ArticlePDF Available

Strengthening mechanisms in direct metal laser sintered AlSi10Mg: Comparison between virgin and recycled powders

Authors:

Abstract and Figures

Rod shaped samples of AlSi10Mg additively manufactured using recycled powder through direct metal laser sintering (DMLS) process showed higher quasi-static uniaxial tensile strength in both horizontal and vertical build directions than those of cast counterpart alloy. In addition, they offered mechanical properties within the range of other additively manufactured counterparts. TEM showed that the microstructure of the as-built samples comprised of cell-like structures featured by dislocation networks and Si precipitates. HRTEM studies revealed the semi-coherency characteristics of the Si precipitates. After deformation, the dislocation density increased as a result of generation of new dislocations due to dislocation motion. The dislocations bypassed the precipitates by bowing around them and penetrating the semi-coherent precipitates. Strengthening of recycled DMLS-AlSi10Mg alloys manufactured in both directions was attributed to Orowan mechanism (due to existence of Si precipitates), Hall-Petch effect (due to eutectic cell walls), and dislocation hardening (due to pre-existing dislocation networks). Due to the slightly different microstructure, the contribution of each strengthening mechanism was slightly different in identical samples made with virgin powder.
Content may be subject to copyright.
1
Strengthening mechanisms in direct metal laser sintered AlSi10Mg:
comparison between virgin and recycled powders
Amir Hadadzadeh1, 2
, Carter Baxter1, Babak Shalchi Amirkhiz2, Mohsen Mohammadi1
1Marine Additive Manufacturing Centre of Excellence (MAMCE), University of New Brunswick, Fredericton, NB,
E3B 5A1, Canada
2CanmetMATERIALS, Natural Resources Canada, 183 Longwood Road South, Hamilton, ON, L8P 0A5, Canada
Abstract
Rod shaped samples of AlSi10Mg additively manufactured using recycled powder through
direct metal laser sintering (DMLS) process showed higher quasi-static uniaxial tensile strength in
both horizontal and vertical build directions than those of cast counterpart alloy. In addition, they
offered mechanical properties within the range of other additively manufactured counterparts.
TEM showed that the microstructure of the as-built samples comprised of cell-like structures
featured by dislocation networks and Si precipitates. HRTEM studies revealed the semi-coherency
characteristics of the Si precipitates. After deformation, the dislocation density increased as a result
of generation of new dislocations due to dislocation motion. The dislocations bypassed the
precipitates by bowing around them and penetrating the semi-coherent precipitates. Strengthening
of recycled DMLS-AlSi10Mg alloys manufactured in both directions was attributed to Orowan
mechanism (due to existence of Si precipitates), Hall-Petch effect (due to eutectic cell walls), and
dislocation hardening (due to pre-existing dislocation networks). Due to the slightly different
microstructure, the contribution of each strengthening mechanism was slightly different in
identical samples made with virgin powder.
Keywords: Direct Metal Laser Sintering (DMLS); Additive manufacturing; Transmission electron
microscopy (TEM); Dislocation; Strengthening mechanism.
Corresponding author: Email: amir.hadadzadeh@unb.ca; Tel: +1 506 458 7104; Fax: +1 506 453 5025.
2
1. Introduction
Metal additive manufacturing (AM) is a process that uses metallic powder or wire to build
a part based on a 3D model, layer by layer until completion [1]. This way of manufacturing is very
attractive in comparison to other possibilities because of its ability to accommodate complex
design and cost saving [2] [3]. Amongst the available AM processes, powder bed fusion (PBF) is
the most commonly-used [4]. In PBF, a focused energy source selectively sinters or melts layers
of powder bed on top of each other [5]. PBF-laser based process, which is known as selective laser
melting (SLM) or direct metal laser sintering (DMLS), has specifically received significant
attention in recent years [1] [6] [7] [8] [9] [10] [11] [12].
SLM/DMLS of Al-Si and Al-Si-Mg alloys have been involved in multiple studies in recent
years, since these alloys possess light weight, superior specific strength, and outstanding corrosion
resistance [13] [14] [15] [16] [17]. Li et al. [18] and Suryawanshi et al. [19] studied SLM of Al-
12Si alloy and reported concurrent improvement of strength and ductility compared to cast
counterparts, due to very fine microstructure resulted from SLM. Evolution of very fine and unique
microstructure during SLM of aluminum alloys is a result of very high cooling rates in the range
of 103-108 °C/s [11] [20] [21] [22]. Similarly, higher strength and ductility obtained for
SLM/DMLS-AlSi10Mg under quasi-static loading conditions, reported in various studies [15] [23]
[24] [25]. Wu et al. [26] attributed the high strength of SLM-AlSi10Mg to the very fine cell-like
structure of the alloy where fine Si precipitates and eutectic Si boundaries inhibited dislocation
motion during plastic deformation. While Chen et al. [27] attributed the high strength of SLM-
AlSi10Mg mainly to the Orowan strengthening mechanism, Li et al. [28] described the strength
increment as a result of both Orowan and Hall-Petch mechanisms. Higher strength of SLM-
AlSi10Mg under high strain rate loading conditions compared to the same cast alloy was reported
by Zaretsky et al. [29]. Similar to the quasi-static loading conditions, both Si precipitates and
eutectic Si contributed to dislocation motion inhibition and hardening mechanism during dynamic
impact loading of DMLS-AlSi10Mg [30].
To make the SLM/DMLS process more affordable to produce aluminum parts, it is
essential to develop knowledge on the possibility of recycling the unused powder and reusing it
[31] [32]. Since a limited fraction of the powder is consumed during SLM process, the powder
outside the consolidation area can be recycled and reused [33] [34]. On the other hand, the quality
3
of the products in terms of microstructure and mechanical properties should not be compromised
using the recycled powder, specifically for sensitive applications such as aerospace industry [31].
In addition, growing sectors such as marine, off shore oil and gas, and shipbuilding industries have
asked for less expensive ways to adopt AM, where one sustainable way is to employ recycled
powder for additive manufacturing of parts regardless of specific AM technology. Therefore, it is
imperative to study the effect of recycled powder on the microstructure evolution, strengthening
mechanism, and mechanical properties of the products fabricated using SLM/DMLS process.
To do so, horizontal and vertical DMLS-AlSi10Mg samples were built using recycled
AlSi10Mg powder. The samples were printed using the same process parameters and then
subjected to uniaxial tensile test under quasi-static loading condition. The microstructure of the
samples and deformation mechanism were then studied in detail using advanced electron
microscopy characterization techniques. A comparison of mechanical properties, strengthening
mechanisms, and microstructures of the samples made using recycled and virgin AlSi10Mg
powder was finally presented.
2. Experimental procedure
2.1. AlSi10Mg powder
AlSi10Mg powder provided by EOS GmbH was the subject of study in this work. After
one build cycle of DMLS process and removing the parts, the unused powder was sieved using a
sieving equipment provided by the EOS Company with 60 µm mesh size to prepare the recycled
powder. Recycled AlSi10Mg powder (one time used), without mixing with the virgin powder, was
used to additively manufacture the samples. In addition, for comparison purposes complementary
samples were also made using virgin (unused) AlSi10Mg powder. The chemical composition of
the virgin powder is listed in Table 1. Analysis of the chemical composition of the recycled powder
showed that the level of the major alloying elements did not change [31], which is consistent with
other studies on the recycled AlSi10Mg powder [32] [35] [36]. The recycled and virgin powders
possessed spherical or near-spherical morphology with an average size of 10 ± 8 µm and 9 ± 7
µm, respectively [31]. Two types of powders (virgin and recycled) were very similar in terms of
powder size distribution (PSD), since they were sieved before using. Similarity of PSD was
4
reported for the recycled AlSi10Mg in the previous studies [32] [36]; however, a slight increase in
the average size of the recycled powder is due to agglomeration during heating cycles in the first
DMLS process.
Table 1- Chemical composition of virgin AlSi10Mg powder used in this study
Element
Si
Mg
Fe
Weight (%)
10.8
0.35
0.55
2.2. DMLS process
The samples prepared for this study were additively manufactured using an EOS M290
machine through DMLS technique with the specifications shown in Table 2. The laser power,
scanning speed, hatching distance, and powder layer thickness values are the same as the ones
used for virgin powder and it yields the least porosity and the best mechanical properties, where
they are listed in Table 3. The samples were manufactured under Ar-0.1%O2 atmosphere using the
strip scanning strategy where the laser beam was rotated 67 degrees between consecutive layers.
Table 2- EOS M290 machine specifications
Laser Type
Beam Spot Size
m)
Building plate dimensions (mm ×
mm × mm)
Preheat temperature
(°C)
400 W Yb-fiber
laser
100
250 × 250 × 325
200
Table 3- Process parameters used to build the samples
Laser Power (W)
Scan Speed (mm/s)
Hatch Distance (µm)
Layer Thickness (µm)
370
1300
190
30
To investigate the impact of building direction on the microstructure evolution and
mechanical properties of DMLS-AlSi10Mg, rod shaped samples with 120 mm height and 12 mm
diameter were manufactured in both vertical and horizontal directions (three samples in each
direction) as shown in Figure 1. The longitudinal axes in the vertical and horizontal samples were
parallel and perpendicular to the building direction, respectively. In order to acquire a deeper
understanding of the microstructure of recycled DMLS-AlSi10Mg alloy built in horizontal and
5
vertical directions, similar rod shape samples with the same geometry were fabricated using virgin
AlSi10Mg powder with the same process parameters [31].
Figure 1- Schematic of horizontal and vertical samples.
2.3. Transmission electron microscopy
Details of the microstructures of as-built and deformed samples were studied using an FEI
Tecnai Osiris TEM. The TEM facility was equipped with a 200 keV X-FEG gun. The precipitates
were analyzed using the Super-EDS X-ray detection system combined with the high current
density electron beam in the scanning transmission electron microscopy (STEM) mode. Spatial 1
nm resolutions were obtained during EDS elemental mapping using a sub-nanometer electron
probe. Electron transparent samples for TEM characterization were prepared using ion milling
method. Initial slices of samples were cut with < 500 µm using a diamond wafering blade followed
by punching of 3 mm diameter disks using a Gatan puncher. The disks were then polished to a
thickness of 80-90 µm followed by dimpling using alumina suspension down to about 10 μm at
sample center. The final ion milling of the dimpled disk was done using a Gatan 691 PIPS with
liquid nitrogen cooling at 5, 3, and 1 keV and gun angle of 4° for 130, 10 and 30 minutes average
time at each step respectively, until the perforation [37].
Microstructural analyses were performed perpendicular to the longitudinal axes of the rods
and tensile specimens after deformation, from the necking area. Such a configuration was chosen
to systematically study the deformation mechanism in planes normal to the tensile direction.
6
Therefore, the microstructure of the horizontal sample was studied in the y-z plane, where for the
vertical sample the microstructure was analyzed in the x-y plane (see Figure 1 for planes).
2.4. Uniaxial tensile test
Mechanical properties of the vertical and horizontal samples were evaluated at room
temperature under quasi-static tensile loading. Cylindrical tensile samples (with a gauge length of
24 mm and a diameter of 6 mm) were machined from the as-built rods based on ASTM E8-15a
standard [38]. The uniaxial tensile tests were conducted using an Instron Model 1332 universal
testing machine at a strain rate of 9×10-4 s-1. Three tensile tests were performed for each building
direction and engineering stress-strain curves were recorded.
3. Results & Discussion
3.1. Microstructure of as-built DMLS-AlSi10Mg
Figure 2 shows the STEM bright field (STEM-BF) microstructure of the as-built samples,
taken from the inside of the melt pools, since the overall structure of DMLS-AlSi10Mg is
dominated by the features inside the melt pools [18] [19] [39]. Similar to typical SLM/DMLS-
AlSi10Mg, recycled AlSi10Mg microstructure consisted of fine cell-like primary aluminum (α-
Al) developed, which then bounded by a continuous network of eutectic Si [27] [40]. Referring to
EDS elemental maps of Fe and Mg, some intermetallic phases formed at the cell boundaries, which
are mainly Al8FeMg3Si6 and Al3FeSi [30].
Microstructure of horizontal sample is featured by a combination of equiaxed and columnar
cells as the columnar cells evolved along the building direction due to epitaxial growth [41]. On
the other hand, equiaxed cells observed in the microstructure of vertical sample are the cross-
sections of the columnar dendrites, since the microstructure is studied in the x-y plane. The average
cell size (eutectic Si wall to wall distance) in horizontal and vertical samples is 0.57 µm and 0.55
µm, respectively. This is close to the observations made earlier for SLM-AlSi10Mg fabricated
using virgin powder [27].
7
(a)
(b)
Figure 2- STEM-BF microstructure of (a) vertical and (b) horizontal samples along with the
corresponding EDS elemental maps.
Details of submicron characteristics of primary Al dendrites are shown in Figure 3. The
unique feature observed inside the cells is the entangled network of dislocations interacting with
both the eutectic Si and Si precipitates, referring to the EDS elemental maps of Si [30]. These
networks evolved due to rapid solidification and associated thermal stresses during DMLS process
[42] [43] [44]. The number of dislocations observed in these two samples was different, where
higher number of free dislocations along with dislocation networks was presented in the horizontal
sample. Several locations on the vertical and horizontal TEM samples were studied to confirm this
observation.
By measuring the TEM foils thickness using electron-energy loss spectroscopy (EELS),
log ratio technique [45], and the method reported by Pesicka et al. [46] and Rojas et al. [47], the
dislocation density ( in m-2) was calculated for both samples. By conducting the measurements
over various locations, the dislocation density was calculated as 2.15 × 1014 m-2 and 1.41 × 1014
8
m-2 in horizontal and vertical samples, respectively. Different dislocation density values in the
horizontal and vertical samples were due to various heating and cooling cycles applied during the
building process. In fact, by changing the building direction, the geometry of the sample with
respect to the building plate and the contact area between the sample and plate changed
significantly. Therefore, thermal boundary conditions and thermal gradients changed in a way that
the dislocation density varied between the horizontal and vertical samples [48].
(a)
(b)
(c)
(d)
Figure 3- STEM-BF microstructure of (a) vertical and (b) horizontal samples. The
corresponding EDS elemental maps of Si are shown in (c) and (d). All images were taken with
the beam direction along the low-index zone axis orientation of Al.
To further investigate the Si precipitates in the Al matrix, high resolution TEM (HRTEM)
of these precipitates were investigated as shown in Figure 4. Fast Fourier transform (FFT) patterns
of the matrix and precipitates are also shown in the same figure. Comparison between the FFT
patterns of the matrix and precipitates in both vertical and horizontal samples shows that there is
an orientation relationship between the precipitates and the matrix. However, it was observed that
9
the interface between the precipitates and the matrix was not fully coherent implying semi-
coherent precipitates for both vertical and horizontal recycled DMLS-AlSi10Mg alloys.
The size of precipitates in different locations on the TEM foils for both samples were
studied afterwards, where the average representative precipitate size in the horizontal and vertical
samples was 61 nm and 65 nm, respectively. This is in the range of coherent nano-Si precipitates
size reported by Li et al. [28] using virgin AlSi10Mg powder; however, the precipitates in the
current study were not coherent with the matrix for both vertical and horizontal samples. This is
mainly due to specific thermal boundary conditions applied during DMLS process in the current
study (which are different from those applied in [28]), where a semi-coherent interface between
the Si precipitates and the Al matrix evolved. By changing the thermal boundaries and heating-
cooling cycles, the precipitates could be exposed to high temperatures for longer time, which then
leads to evolution of semi-coherent interface.
(a)
(b)
Figure 4- HRTEM images of Si precipitates in (a) vertical and (b) horizontal samples along
with the corresponding FFT patterns of matrix (box 1) and precipitates (box 2).
10
3.2. Mechanical properties
Typical engineering stress-strain curves of recycled DMLS-AlSi10Mg built in horizontal
and vertical directions under uniaxial tensile loading are shown in Figure 5, where the two samples
exhibited consistent flow behavior. However, the main difference between them is the tensile
elongation. While the horizontal sample elongated up to ~7.2%, the vertical sample failed at a
strain of ~5.5%. Ductility of DMLS components is sensitive to defects such as voids and cracks
rather than microstructural characteristics [44]. Evolution of such defects is strongly dependent on
the processing conditions. Since both samples were fabricated using the same process parameters,
it seems that the configuration of defects was different in the two samples so that the vertical
sample was more prone to fail at lower fracture strains [31]. Yield strength (YS) and ultimate
tensile strength (UTS) of the horizontal sample was 235 MPa and 386 MPa, respectively. The
vertically built sample possessed YS and UTS of 210 MPa and 392 MPa, respectively.
Tensile tests of recycled DMLS-AlSi10Mg revealed that both YS and elongation of the
vertical and horizontal samples are higher than those of cast counterpart alloy (for mechanical
properties of die cast A360 from see [49] [50]), as shown in Figure 6. Such an enhancement in the
mechanical behavior is due to the very fine microstructure of the DMLS alloy (cell size of <1 µm
in DMLS parts in comparison to typical 100-300 µm in A360 [51] [52]). In addition, YS and
elongation of SLM/DMLS-AlSi10Mg produced from virgin powder were summarized from
literature and shown in the same graph. The uniaxial tensile properties of recycled DMLS-
AlSi10Mg are within the range of those of virgin SLM/DMLS-AlSi10Mg, which indicates the
feasibility of producing AM parts from recycled powder without compromising the mechanical
properties. It should be noted that, the mechanical properties from literature reported in Figure 6
were resulted from using various process parameters including different laser power, powder layer
thickness, scan speed, and powder bed fusion machine.
11
Figure 5- Engineering stress-strain curves of the horizontal and vertical recycled AlSi10Mg
samples.
Figure 6- The yield strength and ductility data of SLM/DMLS-AlSi10Mg and cast counterpart
alloy (A360) from literature [11] [14] [15] [24] [27] [31] [53] [54] [49] [50].
3.3. Deformation behavior
Figure 7 shows the deformed microstructure of horizontal and vertical samples in STEM-
BF mode. The microstructure of deformed DMLS-AlSi10Mg in both directions is featured with
heavily entangled dislocation networks, developed toward the cell walls (eutectic Si) and in
12
interaction with the Si precipitates [55]. Obviously, dislocation glide during plastic deformation
resulted in the generation of new dislocations and increase of dislocation density [27] [30].
Therefore, an entangled network was developed. Moreover, the cell walls acted as barrier against
dislocation motion and hindered their further dynamics [44]. Referring to the EDS elemantal maps
of the samples, morphology of the eutectic Si networks did not change during plastic deformation.
Moreover, the dislocation density is higher at the vicinity of the eutectic Si walls [55].
Using the method described earlier, the dislocaton density was calculted for the deformed
DMLS-AlSi10Mg alloys in both directions. The calculated dislocation density was 2.4 × 1014 m-2
in the horizontal sample and 2.7 × 1014 m-2 in the vertical one. While the dislocation density in the
as-built horizontal sample was ~1.5 times higher than of that in the vertical sample, the post-
deformed dislocation density in two samples was very close. Under quasi-static cold deformation
of DMLS-AlSi10Mg, the dislocation density increased up to a near-saturation level, followed by
final fracture. Since the initial dislocation density in the vertical sample was lower, the new
dislocations generated more rapidly than the horizontal sample. As the dislocation density was
increased in the material and some dislocations were piled-up [56], a “back stress” developed in
the matrix and inhibited further motion and generation of dislocations [57] [58].
13
(a)
(b)
Figure 7- STEM-BF images of deformed (a) vertical and (b) horizontal samples along with the
corresponding EDS elemental maps.
Despite the similarity in the amount of the alloying elements in the virgin and recycled
powders and low propensity of recycled AlSi10Mg powder to absorb high level of oxygen [32]
[35] [36], the deformed microstructure was analyzed in terms of possible oxides, and the
corresponding oxygen and magnesium EDS maps are shown in Figure 8. As seen, the overall
oxygen in the microstructure is not significant; however, some local oxygen concentration is
observed in both samples. In both vertical and horizontal samples, it appears that oxygen tended
to form magnesium oxide. Presence of oxygen in the recycled AlSi10Mg powder is in the form of
thin films of MgAl2O5 compounds on the surface of the powders [32]. During the DMLS process
and in the presence of heat, it appears that magnesium oxides formed and distributed mainly at the
14
vicinity or along the eutectic Si walls. Presence of oxides can have a detrimental effect on the
elongation of the samples [35], with no significant effect on the strength.
(a)
(b)
Figure 8- EDS elemental maps of oxygen and magnesium corresponding to STEM-BF images
shown in Figure 7, (a) vertical and (b) horizontal samples. Oxides are marked by white circles
in the O map.
3.4. Strengthening mechanism
Existence of particles and precipitates in the matrix of alloys is a common mechanism of
strengthening, since they act as obstacles against the motion of dislocations [59]. Dislocations
bypass precipitates by Orowan looping, cross-slip or particle shearing (cutting/penetrating) [60]
depending on the characteristics of the precipitates. Since the Si precipitates in recycled DMLS-
AlSi10Mg alloys both in the vertical and horizontal directions are semi-coherent (Figure 4), a
misfit exists at the precipitate-matrix interface. Therefore, a misfit strain field was shared between
the particle and the matrix [61], which affected the precipitate-dislocation interaction.
Details of dislocation-Si precipitate interaction and how dislocations bypass the particles
are shown in Figure 9 for the vertical and Figure 10 for the horizontal sample. In the vertical
15
sample where the plastic deformation (elongation) was less than the horizontal sample (~5.5%
versus ~7.2%), an evidence of the early stage of dislocation-precipitate interaction can be observed
(Figure 9). In this stage, the dislocation bypassed the Si precipitates by bowing around them.
Similarly, in the horizontal sample, bypassing of dislocation from the precipitates initiated with
the bowing procedure as seen in Figure 10 (a). As deformation continued, the dislocation passed
the precipitate and bowed toward it (Figure 10 (b)). In some cases, bowing away from the
precipitates after bypassing them is possible as well.
Bowing of dislocation toward or away the precipitate is due to the offset position of the
dislocation slip plane from the precipitate’s center [62]. If the slip plane of the dislocation is
slightly below the precipitate center, the dislocation bows toward the precipitate after bypassing
and if the slip plane of the dislocation is slightly above the precipitate center, the dislocation bows
away from the precipitate after bypassing [62]. In the case that the precipitate center is on the slip
plane of the dislocation, the dislocation remains straight after bypassing [62]. Since no evidence
of dislocation loop formation around the precipitates after bypassing was observed, it seems that
the mechanism of dislocation bypassing in the current study for recycled DMLS-AlSi10Mg is
particle penetrating [62]. Both horizontal and vertical recycled DMLS-AlSi10Mg TEM foils were
scanned in multiple locations for other strengthening mechanisms and dislocation-Si precipitates
interactions, where solely the same mechanism reported in Figure 9 and Figure 10 was observed
through out the foils. Based on the observation in many locations, a simplistic schematic of
dislocation-Si precipitate interaction in recycled DMLS-AlSi10Mg for both vertical and horizontal
samples is proposed and shown in Figure 11, for the cases that the slip plane of the dislocation is
slightly below or above the precipitate center. Based on the proposed mechanism, the dislocation
moves in the matrix as a result of applying stress on the material. Once the dislocation reaches the
vicinity of the precipitate, the interaction between the dislocation and the particle initiates. Due to
the misfit stress around the semi-coherent precipitate, the dislocation does not remain straight as
it bypasses the particle. After bypassing, the misfit stress modifies the glide force in a way that the
dislocation either bows away or toward the particle, depending on the relative position of particle
center and the dislocation slip plane.
16
Figure 9- STEM-BF images of dislocation-Si precipitate interaction (bowing) in deformed
vertical sample.
(a)
(b)
Figure 10- STEM-BF images of dislocation-Si precipitate interaction in deformed horizontal
sample in (a) early stage of interaction (bowing) and (b) after bypassing.
17
Figure 11- Schematic interaction of dislocation-Si precipitate in DMLS-AlSi10Mg where the
slip plane of the dislocation is slightly (a) below and (b) above the precipitate center.
Considering the characteristics of the deformed microstructure of recycled DMLS-
AlSi10Mg (Figure 7-Figure 10), the strengthening mechanism in the material can be attributed to
Orowan mechanism (due to existence of Si precipitates), Hall-Petch effect (due to eutectic cell
walls), and dislocation hardening (due to pre-existing dislocation network) [63]. Therefore, the
yield strength () of both alloys can be calculated as follows
      
(1)
where is an internal friction stress. The Orowan mechanism in Al-Si binary system can
be estimated using Eq. (2) [28] written as follows
 
 

(2)
where is a material constant, is the shear modulus, is the Burgers vector,  is the
diameter of Si precipitates, and  is the volume fraction of Si precipitates. The Hall-Petch
strengthening in AM parts (by considering the cellular structure) is estimated using Eq. (3) [64]
given as below
18
  
(3)
where is a material constant and is the aluminum cell size. The dislocation hardening
is estimated using Eq. (4) [65], which can be written as
  
(4)
where is a material constant, is the Taylor factor, and is the dislocation density.
The parameters used in Eqs. (1)-(4) are all listed in Table 4. In addition, the calculated values of
each strengthening component and predicted yield strength of recycled DMLS-AlSi10Mg in two
directions are summarized in Table 5. The percent error between the calculated and measured yield
strength in horizontal and vertical samples was 5.5% and 2.4%, respectively. The good agreement
between the calculated and measured yield strength values implies the reliability of proposed
strengthening mechanisms to calculate the strength of DMLS_AlSi10Mg.
Table 4- Parameters used in Eqs. (1)-(4)
Parameter
Units
Value
Reference
/Friction stress
MPa
72
[66]
/Material constant
-
0.4*
[67] [68]
/Shear modulus of
Al matrix
GPa
26.5
[27]
/Burgers vector of
Al
nm
0.286
[27]
/Diameter of Si
precipitates
nm
61 (horizontal)-65
(vertical)
Current study
/Volume fraction
of Si precipitates
vol.%
2.5
[28]
/Material constant
MPa.m1/2
0.04
[69]
/Cell size
µm
0.57 (horizontal)-0.55
(vertical)
Current study
/Material constant
-
0.16
[70]
/Taylor factor
-
3.06
[27]
/Dislocation
density
m-2
2.15 × 1014
(horizontal)-1.41 ×
1014 (vertical)
Current study
* Value for semi-coherent precipitates.
19
Table 5- Calculated strengthening components in recycled DMLS-AlSi10Mg

(MPa)

(MPa)

(MPa)
(MPa)
Measured YS
(MPa)
Error (%)
Horizontal
80
53
17
222
235
5.5
Vertical
75
54
14
215
210
2.4
3.5. Comparison with virgin DMLS-AlSi10Mg
Using the same advanced microscopy techniques as described before, the microstructure
of the virgin DMLS-AlSi10Mg in both directions were then characterized. The overall morphology
of aluminum cells in the virgin DMLS-AlSi10Mg was similar to that of recycled DMLS-
AlSi10Mg. The virgin DMLS-AlSi10Mg microstructure was consisted of fine cellular aluminum
(α-Al) bounded by a continuous network of eutectic Si. However, the average cell size in the virgin
DMLS-AlSi10Mg alloy was higher than those in the recycled samples, where it was 0.77 µm in
the horizontal sample and 0.86 µm in the vertical one.
Moreover, the submicron characteristics of the microstructure were different in the virgin
DMLS-AlSi10Mg samples. Figure 12 shows the STEM-BF images of virgin vertical and
horizontal DMLS-AlSi10Mg samples. Similar to the recycled samples, networks of dislocation
developed in the microstructure of virgin DMLS-AlSi10Mg. However, the dislocation density in
virgin vertical and horizontal samples was 3.05 × 1014 m-2 and 1.14 × 1014 m-2, respectively. In
contrasts to the recycled samples, the dislocation density in virgin vertical sample was higher than
that of virgin horizontal sample.
Figure 13, shows the HRTEM images of Si precipitates in virgin DMLS-AlSi10Mg
samples in vertical and horizontal directions along with the corresponding FFT patterns of matrix
and precipitates. The Si precipitates in the virgin samples developed with different characteristics
in comparison with the recycled ones. As shown in Figure 13, while fine and coherent Si
precipitates formed in the virgin vertical sample, coarse and semi-coherent ones developed in the
virgin horizontal sample. The average diameter of Si precipitates in vertical and horizontal virgin
samples was 25 nm and 62 nm, respectively.
The main reason of differences between the submicron characteristics of virgin and
recycled DMLS-AlSi10Mg can be related to the particle size. The recycled powder possessed
slightly larger particle size than the virgin powder since the particles absorbed energy in the
20
previous DMLS process and slightly enlarged [31]. Since finer particles provide a larger surface
area, they can absorb more laser energy and consequently their working temperature and sintering
kinetics are higher [71]. On the other hand, the effective thermal conductivity of the powder bed
is dependent on the amount of gas in the inter-particle pores, which is controlled by the particle
size [72]. Finer particles will provide less pores and higher effective thermal conductivity. The
complex interaction between the thermal boundaries, sample geometry, working temperature and
sintering kinetics led to differences in the submicron characteristics of virgin and recycled DMLS-
AlSi10Mg. Consequently, the variation in the thermal boundaries alters the thermal cycles
experienced by the virgin and recycled samples, which then leads to evolution of semi-coherent
precipitates in the recycled and coherent ones in the virgin vertical sample.
(a)
(b)
(c)
(d)
Figure 12- STEM-BF microstructure of virgin (a) vertical and (b) horizontal samples. The
corresponding EDS elemental maps of Si are shown in (c) and (d). All images were taken
with the beam direction along the low-index zone axis orientation of Al.
21
(a)
(b)
Figure 13- HRTEM images of Si precipitates in virgin (a) vertical and (b) horizontal samples
along with the corresponding FFT patterns of matrix (box 1) and precipitates (box 2).
Due to more pronounce difference in the microstructure and submicron characteristics of
virgin and recycled vertical samples, the mechanical properties of virgin vertical DMLS-
AlSi10Mg were measured for further discussion. Figure 14 shows the comparison between the
uniaxial tensile stress-strain curves of virgin and recycled DMLS-AlSi10Mg in vertical
configuration. The test was performed under the same quasi-static strain rate for consistency and
comparison. The yield strength of the two samples was very close, where, a slight difference
between the UTS of the samples was observed. Moreover, the elongation of the virgin sample was
slightly higher than that of the recycled one. The virgin samples were made using a slightly finer
powder (with an average particle size of ~9 µm in comparison with the average particle size of
~10 µm for the recycled powder), which can lead to better mechanical properties both in terms of
ultimate strength and ductility. Moreover, finer particles possess better densification properties
[71]; therefore, it is very likely that the porosity percentage in the recycled DMLS-AlSi10Mg was
slightly higher than the virgin one, which then resulted in lower elongation. In addition, presence
22
of magnesium oxide particles in the microstructure of recycled AlSi10Mg (Figure 8) is another
reason for lower elongation of the recycled sample.
Despite the dissimilarities in the microstructure of virgin and recycled samples, the yield
stresses were similar, in the vertical sample. Using the strengthening mechanism equations
developed earlier the similarities in terms of yield strength can be described. The yield strength of
vertical virgin DMLS-AlSi10Mg was calculated, where comparable results to the recycled one
were obtained, as shown in Table 6. Competing strengthening mechanisms in the virgin and
recycled samples contributed at different levels (Table 5 and Table 6). However, the overall
strength of the virgin and recycled samples, resulted from the overall contributions was
comparable.
Table 6- Calculated strengthening components in virgin DMLS-AlSi10Mg

(MPa)

(MPa)

(MPa)
(MPa)
Vertical
73*
43
20
209
* -constant in Eq. (2) for coherent precipitates is 0.15 [68].
Figure 14- Engineering stress-strain curves of the recycled and virgin vertical samples.
23
4. Conclusions
In the current study, feasibility of fabrication of DMLS-AlSi10Mg components using
recycled powder was investigated. Using both virgin and recycled powders, rod shaped samples
of AlSi10Mg were additively manufactured through DMLS process in vertical and horizontal
directions. Both recycled vertical and horizontal samples exhibited superior mechanical properties
compared to the cast counterpart. Moreover, the strength and elongation of the recycled DMLS-
AlSi10Mg were in the range of other SLM/DMLS-AlSi10Mg counterparts. The followings were
the main findings of the current study:
The microstructure of the as-built recycled DMLS-AlSi10Mg samples in both
directions comprised cell-like structure (primary α-Al bounded by eutectic Si walls)
and featured by pre-existing networks of dislocations and semi-coherent Si
precipitates.
After tensile tests, the dislocation density in both microstructures was increased and
saturated to almost the same order due to the motion of dislocations and generation
of new ones. Consequently, heavily entangled networks of dislocations evolved in
the microstructures.
It was observed that the dislocations interacted with the eutectic Si walls and Si
precipitates. Analysis of the dislocations-Si precipitates interaction in recycled
samples revealed that the dislocations bypassed the precipitates by bowing around
them and penetrating the precipitates due to the semi-coherency of the precipitates.
Strengthening of recycled DMLS-AlSi10Mg was attributed to Orowan mechanism
(due to existence of Si precipitates), Hall-Petch effect (due to eutectic cell walls),
and dislocation hardening (due to pre-existing dislocation network).
Modeled yield strength of the recycled DMLS-AlSi10Mg samples (considering
different strengthening mechanisms) was in good agreement with the actual
measured one in both directions.
The comparison between the microstructure of virgin and recycled DMLS-
AlSi10Mg samples showed differences in terms of cell size, Si precipitate
characteristics, and dislocation density. However, the calculated strength of the
samples was comparable as confirmed by tensile test of virgin samples.
24
Acknowledgement
The Authors would like to thank Natural Sciences and Engineering Research Council of
Canada (NSERC) grant number RGPIN-2016-04221 and New Brunswick Innovation Foundation
(NBIF) grant number RIF2017-071 for providing sufficient funding to execute this work. The
authors would also like to acknowledge AMM for fabricating the DMLS samples, Dr. Mark
Kozdras at CanmetMATERIALS for facilitating the research and Catherine Bibby for TEM
sample preparations.
References
[1]
D.D. Gu, W. Meiners, K. Wissenbach, R. Poprawe, Laser additive manufacturing of metallic
components: materials, processes and mechanisms. Int. Mat. Rev. 57 (2012) 133-164,
https://doi.org/10.1179/1743280411Y.0000000014.
[2]
H. Asgari, M. Mohammadi, Microstructure and mechanical properties of stainless steel CX
manufactured by Direct Metal Laser Sintering. Mater. Sci. Eng. A 709 (2018) 8289,
http://dx.doi.org/10.1016/j.msea.2017.10.045.
[3]
W.E. Frazier, Metal Additive Manufacturing: A Review. J. Mater. Eng. Perform. 23 (2014)
1917-1928, http://doi.org/10.1007/s11665-014-0958-z.
[4]
L.X. Lu, N. Sridhar, Y.W. Zhang, Phase field simulation of powder bed-based additive
manufacturing. Acta Mater. 144 (2018) 801-809,
https://doi.org/10.1016/j.actamat.2017.11.033.
[5]
W.J. Sames, F.A. List,, S. Pannala, R.R. Dehoff, S.S. Babu, The metallurgy and processing
science of metal additive manufacturing. Int. Mater. Rev. 61 (2016) 315-360,
https://doi.org/10.1080/09506608.2015.1116649.
25
[6]
N.T. Aboulkhair, N.M. Everitt, I. Ashcroft, C. Tuck, Reducing porosity in AlSi10Mg parts
processed by selective laser melting. Addit. Manuf. 1-4 (2014) 77-86,
http://dx.doi.org/10.1016/j.addma.2014.08.001.
[7]
J.H. Martin, B.D. Yahata, J.M. Hundley, J.A. Mayer, T.A. Schaedler, T.M. Pollock, 3D
printing of highstrength aluminium alloys, Nature 549 (2017) 365369,
https://doi.org/10.1038/nature23894.
[8]
A.B. Spierings, K. Dawson, P. Dumitraschkewitz, S. Pogatscher, K. Wegener,
Microstructure characterization of SLM-processed Al-Mg-Sc-Zr alloy in the heat treated and
HIPed condition. Addit. Manuf. 20 (2018) 173181,
https://doi.org/10.1016/j.addma.2017.12.011.
[9]
C. Galy, E.L Guen, E. Lacoste, C. Arvieu, Main defects observed in aluminum alloy parts
produced by SLM: From causes to consequences. Addit. Manuf. 22 (2018) 165-175,
https://doi.org/10.1016/j.addma.2018.05.005.
[10]
S.Z. Uddin, L.E. Murr, C.A. Terrazas, P. Morton, D.A. Roberson, R.B. Wicker, Processing
and characterization of crack-free aluminum 6061 using high temperature heating in laser
powder bed fusion additive manufacturing. Addit. Manuf. 22 (2018) 405-415 ,
https://doi.org/10.1016/j.addma.2018.05.047.
[11]
D. Herzog, V. Seyda, E. Wycisk, C. Emmelmann, Additive manufacturing of metals. Acta
Mater. 117 (2016) 371-392, http://dx.doi.org/10.1016/j.actamat.2016.07.019.
[12]
T. DebRoy, H.L. Wei, J.S. Zuback, T. Mukherjee, J.W. Elmer, J.O. Milewski, A.M. Beese,
A. Wilson-Heid, A. De, W. Zhang, Additive manufacturing of metallic components
Process, structure and properties, Prog. Mater. Sci. 92 (2018) 112224.
https://doi.org/10.1016/j.pmatsci.2017.10.001.
[13]
K.G. Prashanth, S.Scudino, H.J.Klauss, K.B.Surreddi, L.Löber, Z.Wang, A.K.Chaubey,
U.Kühn, J.Eckert, Microstructure and mechanical properties of Al12Si produced
26
byselective laser melting: Effect of heat treatment. Mater. Sci. Eng. A 590 (2014) 153-160,
http://dx.doi.org/10.1016/j.msea.2013.10.023.
[14]
N. Read, W. Wang, K. Essa, M.M. Attallah, Selective laser melting of AlSi10Mg alloy:
process optimisation and mechanical properties development, Mater. Des. 65 (2015) 417
424, https://doi.org/10.1016/j.matdes.2014.09.044.
[15]
W. Li, S. Li, J. Liu, A. Zhang, Y. Zhou, Q. Wei, C. Yan, Y. Shi, Effect of heat treatment on
AlSi10Mg alloy fabricated by selective laser melting: microstructure evolution, mechanical
properties and fracture mechanism, Mater. Sci. Eng. A 663 (2016) 116-125,
https://doi.org/10.1016/j.msea.2016.03.088.
[16]
Y. Bai, Y. Yang, Z. Xiao, M. Zhang, D. Wang, Process optimization and mechanical
property evolution of AlSiMg0.75 by selective laser melting. Mater. Des. 140 (2018) 257-
266, https://doi.org/10.1016/j.matdes.2017.11.045.
[17]
P. Fathi, M. Mohammadi, X. Duan, A.M. Nasiri, A comparative study on corrosion and
microstructure of direct metal laser sintered AlSi10Mg_200C and die cast A360.1 aluminum.
J. Mater. Process. Technol. 259 (2018) 1-14,
https://doi.org/10.1016/j.jmatprotec.2018.04.013.
[18]
X.P. Li, X.J. Wang, M. Saunders, A. Suvorova, L.C. Zhang, Y.J. Liu, M.H. Fang, Z.H.
Huang, T.B. Sercombe, A selective laser melting and solution heat treatment refined Al
12Si alloy with a controllable ultrafine eutectic microstructure and 25% tensile ductility,
Acta Mater. 95 (2015) 7482, http://dx.doi.org/10.1016/j.actamat.2015.05.017.
[19]
J. Suryawanshi, K.G. Prashanth, S. Scudino, J. Eckert, O. Prakash, U. Ramamurty,
Simultaneous enhancements of strength and toughness in an Al-12Si alloy synthesized using
selective laser melting. Acta Mater. 115 (2016) 285-294,
http://dx.doi.org/10.1016/j.actamat.2016.06.009.
[20]
D. Gu, Y.C. Hagedorn, W. Meiners, G. Meng, R.J.S. Batista, K. Wissenbach, R. Poprawe,
Densification behavior, microstructure evolution, and wear performance of selective laser
27
melting processed commercially pure titanium. Acta Mater. 60 (2012) 3849-3860,
https://doi.org/10.1016/j.actamat.2012.04.006.
[21]
L. Thijs, K. Kempen, J.P. Kruth, J.V. Humbeeck, Fine-structured aluminium products with
controllable texture by selective laser melting of pre-alloyed AlSi10Mg powder. Acta Mater.
61 (2013) 18091819, http://dx.doi.org/10.1016/j.actamat.2012.11.052.
[22]
Y.J. Liu, Z. Liu, Y. Jiang, G.W. Wang, Y. Yang, L.C. Zhang, Gradient in microstructure and
mechanical property of selective laser melted AlSi10Mg. J. Alloy. Compd. 735 (2018) 1414-
1421, https://doi.org/10.1016/j.jallcom.2017.11.020.
[23]
K. Kempen, L.Thijs, J. Van Humbeeck, J.-P. Kruth, Mechanical properties of AlSi10Mg
produced by Selective Laser Melting, Physics Procedia 39 ( 2012 ) 439-446,
https://doi.org/10.1016/j.phpro.2012.10.059.
[24]
N.T. Aboulkhair, I. Maskery, C. Tuck, I. Ashcroft, N.M. Everitt, The microstructure and
mechanical properties of selectively laser melted AlSi10Mg: the effect of a conventional T6-
like heat treatment, Mater. Sci. Eng. A 667 (2016) 139146,
https://doi.org/10.1016/j.msea.2016.04.092.
[25]
R. Chou, A. Ghosh, S.C. Chou, M. Paliwal, M. Brochu, Microstructure and mechanical
properties of Al10SiMg fabricated by pulsed laser powder bed fusion, Mater. Sci. Eng. A
689 (2017) 5362, https://doi.org/10.1016/j.msea.2017.02.023.
[26]
Wu, X.Q. Wang, W. Wang, M.M. Attallah, M.H. Loretto, Microstructure and strength of
selectively laser melted AlSi10Mg. Acta Mater. 117 (2016) 311-320,
http://dx.doi.org/10.1016/j.actamat.2016.07.012.
[27]
B. Chen, S.K. Moon, X. Yao, G. Bi, J. Shen, J. Umeda, K. Kondoh, Strength and strain
hardening of a selective laser melted AlSi10Mg alloy. Scr. Mater. 141 (2017) 4549,
http://dx.doi.org/10.1016/j.scriptamat.2017.07.025.
[28]
X.P. Li, G. Ji, Z. Chen, A. Addad, Y. Wu, H.W. Wang, J. Vleugels, J. Van Humbeeck, J.P.
Kruth, Selective laser melting of nano-TiB2 decorated AlSi10Mg alloy with high fracture
28
strength and ductility. Acta Mater. 129 (2017) 183-193 ,
http://dx.doi.org/10.1016/j.actamat.2017.02.062.
[29]
E. Zaretsky, A. Stern, N. Frage, Dynamic response of AlSi10Mg alloy fabricated by selective
laser melting, Mater. Sci. Eng. A 688 (2017) 364370,
https://doi.org/10.1016/j.msea.2017.02.004.
[30]
A. Hadadzadeh, B. Shalchi Amirkhiz, J. Li, A. Odeshi, M. Mohammadi, Deformation
mechanism during dynamic loading of an additively manufactured AlSi10Mg_200C. Mater.
Sci. Eng. A 722 (2018) 263-268, https://doi.org/10.1016/j.msea.2018.03.014.
[31]
H. Asgari, C. Baxter, K. Hosseinkhani, M. Mohammadi, On microstructure and mechanical
properties of additively manufactured AlSi10Mg_200C using recycled powder. Mater. Sci.
Eng. A 707 (2017) 148158, https://doi.org/10.1016/j.msea.2017.09.041.
[32]
A.H. Maamoun, M. Elbestawi, G.K. Dosbaeva, S.C. Veldhuis, Thermal post-processing of
AlSi10Mg parts produced by Selective Laser Melting using recycled powder. Addit. Manuf.
21 (2018) 234247, https://doi.org/10.1016/j.addma.2018.03.014.
[33]
A. Strondl, O. Lyckfeldt, H. Brodin, U. Ackelid, Characterization and Control of Powder
Properties for Additive Manufacturing. JOM 67 (2015) 549-554,
https://doi.org/10.1007/s11837-015-1304-0.
[34]
H.P. Tang, M. Qian, N. Liu, X.Z. Zhang, G.Y. Yang, J. Wang, Effect of powder reuse times
on additive manufacturing of Ti-6Al-4V by selective electron beam melting, JOM 67 (2015)
555563, https://doi.org/10.1007/s11837-015-1300-4.
[35]
U. Tradowsky, J. White, R.M. Ward, N. Read, W. Reimers, M.M. Attallah, Selective laser
melting of AlSi10Mg: Influence of post-processing on the microstructural and tensile
properties development. Mater. Des. 105 (2016) 212-222,
http://dx.doi.org/10.1016/j.matdes.2016.05.066.
[36]
F. Del Re, V. Contaldi, A. Astarita, B. Palumbo, A. Squillace, P. Corrado, P. Di Petta,
Statistical approach for assessing the effect of powder reuse on the final quality of AlSi10Mg
29
parts produced by laser powder bed fusion additive manufacturing, Int. J. Adv. Manuf. Tech.
97 (2018) 2231-2240, https://doi.org/10.1007/s00170-018-2090-y.
[37]
A.Hadadzadeh, F.Mokdad, B. Shalchi Amirkhiz, M.A. Wells, B.W. Williams, D.L.Chen,
Bimodal grain microstructure development during hot compression of a cast-homogenized
Mg-Zn-Zr alloy. Mater. Sci. Eng. A 724 (2018) 421-430,
https://doi.org/10.1016/j.msea.2018.03.112.
[38]
E8/E8M 15a Standard Test Methods for Tensions Testing of Metallic Materials, ASTM,
2015.
[39]
M. Mohammadi, H. Asgari, Achieving low surface roughness AlSi10Mg 200C parts using
direct metal laser sintering, Addit. Manuf. 20 (2018) 2332,
https://doi.org/10.1016/j.addma.2017.12.012.
[40]
M. Fousová, D. Dvorský, A. Michalcová, D. Vojtěcha, Changes in the microstructure and
mechanical properties of additively manufactured AlSi10Mg alloy after exposure to elevated
temperatures. Mater. Charact. 137 (2018) 119-126 ,
https://doi.org/10.1016/j.matchar.2018.01.028.
[41]
A. Basak, S. Das, Epitaxy and Microstructure Evolution in Metal Additive Manufacturing.
Annu. Rev. Mater. Res. 46 (2016) 125149, https://doi.org/10.1146/annurev-matsci-
070115-031728.
[42]
L.E. Murr, S.M. Gaytan, D.A. Ramirez, E. Martinez, J. Hernandez, K.N. Amato, P.W.
Shindo, F.R. Medina, R.B. Wicker, Metal Fabrication by Additive Manufacturing Using
Laser and Electron Beam Melting Technologies. J. Mater. Sci. Technol 28 (2012) 1-14,
https://doi.org/10.1016/S1005-0302(12)60016-4.
[43]
N. Takata, H. Kodaira, K. Sekizawa, A. Suzuki, M. Kobashi, Change in microstructure of
selectively laser melted AlSi10Mg alloy with heat treatments, Mater. Sci. Eng. A 704 (2017)
218228, http://dx.doi.org/10.1016/j.msea.2017.08.029.
30
[44]
L. Liu, Q. Ding, Y. Zhong, J. Zou, J. Wu, Y.L. Chiu, J. Li, Z. Zhang, Q. Yu, Z. Shen,
Dislocation network in additive manufactured steel breaks strengthductility trade-off.
Mater. Today 21 (2018) 354-361, https://doi.org/10.1016/j.mattod.2017.11.004.
[45]
T. Malis, S.C. Cheng, R.F. Egerton, EELS log-ratio technique for specimen-thickness
measurement in the TEM., J. Electron Microsc. Tech. 8 (1988) 193200,
https://doi.org/10.1002/jemt.1060080206.
[46]
J. Pesicka, R. Kuzel, A. Dronhofer, G. Eggeler, The evolution of dislocation density during
heat treatment and creep of tempered martensite ferritic steels. Acta Mater. 51 (2003) 4847
4862, https://doi.org/10.1016/S1359-6454(03)00324-0.
[47]
D. Rojas, J. Garcia, O. Prat, L. Agudo, C. Carrasco, G. Sauthoff, a. R. Kaysser-Pyzalla,
Effect of processing parameters on the evolution of dislocation density and sub-grain size of
a 12%Cr heat resistant steel during creep at 650°C, Mater. Sci. Eng. A. 528 (2011) 1372
1381, https://doi.org/10.1016/j.msea.2010.10.028.
[48]
Y.L. Kuo, S. Horikawa, K. Kakehi, Effects of build direction and heat treatment on creep
properties of Ni-base superalloy built up by additive manufacturing. Scr. Materi. 129 (2017)
7478, http://dx.doi.org/10.1016/j.scriptamat.2016.10.035.
[49]
ASTM B85, Standard Specification for Aluminum-Alloy Die Castings, ASTM International,
West Conshohocken, PA, 2013.
[50]
J.G. Kaufman, Properties of Aluminum Alloys: Fatigue Data and the Effects of Temperature,
Product Form, and Processing, ASM International, Materials Park, Ohio, 2008, p. 15.
[51]
P. Bassani, B. Previtali, A. Tuissi, B. M. Vedani, G. Vimercati, S. Arnaboldi, Solidification
behaviour and microstructure of A360-SIC P cast composites. Metallurgical Science and
Technology 23 (2005) 3-10.
31
[52]
Z.-W. Chen, X.-L. Hao, J. Zhao, C.-Y. Ma, Kinetic nucleation of primary α(Al) dendrites in
Al-7%Si-Mg cast alloys with Ce and Sr additions. Trans. Nonferrous Met. Soc. China 23
(2013) 3561-3567, https://doi.org/10.1016/S1003-6326(13)62901-5.
[53]
D. Manfredi, F. Calignano, M. Krishnan, R. Canali, E. Ambrosio, E. Atzeni,From Powders
to Dense Metal Parts: Characterization of a Commercial AlSiMg Alloy Processed through
Direct Metal Laser Sintering. Materials 6 (2013) 856-869,
https://doi.org/10.3390/ma6030856.
[54]
L. HitzlerC. Janousch, J. Schanz, M. Merkel, B. Heine, F. Mack, W. Hall, A. Öchsner,
Direction and location dependency of selective laser melted AlSi10Mg specimens. J. Mater.
Process. Technol. 243 (2017) 48-61 , http://dx.doi.org/10.1016/j.jmatprotec.2016.11.029.
[55]
J. Wu, X.Q. Wang, W. Wang, M.M. Attallah, M.H. Loretto, Microstructure and strength of
selectively laser melted AlSi10Mg. Acta Mater. 117 (2016) 311-320,
http://dx.doi.org/10.1016/j.actamat.2016.07.012.
[56]
F. Mompiou, D. Caillard, M. Legros, H. Mughrabi, In situ TEM observations of reverse
dislocation motion upon unloading in tensile-deformed UFG aluminium. Acta Mater. 60
(2012) 3402-3414, https://doi.org/10.1016/j.actamat.2012.02.049.
[57]
G. E. Dieter, Mechanical Metallurgy. 2nd Ed. McGraw-Hill, 1961.
[58]
H. Mughrabi, Deformation-induced long-range internal stresses and lattice plane
misorientations and the role of geometrically necessary dislocations. Philos. Mag. 86 (2006)
4037-4054, https://doi.org/10.1080/14786430500509054.
[59]
A.J. Ardell, Precipitation Hardening. Metall. Trans. A 16 (1985) 2131-2165,
https://doi.org/10.1007/BF02670416.
[60]
T. Gladman, Precipitation hardening in metals. Mater. Sci. Technol. 15 (1999) 30-36,
https://doi.org/10.1179/026708399773002782.
32
[61]
S.S. Quek, Y. Xiang, D.J. Srolovitz, Loss of interface coherency around a misfitting spherical
inclusion. Acta Mater. 59 (2011) 5398-5410, https://doi.org/10.1016/j.actamat.2011.05.012.
[62]
Y. Xiang, D.J. Srolovitz, L.-T. Cheng, E. Weinan, Level set simulations of dislocation-
particle bypass mechanisms. Acta Mater. 52 (2004) 1745-1760 ,
https://doi.org/10.1016/j.actamat.2003.12.016.
[63]
N.T. Aboulkhair, C. Tuck, I. Ashcroft, I. Maskery, N.M. Everitt, On the Precipitation
Hardening of Selective Laser Melted AlSi10Mg. Metall. Mater. Trans. A 46 (2015) 3337-
3341, https://doi.org/10.1007/s11661-015-2980-7.
[64]
Y.M. Wang, T. Voisin, J.T. McKeown, J. Ye, N.P. Calta, Z. Li, Z. Zeng, Y. Zhang, W. Chen,
T.T. Roehling, R.T. Ott, M.K. Santala, P.J. Depond, M.J. Matthews, A.V. Hamza, T. Zhu,
Additively manufactured hierarchical stainless steels with high strength and ductility, Nature
Materials 17 (2018) 63-70. https://doi.org/10.1038/NMAT5021.
[65]
P. Rodriguez, Sixty years of dislocations. Bull. Mater. Sci. 19 (1996) 857 872,
https://doi.org/10.1007/BF02744623.
[66]
E. Ghassemali, M. Riestra, T. Bogdanoff, B.S. Kumar, S. Seifeddine, Hall-Petch equation in
a hypoeutectic Al-Si cast alloy: grain size vs. secondary dendrite arm spacing. Procedia
Engineering 207 (2017) 1924, https://doi.org/10.1016/j.proeng.2017.10.731.
[67]
J. Friedel, Dislocations, 70 Oxford, 1964, pp. 1524.
[68]
E.D. Cyr, A. Brahme, M. Mohammadi, R.K. Mishra, K. Inal, A new crystal plasticity
framework to simulate the large strain behaviour of aluminum alloys at warm temperatures.
Mater. Sci. Eng. A 727 (2018) 1128, https://doi.org/10.1016/j.msea.2018.04.020.
[69]
N. Hansen, HallPetch relation and boundary strengthening. Scr. Mater. 51 (2004) 801806,
https://doi.org/10.1016/j.scriptamat.2004.06.002.
33
[70]
S.K. Shaha, F. Czerwinski, D.L. Chen, W. Kasprzak, Dislocation slip distance during
compression of AlSiCuMg alloy with additions of TiZrV. Mater. Sci. Technol. 31
(2015) 63-72, https://doi.org/10.1179/1743284714Y.0000000606.
[71]
A. Simchi, Direct laser sintering of metal powders: Mechanism, kinetics and microstructural
features. Mater. Sci. Eng. A 428 (2006) 148-158,
https://doi.org/10.1016/j.msea.2006.04.117.
[72]
A.V. Gusarov, I. Yadroitsev, P. Bertrand, I. Smurov, Heat transfer modelling and stability
analysis of selective laser melting. Appl. Surf. Sci. 254 (2007) 975979,
https://doi.org/10.1016/j.apsusc.2007.08.074.
... where β is a material constant, M is the Taylor factor, G is shear modulus, b is the burgers vector of Al. Their values are given as follows, β = 0:16, M = 3:06, G = 26:5 GPa and b = 0.286 nm 23,53 . Orowan strengthening is also secondary as the morphology of dislocations are not those typical to Orowan strengthening, which would be manifested by dislocation loops encircling intra-granular precipitates. ...
Article
Full-text available
Light-weight, high-strength, aluminum (Al) alloys have widespread industrial applications. However, most commercially available high-strength Al alloys, like AA 7075, are not suitable for additive manufacturing due to their high susceptibility to solidification cracking. In this work, a custom Al alloy Al92Ti2Fe2Co2Ni2 is fabricated by selective laser melting. Heterogeneous nanoscale medium-entropy intermetallic lamella form in the as-printed Al alloy. Macroscale compression tests reveal a combination of high strength, over 700 MPa, and prominent plastic deformability. Micropillar compression tests display significant back stress in all regions, and certain regions have flow stresses exceeding 900 MPa. Post-deformation analyses reveal that, in addition to abundant dislocation activities in Al matrix, complex dislocation structures and stacking faults form in monoclinic Al9Co2 type brittle intermetallics. This study shows that proper introduction of heterogeneous microstructures and nanoscale medium entropy intermetallics offer an alternative solution to the design of ultrastrong, deformable Al alloys via additive manufacturing.
... where H 0 is the intrinsic hardness of the alloy, D is the average grain size, and the K value is the Hall-Petch coefficient ( Ref 27,28). The distance between Si particles is also essential in terms of strength. ...
Article
In this study, alternative solution heat treatments, applied at lower temperatures and shorter times, were applied to AlSI10Mg samples produced with the powder bed laser fusion technique. Thermocalc software was used to determine the temperatures of the NT6 solution heat treatment. The samples were analyzed by optical microscope, scanning electron microscope, energy-dispersive x-ray analysis, x-ray diffractometer and their mechanical behavior was examined. The results were compared with traditional T6 heat treatments and it was determined that the samples showed better ductility after NT6 heat treatments while maintaining the strength values in the traditional method.
Article
Full-text available
Metal‐based additive manufacturing can make complicated parts that are complex or expensive to cast and process. Rapid cooling rates increase laser powder bed fusion (LPBF's) mechanical properties during manufacturing. The objective of this study is to examine the impact of process parameters in the L‐PBF technique on the characteristics of microstructure and mechanical properties, specifically, on the nanohardness influenced by Si segregation. The microstructures of the produced specimens are examined using field‐emission scanning electron microscopy and the analysis identifies the existence of bimodal equiaxed α‐Al grains, accompanied by Si phases located within their grain boundaries. In addition, the solidified sample exhibits the segregation of secondary precipitates, particularly Mg2Si, which results in enhanced mechanical properties. Both cellular walls and Si precipitates impede the motion and generation of dislocations, thereby influencing the overall behavior of dislocations. The examination of segregation at the top layer is conducted in a comprehensive manner, subsequently using energy‐dispersive X‐ray spectroscopy for analysis. The presence of Mg2Si, Al2MgSi2, and other phases in all samples is confirmed through X‐ray diffraction. The as‐built samples’ residual stress under different process conditions is also investigated. Additionally, the obtained microstructure is compared to a phase‐field model to forecast the evolution of the microstructure.
Article
Full-text available
Hot deformation of a cast-homogenized ZK60 alloy was studied by compression at a temperature of 450 °C and a strain rate of 0.001 s⁻¹ to investigate microstructural evolution. The deformed microstructure was characterized using electron backscatter diffraction (EBSD) and high resolution transmission electron microscopy (HRTEM). EBSD observations of the deformed microstructure showed that hot deformation of this alloy resulted in a bimodal grain microstructure consisting of large pancaked unrecrystallized dendrites surrounded by recrystallized equiaxed fine grains. HRTEM studies revealed the presence of nano-(Zn-Zr)-precipitates in the deformed microstructure. Due to the coherency of precipitates/matrix, the dislocations were pinned by the nano-precipitates inside the unrecrystallized grains and the dislocation motion inside the grains was impeded, hence, a substructure evolved. Consequently, dynamic recrystallization (DRX) was suppressed and deformation was concentrated inside the DRXed region.
Article
Full-text available
The performance enhancement of parts produced using Selective Laser Melting (SLM) is an important goal for various industrial applications. In order to achieve this goal, obtaining a homogeneous microstructure and eliminating material defects within the fabricated parts are important research issues. The objective of this experimental study is to evaluate the effect of thermal post-processing of AlSi10Mg parts, using recycled powder, with the aim of improving the microstructure homogeneity of the as-built parts. This work is essential for the cost-effective additive manufacturing (AM) of metal optics and optomechanical systems. To achieve this goal, a full characterization of fresh and recycled powder was performed, in addition to a microstructure assessment of the as-built fabricated samples. Annealing, solution heat treatment (SHT) and T6 heat treatment (T6 HT) were applied under different processing conditions. The results demonstrated an improvement in microstructure homogeneity after thermal post-processing under specific conditions of SHT and T6 HT. A micro-hardness map was developed to assist in the selection of the optimized post-processing parameters in order to satisfy the design requirements of the part.
Article
During solidification of many so-called high-performance engineering alloys, such as 6000 and 7000 series aluminum alloys, which are also unweldable autogenously, volumetric solidification shrinkage and thermal contraction produces voids and cracks. During additive manufacturing processing, these defects can span the length of columnar grains, as well as intergranular regions. In this research, laser powder bed fusion (LPBF) of aluminum alloy (AA) 6061 used powder bed heating at 500 °C in combination with other experimentally determined processing parameters to produce crack-free components. In addition, melt-pool banding, which is a normal solidification feature in LPBF, was eliminated, illustrating solidification process modification as a consequence of powder bed heating. Corresponding microindentation hardness and tensile testing of the as-fabricated AA6061 components indicated an average Vickers hardness of HV 54, and tensile yield, ultimate strength, and elongation values of 60 MPa, 130 MPa, and 15%, respectively. These mechanical properties and those of heat treated parts showed values comparable to annealed and T6 heat treated wrought products, respectively. X-ray diffraction and optical microscopy revealed columnar grain growth in the build direction with the as-fabricated, powder-bed heated product microstructure characterized by [100] textured, elongated grains (∼ 25 μm wide by 400 μm in length), and both intragranular and intergranular, noncoherent Al-Si-O precipitates which did not contribute significantly to the mechanical properties. The results of this study are indicative that powder bed heating may be used to assist with successful fabrication of AA6061 and other alloy systems susceptible to additive manufacturing solidification cracking.
Article
In recent years, the SLM process has been studied for the production of aluminum alloy parts, as these alloys demonstrate significant potential for the future, notably due to their low density which allows a considerable reduction in mass. The aim of this bibliographical study is to identify and classify the parameters and phenomena which influence the appearance of defects in aluminum alloy parts produced using the SLM process and hence the final properties of these parts. To do this, a cause tree diagram was created. For each defect or consequence identified (porosities, defects linked with hot cracking phenomena, anisotropy in the material and surface quality), we revealed the potential sources of the appearance of this defect, going back to the initial causes.
Article
To improve metal formability, high temperature forming has become a desired manufacturing process. Phenomenological plasticity models are widely used in this application, however lack good predictive capability concerning microstructure evolution during forming. Many crystal plasticity hardening models have been developed to predict deformation phenomena of metals during high temperature forming; however, few have thermodynamic self-hardening formulations based on deformation mechanisms. This work presents a crystal plasticity based analysis with a Taylor polycrystal averaging scheme of warm forming employing a new microstructure and dislocation based strain hardening model to simulate deformation behaviour. The hardening model is derived from energy balance between dislocation storage, dislocation accumulation, and dislocation recovery, based on remobilization of immobile dislocations, due to thermal activation. The constitutive formulation is extended to include alloying effects due to solute strengthening of Mg. The material hardening properties of AA5754 are characterized for a range of temperatures at constant strain-rates. A formulation for the kinematics of dynamic strain aging is presented and employed for room-temperature simulations. The hardening characterization is then used to predict stress-strain behaviour of AA5182 for similar conditions. The model shows excellent predictability of experimental results. An analysis on the microstructural connection between temperature and stress-strain response is presented.
Article
Direct metal laser sintering (DMLS) is an additive manufacturing technique that creates near-net-shape functional components by selectively melting metal powders in two dimensions layer by layer using a high power laser as a heat source. This technique offers to create parts with complex net-shape structures at an affordable cost with the least lead time. The main purpose of this study is to investigate the corrosion behavior and microstructure of AlSi10Mg_200C manufactured using DMLS compared with its die cast counterpart (A360.1 die cast Al alloy). The impact of the alloy's surface finish, i.e. as-printed surface versus as-ground one, on the corrosion performance was also investigated. Several AlSi10Mg_200C cube samples were additively manufactured through DMLS technique. In addition, the same size cubes were cut from an aluminum A360.1 die cast ingot. The corrosion behavior of the two alloys was analyzed utilizing potentiodynamic polarization testing and electrochemical impedance spectroscopy in aerated 3.5 wt.% NaCl solution to mimic sea water environment at 25 °C. Further, the microstructures and composition of the samples before and after corrosion testing were investigated using Optical Microscopy (OM), Scanning Electron Microscopy (SEM), and Energy Dispersive X-ray (EDX) spectroscopy. The results confirmed that the corrosion resistance of the alloy processed through DMLS was significantly better than the cast counterpart. This was attributed to the fine microstructure produced by DMLS, uniform distribution of the fine Si particles without formation of any intermetallic, due to the extremely rapid cooling and solidification rate during DMLS process and slightly lower Fe and Cu concentration of the AlSi10Mg alloy. In contrast, the A360.1 cast Al alloy samples experienced severe localized corrosion of the Al matrix in the periphery of the Fe containing IMC and Si flakes. The results also highlighted improved corrosion resistance of the as-printed DMLS sample compared with that of the as-ground one.
Article
The additive manufacture (AM) of the AlSi10Mg alloy has become the subject of considerable attention, especially for production of complex parts in engines. Because such parts can be exposed to elevated temperatures during operation, material stability is very important, but as yet little is known about it in relation to AM. Here, we studied changes of the AlSi10Mg alloy produced by selective laser melting (SLM) after its exposure to temperatures between 120 and 180 °C. At each temperature, hardness evolution was measured, with hardness increasing over time. The maximum hardness state obtained at 160 °C was selected for further studies comprising microstructural analysis by scanning and transmission electron microscopy, chemical composition analysis and mechanical properties assessment. Transmission microscopy revealed nano-scale acicular precipitates that caused a slight increase in the yield strength of the alloy together with a significant drop in elongation. Electron energy loss spectroscopy (EELS) and energy dispersive spectroscopy (EDS) showed that the precipitates surprisingly consisted of pure Si. To provide a comparison, conventional regimes of heat treatment (stress-relief and T6) were applied. Despite a considerable loss in mechanical performance, thermal instability was no longer observed. Overall, our results indicate that operating temperatures are a key factor in ensuring the smooth operation of AM parts of the AlSi10Mg alloy. In respect to that, we offer recommendations for their industrial use.
Article
JOURNAL: Additive Manufacturing // Abstract: Laser sintered aluminum alloys produced by metal 3D printers can replace cast aluminum alloys in aerospace, defense, and marine industries by offering better mechanical properties, less porosity, and competitive fatigue characteristics. One of the major issues currently is the considerable surface roughness of additively manufactured aluminum alloys demanding post-processing procedures such as bid blasting or machining. In the current study, the process parameters such as laser power, scan speed, and hatch spacing were altered such that better surface roughness could be achieved for AlSi10Mg_200C using a Direct Metal Laser Sintering (DMLS) system. The process parameters were chosen such that three samples with the same core properties but different upskin characteristics were produced. The achieved surface roughness of the additively manufactured aluminum samples were almost as low as one fifth of the regular samples manufactured using standard process parameters. The microstructure and the porosity level of the samples printed by different process parameters were studied to reveal the causality of the low surface roughness for the proposed process.
Article
Quality control in parts built layer-by-layer via Additive Manufacturing (AM) process is still a challenge since critical features, such as the control in porosity, residual stress and deformation, surface roughness and microstructure, are yet to be fully addressed. In this work, we develop a phase field model to simulate powder bed-based AM, focusing on the effects of two major process parameters, the beam power and scanning speed, on the melt pool size and shape, porosity and grain structure. The model reproduces many important phenomena observed experimentally and reveals scaling relations for the depth and length of melt pool, the porosity and the grain density on process parameters. We find critical power densities below which the grain density or the porosity increases rapidly and the grain structure is controlled by different mechanisms in different power density regimes which allows for the possibility of controlling the grain structure. The present work could serve as a useful reference for accurate control of defects and microstructure in AM build.