Sunteți pe pagina 1din 11

Additive Manufacturing 25 (2019) 286–296

Contents lists available at ScienceDirect

Additive Manufacturing
journal homepage: www.elsevier.com/locate/addma

Full Length Article

Characterisation of porosity, density, and microstructure of directed energy T


deposited stainless steel AISI 316L

Zhi’En Eddie Tana,b, John Hock Lye Panga, , Jacek Kaminskib, Helene Pepinb
a
Nanyang Technological University, School of Mechanical and Aerospace Engineering, 50 Nanyang Avenue, 639798, Singapore
b
Advanced Remanufacturing and Technology Centre, 3 CleanTech Loop, CleanTech Two, 637143, Singapore

A R T I C LE I N FO A B S T R A C T

Keywords: Directed Energy Deposition (DED) was used to form a Stainless Steel AISI 316 L steel block component on a Mild
Additive manufacturing Steel S235JR substrate. Porosity, density, and defect were characterised at 4 localities within the DED compo-
Laser metal deposition nent by microscopy and x-ray tomography. Three-dimensional (3D) reconstruction of the x-ray tomographic
Directed energy deposition image sequences focused at select porosities is presented. The element composition and Vickers microhardness
Stainless steel 316L
measurements were taken at the fusion lines and track body locations to characterise the differences in materials
Porosity
and mechanical properties at the 2 locations. Lastly, an element mapping analysis was conducted to determine
Density
Microstructure the solidification mode for the DED component. Sources for defects were proposed based on the characteristics of
the porosity analysis and conclusions were made about the solidification behaviour of the DED component.

1. Introduction the deposited track [3]. Antony et al. found that process parameters like
laser speed, beam size, and laser power on the deposition geometry
Laser Metal Deposition (LMD) is a subset of Directed Energy have a significant role to play in determining the deposition geometry
Deposition (DED) method of Additive Manufacturing (AM) technology and profile [4]. A. Yadollahl et al. studied the effects of inter-layer time
that has numerous industry applications, including: application of interval and heat treatment on LMD Stainless Steel AISI 316 L part
coatings, repair, and 3D build ups [1,2]. LMD uses a combination of quality. Coarse columnar grains were observed toward the boundary
technologies: computer-aided design (CAD), computer-aided manu- region, while fine equiaxed grains were observed toward the centre
facturing (CAM), lasers, powder metallurgy, and sensors, in its process. region [5]. Thijs et al. identified the thermal gradient and direction of
LMD is considered beneficial in medium-to-large sized, small batch cooling to be key process parameters that influence the microstructural
manufacturing by employing a near net shape manufacturing process. evolution in laser based additive manufacturing [6]. Ma et al. studied
In contrast to Selective Laser Melting (SLM), which employs a powder the effect of higher scanning speeds and solidification rates, and their
bed method whereby a laser selectively melts a bed of metal powder respective effects on the homogeneity and finer microstructure evolu-
feed, LMD directly feeds the metal powder material to a target location tion [7]. Song et al. associated the mechanical strength of laser based
while using a focused laser beam to melt the powder feed. In LMD, a additive manufactured parts to the fine grain microstructure that is a
Computer Numerical Control (CNC) system is used to control the laser result from the rapid cooling rates [8]. Wang et al. found that the laser
beam path from the CAM model input to manufacture a three-dimen- energy density used in the process has a strong influence over the grain
sional component build layer-by-layer. Control of parameter input such size and the part density, which hence affects its mechanical properties
as: laser spot size, laser power, scanning speed and powder feed rate, [9]. Bartolomeu et al. correlated the differences in the microstructure of
are crucial in the control of the build’s dimensional geometry and AM parts and conventionally manufactured parts with the differences in
properties. mechanical properties like hardness and wear properties [10]. Abe
Yadriotsev et al. studied the influence of the process parameters like et al. correlated the large size and amount of porosities to the lower
laser scanning speed and preheating temperature on the geometry and ductility in AM parts [11].
microstructure of a single-track metal deposition of Stainless Steel AISI Detrimental defects in LMD-created structures such as foreign par-
316 L. The preheating was found to control the contact angle and ticles, porosity, interface delamination, have been proven to be crucial
height, and scanning speed controlled the contact angle and height of contributors to a build’s propensity to rupture, by acting as stress


Corresponding author.
E-mail address: mhlpang@ntu.edu.sg (J.H.L. Pang).

https://doi.org/10.1016/j.addma.2018.11.014
Received 3 October 2017; Received in revised form 12 November 2018; Accepted 14 November 2018
Available online 16 November 2018
2214-8604/ © 2018 Elsevier B.V. All rights reserved.
Z.E. Tan et al. Additive Manufacturing 25 (2019) 286–296

concentrators and crack growth locations. For example, porosities can Table 1
affect the overall build quality in LMD built components as the con- Build parameters employed in LMD process.
joining of adjacent discontinuations have a severe effect on mechanical Parameter Value
properties like tensile, fatigue, and fracture strengths [12–14]. Li et al.
found that by increasing the scan speeds, larger pores are observed in 1. Laser Spot Diameter 3 mm
2. Laser Power 1.4 kW
the structure [15]. Stainless Steel AISI 316 L’s mechanical properties are
3. Deposition Rate 14 g/min
largely dependent on the grain boundary strengthening effect, where 4. Scanning Speed 1000 mm/min
the grain sizes have a direct effect on the yield strength of the material. 5. Track Overlap 60%
The rate of cooling of the melt pool plays a significant role in the size of
the grain growth. A higher cooling rate generally produces a smaller
grain size [7]. The LMD process has a smaller melt pool size, high 2.2. Build evaluation
temperature gradients, and a rapid cooling rate, compared to tradi-
tional manufacturing techniques like casting [16]. The cross-sectional morphology of the specimen is chemically
The objective of this study is to investigate and characterize the etched, and the deposition track widths were measured by microscopy
porosity, microstructural and mechanical differences between the fu- as displayed in Fig. 3. All track widths were measured for each visible
sion lines and track bodies of an LMD built Stainless Steel AISI 316 L layer at the Top Zone: YZ-Plane, Bottom Zone: YZ-Plane, Top Zone: XZ-
component. Microscopy and X-ray tomography were employed to study Plane, Bottom Zone: XZ-Plane cross-sections.
the characteristics of porosity population in four internal target zones of
the build component in alternate X–Y and Top–Bottom zones. Vickers 2.3. Porosity and density analysis
microhardness, SEM and EDX methods were used to characterise the
material and mechanical differences between the fusion lines and de- These specimens were hot mounted in resin under heat and high
position body locations. SEM and EDX were also used for the element pressure, and polished for density and porosity analysis by microscopy.
mapping of track body site within the LMD component. Observations Microscope images were captured, stitched, and analysed using a Zeiss
about the solidification mode were provided. Light Microscope as seen in Fig. 4. Specimen density was measured by
thresholding the pores detected in the microscope images and com-
paring the pores area with the total measured area of the sample. This
2. Experimental procedure measurement was done in Zeiss software and porosity characteristics
were analysed using ImageJ software. X-ray tomography measurements
2.1. Laser metal deposition process were used to capture porosities within the LMD block using a Nikon XT
H 225 with a 14 μm/pixel voxel size, 210 kV voltage, 72 μA current,
Stainless Steel 316 L powder feed material with a 45 μm – 105 μm 1 mm filter, and 3141 projections. The image sequences were subse-
particle size range was used as feedstock material. LaserTec65 3D hy- quently processed and analysed using Fiji software.
brid was used to fabricate the LMD samples. An SEM image of the
powder feed is shown in Fig. 1. The build parameters employed in this 2.4. Microstructure and morphology characterisation
build are specified in Table 1.
A diode laser was used as a laser source, and argon gas as the Specimens were chemically etched using an etchant solution of 5%
shielding gas. A bidirectional deposition pattern (raster scan) and bi- nitric acid, 5% hydrochloric acid, 90% ethanol composition (% vol.).
directional layer orientation (90° rotation per layer) build strategy was Micrographs were captured and stitched using Zeiss Microscope Stemi
used to additively manufacture a Stainless Steel AISI 316 L block 2000-C. SEM images were captured using Zeiss EVO HD25 SEM
component of 30 mm × 50 mm × 50 mm dimensions, built on a Mild equipment, using the SE mode, at EHT 15 kV, and analysed with Inca
Steel S235JR substrate of 100 mm × 100 mm × 12 mm dimensions. software. Vickers microhardness tests were conducted on the samples
The specimen was subsequently sectioned using a bench saw, extracting using an Innova Test Falcon 500 microhardness tester. A load of 500 g
4 specimens at 4 target surface locations: Top Zone: YZ-Plane, Top and a 15 s dwell time were used.
Zone: XZ-Plane, Bottom Zone: YZ-Plane, Bottom Zone: XZ-Plane as il-
lustrated in Fig. 2(a). The bidirectional raster scan build strategy is 3. Results and discussion
shown in Fig. 2(b).
3.1. Build evaluation

In order to evaluate the LMD Stainless Steel AISI 316 L build, track
widths were measured across the Top Zone: YZ-Plane, Bottom Zone: YZ-
Plane, Top Zone: XZ-Plane, Bottom Zone: XZ-Plane cross-sections as
designated in Fig. 2(a). 10 track width measurements were taken for
each of the 8 measurement layers per 4 specimen cross-sections. The
mean track width measured for each of measurement layers is displayed
in Fig. 5. The Top Zone and Bottom Zone’s mean and standard deviation
track width measurement values are collated in Table 2.
From Fig. 5 and Table 2, the Bottom Zone locations of the specimen
exhibited a lower track width mean value than the Top Zone locations.
This is largely due to the differences in the thermal gradient and cooling
rates between the Bottom Zone and Top Zone. A smaller thermal gra-
dient and a faster cooling rate at the start of the process compared to
toward the end results in a larger melt pool size toward the Top Zone of
the LMD Stainless Steel AISI 316 L specimen. This is consistent with the
larger track width size measured at the Top Zone compared to the
Fig. 1. SEM image of Stainless Steel 316 L powder feed material used in LMD Bottom Zone zones. Manvatkar et al. developed a transient heat transfer
process. and fluid flow model for laser assisted deposition of stainless steel, and

287
Z.E. Tan et al. Additive Manufacturing 25 (2019) 286–296

Fig. 2. (a) LMD Stainless Steel AISI 316 L build dimensions and specimen section locations, (b) bidirectional scanning path strategy.

Fig. 3. Track width measurements for cross section micrograph of specimen.

found that the melt pool dimensions increases while the cooling rate bottom zone measurement areas.
decreases toward the upper layers [17]. Aggarangsi et al. found that in A difference in top and bottom zone measurements for part density
order to control for the track width size consistency within a build, the and porosity sizes can be attributed to the dissimilar rate of melting and
process scaling of laser powers need to be monitored and controlled solidification experienced at the two zones. A lower object count was
[18]. observed at the bottom zone where larger pore sizes were detected,
compared to a higher object count observed at the top zone where
3.2. Density and porosity evaluation smaller pore sizes were detected. X-ray tomography was used to char-
acterise the porosities within the LMD part. 3D tomographic re-
The LMD Stainless Steel AISI 316 L’s object count distribution per constructions of the pores located at the first deposition layer and other
porosity size class at 4 internal target measurement areas are compiled interlayer porosities are given in Fig. 8. The reconstruction is threshold
in Fig. 6, and overall density and object count per target measurement into solid and pore phases, indicated in black and white respectively.
area is consolidated in Table 3. Most of the objects detected con- A method to narrow down the possible sources of defect for the
centrated within the Class 1 (0.0 μm – 20. 0 μm) and Class 2 porosity measurement is to characterise the form of the detected ob-
(21.0 μm – 40.0 μm) range. The four target zones exhibited a high de- jects. By determining the pore form and size, an inference of the source
gree of part density of more than 99.80%. Top zone measurement areas of defect can be made. There are 2 quantifiable characteristics for the
exhibited a slightly higher degree of part density, and a larger object sphericity measurement of objects measured in a 2D plane like micro-
count distribution of small pore sizes, and larger object count, than scopy: circularity and roundness according to ImageJ [19]. Circularity

288
Z.E. Tan et al. Additive Manufacturing 25 (2019) 286–296

Fig. 4. Stitched microscopy images of cross-sectioned LMD Stainless Steel AISI 316 L (a) Top Zone: YZ-Plane, (b) Top Zone: XZ-Plane, (c) Bottom Zone: YZ-Plane, (d)
Bottom Zone: XZ-Plane.

takes into account the object’s form and smoothness, while roundness originate from defects like, moisture evaporation or local voids after
takes into account the object’s ovality and aspect ratio. The closer these powder-layer deposition. For objects that exhibit strong degrees of
2 characteristic values are to 1.0, the greater the resemblance to a circularity and low degrees of roundness, a possible source of these
sphere. Each detected pore was measured for its roundness and circu- fabrication problems are bonding defects due to incomplete melting.
larity based on Eqs. (1) and (2). Results of the findings are graphed and Since these 2 categories form the bulk of the object population, it can be
compiled in Fig. 7. inferred that the defects associated with these characteristics are
dominant in this LMD Stainless Steel AISI 316 L build. Monroy et al.
Area
Circularity = 4π × studied the porosity features and correlated round pores with smooth
Perimeter 2 (1)
features with gas entrapments [20]. Hu et al. found that when the scan
velocity is high, gaseous bubbles get trapped within the melt pool and
Area
Roundness= 4 × settle to form round pores, and that poor interlayer bonding produces
π × Major Axis2 (2) pore defects that are more irregular in form [21]. Ziolkowski et al., in
For objects that exhibit both strong degrees of circularity and their porosity study, concluded that by diagnosing the porosities’ forms
roundness, there is a strong likelihood that these are voids that is useful in determining the course of cracks and pin point locations for

289
Z.E. Tan et al. Additive Manufacturing 25 (2019) 286–296

Fig. 5. Mean track width measurement for Bottom Zone and Top Zone locations.

Table 2 Table 3
Track width measurements. Density and object count detection per measurement area.
Measurement Zone Mean Standard Deviation Measurement Area Object Count Density

Bottom Zone 3.22 mm 0.24 mm 1. Top Zone: XZ-Plane 730 99.87%


Top Zone 3.57 mm 0.26 mm 2. Top Zone: YZ-Plane 1313 99.91%
3. Bottom Zone: XZ-Plane 220 99.80%
4. Bottom Zone: YZ-Plane 101 99.80%
crack initiation sites, especially in fatigue [22].
The porosities measured by X-ray tomography at the first deposition
layer are distinctly identifiable by the irregular profile, as seen from the measurement of smaller sized pores due to the spatial resolution of
Fig. 8(a). The cooler substrate at the deposition initiation result in a the X-ray tomography [24].
lack of fusion, producing a lower part density measurement toward the
first layer. Other large interlayer porosities and located toward the
3.3. Microstructure and composition analysis
middle of the block were also measured, as seen from Fig. 8(b). These
large pores exhibit a flat and spheroid profile and are > 150 μm in
An optical micrograph of the microstructure morphology is shown
width and can be attributed to the interlayer air entrapments that result
in Fig. 9, where the transition points between the fusion line and track
in pores of this specific form. Smaller porosities are not able to be
body locations are indicated. The microstructure morphology primarily
captured with this technique due to the spatial resolution limitations of
comprises of columnar grains that grow epitaxially from the base of the
X-ray tomography. Maskery et al. found that larger pores measured by
LMD build. This is due to the direction of heat flow through previously
X-ray tomography, exhibited a similar disc-like profile, with aspect
deposited layers and the substrate that results from the layer-by-layer
ratios extending to high values, and proposed that the layer-by-layer
deposition. At the fusion line locations, the epitaxially grown columnar
manufacture of AM parts is responsible for the large disc-like pore
grains transit to a fine equiaxed grain growth, where cooling rates are
profiles observed [23]. Van Bael et al. also cited similar limitations in
expected to be highest. The rapid flow of heat out of the melt zone into

Fig. 6. Object count distribution per pore size class for each target zone surface.

290
Z.E. Tan et al. Additive Manufacturing 25 (2019) 286–296

Fig. 7. Object count distribution for (a) circularity and (b) roundness.

Fig. 8. X-ray tomography (a) image slice at first deposition layer, (b) 3D image sequence reconstruction of porosities at first deposition layer, (b) image slice of large
interlayer porosities, (d) 3D image sequence reconstruction of large interlayer porosities.

surround metal creates a thin layer of fine equiaxed grains at the fusion to the Hall-Petch relation:
line locations that trail the laser spot scan movement.
Vickers microhardness test was conducted to characterise the me- k
σy = σ0 +
chanical differences at the fusion line and track body locations. Images d (3)
of some of the indent locations are shown in Fig. 10. 16 indentation
points were measured at the fusion line locations, and another 16 in- Where σy is the yield strength, σ0 is a materials constant that governs the
dentation points were measured at the track body locations. Results of stresses for grain dislocation movement, k is the Hall-Petch slope
the microhardness measurements for the fusion line and deposition coefficient, and d is the mean grain size of the material. The Hall-Petch
body are shown in Fig. 11. relates the material’s yield strength with its grain size where a smaller
The fusion line locations exhibited a higher Vickers hardness than grain size would produce a higher yield strength and hardness. The
the track body locations as seen in Fig. 11. The mean microhardness higher microhardness measured of the fine equiaxed grains at the fusion
value at the track body and fusion locations were measured to be line locations is reciprocal of the square root of the grain width.
182.24 HV and 187.10 HV respectively. This observation is attributed In order to correlate the grain sizes with the microhardness mea-
surements, a Hall-Petch plot was used to chart the 2 sets of data as seen

291
Z.E. Tan et al. Additive Manufacturing 25 (2019) 286–296

Fig. 9. Micrograph of microstructure morphology within LMD build. Dashed lines indicate the transition between track body and fusion line locations.

in Fig. 14. From the Hall-Petch plot, the σ0 and k constants are ap-
proximated by applying a linear fit and using the general relationship
between hardness and yield stress [25]:

Δσy = 3.03 ΔHv (4)

Where Hv is Vickers hardness. The σ0 and k constants of at the fusion


lines are hence approximated to values of σ0 = 534.63 MPa and
k = 345.40 MPa and the track body to values of σ0 = 494.76 MPa and
k = 334.89 MPa. These σ0 and k constants are consistent with stainless
steel’s Hall-Petch data [26]. Slight differences observed between the 2
sets of data can be attributed to the uneven distribution and variation in
the grain sizes morphologies, typical of an additively manufactured
part. Huang et al. found that the hardness distribution across a section
of a AM part produced a periodical fluctuation in the hardness values Fig. 11. Vickers microhardness measurements for track body and fusion line.
that was attributed to the bonding surface locations [27].
SEM images were taken at the fusion lines and track body locations
carbon and oxygen is possible mainly because of a Super Atmospheric
and the areas element constituents were measured by EDX as seen in
Thin Window (SATW) configuration, which has a greater sensitivity to
Fig. 12. The measured element composition for the target track body
lighter elements.
and fusion line areas is shown in Fig. 13. The elemental mapping of

Fig. 10. Vickers microhardness indentation locations at (a) fusion line, (b) track body.

292
Z.E. Tan et al. Additive Manufacturing 25 (2019) 286–296

Fig. 12. EDX measurement sites at (1) fusion line and (2) track body.

The element Oxygen (O) was observed in the characteristic X-ray points were characterised for their elemental compositions, 3 at the
peaks for the fusion line location, not seen at track body location. columnar boundary and 3 at the columnar body, as seen in Molybde-
During the deposition process, an oxide layer forms at the surface of the num’s element map Fig. 15(b). The element compositions are compiled
deposition track, occurring as each LMD track is deposited. Sporadic in Table 4.
oxide deposits can also be observed within track body locations. This is The element compositions were measured and calculated for their
due to same reaction occurring at the powder feed particle level during Nickel and Chromium equivalents using Eqs. (5) and (6), and their
the deposition process. ferrite composition was derived using Schaeffler’s diagram as seen in
An EDX element mapping of a deposition track body location was Fig. 16a. [28]. From Schaeffler’s diagram, the target columnar
conducted and is displayed in Fig. 15, with (a) showing the layered boundary area has a δ-ferrite composition of approximately 80%, while
element image, and (b) – (h) showing the detected elemental con- the columnar body area has a wholly austenitic composition. This
stituents. From this measurement, columnar microstructures were ob- shows that the columnar boundary area composition is primarily δ-
served in Molybdenum, Chromium, and Iron element maps. To under- ferritic, while the columnar body area is austenitic. From the overall
stand the composition of these columnar microstructures, 6 target scanned area element composition, the Creq/Nieq ratio was derived at

Fig. 13. Characteristic X-ray peaks for (a) fusion line and (b) track body sites.

293
Z.E. Tan et al. Additive Manufacturing 25 (2019) 286–296

−½
Fig. 14. Vickers microhardness versus d , where d is the mean grain width size adjacent to microindentation location.

Fig. 15. Element mapping for deposition body area: (a) Layered element image, (b) Molybdenum and target body, boundary element composition points, (c)
Chromium, (d) Silicon, (e) Iron, (f) Nickel, (g) Carbon, (h) Manganese.

Table 4
Element composition for target (a) white columnar boundary (b) columnar body areas (c) overall scanned area.
(a) Element 1 2 3 Mean wt% (b) Element 4 5 6 Mean wt% (c) Element Overall
wt%

C 0.0 0.0 0.0 0.0 C 0.1 0.1 0.0 0.1 C 0.1


Si 0.5 0.5 0.4 0.5 Si 0.5 0.5 0.5 0.5 Si 0.9
Cr 26.1 25.7 25.3 25.7 Cr 17.2 16.4 16.2 16.6 Cr 16.6
Mn 1.3 1.3 1.6 1.4 Mn 1.5 1.3 1.4 1.4 Mn 1.6
Fe 59.6 58.6 59.7 59.3 Fe 69.3 70.2 70.5 70.0 Fe 69.3
Ni 10.3 11.7 10.7 10.9 Ni 10.0 10.4 10.0 10.1 Ni 10.8
Mo 2.3 2.1 2.2 2.2 Mo 1.4 1.3 1.4 1.4 Mo 0.9
Nieq : 11.6 Nieq : 13.8 Nieq : 14.6
Creq: 28.7 Creq: 18.6 Creq: 18.8

1.3, and hence falls within the upper range of the Austenitic-Ferritic formed at the sub-grain boundaries as a result of its enrichment of
(AF) solidification mode, as determined in Table 4 [29,30]. In an AF chromium and molybdenum, and the impoverishment of nickel [32]
solidification mode, austenitic steels form a microstructure con- (Table 5).
stituency of predominantly austenite with columnar δ-ferrite forma-
Ni eq = Ni + (30 × C) + (0.5 × Mn) (5)
tions, along the solidification direction as seen in Fig. 16b. This is
consistent with the experimental findings from Fig. 15 where columnar Creq = Cr + Mo + (1.5 × Si) + (0.5 × Nb) (6)
δ-ferrite formations were observed, parallel to the deposition direction.
Milton et al. investigated laser additive manufacture of Stainless
Steel AISI 316 L and found that Creq and Nieq contents fell within the 4. Conclusion
5–10% δ-ferrite range and is consistent with the findings of this study
[31]. Michal et al. studied the Laser Engineered Net Shaping (LENS) A Stainless Steel AISI 316 L block was built by LMD. The part’s
manufacture of Stainless Steel AISI 316 L and found that δ-ferrite and porosity, density, defect, and microstructure were characterised by
austenite formations were present, and that the intercellular δ-ferrite microscopy, X-ray tomography, Vickers microhardness, SEM, and EDX

294
Z.E. Tan et al. Additive Manufacturing 25 (2019) 286–296

Fig. 16. (a) Schaeffler's diagram [33] and (b) Austenitic steel solidification diagram [34].

Table 5 were observed in the Molybdenum, Chromium, and Iron element


Solidification mode identification. maps. This observation is the result of an AF solidification mode that
Solidification Mode Creq/Nieq ratio
leads to the formation of columnar δ-ferrites parallel to the solidi-
fication direction.
Ferritic (F) Mode 1.95 ≤ Creq/Nieq
Ferritic-Austenitic (FA) Mode 1.48 ≤ Creq/Nieq ≤ 1.95 Acknowledgements
Austenitic (A), Austenitic-Ferritic (AF) Mode Creq/Nieq ≤ 1.14

Collaborative work between Nanyang Technological University


methods. Conclusions about the likely sources for defects, fusion line (NTU), School of Mechanical and Aerospace Engineering (MAE) and
and track body compositions, and austenitic and ferritic characterisa- Agency for Science, Technology, and Research (A*STAR), Advanced
tions are offered. Remanufacturing and Technology Centre (ARTC). The authors would
also like to express gratitude for the technical input and support from
1 Track widths toward the Top Zone of the specimen were measured DMG MORI.
to be larger than the Bottom Zone locations. This is due to higher in
the thermal gradient and lower cooling rates at the Top Zone that References
result in the larger melt pool and track width sizes compared to the
[1] L. Thivillon, et al., Potential of direct metal deposition technology for manu-
Bottom Zone location.
facturing thick functionally graded coatings and parts for manufacturing thick
2 The majority of pores detected fell within the 0.0 μm – 20.9 μm functionally graded coatings and parts for reactors components, J. Nucl. Mater. 385
range, and part densities were measured to be 99.8% and above. A (2009) 236–241.
larger population of small pore sizes, and slightly higher part den- [2] S. Nowotny, et al., Laser beam build-up welding: precision in repair, surface clad-
ding, and direct 3D metal deposition, J. Therm. Spray Technol. 16 (3) (2007) 344.
sities were measured in the top zone area than the bottom zone area. [3] I. Yadroitsev, et al., Energy input effect on morphology and microstructure of se-
To achieve a more consistent build between the bottom and top lective laser melting single track from metallic powder, J. Mater. Process. Technol.
zones, further optimisation can be employed in the dynamic control 213 (4) (2013) 606–613.
[4] K. Antony, N. Arivazhagan, K. Senthilkumaran, Numerical and experimental in-
of laser power input, or a pre-heating to the substrate prior to the vestigations on laser melting of stainless steel 316L metal powders, J. Manuf.
build can be applied Process. 16 (3) (2014) 345–355.
3 Most of the pores detected exhibited a high degree of circularity and [5] A. Yadollahi, et al., Effects of process time interval and heat treatment on the
mechanical and microstructural properties of direct laser deposited 316L stainless
a mixed degree of roundness. This suggests that the main source for steel, Mater. Sci. Eng. A 644 (2015) 171–183.
defects in this build is air entrapment from moisture evaporation, [6] L. Thijs, et al., Fine-structured aluminium products with controllable texture by
local voids after powder layer deposition, and incomplete melting. selective laser melting of pre-alloyed AlSi10Mg powder, Acta Mater. 61 (5) (2013)
1809–1819.
4 X-ray tomography image sequencing and 3D reconstruction techni-
[7] M. Ma, et al., Control of shape and performance for direct laser fabrication of
ques were used to characterise areas within the LMD block that precision large-scale metal parts with 316L stainless steel, Opt. Laser Technol. 45
exhibited severe porosity defects. Large irregular profile porosities (2013) 209–216.
[8] B. Song, et al., Microstructure and tensile properties of iron parts fabricated by
due to a lack of fusion were detected at the first layer, and large disc-
selective laser melting, Opt. Laser Technol. 56 (2014) 451–460.
like interlayer porosities due to air entrapments were detected to- [9] D. Wang, et al., Investigation of crystal growth mechanism during selective laser
ward the middle of the LMD block. melting and mechanical property characterization of 316L stainless steel parts,
5 Differences in the Vickers microhardness measurement values are Mater. Des. 100 (2016) 291–299.
[10] F. Bartolomeu, et al., 316L stainless steel mechanical and tribological behavior—A
attributed to the larger grain sizes found at the track body than comparison between selective laser melting, hot pressing and conventional casting,
fusion line sites. The Hall-Petch relation’s constants are found to be Addit. Manuf. 16 (2017) 81–89.
consistent with Stainless Steel’s data. [11] F. Abe, et al., Influence of forming conditions on the titanium model in rapid
prototyping with the selective laser melting process, Proc. Inst. Mech. Eng. Part C J.
6 In the EDX analysis, oxides were detected at the fusion line area that Mech. Eng. Sci. 217 (1) (2003) 119–126.
is not seen in the track body area. This is due to the natural for- [12] R.A. Hardin, C. Beckermann, Effect of porosity on the stiffness of cast steel, Metall.
mation of a protective oxide layer after each track of AISI Stainless Mater. Trans. A 38 (12) (2007) 2992–3006.
[13] R.A. Hardin, C. Beckermann, Prediction of the fatigue life of cast steel containing
Steel 316 L is deposited. shrinkage porosity, Metall. Mater. Trans. A 40 (581) (2009).
7 An SEM image was also taken within a deposited track body area. In [14] N. Kurgan, Effect of porosity and density on the mechanical and microstructural
the EDX mapping analysis for this image, columnar microstructures properties of sintered 316L stainless steel implant materials, Mater. Des. 55 (2014)

295
Z.E. Tan et al. Additive Manufacturing 25 (2019) 286–296

235–241. A 528 (24) (2011) 7423–7431.


[15] R. Li, et al., Densification behavior of gas and water atomized 316L stainless steel [25] J.T. Busby, M.C. Hash, G.S. Was, The relationship between hardness and yield stress
powder during selective laser melting, Appl. Surf. Sci. 256 (13) (2010) 4350–4356. in irradiated austenitic and ferritic steels, J. Nucl. Mater. 336 (2-3) (2005) 267–278.
[16] W. Hofmeister, M. Griffith, Solidification in direct metal deposition by LENS pro- [26] K. Singh, S. Sangal, G. Murty, Hall–Petch behaviour of 316L austenitic stainless
cessing, JOM 53 (9) (2001) 30–34. steel at room temperature, Mater. Sci. Technol. 18 (2) (2002) 165–172.
[17] V. Manvatkar, A. De, T. DebRoy, Spatial variation of melt pool geometry, peak [27] F. Huang, et al., Microstructure and properties of thin wall by laser cladding
temperature and solidification parameters during laser assisted additive manu- forming, J. Mater. Process. Technol. 209 (11) (2009) 4970–4976.
facturing process, Mater. Sci. Technol. 31 (8) (2015) 924–930. [28] M.A.V. Bermejo, Predictive and measurement methods for Delta ferrite determi-
[18] P. Aggarangsi, J.L. Beuth, M.L. Griffith, Melt pool size and stress control for laser- nation in stainless steels, Weld. J. (2012) 113–121.
based deposition near a free edge, Solid Freeform Fabrication Proceedings. Proc. [29] N. Suutala, T. Takalo, T. Moisio, Single-phase ferritic solidification mode in aus-
2003 Solid Freeform Fabrication Symposium (2003). tenitic-ferritic stainless steel welds, Metall. Trans. A 10A (1979) 1183–1190.
[19] Health, N.I.O.M., ImageJ, in Circularity and Roundness. https://imagej.nih.gov/ij/ [30] N. Suutala, T. Takalo, T. Moisio, Ferritic-austenitic solidification mode in austenitic
docs/guide/146-30.html. stainless steel welds, Metall. Mater. Trans. A 11A (1980) 717–725.
[20] K. Monroy, J. Delgado, J. Ciurana, Study of the pore formation on CoCrMo alloys by [31] M.S.F. de Lima, S. Sankaré, Microstructure and mechanical behavior of laser ad-
selective laser melting manufacturing process, Procedia Eng. 63 (2013) 361–369. ditive manufactured AISI 316 stainless steel stringers, Mater. Des. 55 (2014)
[21] Z. Hu, et al., Experimental investigation on selective laser melting of 17-4PH 526–532.
stainless steel, Opt. Laser Technol. 87 (2017) 17–25. [32] M. Ziętala, et al., The microstructure, mechanical properties and corrosion re-
[22] G. Ziółkowski, et al., Application of X-ray CT method for discontinuity and porosity sistance of 316L stainless steel fabricated using laser engineered net shaping, Mater.
detection in 316L stainless steel parts produced with SLM technology, Arch. Civ. Sci. Eng. A 677 (2016) 1–10.
Mech. Eng. 14 (4) (2014) 608–614. [33] D. Kotecki, T. Siewert, WRC-1992 constitution diagram for stainless steel weld
[23] I. Maskery, et al., Quantification and characterisation of porosity in selectively laser metals: a modification of the WRC-1988 diagram, Weld. J. 71 (5) (1992) 171–178.
melted Al–Si10–Mg using X-ray computed tomography, Mater. Charact. 111 (2016) [34] N. Suutala, T. Takalo, T. Moisio, The relationship between solidification and mi-
193–204. crostructure in austenitic and austenitic-ferritic stainless steel welds, Metall. Trans.
[24] S. Van Bael, et al., Micro-CT-based improvement of geometrical and mechanical A 10 (4) (1979) 512–514.
controllability of selective laser melted Ti6Al4V porous structures, Mater. Sci. Eng.

296

S-ar putea să vă placă și