Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Materials Science & Engineering A 696 (2017) 113–121

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Mechanical behavior of selective laser melted 316L stainless steel MARK


a b,c a,⁎
Jyoti Suryawanshi , K.G. Prashanth , U. Ramamurty
a
Department of Materials Engineering, Indian Institute of Science, Bangalore 560012, India
b
Erich Schmid Institute of Materials Science, Austrian Academy of Sciences, Jahnstraße 12, A-8700 Leoben, Austria
c
Department of Manufacturing and civil Engineering, Norwegian University of Science and Technology, Teknogivegen 22, Gjovik, Norway

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

Keywords: The tensile, fracture, and fatigue crack growth properties of 316L stainless steel (SS) produced using the selective
Selective laser melting laser melting (SLM) technique were evaluated and compared with those of conventionally manufactured (CM)
Microstructure austenitic SSs. For SLM, both single melt (SM) and checker board (CB) laser scanning strategies were employed,
Mechanical properties so as to examine the effect of scanning strategy on the mechanical properties. The experimental results show that
Fatigue
the SLM alloys' yield strength is significantly higher than that of CM 316L SS, a result of the substantial
Fracture toughness
refinement in the microstructure. In contrast, only a marginal improvement in the ultimate tensile strength and a
marked reduction ductility, which are a result of the loss of work hardening ability, are attributed to the absence
of stress induced martensitic transformation common in CM austenitic SSs. In spite of these, the fracture
toughness, which ranges between 63 and 87 MPa m0.5, of the SLM alloys is good, which is a result of the
mesostructure induced crack tortuousity. The SLM process was found to marginally reduce the threshold stress
intensity factor range for fatigue crack growth initiation and enhance the Paris exponent within the steady state
crack growth regime. Both tensile and toughness properties were found to be anisotropic in nature. SLM with CB
scanning strategy improves both these properties. All these observations on the mechanical properties are
rationalized by recourse to micro- and meso-structures seen these alloys.

1. Introduction such as laser power, scanning speed, hatch distance, etc., on the
evolution of microstructure and defects during solidification of SLM
Additive manufacturing (AM) of metallic components using techni- parts are widely reported [6–11]. It is reported that the balling
ques such as the selective laser melting (SLM) of powders has gained tendency associated with the SLM process can be reduced by increasing
prominence in the recent past as it offers a number of advantages [1]. the laser power and/or lowering the scan speed, or by decreasing
The high cooling rates (~103–108 K/s [2,3]) inherent to the SLM powder layer thickness [12–14]. Casati et al. [15] reported reduced
process result in the formation of non-equilibrium phases with the tensile strength with a wide scatter in the elongation to failure in the
extended composition range [4,5] as well as highly refined micro- SLM 316L SS and attributed them to the presence of partially melted
structures. The overlapping of the hatch spacing and/or powder layers powder particles in the microstructure. In order to address this, laser
imparts the common characteristic of a mesostructure to these alloys. remelting, i.e. re-scanning of powder layer twice before spreading the
As a consequence, the mechanical behavior of the SLM alloys tends to next powder layer, was suggested as an alternative. Such a method was
be significantly different from their conventionally manufactured (CM) shown to improve to the density of the produced parts and reduce the
counterparts. Therefore, firm establishment of process-structure-prop- surface roughness, which may in turn improve fatigue properties [16].
erty correlations is essential before such SLM metallic components are It is important to note, however, that such laser remelting technique
widely deployed, especially in industries such as aerospace where there could markedly reduce the productivity and hence may not be
is a great advantage in using AM but prevention of in-service failure is desirable. Riemer et al. [17] examined the unnotched high cycle fatigue
absolutely critical. The primary objective of the research reported in the behavior as well as anisotropy in the fatigue crack growth character-
current paper is to study such structure-property correlations in 316L istics in these alloys.
stainless steel (SS), which is a widely used alloy. Most of the work on SLM-SS that has been published hitherto has
A number of papers are already available in literature that report been on the optimization of processing parameters and their effect on
research on the SLM SS samples. The effect of processing parameters, the quasi-static tensile properties. However, the fracture behavior of


Corresponding author.
E-mail addresses: jyoti.suryawanshi01@gmail.com (J. Suryawanshi), ramu@materials.iisc.ernet.in (U. Ramamurty).

http://dx.doi.org/10.1016/j.msea.2017.04.058
Received 9 January 2017; Received in revised form 11 April 2017; Accepted 13 April 2017
Available online 15 April 2017
0921-5093/ © 2017 Elsevier B.V. All rights reserved.
J. Suryawanshi et al. Materials Science & Engineering A 696 (2017) 113–121

Table 1 size 0.25 µm. The densities of the SLM samples were estimated from the
Chemical composition (in wt%) of 316L stainless steel powder. image analysis method using the methodology described in [19] as well
as the Archimedes' principle. The overall density was measured by
Fe Cr Ni Mo Mn Si N C P,S
Archimedes method as it considers whole volume of the part. It was
67–69 16–18 10–14 2–3 2 0.75 < 0.1 < 0.03 < 0.045 calculated by weighting three samples for each SLM alloy in air and
distilled water as per ASTM B962-15 standard [20]. The average
density of cross-sectional views using IAM was measured by capturing
SLM-SS was not reported hitherto, which assumes particular signifi- ten micrographs from three different locations of the same view.(Here,
cance in view of our recent finding on SLM prepared Al-12 wt% Si alloy the relative density was measured with respect to material's bulk
[18]. Through that study, we have demonstrated the possibility of density of 8 g/cm3).
simultaneously improving the strength and fracture toughness by
adopting the SLM process. While the strength enhancement is a result
of the refinement in the microstructure as well as solid solution 2.2. Mechanical testing
strengthening, the toughness improvement is a result of the mesoscopic
structure and the resultant crack tortuosity. "Does such simultaneous The samples for the fracture toughness (FT) and fatigue crack
enhancement in properties also occur in SLM-SS?" is the question we growth (FCG) tests were in the form of blocks with 32×32 mm2 cross
seek to address through this study. section (BD-TD plane) and 14 mm thickness (in SD direction). For the
uniaxial tensile tests, plate type dog-bone samples were cut from these
blocks. The gauge length, thickness, and width of the samples were 6,
2. Materials and experiments 0.5, and 2 mm respectively, and were according to the ASTM E8/E8M-
11 [21] standard. A nominal stain rate of 10−3 s−1 was employed.
2.1. Materials These tests were conducted with the loading axis either parallel or
perpendicular to the laser tracks present in BD-TD plane, as shown in
The SLM 316L SS samples examined in this study were fabricated by Fig. 1, which will be referred to as ∥ and ⊥ respectively, here
using the Concept laser machine equipped with a YAG fibre laser afterwards.
having built plate dimensions of 89.5×89.5 mm2. The nominal chemi- The FT and FCG tests were carried out on the mirror polished
cal composition of the pre-alloyed SS powder (size range 20–50 µm) is Chevron-notched half compact tension [C(T)] samples, having 24 mm
given in Table 1. The processing parameters used to prepare samples width (w) and 12 mm thickness (t), as per the ASTM-647-15 [22] and
are as following: laser power −90 W, scanning speed −1000 mm/s, ASTM-399-12 [23] standards, respectively. For these tests, a travelling
hatch spacing −150 µm (30% overlap), layer thickness −30 µm, and microscope with resolution 0.01 mm was utilized to monitor the
scan rotation between successive layers −90°. These parameters are the initiation of a sharp crack at the notch root and its subsequent growth.
standard parameters set by the machine manufacturer – Concept Laser For the fatigue crack initiation and growth, a load ratio of 0.1 and
for the MLAB device,for obtaining samples with near full density. Two sinusoidal waveform frequency at 10 Hz were employed on a servo-
different scanning strategies were employed: single melt (SM) and hydraulic testing machine was used. For FT tests, the C(T) samples were
checker board (CB), which are illustrated schematically in Fig. 1(a) and first pre-cracked using cyclic loading such that a/w ~0.4–0.5 (where a
(b), respectively. The scanning, building, and transverse directions are is the total crack length). Subsequently, they were loaded at the cross-
referred hereafter as SD, BD, and TD, respectively. After building the head displacement of 0.1 mm/s until fracture. The thickness of the
coupons, they were given a stress-relieving annealing treatment at sample required for satisfying the plane-strain conditions is estimated
500 °C for 1 h. to be ~50 mm. Since the sample thickness is smaller, the fracture
Microstructural analysis was carried out on the mirror polished and toughness values, estimated using the standard empirical equations, can
etched (using V2A etchant, i.e. 100 ml distilled water, 100 ml HCl, only be considered as conditional and hence are labeled as Kq.
10 ml HNO3) samples using optical and scanning electron microscopies For the FCG tests, the C(T) samples were first fatigue pre-cracked
(SEM). The samples for microstructural analysis were extracted from such that a/w ~0.3. Subsequent to a sharp crack initiation, the stress
the centre of the coupons. Electron backscattered diffraction (EBSD) intensity factor range, ΔK, was lowered by 10% after every 0.25 mm
was performed on a SEM that is attached with an EBSD detector. The crack growth to get threshold stress intensity factor range, ΔKth, as per
EBSD scans were carried out on SM-SS and CB-SS samples with a step the load shedding method [22]. Since a well defined ΔKth could not be

Fig. 1. Schematic diagram illustrating build (BD), scanning (SD), and transverse (TD) directions of the selective laser melting process employed to fabricate the 316L SS samples examined
in this work. SD indicates the direction of the scanning vector, whereas BD signifies the direction of deposition of powder layers. While figure (a) illustrates single melt (SM) hatch style
with 90° scan rotation between successive layers, (b) illustrates checker board (CB) hatch style.

114
J. Suryawanshi et al. Materials Science & Engineering A 696 (2017) 113–121

Fig. 2. (a) Representative optical micrographs of the SLM 316L SS on the three orthogonal planes showing the BD-TD, BD-SD, and SD-TD views of the sample. Figures (b) and (c) show the
higher magnification images BD-TD plane's microstructure showing the finer cellular microstructure and hatch intersections.

Fig. 3. Inverse pole figure orientation maps obtained through EBSD on the BD-TD plane of (a) SM and (b) CB alloys. Black dotted lines are utilized to highlight some of the melt pool
boarders.

115
J. Suryawanshi et al. Materials Science & Engineering A 696 (2017) 113–121

observed from the experimentally measured ΔK vs. da/dN plots, the Images of the pores observed in SM and CB samples on the BD-TD
value ΔK that corresponds to a crack growth rate of ~ 1×10−9 m/cycle and the SD-TD planes are displayed in Fig. 4. Such images were utilized
is taken as ΔKth. After obtaining ΔKth, ΔK was increased by 10% after to estimate the porosity levels in the SLM samples. They are 1.4% and
every 0.25 mm crack growth until fracture. The crack growth rate per 4.8% (area fractions expressed in %) on BD-TD and SD-TD planes of the
cycle, da/dN, vs. ΔK data in the steady state crack growth regime were SM alloy, respectively whereas in the CB alloy, they are 3.1% and 2.4%
utilized to estimate the Paris exponent (m) by using the Paris law, i.e. respectively. Additionally, the pores in the SM alloy appear to be
da/dN = c(ΔK)m, where c is the Paris constant. Fractographic analysis irregularly shaped and interconnected across multiple layers whereas
was carried out on all the tensile and fatigue tested samples using a they are relatively homogeneous distributed in the CB alloy. It was
SEM. suggested that the uniform distribution of pores in SLM alloys proces-
sing using the CB scanning strategy are due to the reduced consolida-
tion areas (~1×1 mm2 each) resulting in the minimization of the
3. Results thermal inhomogeneity [29,30]. The relative porosities estimated using
the Archimedes' principle are 2.3% and 1.9% for SM and CB alloys.
3.1. Microstructure These lower levels of the porosity vis-á-vis image analysis results is due
to the fact that the unmelted powder particles that reside inside pores
An 3D optical micrograph showing the BD-TD, SD-TD, and SD-BD also contribute to the mass of the sample and hence enhance the
planes of the SLM-SS alloy is displayed in Fig. 2(a). It shows the measured density when estimated using the Archimedes' principle [31].
mesoscopic structure, with the hatch overlapped regions in the form of This experiment artifact is the reason for adapting the image analysis
half cylindrical contours, that is a result of the SLM process. The method for estimating the density of the AM parts as it not only gives a
straight lines on the SD-TD plane form because heat moves radially and more accurate porosity estimate but also gives information about
away from the centre of the laser beam. High magnification SEM distribution, types, and morphology of pores in the SLM parts.
images taken from BD-TD plane of the alloy are displayed in Fig. 2(b)
and (c). They show a fine cellular morphology, which is due to the high
cooling rate associated with rapid quenching of the molten metal due to 3.2. Mechanical properties
localized laser melting. This morphology indicates to cellular solidifica-
tion, which primarily depends on two parameters: the thermal gradient Representative engineering stress (σ) vs. strain (e) responses of the
in the liquid, G, and the solidification front growth rate, R, achieved both SM and CB samples tested in both ∥ and ⊥ directions are displayed
during solidification of the alloy [24]. The G/R ratio controls the in Fig. 5. Tensile properties such as yield strength (YS), ultimate tensile
morphology of the alloy, while the product G ×R determines the cooling strength (UTS), and strain to failure (ef) extracted from these plots are
rate and hence the refinement in the microstructure. In contrast to the listed in Table 2. For comparison, literature data on CM 316L SS is also
SLM prepared steel, the CM alloy with similar composition would listed. A substantial improvement (of up to ~60%) in YS due to SLM
contain FCC austenite grains of size 30–80 µm [25–27]. process can be noted. In contrast, the enhancement in UTS is only
The EBSD results obtained from the SD-BD plane of the both SM and marginal at ~10%. These improvements in strength come at the
CB alloys are displayed in Fig. 3. They show several melt pools, which expense of ductility, which is reduced by about 50% in ∥ direction
are highlighted with the aid of dotted lines. It was widely reported in and even more markedly by ~72% in ⊥ direction. Amongst the SM and
the literature on SLM process that < 001 > texture of the cubic CB samples, the latter were found to exhibit marginally better proper-
material evolves favorably due to the epitaxial growth during solidifi- ties. In both the cases, better properties were obtained in ∥ direction
cation in [17,18,24,28]. However, the fraction of < 001 > orientated than in ⊥ direction, indicating to significant levels of anisotropy in the
grains (represented by red color) is relatively low. This is possibly a mechanical performance of the SLM 316L SS.
result of the scan rotation and scanning strategies adopted during the Results of FT tests are listed in Table 2, which shows that Kq of the
SLM process. SLM alloys ranges between 63 and 87 MPa m0.5, which is lower than

Fig. 4. Optical micrographs of the SM (a and b) and CB (c and d) alloys showing the nature of pores and their distribution with the microstructure. Figures (a) and (c) are obtained on the
BD-TD plane whereas figure (b) and (d) from the SD-TD plane.

116
J. Suryawanshi et al. Materials Science & Engineering A 696 (2017) 113–121

be considered reasonably ductile. Moreover, since the SLM process


allows for near-net shaped final product, i.e., there would be no further
forming steps such as rolling or forging on the components, the
reduction in ductility need not necessarily viewed as a serious detri-
ment when it comes to the application potential of the SLM 316L SS
components.
The substantial enhancement in YS due to the SLM of 316L SS can
be attributed to the marked refinement in the microstructure, which is a
result of the high cooling rates (~103–108 K/s) achieved during the
SLM processing. For example, the size of the cells within the cellular
microstructure is about 1 µm, see Fig. 2(c), whereas the grain size in a
CM austenitic SS tends to be in the range of hundreds of μm. We do not
expect any difference in the solid solubility of the alloying elements
between SLM and CM alloys, as all those are in solid solution in both the
alloys. Therefore, it is unlikely that the differences in the YS observed
between SLM and CM alloys can be attributed to solid solution
strengthening (as was done in the SLM Al-Si alloy, for example). In
the CM 316L SS alloys, substantial strain hardening occurs due to stress-
induced austenite to martensite transformation (SIMT). Then, the low
strain hardening in SLM 316L SS can be construed as due to the absence
Fig. 5. Representative engineering stress (σ) vs. strain (e) plots obtained through tensile
of SIMT. It is possible that the extremely fine microstructure in the SLM
testing the SLM 316L SS manufactured by using SM and CB scanning strategies, and in
two different orientations with respect to the build direction. alloy elevates the YS and in the process prevents SIMT.
The high YS combined with the possible absence of SIMT –and
that reported in literature for the CM 316L SS. Thus, a simultaneous hence work hardening– in the SLM SS samples is possibly the reason for
enhancement in both strength and toughness, as observed in SLM Al- the observed low fracture toughness in them. A low YS allows for plastic
12 wt% alloy, was not seen in the SLM 316L SS. In between SM and CB flow at the crack tip, in turn, blunting it and requiring more energy or
alloys, the latter exhibits better toughness. Similar to the results of stress to fracture the sample. Such a dissipation mechanism appears to
tensile properties, Kq is also anisotropic with maximum toughness in ∥ be limited in the case of SLM samples; a relatively lower toughness is
direction. therefore obtained. It is worth noting here that while the Kq values are
The da/dN vs. ΔK plots obtained for the SM and CB alloys are indeed lower than those reported for the CP austenitic SS, the toughness
displayed in Fig. 6. Values of ΔKth and m extracted from these plots are values of the SLM 316L SS, ranging between 63 and 87 MPa m0.5, are
listed in Table 2, along with those inferred from literature for CM quite good when viewed on 'stand alone' basis, and hence may not be
alloys. Like the toughness, ΔKth of the SLM alloys are slightly inferior to pose a major drawback in terms of application potential.
those of CM alloy. For the SLM alloys, the FCG also occurs faster, with In the context of fracture toughness, and as noted already, we did
all the m values around 4, whereas m reportedly ranges between 2 and 3 not observe a simultaneous enhancement in both strength and tough-
for CM alloys. While ΔKth of SM alloy were slightly better than the CB ness as seen in SLM Al-12Si alloy [18]. In that case, the Si preferentially
alloy, superior ΔKth was obtained in ⊥ direction as compared to ∥ segregates to the laser tracks and provides for a less resistant crack
direction in both the cases. path. However, the crack deflection that occurs, as the crack meanders
along the laser tracks, results in lowered crack driving force at the crack
tip, imparting higher fracture toughness. This is also the reason for the
4. Discussion observance of significant anisotropy in fracture toughness values in
SLM Al-12Si alloy. In the case of SLM 316L, (a) there are no brittle
The mechanical property data, listed in Table 2, clearly indicates phase decorating the laser tracks, i.e., there are no facilitators for
that the when the SLM process is employed for manufacturing 316L SS pronounced crack deflection, and (b) the melt pool height is much
components, a significant benefit in terms of elevating YS can be smaller (~30 µm) vis-á-vis ~120–150 µm in Al-12Si. Thus, crack
derived. However, the increase in UTS as compared to that of the CM deflection is relatively smaller. Nevertheless, the crack tip tortuousity
alloy is relatively smaller, which indicates that the SLM alloys do not does exist, which will be described later, that contributes partly to the
exhibit the same level of work hardening as the CM austenitic SS. toughness of the SLM alloys.
Indeed, the tensile stress-strain plots displayed in Fig. 5 indicate that The possible absence of roughness induced crack closure, due to the
the SLM alloys' stress-strain response is nearly elastic-perfectly plastic, lack of pronounced crack deflection, could also be one of the reasons for
with only a marginal strain hardening. The lack of work hardening the observed lower ΔKth in the SLM-SS alloy, as compared to CM alloy.
could also be the reason for the significant reduction in tensile ductility The high YS, and hence smaller cyclic plastic zone, and lower levels of
of SLM alloys, as per the Considère's criterion for necking. It is worth plasticity induced crack closure could also be a contributing factor for
noting here that such reductions in ductility do not make the alloy ΔKth of SLM alloy being lower than CM 316L SS alloy. The higher values
necessarily brittle. In fact, the lowest ef, at 11.6% for the CB-∥ case, can

Table 2
Summary of the mechanical properties of 316L stainless steel synthesized by selective laser melting and conventional manufacturing process routes.

YS (MPa) UTS (MPa) ef (%) Kq (MPa m0.5) ΔKth (MPa m0.5) m

SS-SM-∥ 511.6 ± 14 621.7 ± 12 20.4 ± 3 72.3 9.1 4.0


SS-SM-⊥ 430.4 ± 11 509.0 ± 3 12.4 ± 1 62.9 9.9 4.1
SS-CB-∥ 536.4 ± 4 668.4 ± 5 24.7 ± 2 86.8 7.8 3.7
SS-CB-⊥ 448.5 ± 20 527.9 ± 7 11.6 ± 1 79.6 8.1 3.8
SS-CMa 220–270 520–680 40–45 112–278 (KIC) 10–15 2.2–2.8

a
YS, UTS, and ef of CM 316L SS are obtained from Ref [11]. whereas KIC, Kth and m values are obtained from Ref [45].

117
J. Suryawanshi et al. Materials Science & Engineering A 696 (2017) 113–121

Fig. 6. Typical fatigue crack growth rate per cycle (da/dN) as a function of the applied stress intensity factor range (ΔK) for 316L SS in (a) SM (b) CB scanning strategies. Inset images
schematically illustrate the crack plane and growth direction with respect BD, SD, and TD orientations.

Fig. 7. Mesoscopic structure of the SLM alloy in (a) ∥ and (b) ⊥ loading conditions. The loading directions with respect to the mesoscopic structure during tensile testing are also
illustrated.

of m, which imply the fatigue cracks grow relatively faster, in SLM alloy
are possibly because of the lower levels of plasticity (as reflected by
lower ductility levels) in it.
After the comparison between SLM and CM alloys, we turn attention
to the effect of scanning strategy, i.e., a comparison of the properties of
SM and CB SLM alloys and possible reasons for the differences. Overall,
the experimental results indicate that the tensile properties of the CB
alloy are only slightly better than the respective properties of SM alloys.
The toughness values, however, are significantly better with CB-⊥'s
toughness being ~26% higher than that of SM-⊥'s. The CB alloy's
resistance to FCG is similar overall with that of the SM alloy, with
slightly better m values but with lower ΔKth. The inferior toughness of
the SM alloy is possibly due the presence of interconnected porosity
present in it (Fig. 4). These pores appear similar to the characteristic
keyhole pores that are due to insufficient energy density required for
complete melting during the SLM process [32–34], particularly due to
increased scan speed. Published literatures indicate that the energy
density required to synthesize fully dense samples for SS alloy ranges
between 60 and 120 J/mm3 [35,36]. The energy used to fabricate the
samples of the current study was much lower at 28.6 J/mm3. Further-
more, and in contrast to the SM scanning strategy, the CB-scanning
Fig. 8. Low magnification fractography images of the tensile tested samples of SM 316L
SS that were tested in (a) ∥ (b) ⊥ loading conditions.
strategy modifies thermal dynamics and minimizes temperature inho-

118
J. Suryawanshi et al. Materials Science & Engineering A 696 (2017) 113–121

Fig. 9. Representative fractographs obtained from FT tested samples on SM 316L SS samples tested in (a) ∥ (b) ⊥ loading conditions. Figure (c) shows a high magnification image of (b), to
highlight pockets of unmelted powder particles on the fracture surface (pointed by the arrow).

Fig. 10. Representative images showing fatigue crack growth interaction with microstructure for SM 316L SS sample tested in (a) ∥ and (b) ⊥ loading conditions.

mogeneity [30]; this, in turn, reduces the propensity for defect are oriented perpendicular to the loading direction, may not provide
formation in the CB alloy. This could also be the reason for the slightly such constraint. Furthermore, since porosity is confined to such hatch
reduced values of m in the CB alloy. In contrast, the slight enhancement tracks, initiation microcracks at the triple points due to stress concen-
in ΔKth of the SM alloy is possibly because of the occurrence of tration at those locations, could lead to lower ductility. This was further
plasticity-induced near-threshold crack closure phenomena as a result confirmed by fractography on tensile tested samples (Fig. 8). SLM alloys
of decreased YS [37–39]. tested in ∥ configuration exhibit a relatively a more jagged fracture
Finally, we examine the anisotropy in properties of the SLM alloys. surface (Fig. 8(a)) compared to the samples that were tested in ⊥
Significant anisotropy was observed in the tensile properties of both the configuration (Fig. 8(b)). These images also show unmelted powder
SM and CB alloys. While there was some anisotropy in toughness as particles on the fracture surface (illustrated by red arrows). Fig. 9(a),
well, it is relatively small. These features are in complete contrast to which is an fractograph obtained on the fracture toughness tested SM
that reported by us [18] in the Al-12Si alloy manufactured through SLM sample (∥ configuration), a significant increase in surface roughness as
route. In that case, no anisotropy was detected in tensile properties compared to that seen in Fig. 9(b), which is the fractograph obtained on
whereas the fracture toughness was anisotropic. The anisotropy in YS of the sample tested in ⊥ orientation that exhibits rather a flat fracture
SLM 316L SS can be rationalized by recourse to Fig. 7, wherein the crack path. The relatively more crack deflection due to the crack tip
micrographs with hatch orientation with respect to the loading direc- interaction with the laser tracks. The presence of some un-melted
tions for both ∥ and ⊥ configurations are displayed. As seen, the laser powder particles along the laser tracks, as shown in high magnification
tracks, which appear aligned due to low layer thickness (~30 µm) but image of the FT tested SM-⊥ fracture surface Fig. 9(c), which facilitate
high lateral width (~120 µm), are parallel to the loading direction in easy cracking due to weak (or no) metallurgical bonding between layers
specimens with ∥ orientation. The constraint imposed by these to plastic due to incomplete consolidation during fabrication, also helps in crack
flow is possibly the reason for the higher YS and UTS in ∥ configuration. deflection.
In contrast, when tested in the ⊥ configuration, the melt pools, which No enhancement in the ΔKth of the SLM-SS alloy when tested in the

119
J. Suryawanshi et al. Materials Science & Engineering A 696 (2017) 113–121

∥ direction is noted. This is possibly associated with the roughness Microstructure and mechanical properties of Al–12Si produced by selective laser
melting: Effect of heat treatment, Mater. Sci. Eng.: A. 590 (2014) 153–160.
induced crack closure mechanism due to preferential cellular growth [5] E. Brandl, U. Heckenberger, V. Holzinger, D. Buchbinder, Additive manufactured
direction during solidification. Due to very high thermal gradient and AlSi10Mg samples using selective laser melting (SLM): microstructure, high cycle
as the heat moves radial away from the centre of a laser beam, the fatigue, and fracture behavior, Mater. Des. 34 (2012) 159–169.
[6] M. de Lima, S. Sankaré, Microstructure and mechanical behavior of laser additive
directional solidification takes place in the melt pool which results in a manufactured AISI 316 stainless steel stringers, Mater. Des. 55 (2014) 526–532.
growth of cellular grains perpendicular to the melt-pool boundaries, i.e. [7] G. Miranda, S. Faria, F. Bartolomeu, E. Pinto, S. Madeira, A. Mateus, et al.,
laser tracks [24,40], seen in Fig. 2(b). Fig. 10 displays the fatigue crack Predictive models for physical and mechanical properties of 316L stainless steel
produced by selective laser melting, Mater. Sci. Eng. 657 (2016) 43–56.
growth images captured when crack growth rate was ~10−9 m/cycle, [8] P. Hanzl, M. Zetek, T. Bakša, T. Kroupa, The influence of processing parameters on
showing the tendency of the fatigue crack to grow along the weak the mechanical properties of SLM parts, Procedia Eng. 100 (2015) 1405–1413.
interfaces of cellular boundaries and matrix. Consequently, for ∥ [9] Q. Wei, S. Li, C. Han, W. Li, L. Cheng, L. Hao, et al., Selective laser melting of
stainless-steel/nano-hydroxyapatite composites for medical applications: micro-
loading condition, ΔKth is reduced as the crack can easily grow along
structure, element distribution, crack and mechanical properties, J. Mater. Process
the cellular boundaries [41–44]. Technol. 222 (2015) 444–453.
[10] L. Hao, S. Dadbakhsh, O. Seaman, M. Felstead, Selective laser melting of a stainless
5. Summary steel and hydroxyapatite composite for load-bearing implant development, J.
Mater. Process Technol. 209 (2009) 5793–5801.
[11] I. Tolosa, F. Garciandía, F. Zubiri, F. Zapirain, A. Esnaola, Study of mechanical
The microstructure and the mechanical properties of the 316L SS properties of AISI 316 stainless steel processed by “selective laser melting”,
fabricated using SLM technique with the two different scanning following different manufacturing strategies, Int. J. Adv. Manuf. Technol. 51 (2010)
639–647.
strategies were evaluated and compared with those of the same alloy, [12] D. Gu, Y. Shen, Balling phenomena in direct laser sintering of stainless steel powder:
but manufactured using conventional methods. The following are the metallurgical mechanisms and control methods, Mater. Des. 30 (2009) 2903–2910.
key observations. [13] T. Grünberger, R. Domröse, Direct metal laser sintering, Laser Tech. J. 12 (2015)
45–48.
[14] J.P. Kruth, L. Froyen, J. Van Vaerenbergh, P. Mercelis, M. Rombouts, B. Lauwers,
1. A significant enhancement in YS of the SLM alloys is noted. This is Selective laser melting of iron-based powder, J. Mater. Process. Technol. 149 (2004)
attributed to the refinement in the microstructure of the alloy, 616–622.
[15] R. Casati, J. Lemke, M. Vedani, Microstructure and fracture behavior of 316L
which in turn is a result of the rapid solidification process that
austenitic stainless steel produced by selective laser melting, J. Mater. Sci. Technol.
occurs during SLM. The improvement in UTS is only marginal and 32 (2016) 738–744.
the elongation to failure is markedly lower. These are due the [16] E. Yasa, J.-P. Kruth, Microstructural investigation of selective laser melting 316L
stainless steel parts exposed to laser re-melting, Procedia Eng. 19 (2011) 389–395.
absence of notable work hardening, which is possibly a result of the
[17] A. Riemer, S. Leuders, M. Thöne, H.A. Richard, T. Tröster, T. Niendorf, On the
lack of stress induced austenite-to-martensite that is typical in fatigue crack growth behavior in 316L stainless steel manufactured by selective
austenitic SSs such as 316L. laser melting, Eng. Fract. Mech. 120 (2014) 15–25.
2. In comparison to the CM 316L SS, the fracture toughness of the SLM [18] J. Suryawanshi, K.G. Prashanth, S. Scudino, J. Eckert, O. Prakash, U. Ramamurty,
Simultaneous enhancements of strength and toughness in an Al-12Si alloy
alloys is also lower, which is possibly due to the reduced ductility synthesized using selective laser melting, Acta Mater. 115 (2016) 285–294.
and/or absence of SIMT-induced toughening in the latter. Yet, the [19] A.B. Spierings, M. Schneider, R. Eggenberger, Comparison of density measurement
SLM alloys cannot be termed brittle as their toughness values are techniques for additive manufactured metallic parts, Rapid Prototyp. J. 17 (2011)
380–386.
equal to or above 63 MPa m0.5. This is due to the crack tortuousity [20] ASTM B962-15: Standard test methods for density of compacted or sintered powder
imparted by the mesostructure of the SLM alloys. metallurgy (PM) products using Archimedes' principle.
3. The fatigue crack growth characteristics of the SLM alloys are [21] ASTM E8/E8M-11: Standard test methods for tension testing of metallic materials,
ASTM International, 3.01 (2011).
similar, overall, to those of CM alloys. A slightly lower ΔKth and [22] ASTM E647-15: Standard test method for measurement of fatigue crack growth
higher values of m in the SLM alloys are due to reduced plasticity- rates, ASTM International 3.01 (2015).
induced crack closure and lowered ductility respectively. [23] ASTM E 399-12. Standard test method for linear elastic plane-strain fracture
toughness KIC of metallic materials, ASTM International 3.01 (2012).
4. Among the two scanning strategies employed, the checker board [24] L. Thijs, K. Kempen, J.-P. Kruth, J. Humbeeck, Fine-structured aluminium products
strategy yields components with better mechanical properties than with controllable texture by selective laser melting of pre-alloyed AlSi10Mg
the single melt strategy, which is possibly due to higher porosity powder, Acta Mater. 61 (2013) 1809–1819.
[25] D. Samantaray, V. Kumar, A.K. Bhaduri, P. Dutta, Microstructural evaluation and
content in the latter.
mechanical properties of type 304L stainless steel processed in semi-solid state, Int.
5. The mechanical properties of both SM and CB alloys were signifi- J. Metall. Eng. 2 (2) (2013) 149–153.
cantly anisotropic in nature, which is attributed to the anisotropic [26] S. Hao, P. Wu, J. Zou, T. Grosdidier, C. Dong, Microstructure evolution occurring in
mesostructure that is, in turn, due to high melt pool width to layer the modified surface of 316L stainless steel under high current pulsed electron beam
treatment, Appl. Surf. Sci. 253 (2007) 5349–5354.
thickness ratio. [27] G. Liu, J. Lu, K. Lu, Surface nanocrystallization of 316L stainless steel induced by
ultrasonic shot peening, Mater. Sci. Eng. 286 (2000) 91–95.
Acknowledgements [28] T. Niendorf, S. Leuders, A. Riemer, H. Richard, T. Tröster, D. Schwarze, Highly
anisotropic steel processed by selective laser melting, Metall. Mater. Trans. B. 44
(2013) 794–796.
This work at IISc was sponsored by the Boeing Company through a [29] T. Gustmann, A. Neves, U. Kühn, P. Gargarella, C.S. Kiminami, C. Bolfarini, et al.,
research grant (PC36035). We are grateful to Dr. Om Prakash of Boeing Influence of processing parameters on the fabrication of a Cu-Al-Ni-Mn shape-
memory alloy by selective laser melting, Addit. Manuf. 11 (2016) 23–31.
India R & D for many useful discussions. UR acknowledges the support [30] J. Jhabvala, E. Boillat, T. Antignac, R. Glardon, On the effect of scanning strategies
of his research by the Department of Science & Technology, in the selective laser melting process, Virtual Phys. Prototyp. 5 (2010) 99–109.
Government of India through a J.C. Bose National Fellowship. [31] J.-P. Kruth, J. Deckers, E. Yasa, R. Wauthlé, Assessing and comparing influencing
factors of residual stresses in selective laser melting using a novel analysis method,
Proc. Inst. Mech. Eng. Part B J. Eng. Manuf. 226 (2012) 980–991.
References [32] N. Aboulkhair, A. Stephens, I. Maskey, C. Tuck, I. Ashcroft, N. Everitt, Mechanical
properties of selective laser melted AlSi10Mg: Nano, micro, and macro properties
(2015).
[1] H. Chen, D. Gu, D. Dai, C. Ma, M. Xia, Microstructure and composition homo-
[33] N. Aboulkhair, N. Everitt, I. Ashcroft, C. Tuck, Reducing porosity in AlSi10Mg parts
geneity, tensile property, and underlying thermal physical mechanism of selective
processed by selective laser melting, Addit. Manuf. 1 (2014) 77–86.
laser melting tool steel parts, Mater. Sci. Eng. 682 (2017) 279–289.
[34] K. Kempen, E. Yasa, L. Thijs, J.-P Kruth, J. van Humbeeck, Microstructure and
[2] L. Zheng, Y. Liu, S. Sun, H. Zhang, Selective laser melting of Al–8.5Fe–1.3V–1.7Si
mechanical properties of selective laser melted 18Ni-300 steel, Phys. Procedia 12
alloy: investigation on the resultant microstructure and hardness, Chin. J. Aeronaut.
(2011) 255–263.
28 (2015) 564–569.
[35] J.-P. Choi, G.-H. Shin, M. Brochu, Y.-J. Kim, S.-S. Yang, K.-T. Kim, et al.,
[3] T. Vilaro, C. Colin, J.D. Bartout, L. Nazé, M. Sennour, Microstructural and
Densification behavior of 316L stainless steel parts fabricated by selective laser
mechanical approaches of the selective laser melting process applied to a nickel-
melting by variation in laser energy density, Mater. Trans. 57 (2016) 1952–1959.
base superalloy, Mater. Sci. Eng. 534 (2012) 446–451.
[36] J. Cherry, H. Davies, S. Mehmood, N. Lavery, S. Brown, J. Sienz, Investigation into
[4] K.G. Prashanth, Scudino, H.J. Klauss, K.B. Surreddi, L.öber Wang, et al.,

120
J. Suryawanshi et al. Materials Science & Engineering A 696 (2017) 113–121

the effect of process parameters on microstructural and physical properties of 316L [41] D. Lados, D. Apelian, J. Donald, Fatigue crack growth mechanisms at the
stainless steel parts by selective laser melting, Int. J. Adv. Manuf. Technol. 76 microstructure scale in Al–Si–Mg cast alloys: mechanisms in the near-threshold
(2015) 869–879. regime, Acta Mater. 54 (2006) 1475–1486.
[37] R.O. Ritchie, Mechanism of fatigue-crack propagationin ductile and brittle solids, [42] D. Lados, D. Apelian, Fatigue crack growth characteristics in cast Al–Si–Mg alloys,
Int. J. Fract. 100 (1999) 55–83. Mater. Sci. Eng. 385 (2004) 187–199.
[38] K. Makhlouf, J.W. Jones, Near-threshold fatigue crack growth behaviour of a ferritic [43] D. Lados, D. Apelian, J. Major, Fatigue crack growth mechanisms at the micro-
stainless steel at elevated temperatures, Int. J. Fatigue 14 (1992) 97–104. structure scale in Al-Si-Mg cast alloys: mechanisms in regions II and III, Metall.
[39] S. Suresh, Fatigue of Materials, second ed., Cambridge University Press, 1998. Mater. Trans. 37 (2006) 2405–2418.
[40] L. Thijs, F. Verhaeghe, T. Craeghs, J. Humbeeck, J.-P. Kruth, A study of the [44] S. Suresh, R. Ritchie, A geometric model for fatigue crack closure induced by
microstructural evolution during selective laser melting of Ti–6Al–4V, Acta Mater. fracture surface roughness, Metall. Trans. 13 (1982) 1627–1631.
58 (2010) 3303–3312. [45] H. Boyer, Atlas of Fatigue, ASM international.

121

You might also like