeng, Eect of the properties of the polymer
materials on the quality of selective laser sintering parts, : Proc. Inst. Mech. Eng.
Part L J. Mater. Des. Appl. 218 (2004) 247252. http://dx.doi.org/10.1177/
146442070421800308.
[17] W. Zhua, C. Yana, Y. Shi, S. Wena, J. Liu, Y. Shi, Investigation into mechanical and
microstructural properties of polypropylene manufactured by selective laser
sintering in comparison with injection molding counterparts, Mater. Des. 82 (2015)
3745. http://dx.doi.org/10.1016/j.matdes.2015.05.043.
[18] Y. Kong, J.N. Hay, The enthalpy of fusion and degree of crystallinity of polymers as
measured by DSC, Eur. Polym. J. 39 (2003) 17211727. http://dx.doi.org/
10.1016/S0014-3057(03)00054-5.
[19] T. Laumer, K. Wudy, M. Drexler, P. Amend, S. Roth, D. Drummer, M. Schmidt,
Fundamental investigation of selective laser sintering of polymers for additive
manufacture, J. Laser Appl. 26 (2014) 042003. http://dx.doi.org/10.2351/
1.4892848.

samples built during a Round Robin initiative allowed valuing the


inuence of pore formation and residual particle and cores on
mechanical properties.
The mechanical tensile testing revealed a high variability of the
results depending on the machine conguration. Literature has been
conrmed which shows that the mechanical properties are aected by
the anisotropic nature of the SLS process, resulting mostly in quite
brittle characteristic along the building direction. However, the results
achieved with machine M2 exposed that high tensile strength and high
ductility is possible also for vertically built specimen made out of PA
12.
Microstructural analysis showed that both features, pores and
residual particles, contribute to the mechanical properties signicantly.
Thus anisotropic pore morphology resembled by elongated pores in
coplanar direction as well as coplanar residual particle arrangements
limit the total elongation values along the build direction. However,
results hint also that both features are not necessarily linked since they
can vary independently in a certain range depending on the machine
conguration. Thus, applicants are advised to control and optimize
both microstructural characteristic by developing adequate process
routines in order to ensure polymeric laser sintered parts with reliable
mechanical properties. Only if low pore concentration, the absence of
coplanar pore or residual particle arrangements and high degree of
particle melting is achieved, samples with advanced ductility over 20%
even along the building direction can be expected.
In future work, the back tracing of the sample features to the
production parameters will be performed on basis of the ndings of
this paper and published elsewhere which shall concretely help users of
Selective Laser Sintering optimizing their process routines and parameters.
Acknowledgment
The authors gratefully acknowledge the German Research
Foundation (DFG) (SFB 814 / 2) for the nancial support of this work
which is part (part project B6) of the collaborative research center SFB
814 Additive Fertigung. The authors are also thankful to the Lehrstuhl
fr Kunststotechnik (LKT) of the University Erlangen for supplying
DSC measurements.
References
[1] J.-P. Kruth, Material incress manufacturing by rapid prototyping techniques, CIRP

40

You might also like