ArticlePDF Available

The effect of build orientation on tensile properties and corrosion resistance of 316L stainless steel fabricated by laser powder bed fusion

Authors:

Abstract

The effects of build orientation on tensile property and corrosion behavior of laser powder bed fusion 316L stainless steel were investigated. The laser powder bed fusion 316L stainless steel in normal direction exhibited lower tensile strength and elongation than one in build direction. In view of the fact that there was no obvious martensitic transformation in the two samples, high density dislocation and more low-angle grain boundaries in laser powder bed fusion 316L stainless steel in build direction both enhanced tensile yield and tensile strength. More deformed twins in the sample in build direction induced better plasticity. In addition, laser powder bed fusion 316L stainless steel in build direction exhibited better corrosion resistance than one in normal direction in borate buffer and sodium chloride solutions. This was attributed to the larger grain and more densely packed crystallographic faces in build direction for laser powder bed fusion 316L stainless steel.
Journal of Manufacturing Processes 106 (2023) 363–369
Available online 12 October 2023
1526-6125/© 2023 Published by Elsevier Ltd on behalf of The Society of Manufacturing Engineers.
The effect of build orientation on tensile properties and corrosion resistance
of 316L stainless steel fabricated by laser powder bed fusion
Lv Jinlong
a
,
*
,
1
, Zhou Zhiping
a
,
1
, Wang Zhuqing
b
,
*
, Yida Xiong
c
a
Sino-French Institute of Nuclear Engineering and Technology, Sun Yat-Sen University, Zhuhai 519082, Guangdong, China
b
School of Mechanical Engineering, Sichuan University, Chengdu 610065, China
c
School of Mechanical and Aerospace Engineering, Nanyang Technological University, Singapore 639798, Republic of Singapore
ARTICLE INFO
Keywords:
Stainless steel
Laser powder bed fusion
Build orientation
Tensile
Corrosion
ABSTRACT
The effects of build orientation on tensile property and corrosion behavior of laser powder bed fusion 316L
stainless steel were investigated. The laser powder bed fusion 316L stainless steel in normal direction exhibited
lower tensile strength and elongation than one in build direction. In view of the fact that there was no obvious
martensitic transformation in the two samples, high density dislocation and more low-angle grain boundaries in
laser powder bed fusion 316L stainless steel in build direction both enhanced tensile yield and tensile strength.
More deformed twins in the sample in build direction induced better plasticity. In addition, laser powder bed
fusion 316L stainless steel in build direction exhibited better corrosion resistance than one in normal direction in
borate buffer and sodium chloride solutions. This was attributed to the larger grain and more densely packed
crystallographic faces in build direction for laser powder bed fusion 316L stainless steel.
1. Introduction
316L stainless steel (SS) was widely used as structural metallic ma-
terials in several important industrial applications due to its excellent
mechanical properties and excellent pitting corrosion resistance [1,2].
Recently, innovation laser powder bed (L-PBF) technology has gained
signicant attention due to its exible fabrication and time saving ad-
vantages [3,4]. The new and advanced L-PBF technique can use a high-
energy laser beam to melt some metal powders by way of progressive
layers and prepare an integral components with complex shape.
The evident effects of the L-PBF parameters variations on micro-
structural characteristics and mechanical properties are still not well
understood. It was demonstrated that higher laser power promoted to
induce strong crystallographic textured columnar coarse-grains and also
signicantly improved ductility of L-PBFed 316L SS [5]. In addition,
Maria et al. [6] found that the lower power facilitated the formation of
more random and some ner cell-like structures in L-PBFed 316L SS.
Larimian et al. [7] found that the higher scanning speed during the L-
PBF process improved densication of L-PBFed 316L SS and thus
induced excellent mechanical strengthening. It was found that lower
scanning speed promoted the higher elongation of L-PBFed 316L SS [8].
The scanning strategy signicantly affected the characteristic size of
cell-like structures and grains which both could determine the tensile
strength for L-PBFed 316L SS [9]. Moreover, Hong et al. [10] observed
that the L-PBF process severely suppressed the stain-induced
α
martensitic transformation compared with the conventional 316L SS,
and low-angle grain boundaries (LAGBs) and ne cellular microstruc-
tures synergistically suppressed dislocation slip and deformed twinning
during deformation, which resulted in the reduction in the nucleation
sites of
α
-martensite. However, Wang et al. [11] found that the cellular
structures in L-PBFed 316L SS facilitated the generation of shear bands
for the nucleation sites of
α
-martensite and nally promoted the phase
transformation. Therefore, the initial microstructures of L-PBFed
316LSS signicantly affected the deformation during tension process
[12]. Therefore, the presence of a lot of dislocation cellular structures
should be very important reason for high yield and tensile strength of L-
PBFed 316L SS [13]. However, a weak dislocation barrier ability was
also found based on experiments and calculations [14].In addition, high
residual internal stresses in the SSs due to rapid solidication also
induced dislocation cells [15]. The mechanical properties of L-PBFed SSs
were affected by residual stresses [1620]. The importance of critical
laser energy density was emphasized during L-PBF processing [16,17].
The build orientation and dislocation density both affected signicantly
mechanical properties of L-PBFed 316L SS [21]. It was found that BD
* Corresponding authors.
E-mail addresses: ljltsinghua@126.com (L. Jinlong), wzhuqing@scu.edu.cn (W. Zhuqing).
1
These authors contributed equally to this work
Contents lists available at ScienceDirect
Journal of Manufacturing Processes
journal homepage: www.elsevier.com/locate/manpro
https://doi.org/10.1016/j.jmapro.2023.10.010
Received 3 April 2023; Received in revised form 4 October 2023; Accepted 5 October 2023
Journal of Manufacturing Processes 106 (2023) 363–369
364
(denoted BD, parallel to building direction) L-PBFed biodegradable high
manganese steel showed lower ultimate tensile strength than ND
(denoted ND, perpendicular to building direction) L-PBFed one [22].
Recent studies showed that build orientation also affected the crystallite
size of the samples and further changed tensile strength [23].
The understanding of corrosion resistance of the passive lms formed
on the surface of L-PBFed 316L SS should also be urgently needed [24].
The molten pool boundary was the most weakened position for the
pitting corrosion in L-PBFed 316L SS [25]. However, it was demon-
strated that better pitting corrosion resistance of L-PBFed 316L in NaCl
solution was attributed to the rened MnS inclusion [26]. Moreover, it
also was found that compressive residual stress for L-PBFed 316L SS
could effectively decrease passive current density and donor density in
the passive lms [27]. The grain size distribution characteristic strongly
depended on the building direction during the L-PBF processing [28].
Previous investigations indicated evident differences in the corrosion
resistance for two faces which were perpendicular and parallel to
building direction in L-PBFed 316L SSs [29,30]. It was found that
crystallographic lamellar microstructure showed superior corrosion
resistance [31]. Moreover, it was also found that the face parallel to
building direction showed better corrosion resistance than the face
perpendicular to building direction for L-PBF maraging steel [32].
Evidently, the relationships between mechanical property and
building direction are still not clear, and the main reason behind the
improved properties of L-PBF 316L SS should be further claried. In
order to break the strengthductility trade-off, the effect of build
orientation on mechanical properties should be investigated. In addi-
tion, the effect of building direction on compactness of the passive lms
on the surface of L-PBF 316L SS also should be compared and evaluated.
Thus, this paper aims to investigate the effect of BD and ND on
microstructures, mechanical properties and corrosion resistance of L-
PBFed 316L SSs.
2. Experimental methods
Table 1 shows the chemical compositions of the 316L SS powers, and
the diameter of 316L SS powers is between 15
μ
m and 55
μ
m. This
powder size has good ow ability. Cubic L-PBFed specimens were
fabricated using the EOS M290 (Germany) system with a 400 WYb,
operating by laser wavelength (YAG ber laser) of 1070 nm at the
continuous mode. Laser power from 275 W to 315 W, scanning speed
from 760 mm/s to 1160 mm/s and hatch distance from 90
μ
m to 110
μ
m
were used. Each layer was scanned unidirectionally and the scanning
direction was rotated by 67for the next layer, as shown in Fig. 1a-c.
The specic main manufacturing process parameters are shown in
Table 2. These set of parameters are ultimately optimized for 316L SS
with the minimum defects for the L-PBF process. The cut dog-bone
tensile samples from cubic structures (50 mm side length) were cut in
Table 1
The composition of the 316L SS powers (wt%).
Sample C S P Cr Mo Ni Mn Si Fe
Power 0.022 0.01 0.034 17.16 2.71 12.2 1.45 0.47 Balance
Fig. 1. (a-c) Schematic of the printing strategy. (d) Dimensions of samples used for tensile testing.
Table 2
Main manufacturing process parameters.
Parameters Values
Laser power, W 275
Scanning speed, mm/s 960
laser spot,
μ
m 80
Layer thickness,
μ
m 40
Hatch distance,
μ
m 110
Scanning strategy 67
Inert gas Argon
L. Jinlong et al.
Journal of Manufacturing Processes 106 (2023) 363–369
365
the required printing directions (BD and ND) in Fig. 1d. Uniaxial tensile
tests for BD and ND L-PBFed specimens were performed on a MTS sys-
tem with an engineering strain rate of 1 ×10
4
s
1
. Before the electron
backscattered diffraction (EBSD) tests, all the specimens were wet
ground, polished with diamond paste to 1
μ
m, and subsequently
vibrated with colloidal alumina (0.05
μ
m) for 4 h. The microstructures
of BD and ND L-PBFed 316L SSs were analyzed by TESCAN MIRA3
scanning electron microscopy.
The specimens welded with Cu wire were embedded in an epoxy
resin (exposed area of 10 mm ×10 mm) for electrochemical test. All
corrosion test evaluations were conducted in borate buffer solution and
0.1 M NaCl solution. A three-electrode CHI 660E electrochemical
equipment was used, consisting of L-PBFed 316L SS as the working
electrode, a large Pt sheet as counter electrode and a saturated calomel
electrode as reference electrode. The electrochemical impedance spec-
troscopy (EIS) was started with an excitation signal amplitude of 5 mV
and a frequency ranging from 10
5
Hz to10
2
Hz. The donor and acceptor
values (defect points densities) in surface passive lms formed on L-
PBFed 316L SSs, determined by the MottSchottky analysis, were ob-
tained by the detection system and calculation. The capacitance (C) of
space charge layer in the depletion region varied with applied potential
(E) according to the following Eq. (1) [33]:
C2=C2
H+C2
SC =±2
ε
S
ε
0qNA,D(EEfb kT
e)(1)
where
ε
s (12) is permittivity of the passive lm,
ε
0 is vacuum dielectric
constant (8.85 ×10
12
F m
1
), e is electron charge (1.6 ×10
19
C), N
D
and N
A
are donor and acceptor densities (cm
3
). E
fb
is at band potential
(V), k is Boltzmann constant (1.38 ×10
23
J K
1
) and T is absolute
temperature (K). N
D
and N
A
both were calculated from tting two slopes
of linear range in the Mott-Schottky relationships.
3. Results and discussion
The microstructures of the ND and BD L-PBFed 316L SSs, including
their grain orientation and the grain-boundary angles, were compared
using inverse pole gures (IPFs) in Fig. 2a and b. The ND face presented
more slender grains than BD face, and average grain size of ND L-PBFed
316L SS and BD L-PBFed 316L SS is 32.5
μ
m and 53.4
μ
m, respectively.
The grain size of ND L-PBFed 316L SS is smaller than that of BD one. This
was attributed to the greater temperature gradient in the build direction
than in the two other directions during the laser melting process. In
addition, LAGBs (56.3 %) in ND L-PBFed 316L SS were less than those
(73.7 %) in BD one. This should be attributed to different heat con-
duction in the different fabrication directions [34,35]. Large grain with
<101>orientation was formed in ND L-PBFed 316L SS, and such grain
morphology was also reported for L-PBFed 316L SS [36]. However, more
<111>orientations were formed in BD L-PBFed 316L SS in Fig. 2b.
The tensile curves of the ND and BD L-PBFed 316L SSs are displayed
in Fig. 3. ND L-PBFed 316L SS shows lower ultimate tensile strength and
yield strength than BD one. Furthermore, the former shows the smaller
elongation after fracture than the later. Güden et al. [37] found that the
variations in the strength and ductility of L-PBFed 316L SSs with
different building directions should be attributed to different ber
texture orientations. It was demonstrated that high hardness of L-PBFed
316L SS should be attributed to a large number of the LAGBs [38],
nanotwins and very ne
α
-martensite [39]. Comparing with traditional
wrought 316L SS, the higher yield and tensile strength of two L-PBFed
316L SSs should come predominantly from the rened grain and the
high density dislocations [40].
Comparing Fig. 4a with Fig. 4b, it can be seen that deformed degree
of grains in BD L-PBFed 316L SS is more evident than that in ND one due
to larger elongation of the former. The ratio of LAGBs of ND and BD L-
PBFed 316L SSs after the fracture is 28.4 % and 57.2 %, respectively.
More LAGBs formed in BD L-PBFed 316L SS should be attributed to the
accumulation and rearrangement of dislocations, and ne grains with
high density dislocations were achieved [41]. More LAGBs during the
deformed processing in BD L-PBFed 316L SS impeded the dislocation
movement and improved the strengthening due to synergistic effect of
dislocation strengthening and grain boundary strengthening. In addi-
tion, more deformation twins are activated in BD L-PBFed 316L SS (8.5
%) compared to ND one (0.6 %). The <111>grain orientation facili-
tated to induce more deformation twins in BD L-PBFed 316L SS [42].
Sun et al. [43] found that higher ratio of <011>textured microstructure
Fig. 2. The IPFs of (a) ND and (b) BD L-PBFed 316L SSs.
Fig. 3. Engineering stressstrain curves for ND and BD L-PBFed 316L SSs.
L. Jinlong et al.
Journal of Manufacturing Processes 106 (2023) 363–369
366
facilitated to form more nano-twins and enhanced the strength and
ductility of L-PBFed 316L SS. Twinning-induced plasticity (TWIP)
resulted into higher tensile elongation of BD L-PBFed 316L SS [44].
Furthermore, it was suggested that hetero-deformation process also
enhanced the strengthening of L-PBFed 316L SS [45]. Some
ε
-martens-
ites were found near the twins in additively manufactured 316L SS
during necking due to decreased stacking fault energy during defor-
mation [46]. The absence of strain-induced
α
-martensite trans-
formation was clearly observed in Fig. 5a and b. There is only a small
amount of strain-induced
α
-martensite (1.2 %) in BD L-PBFed 316L SSs
in Fig. 5b. Evidently, the phenomenon which involved the deformation-
induced transformation following sequence of γ-austenite
ε
-
martensite
α
-martensite was very weak during the tensile process.
Thus, in both L-PBFed 316L SS samples, the deformation behavior was
predominantly governed by dislocation slips and twins. These disloca-
tions induced misorientations can be measured by the EBSD method.
Therefore, the KAM value quanties the averaged local misorientation
in EBSD and can be used to estimate the geometric dislocation density
(GND) in deformed L-PBFed 316L SSs using the formula:
ρ
GND
=
α
KAMavg/
μ
b [47], where
ρ
GND
is the density of the GNDs, and KAMavg is
the average KAM value of the selected area.
μ
and b are the step size (
μ
m)
in the EBSD processing and the magnitude of Burgers vector (nm). The
value of constant (12) depends on dislocation type. The more disloca-
tions are concentrated near the grain boundary in ND L-PBFed 316L SS
Fig. 4. IPF map and grain boundaries distribution for after the fracture for (a) ND and (b) BD L-PBFed 316L SSs.
Fig. 5. (a, b) Phase maps and (a
, b
) KAM maps after the fracture for (a, a
) ND and (b, b
) BD L-PBFed 316L SSs
L. Jinlong et al.
Journal of Manufacturing Processes 106 (2023) 363–369
367
in Fig. 5a, while most of dislocations are formed inside smaller grains in
BD L-PBFed 316L SS in Fig. 5b. Evidently, BD L-PBFed 316L SS has
higher dislocation density than ND one. The high density of dislocations
existed in cellular microstructures in the grains due to the high cooling
rates (10
3
10
8
K/s) during the L-PBF process [48]. It was found that the
cellular dislocation structure could inhibited the strain-induced
α
-martensite transformation of austenite at cryogenic temperature,
which resulted in excellent combination of strength and ductility in L-
PBFed 316L stainless steel [49]. This is consistent with current research.
The high density of dislocations at the cellular boundary and LAGBs
should be contributed to the high strength of BD L-PBFed 316L SS due to
the interfacial strengthening effect by inhibiting the dislocation motion
[50,51]. Lower elongation of ND L-PBFed 316L SS should be attributed
to less dislocation glides in the grain interiors, while the higher uniform
elongation of BD L-PBFed 316L SS should be correlated to the TWIP,
improving working-hardening ability [52].
Fig. 6a and b shows the potentiodynamic polarization curves for ND
and BD L-PBFed 316L SSs in borate buffer solution and 0.1 M NaCl so-
lution. No evident changes of corrosion potentials of two samples were
observed under different orientation faces in two solutions. However,
BD L-PBFed 316L SS exhibited a lower passive current than ND one in
two solutions. In general, passive current reects the dissolution rate of
the passive lm during the corrosion process [53], and its larger passive
current value in ND L-PBFed 316L SS indicates its higher dissolution
Fig. 6. The polarization evolution of ND and BD L-PBFed 316L SSs in (a) borate buffer solution and (b) 0.1 M NaCl solution.
Fig. 7. (a, c) Nyquist and (b, d) Bode plots of two L-PBFed 316L SSs in (a, b) borate buffer solution after passivation at 0.6 V
SCE
for 1 h and in (c, d) 0.1 M NaCl
solution after passivation at 0.1 V
SCE
for 1 h.
L. Jinlong et al.
Journal of Manufacturing Processes 106 (2023) 363–369
368
rate, especially in 0.1 M NaCl solution. This demonstrated that BD L-
PBFed 316L SS exhibited higher corrosion resistance than ND one in the
two solutions. Karimi et al. [54] also found that the corrosion current
density of L-PBFed 316L SS was about one-third of wrought counterpart.
The Nyquist and Bode plots acquired for two L-PBFed 316L SSs in the
both solutions are presented in Figs. 7a-d. The radius of the capacitive
arc in the Nyquist plots is usually used to evaluate anticorrosion ability
of the surface passive lms [55]. Therefore, the corrosion resistance of
BD L-PBFed 316L SS was higher than ND one, as shown in Fig. 7a and c.
Negative phase angle values near 85indicate a main capacitive sur-
face layer in Fig. 7b and d. Thus, the results above indicated that BD L-
PBFed 316L SS showed higher corrosion resistance than ND one, which
agrees with the potentiodynamic polarization results.
Fig. 8a depicts the Mott-Schottky analysis results for the passive lms
formed on two L-PBFed 316L SSs in 0.1 M NaCl solution. The negative
and positive slopes indicate p-type and n-type semiconductor. Mac-
Donald et al. [56] demonstrated that the cation vacancies were the main
point defect for the contribution of p-type semiconductor, while inter-
stitial cations and oxygen vacancies were the main point defect for the
contribution of n-type semiconductor. In Fig. 8b, the density of pre-
dominant point defects in the passive lm on BD L-PBFed 316L SS is
lower than that on ND one based on the calculated results from the
slopes of the Mott-Schottky analyses. Comparing the pitting corrosion
attacks after polarization test in 0.1 M NaCl solution in Fig. 9a and b,
obviously, the extent of attack of BD L-PBFed 316L SS is much less than
that of ND one due to less pitting corrosion of the former. Sander et al.
[57] noted that the protective potential decreased with the increasing
porosity. Kong et al. [58] emphasized that some defects such as melt
pool boundaries and nonequilibrium phases decreased the corrosion
resistance of L-PBFed 316L SS. Sub-surface enclosed pores were partic-
ularly vulnerable for pitting attack, while the cell boundaries remained
relatively intact due to the segregation of corrosion resistant elements
along the boundaries [59]. The residual compressive stress induced by
the L-PBF also could promote the passivation process and decreased
driving force of the repassivation process [27]. This will undoubtedly
improve pitting resistance of 316L SS. It was demonstrated that the ne
cellular and columnar structures improved the durability of surface
passive lms formed on 316L SS obtained by severe plastic deformation
in the corrosive media [60]. However, through microstructures, elec-
trochemical measurement analyses and pitting corrosion attack char-
acteristics in this study, larger grain decreased the dissolution of BD L-
PBFed 316L SS. Moreover, the more densely packed crystallographic
faces of (111) orientation in BD L-PBFed 316L SS resulted in its higher
corrosion resistance in borate buffer solution and sodium chloride so-
lution [61].
4. Conclusions
The tensile behaviors and corrosion characteristics of ND and BD L-
PBFed 316L SSs were investigated. It was found that BD L-PBFed 316L
SS exhibited higher strengthening, larger elongation and better corro-
sion resistance than ND one. Higher yield strengthening and tensile
strengthening of BD L-PBFed 316L SS were attributed to more LAGBs
and higher density dislocation during tensile process. Better plasticity of
BD L-PBFed 316L SS should be attributed to synergetic interaction be-
tween deformed twins and dislocations. The tangled dislocation inter-
acted with deformation twins, which induced superior work hardening
capability and ductility of BD L-PBFed 316L SS. The BD L-PBFed 316L SS
showed less pitting pits than one in ND one in sodium chloride solution,
which was attributed to more compact passive lm formed on the sur-
face of the former. Comparing with ND L-PBFed 316L SS, the larger
grain and more densely packed crystallographic faces in BD L-PBFed
316L SS induced higher corrosion resistance in borate buffer solution
and sodium chloride solution. Very small amounts of deformation twins
Fig. 8. (a) The Mott-Schottky plots after being passivated at 0.1 V
SCE
for 1 h in 0.1 M NaCl solution for two L-PBFed 316L SSs, (b) donor and acceptor concentrations
in the passive lms.
Fig. 9. Optical microscopy images after polarization in 0.1 M NaCl solution for (a) ND and (b) BD L-PBFed 316L SSs.
L. Jinlong et al.
Journal of Manufacturing Processes 106 (2023) 363–369
369
and
α
-martensites were generated in BD L-PBFed 316L SS after tensile
strain. Whether more deformation twins and
α
-martensites can be
induced by the grain renement technique. This is a problem worth
studying in the future for improving mechanical properties of L-PBFed
316L SS. Therefore, Current research revealed the strengthening
mechanism and corrosion optimization mechanism of ND and BD L-
PBFed 316L SSs, which provided theoretical guidance for the applica-
tion of new technology in metallic materials.
CRediT authorship contribution statement
Authors Lv Jinlong and Zhou Zhiping: Planning and performing the
experiments, investigation, analysis, writing original draft. All other
authors: Reviewing and editing.
Declaration of competing interest
The authors declare that they have no known competing nancial
interests or personal relationships that could have appeared to inuence
the work reported in this paper.
Acknowledgment
This work was supported by The Chengdu Xianhe Semiconductor
Technology Co., Lt., China, No. 45000-71020047; and The Chengdu
Rihe Xianrui Technology Co., Ltd., China, No. 45000-71020048.
References
[1] Yan FK, Liu GZ, Tao NR, Lu K. Acta Mater 2012;60:105971.
[2] Shaikh H, Amirthalingam R, Anita T, Sivaibharasi N, Jaykumar T, Manohar P, et al.
Corros Sci 2007;49:74065.
[3] Sutton AT, Kriewall CS, Karnati S, Leu MC, Newkirk JW. Addit Manuf 2020;32:
100981.
[4] Yusuf SM, Chen Y, Yang SF, Gao N. Mater Charact 2020;159:110012.
[5] Niendorf T, Leuders S, Riemer A, Richard HA, Trter T, Schwarze D. Metall Mater
Trans B Process Metall Mater Process Sci 2013;44(4):7946.
[6] Montero-Sistiaga ML, Godino-Martinez M, Boschmans K, Kruth JP, Humbeeck JV,
Vanmeensel K. Addit Manuf 2018;23:40210.
[7] Larimian T, Kannan M, Grzesiak D, AlMangour B, Borkar T. Mater Sci Eng A 2020;
770:38455.
[8] Liu JW, Song YN, Chen CY, Wang XB, Li H, Zhou CA, et al. Mater Des 2020;186:
108355.
[9] Salman OO, Brenne F, Niendorf T, Eckert J, Prashanth KG, Hea T, et al. J Manuf
Process 2019;45:25561.
[10] Hong YJ, Zhou CS, Zheng YY, Zhang L, Zheng JY, Chen XY, et al. Mater Sci Eng A
2019;740741:4206.
[11] Wang C, Lin X, Wang LL, Zhang SY, Huang WD. Mater Sci Eng A 2021;815:141317.
[12] Kale AB, Singh J, Kim BK, Kim DI, Choi SH. J Mater Res Technol 2020;9:886783.
[13] Riabov D, Leicht A, Ahlstrm J, Hryha E. Mater Sci Eng A 2021;822:141699.
[14] Li Z, He B, Guo Q. Scripta 2020;177:1721.
[15] Li MM, Zhang X, Chen WY, Byun TS. J Nucl Mater 2021;548:152847.
[16] Yakout M, Elbestawi MA, Veldhuis SC. J Mater Process Technol 2019;266:
397420.
[17] Yakout M, Elbestawi MA, Veldhuis SC. Addit Manuf 2018;24:40518.
[18] Sun ZJ, Tan XP, Tor SB, Yeong WY. Mater Des 2016;104:197204.
[19] Yakout M, Elbestawi MA, Veldhuis SC. Int J Adv Manuf 2018;95:195374. Mar.
[20] Liverani E, Toschi S, Ceschini L, Fortunato A. J Mater Process Technol 2017;249:
25563.
[21] Dixit S, Liu SY, Murdoch HA, Smith PM. Mater Sci Eng A 2023;880:145308.
[22] Otto M, Pilz S, Gebert A, Kühn U, Hufenbach J. Metals 2021;11:944.
[23] R. S. Thanumoorthy, J. K. Chaurasia, V. A. Kumar, P. I. Pradeep, A. S. S. Balan, B.
Rajasekaran, A. Sahu, S. Bontha, J Mater Eng Perform, online.
[24] Kong DC, Dong CF, Ni XQ, Zhang L, Luo H, Li RX, et al. Appl Surf Sci 2020;504:
144495.
[25] Zhou C, Hu S, Shi Q, Tao H, Song Y, Zheng J, et al. Corros Sci 2020;164:108353.
[26] Sander G, Thomas S, Cruz V, Jurg M, Birbilis N, Gao X, et al. J Electrochem Soc
2017;164:2507.
[27] Cruz V, Chao Q, Birbilis N, Fabijanic D, Hodgson PD, Thomas S. Corros Sci 2020;
164:108314.
[28] Atapour M, Wang XY, F¨
arnlund K, Wallinder IO, Hedberg Y. Electrochim Acta
2020;354:136748.
[29] Zhao CL, Bai YC, Zhang Y, Wang XP, Xue JM, Wang H. Mater Des 2021;209:
109999.
[30] Lin KJ, Gu DD, Xi LX, Yuan LH, Niu SQ, Lv P, et al. Int J Adv Manuf Technol 2019;
104:266979.
[31] Sun SH, Ishimoto T, Hagihara K, Tsutsumi Y, Hanawa T, Nakano T. Scr Mater 2019;
159:8993.
[32] Shahriari A, Khaksar L, Nasiri A, Hadadzadeh A, Amirkhiz BS, Mohammadi M.
Electrochim Acta 2020;339.
[33] Liu Q, Zhou B, Zhang JT, Zhang WF, Zhao MF, Li N, et al. Corros Sci 2022;207:
110569.
[34] Liu YD, Zhang M, Shi WT, Ma YY, Yang J. Opt Laser Technol 2021;138:106872.
[35] Ma JW, Zhang BC, Fu Y, Hu XJ, Cao XF, Pan ZM, et al. Corros Sci 2022;201:
110257.
[36] Sander G, Babu AP, Gao X, Jiang D, Birbilis N. Corros Sci 2021;179:109149.
[37] Güden M, Yavas¸ H, Tanrıkulu AA, Tas¸demirci A, Akın B, Enser S, et al. Mater Sci
Eng A 2021;824:14180.
[38] Tekumalla S, Selvarajou B, Raman S, Gao SB, Seita M. Mater Sci Eng A 2022;833:
142493.
[39] Wang ZD, Sun GF, Lu Y, Chen MZ, Bi KD, Ni ZH. Surf Coat Technol 2020;385:
125403.
[40] Ghosh S, Bibhanshu N, Suwas S, Chatterjee K. Mater Sci Eng A 2021;820:141540.
[41] Wang CL, Yu DP, Niu ZQ, Zhou WL, Chen GQ, Li ZQ, et al. Acta Mater 2020;200:
10115.
[42] Bahl S, Mishra S, Yazar KU, Kola IR, Chatterjee K, Suwas S. Addit Manuf 2019;28:
6577.
[43] Sun ZJ, Tan XP, Tor SB, Chua CK. NPG Asia Mater 2018;10:12736.
[44] Kwok TWJ, Gong P, Rose R, Dye D. Mater Sci Eng A 2022;855:143864.
[45] Kong D, Dong C, Ni X, Liang Z, Man C, Li X. Mater Res Lett 2020;8:3907.
[46] Jaskari M, Ghosh S, Miettunen I, Karjalainen P, J¨
arvenp¨
a¨
a A. Materials 2021;14:
5809.
[47] Mur´
ansky O, Balogh L, Tran M, Hamelin CJ, Park J-S, Daymond MR. Acta Mater
2019;175:297313.
[48] Vilaro T, Colin C, Bartout JD, Naz´
e L, Sennour M. Mater Sci Eng A 2012;534:
44651.
[49] Karthik GM, Kim ES, Sathiyamoorthi P, Zargaran A, Jeong SG, Xiong RL, et al.
Addit Manuf 2021;47:102314.
[50] Pinomaa T, Lindroos M, Walbrühl M, Provatas N, Laukkanen A. Acta Mater 2020;
184:116.
[51] Hong YJ, Zhou CS, Zheng YY, Zhang L, Zheng JY. Mater Sci Eng A 2021;799:
140279.
[52] Chen YN, Li B, Chen B, Xuan FZ. Addit Manuf 2023;61:103319.
[53] Yang GM, Du YF, Chen SY, Ren YS, Ma YL. J Mater Res Technol 2021;15:682840.
[54] Shaeri Karimi MH, Yeganeh M, Alavi Zaree SR, Eskandari M. Opt Laser Technol
2021;138:106918.
[55] de Castro Gir˜
ao D, B´
ereˇ
s M, Jardini AL, Filho RM, Silva CC, de Siervo A, et al. Mater
Sci Eng C 2020;107:110305.
[56] Macdonald DD. J Electrochem Soc 1992;139:343449.
[57] Sander G, Thomas S, Cruz V, Jurg M, Birbilis N, Gao X, et al. J Electrochem Soc
2017;164:C2507.
[58] Kong DC, Ni XQ, Dong CF, Zhang L, Man C, Yao JZ, et al. Electrochim Acta 2018;
276:293303.
[59] Wang GQ, Liu Q, Rao H, Liu HC, Qiu CL. J Alloys Compd 2020;831:154.
[60] Muley SV, Vidvans AN, Chaudhari GP, Udainiya S. Acta Biomater 2016;30:40819.
[61] Lv JL, Luo HY. Appl Surf Sci 2013;280:12431.
L. Jinlong et al.
... The microstructure and porosity of built parts have a strong influence on their mechanical performance. The variations of L-PBF parameters on microstructural characteristics and mechanical properties are still debatable [5]. Niendorf et al. found that a higher laser power promoted the induction of strong crystallographic textured columnar coarse-grains, which also significantly improved the ductility of L-PBFed 316L SS [6]. ...
Article
Full-text available
316L stainless steel is widely used in various industrial fields, but its strength is relatively low. The improvement of its strength has become a research hotspot. In this study, nano titanium carbide (TiC) particles are ball milled with 316L with the addition of 2 wt% and 4 wt%. The composite powder was then used for the fabrication of samples by laser powder bed fusion. The results show that the TiC is uniformly distributed in the microstructure. With the addition of TiC, the average size of the grains is significantly reduced. The strength, hardness, and wear resistance of TiC/316L samples have been greatly improved. The tensile strength of formed 2 wt% TiC/316L is 948 MPa, together with a extension rate of 36.0%, which has been increased by 42.6% and 79.7%, respectively. This study provides an effective way to improve the strength at room temperature and the high temperature of 316L built by laser powder bed fusion.
Conference Paper
This paper explores the potential of N08904 as an alternative alloy to Ni-based alloys in the Oil & Gas industries by characterizing the mechanical and corrosion properties of Laser Beam Powder Bed Fusion (PBF-LB) manufactured samples and comparing the results against rolled flat material of the same alloy. As of today, Oil & gas industries are slowly adopting AM technologies to produce some of their parts. However, regarding material selection, the possibilities are scarce, as the investigations are focused on proving the feasibility of different AM technologies using mainly highly alloyed Ni-based materials or standard stainless steels such as S31603.
Article
Full-text available
A medium Mn steel of composition Fe-4.8Mn-2.8Al-1.5Si-0.51C (wt.%) was processed to obtain two different microstructures representing two different approaches in the hot rolling mill, resulting in equiaxed vs. mixed equiaxed and lamellar microstructures. Both were found to exhibit a simultaneous Twinning Induced Plasticity and Transformation Induced Plasticity (TWIP+ TRIP) mechanism where deformation twins and α′-martensite formed independently of twinning with strain. Interrupted tensile tests were conducted in order to investigate the differences in deformation structures between the two microstructures. A constitutive model was used to find that, surprisingly, twinning contributed relatively little to the strength of the alloy, chiefly due to the fine initial slip lengths that then gave rise to relatively little opportunity for work hardening by grain subdivision. Nevertheless, with lower high-cost alloying additions than equivalent Dual Phase steels (2-3 wt% Mn) and greater ductility, medium-Mn TWIP+ TRIP steels still represent an attractive area for future development.
Article
Full-text available
Additive manufacturing (AM) is an emerging fabrication technology that offers unprecedented potential for manufacturing end-to-end complex shape customized products. However, building products with high performance by AM presents a technological challenge. Inadequate processing parameters, fabrication environment or changes in powder properties may lead to high defect density in the part and poor mechanical properties. Microstructure, defect structure, and mechanical properties of AISI 316L stainless steel pieces, additively manufactured by the laser powder bed fusion method using three different volume energy densities (VEDs), were investigated and compared with those of a commercial wrought AISI 316L sheet. Scanning and transmission electron microscopies were employed for characterization of grain and defect structures, and mechanical properties were determined by tensile testing. It was found that the number of defects such as pores and lack of fusion in AM specimens did not affect the strength, but they impaired the post-uniform elongation, more significantly when processed with the low VED. Twinning was found to be an active deformation mechanism in the medium and high VED specimens and in the commercially wrought material in the later stage of straining, but it was suppressed in the low VED specimens presumably because the presence of large voids limited the strain attained in the matrix.
Article
Full-text available
In-depth understanding of corrosion behaviour is a key aspect regarding the application of additively manufactured parts. In this study, 316L stainless steel was manufactured under different scanning strategies using selective laser melting (SLM). Microstructure characterization and electrochemical tests in NaCl aqueous solution (3.5 wt%), including open circuit potential (OCP), potentiodynamic polarization, and electrochemical impedance spectroscopy (EIS), were conducted to study the influence of scanning strategies on the corrosion behaviour. The microstructure and corrosion on different planes were characterized to reveal the influence of building direction. EBSD analysis shows that the scanning strategy affects the continuity of grain growth through adjacent layers and the growth of grains inside the melt track. Electrochemical tests indicate a clear difference in corrosion resistance perpendicular and parallel to building direction and with different scanning strategies. Pitting corrosion is the main form of corrosion in SLM 316L stainless steel and preferentially initiates on molten pool boundaries.
Article
Full-text available
In the last decade, additive manufacturing technologies like laser powder bed fusion (LPBF) have emerged strongly. However, the process characteristics involving layer-wise build-up of the part and the occurring high, directional thermal gradient result in significant changes of the microstructure and the related properties compared to traditionally fabricated materials. This study presents the influence of the build direction (BD) on the microstructure and resulting properties of a novel austenitic Fe-30Mn-1C-0.02S alloy processed via LPBF. The fabricated samples display a {011} texture in BD which was detected by electron backscatter diffraction. Furthermore, isolated binding defects could be observed between the layers. Quasi-static tensile and compression tests displayed that the yield, ultimate tensile as well as the compressive yield strength are significantly higher for samples which were built with their longitudinal axis perpendicular to BD compared to their parallel counterparts. This was predominantly ascribed to the less severe effects of the sharp-edged binding defects loaded perpendicular to BD. Additionally, a change of the Young's modulus in dependence of BD could be demonstrated, which is explained by the respective texture. Potentiodynamic polarization tests conducted in a simulated body fluid revealed only slight differences of the corrosion properties in dependence of the build design.
Article
Build orientation-induced mechanical anisotropy is critical for additive manufacturing metals. To leverage the existing research on anisotropy between Z and X/Y build orientations, this work focuses on the anisotropic behavior within 2D planes. Therefore, tensile coupons of 316L stainless steel (316L-SS) were fabricated in five different build orientations (X, Y, XY45°, Z, and ZX45°) using the laser powder bed fusion process with a pulsed laser. Different characterization techniques at various length scales were employed to investigate the mechanisms of mechanical anisotropy. According to the electron back-scattered diffraction (EBSD) study, X and Y samples had a larger fraction of <001> and <110> grains, whereas XY45° coupons had a larger fraction of <111> grains along the tensile axis. In contrast, Z samples exhibited a dominant <110> texture parallel to the tensile axis. X-ray diffraction analysis revealed that all the samples had a single austenitic phase but different dislocation densities. The tensile results of all samples showed higher yield strength and comparable ductility with those in the literature. Among all the build orientations, XY45° coupons had the highest yield strength due to the large fraction of <111> grains oriented along the tensile axis and the highest dislocation density of 2.2 × 1015 m−2. In contrast, the highest ductility of Z samples was caused due to twinning-favored <110> crystallographic texture along the tensile axis. The onset and termination of different stages of strain hardening were mainly affected by the cellular sub-grain structure and crystallographic texture, as well as their interactions with dislocation generation and evolution. These findings indicated that crystallographic texture and dislocation density were the dominant contributions to the mechanical anisotropy followed by grain morphology.
Article
Owing to the rapid cooling rates and directional thermal gradients involved, fusion-based additive manufacturing (AM) processes yield complex, fine solidification structures that impart anisotropic mechanical properties in materials, such as stainless steel 316L (SS316L). In this work, we present a comprehensive study of the mechanical anisotropy of SS316L produced by laser powder bed fusion using instrumented nanoindentation. We produce and test near-single crystal samples oriented along the three principal crystallographic axes—namely 〈100〉, 〈110〉, and 〈111〉 using a Berkovich indenter. We find that the 〈111〉 and 〈100〉 orientations exhibit the highest and the lowest hardness, respectively. To decouple the contributions of grain orientation and solidification structure to the alloy's mechanical anisotropy, we compare our experimental results against crystal plasticity finite element (CPFE) simulations. We ascribe the hardness anisotropy in SS316L to the cell spacing along the slip direction (CSSD), which is a novel metric that we introduce to account for the role of the solidification structure on plasticity. Our work provides a universal pathway to understanding the microstructure-property relationships in cubic metallic materials that exhibit solidification structures—such as those commonly imparted by fusion-based AM.
Article
The effects of secondary passivation on the electrochemical behavior of passive films on the surface of 2205 duplex stainless steel were studied by potentiodynamic polarization, EIS and Mott-Schottky analysis. The composition and surface morphology of the passive film were analyzed by ICP-MS, ARXPS and SEM. These results manifested that the passive film formed by secondary passivation has good pitting resistance, due to the increase of the content of oxidized Fe³⁺, Cr³⁺ and Mo⁴⁺ in the passive film, which improves and increases the dissolution and migration rate of metal cations in the passive film, thus improving the compactness and stability of the passive film. The Mott-Schottky plots demonstrated that the five kinds of passive films have different semiconductor properties. The estimated doping density shows that the donor density of the passive film was significantly reduced by secondary passivation. In addition, Cl⁻ and SO4²⁻ are permeated in the passivation layer and their adsorption on the surface of the passive film can be inhibited by the secondary passivation.
Article
The effect of specimen inclination angle with respect to building direction on the tensile properties of a selective laser melt 316L alloy was investigated. Tensile test specimens were fabricated with the angles between 0° to 90° at 15° intervals using a rotation scanning. In addition, 316L alloy test specimens were generated in the ANSYS 2020R1 additive module and tensile tested in LS-DYNA in order to determine the effect of residual stresses on the tensile strengths. The microscopic analysis revealed a strong 〈110〉 fiber texture orientation along the building direction (the loading axis of 0° inclined specimens) and a weak <111> texture or nearly random distribution of directions in the normal to the building direction (tensile loading axis of 90° inclined specimens). The yield and tensile strength increased and ductility decreased with increasing inclination angle. The strength variation with the inclination angle was shown well-fitted with the Tsai-Hill failure criterion. Although, the used numerical models indicated an inclination-dependent residual stress, the difference in the residual stresses was much lower than the difference in the strengths between 0° and 90° inclined specimens. Predictions showed a lower twinning stress in 0° inclined specimens due to 〈110〉 fiber texture orientation in the tensile axis. The fiber texture resulted in extensive twinning; hence, higher ductility and tension-compression asymmetry in 0° inclined specimens. Based on these results, the variations in the strength and ductility of tested SLM-316L specimens with the inclination angle was ascribed to the variations in the angle between the fiber texture orientation and loading axis.
Article
Of the many benefits of the additive manufacturing process, laser powder bed fusion (L-PBF) has specifically been shown to produce hierarchical microstructures that circumvent the common strength-ductility trade-off. Typically, high strength materials have limited ductility, and vice versa. The L-PBF microstructure, consisting of fine cells, is formed during the rapid solidification of the laser powder bed fusion process. The cell boundaries are often characterized by the segregation of alloying elements and a dislocation network. While there are a number of works describing the strengthening mechanisms in L-PBF-produced 316L, there are still some gaps in understanding the effect of stress-relief and annealing at various annealing temperatures (400, 800 and 1200 °C) on the plastic strain accumulation during deformation. In this study, the authors evaluated strain partitioning using electron backscatter diffraction and kernel average misorientation maps. The results show strain partitioning to be dependent on both the annealing temperature and the pre-straining of samples. Further, the results indicated that the dislocation structure was stable until 400 °C, whereas at 800 °C strain was no longer detected at the cell boundaries. Similarly, after the heat treatment at 800 °C, elemental segregation at the cell walls was no longer detectable. Upon straining, the boundaries of as-built and annealed samples at 400 and 800 °C registered accumulation of additional strain as compared to the unstrained states. The results demonstrate that even a weak array of dislocations along the cell walls can successfully pin dislocations, albeit at a reduced capability relative to the co-existent dislocation and segregate structures found in microstructures of the as-built and annealed samples at 400 °C.
Article
Surface severe plastic deformation (S2PD) of additively manufactured/three-dimensional (3D) printed metallic parts is gaining increased attention as a post-manufacturing operation to enhance the material performance in a wide variety of applications. Surface mechanical attrition treatment (SMAT) is an S2PD technique that can yield a nanostructured surface layer induced by compressive stresses and work hardening. In the present study, SMAT was performed on 316L (austenitic) stainless steel (SS) processed by selective laser melting (SLM), and the consequent effects on mechanical response were investigated. The underlying mechanisms of microstructural evolution leading to the formation of nanocrystalline grains resulting from SMAT in SLM 316L SS are elucidated. The interactions between twins and deformation bands act as potential sites for impeding the movement of dislocations, which in turn leads to the formation of stacking faults, twinning, and occasionally transform to a different crystal structure. Twin-twin and/or twin-deformation band intersections sub-divide the matrix grains into smaller cells or low-angle disoriented blocks, which result in the formation of low-angle grain boundaries and finally in nanocrystallization at the surface. The size of nanocrystalline grains increases progressively with depth from the surface to micrometer size grains in bulk. The gradient nanostructure in the additively manufactured alloy after SMAT imparts an unusual combination of strength and ductility that markedly exceeds that of conventional, bulk nanostructured, or even high-performance 316L SS (containing nanoscale deformation twins embedded in submicron-sized austenitic grains obtained by dynamic plastic deformation processes). Analytical models revealed that strengthening results from a combination of grain boundaries and dislocations. The results of the present investigation pave the way for engineering high-performance SS for a variety of engineering applications.