Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Additive Manufacturing 40 (2021) 101902

Contents lists available at ScienceDirect

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

Research Paper

The effect of build direction and heat treatment on atmospheric stress


corrosion cracking of laser powder bed fusion 316L austenitic stainless steel
P. Dong a, F. Vecchiato a, Z. Yang a, P.A. Hooper b, M.R. Wenman a, *
a
Department of Materials and Centre for Nuclear Engineering, Imperial College London, London SW7 2AZ, UK
b
Department of Mechanical Engineering, Imperial College London, London SW7 2AZ, UK

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

Keywords: Austenitic stainless steels are susceptible to stress corrosion cracking (SCC) when exposed to chloride ions and
Laser powder bed fusion tensile stresses. For laser powder bed fusion (LPBF) components, high residual tensile stresses exist from the large
316L thermal gradients during solidification. While post fabrication heat treatments can relieve residual stresses, their
Stress corrosion cracking
effectiveness in improving SCC resistance of LPBF material is undetermined. The effects of different heat
Chloride
treatments (i.e. temperature and cooling method) and build orientations on SCC susceptibility of LPBF 316L were
Heat treatment
explored. Samples were exposed to MgCl2 and maintained at 75 ◦ C and 70% RH for the test duration. Cracks were
highly branched and propagated perpendicularly to build direction even in the presence of an applied perpen­
dicular 60 MPa stress. Vertical as-built specimens exhibited higher cracking densities than horizontal ones, due to
higher surface residual stresses. SCC and corrosion favoured one side for horizontal specimens while the dis­
tribution was random for vertical specimens. After 750 ◦ C (2 h, furnace cooled) heat treatment, SCC susceptibility
reduced for vertical 316L but increased for horizontal 316L. This was attributed to competing effects: reduction
in residual surface tensile stresses (beneficial) and loss of non-equilibrium microstructure (detrimental).
Encapsulation to prevent oxidation during heat treatment negatively impacted SCC resistance of LPBF 316L,
likely due to changes to surface oxide condition. No SCC was observed after 900 ◦ C (0.5 h) heat treatments within
the test duration, however, pitting corrosion persisted.

1. Introduction velocities, dendritic structures similar to that of welds can be expected


[7,9]. Furthermore, the existence of cellular substructures in LPBF 316L
Laser powder bed fusion (LPBF) is a common additive manufacturing has been reported in literature, as a result of elemental redistribution
technique for the fabrication of metallic components. The main benefits during solidification, in particular segregation of Cr and Mo at the cell
of LPBF compared to conventional forging techniques are the ability to walls [10]. This can have important consequences as both Cr and Mo are
produce complex structures, lower production cost (moulds are no critical alloying additions in the corrosion and stress corrosion cracking
longer required) and material efficiency (traditional processes usually (SCC) resistance of austenitic SS [11].
require removal of excess material, while with additive manufacturing, While austenitic SS are widely regarded to possess good corrosion
near net shapes are produced) [1,2]. Compared to traditional forged resistance properties, surface corrosion can still occur in aggressive
material, additively manufactured austenitic SS possess significantly environments. The general corrosion resistance properties of printed
different microstructural and mechanical properties [3–5]. LPBF 316L 316L are usually considered superior to forged material, which has often
have been reported in literature to have lower ductility and reduced been attributed to the formation of a more compact and stable protective
material toughness than wrought material, largely as a result of residual oxide layer [12–14]. It has also been suggested that the cellular
porosity during fabrication [6]. Mechanical properties are also strongly sub-structure may aid in the general corrosion performance of LPBF
anisotropic due to the material being highly textured in the as-built state austenitic SS, where segregation of Mo and Cr at cell walls results in a
[7,8]. The solidification microstructure of additively manufactured more corrosion resistant cell wall that limits the depth of corrosion
316L is dependent on the temperature gradient and the solidification attack [7,14]. Ettefagh et al. have attributed the superior corrosion
velocity, where due to high temperature gradients and fast interface performance of LPBF 316L in chloride solutions to the lack of

* Corresponding author.
E-mail address: m.wenman@imperial.ac.uk (M.R. Wenman).

https://doi.org/10.1016/j.addma.2021.101902
Received 17 September 2020; Received in revised form 5 February 2021; Accepted 8 February 2021
Available online 20 February 2021
2214-8604/© 2021 Published by Elsevier B.V.
P. Dong et al. Additive Manufacturing 40 (2021) 101902

non-equilibrium phases, particularly martensite that can exist in


wrought austenitic SS [13].
Austenitic SS are, however, susceptible to localised corrosion pro­
cesses such as stress corrosion cracking. SCC requires a tensile stress (can
be applied or residual), a susceptible material and a specific chemical
environment [11,15,16]. In the case of transgranular SCC of austenitic
SS, chloride ions, typically found in sea-salts, are particularly aggressive
[17–19]. Due to poor sample surface finish, residual porosity and
Fig. 1. Dimensions of LPBF samples in mm. Thickness = 4 mm. [42].
inherent high residual stresses, SCC may be especially problematic in
LPBF material [12,20–22]. The high residual stresses found in LPBF
316L are likely to be sufficient at initiating SCC without any externally Table 1
applied stress [23,24]. Tensile residual stresses recorded in as-printed Printing parameters used for substrate deposition.
316L components have been reported to be as high as 350 MPa in the
Laser Laser Hatch Points Layer Inter-
build direction (dependent on printing pattern and build orientation of Power Exposure Spacing Distance Height layer
the component), which far exceeds any of the suggested stress thresholds (W) Time (μs) (μm) (μm) (μm) angle (◦ )
for SCC [25]. Corrosion pits are often regarded as a precursor to SCC and 200 80 110 60 50 67
the pitting corrosion resistance of LPBF austenitic SS has been reported
to be worse due to porosity, where pores can act as potential sites for
acidification and concentration of Cl- ions [26–29]. The size and 2. Experimental
morphology of the pores have been reported to influence pitting
corrosion of LPBF austenitic SS [12,30,31]. On the other hand, greater Tensile specimens (dimensions shown in Fig. 1) were printed using a
resistance to pit initiation has been reported by several authors for LPBF Renishaw AM250 LPBF system in an argon atmosphere (≈0.01% oxygen
316L and has been attributed to the inability for MnS inclusions content) using a continuous-wave fibre laser. The printing parameters
(regarded as a site for pit initiation) to form during rapid solidification are given in Table 1. These parameters were optimised for production of
[8]. However, LPBF 316L may still be inferior as after pit initiation, pit high-density substrates (> 99.9%). The thickness of each layer was fixed
growth rates has been reported to be higher due to a combination of at 50 µm and with every successive layer, the scan direction was rotated
porosity, chemical segregation and residual stress/plastic strain effects, by 67◦ . Samples were printed using a stripe hatching pattern as outlined
resulting in the formation of stable pits [8,32]. by the Renishaw “Design for metal AM” guide for a more homogeneous
However, literature regarding Cl- induced transgranular SCC in LPBF distribution of residual stresses [48]. The composition of the AISI 316L
austenitic SS is still rather scarce with most studies on intergranular SCC austenitic SS powder used is given in Table 1. With regards to compo­
[33]. For intergranular SCC in high temperature water environments, nents such as pipe or canisters, only one surface would be exposed to a
Lou et al. have suggested a dependence on SCC crack growth rate and corrosive environment. In this study two different build orientations
increasing porosity/pore size [34]. While MnS inclusions are not typi­ were chosen to provide two different surface microstructures for com­
cally present in LPBF 316L, preliminary studies by Lou et al. and Vec­ parison purposes. The two build orientations of samples chosen for this
chiato have reported the existence of Mn and Si rich oxide inclusions at study were: vertical (printed perpendicularly to the build plate, Fig. 2a)
the cellular level [9,35]. Lou et al. have reported that the Si-rich in­ and horizontal (printed in parallel to the build plate, Fig. 2b). Samples
clusions can act as a site for intergranular SCC initiation in high tem­ were deposited directly onto a build plate via support structures (small
perature water conditions [35]. Evidence of transgranular SCC in LPBF cylindrical struts approximately 5 mm in height) to facilitate easy
316L has been presented by Bruycker et al. under immersed boiling removal of specimens from the base plate.
MgCl2 solutions, however, this is typically regarded as a very aggressive Following removal of supports, both horizontally and vertically
environment and unlikely to be representative of atmospheric condi­ printed specimens were exposed to the various heat treatments outlined
tions [36]. The risk of SCC can be reduced through a reduction of the in Table 2 via a furnace. The samples were then either left to cool in the
residual tensile stress remaining within the fabricated LPBF components furnace or were quenched immediately in water. 750 ◦ C furnace cooled
using post-fabrication heat treatments, however there are currently a treatment was selected to allow some retention of the cellular structure
lack of established parameters for effective stress-relief in LPBF 316L while potentially avoiding sensitisation of grain boundaries. 900 ◦ C heat
[37–39]. Furthermore, due to the beneficial nature of the temperature treatment for 30 min followed by water quenching was selected for
sensitive cellular substructure, allowing for high strength and ductility, comparison as this heat treatment had been previously suggested to be
and the need to avoid component distortion, it would be beneficial to effective for stress relief and recovery in forged 304L material [42]. The
retain the as-printed microstructure during heat treatment. It has been samples were tested with the oxide scale intact to determine if such a
suggested in literature that the cellular structure begins to dissolve material would be industrially suitable: reduced processing steps would
above 600–700 ◦ C [7,38]. This is a concern as limiting heat treatment reduce manufacturing costs. Furthermore machine grinding to remove
temperatures to this range would not only be ineffective at fully the oxide layer can affect the surface tensile residual stress state and can
relieving residual stresses, but since it is also close to the sensitisation have significant impacts on SCC development [49,50]. This was then
temperature in austenitic SS, intergranular SCC may become a risk as compared with components heat treated in an inert atmosphere to avoid
well [40,41]. oxidation by air. Prior to the heat treatment, these samples were
Although SCC of forged, and to an extent welded, austenitic SS ma­ encapsulated in a glass tube that was backfilled with argon and subse­
terial have been widely studied in literature, chloride-induced SCC in quently sealed using an oxypropane torch. (Table 3).
LPBF 316L has not been addressed in significant detail [42–45]. In this After the heat treatment, the specimens were then sprayed with a
study, LPBF 316L tensile specimens that had undergone various heat MgCl2 suspension in isopropanol using an airbrush. The solvent was
treatments were exposed to MgCl2 and maintained in a corrosion evaporated on a hot plate, allowing deposition of a uniform layer of salt
chamber at 75 ◦ C and 70% relative humidity (RH) for 480 h. The chosen onto the test surface. This process was repeated until the desired salt
atmospheric SCC test conditions reflect those used by several previous loading of 0.02 g was reached. A loading of 0.02 g was chosen for all
atmospheric SCC studies on forged 304L and 316L allowing for more samples, for both consistency and for comparison to existing work on
direct comparisons of performance [24,42,46,47]. forged 304 L and 316 L material using an identical corrosion testing
setup [42,47]. Masses of the specimens before and after deposition were
determined using a Fisher Scientific MH-214 analytic scale (accuracy

2
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 2. Diagram of the LPBF samples attached to the build plate: (a) Vertically printed; (b) Horizontally printed. Indication of the build direction (BD) with respect to
the specimen geometry shown using the axes. LD = Longitudinal direction, TD = Transverse direction, ND = Normal direction. The face shown (parallel to BD) was
exposed to chloride salt for SCC testing.

Table 2
Elemental composition of 316L powder by wt%.
Grade Cr Ni Mo Mn Si Cu C P S Fe

316L 18 13 2.5 2.0 0.75 0.50 0.030 0.025 0.010 Bal.

Table 3
List of sample name abbreviations and their corresponding heat treatment conditions.
Sample Heat Treatment

Sample Abbreviation Temperature/◦ C (time/h) Encapsulation Cooling Method Applied Stress at Start of SCC Test/ MPa

AB (as-built) N/A N/A N/A 60


750FC 750 (2) No Furnace Cooled 60
750ENFC 750 (2) Yes Furnace Cooled 60
900FC 900 (0.5) No Furnace Cooled 60
900WQ 900 (0.5) No Water Quenched 60
900ENFC 900 (0.5) Yes Furnace Cooled 60
900ENWQ 900 (0.5) Yes Water Quenched 60
1250WQ (Full recrystallisation) 1250 (1) No Water Quenched 60
AB*(as-built) N/A N/A N/A 0
750FC* 750 (2) No Furnace Cooled 0
900WQ* 900 (0.5) No Water Quenched 0

Samples denoted with a * are from a 2nd test without applied stress during the corrosion experiment.

0.1 mg). The coated specimens were then mounted on bend jigs (as
shown in Fig. 3) that applied an initial upper surface tensile stress of
60 MPa parallel to LD. The jigs are designed to maintain a relatively
consistent stress intensity factor through the use of a compression
spring: as a SCC crack propagates within a sample, the sample will
experience greater bending, resulting in a reduced load being applied by
the compression spring (the increased stress intensity factor due to
greater crack length can to some degree be balanced by the decrease in
applied load) [24]. The required deflection to achieve the desired initial
tensile stress was calculated for each specimen and measured using a
Mitutoyo ABSOLUTE digimatic indicator (accuracy 0.001 mm). The
loaded jigs were then placed in a non-metallic corrosion chamber inside
a furnace, to maintain temperature at 75 ◦ C. A beaker of saturated NaCl
solution was placed alongside the jigs to maintain RH of 70% within the
corrosion chamber.
The specimens were inspected and photographed every 3–4 days for
the 20-day test duration to record the development of corrosion prod­
ucts. A second set of tests was also repeated for a limited selection of the
heat treatments but with the absence of an applied stress. Samples were
Fig. 3. Bend test jig with a loaded test specimen [42]. still loaded onto jigs, to isolate the specimens from the floor of the
corrosion chamber where salt and water could accumulate, but these
were not tightened. Since it was expected that there would be less SCC

3
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 4. Images of the LPBF components at each of the different heat treatments: (a) Vertically printed; (b) Horizontally printed. The face parallel to BD shown here
was SCC tested.

development than with the previous run, the duration of the test was 3. Results and discussion
increased to 30 days. Three samples were tested per build orientation
and heat treatment condition. It is important to note that the samples 3.1. Material characterisation
used in the second test without applied stress were from a different
printing batch to the first set with applied stress. Due to the inherent Following removal of the LPBF specimens from the base plate (by
variations between print runs, differences in properties may be expected cutting the support structures), it was observed that the horizontally
and direct comparisons between the two sets may not be representative. printed specimens had a tendency to bend, with the longer surface
For optical and secondary electron microscopy (SEM) analysis, the corresponding to the side that had been attached to the support struc­
samples were sectioned using a low deformation circular saw with an tures. This bending was likely a result of the presence of high Type I
aluminium oxide blade. The lowest feed rate of 0.005 mm s− 1 was used residual tensile stresses caused by a strong thermal gradient in the
to minimise the introduction of further residual stresses during cutting. longitudinal direction. When the support structures were removed,
Following sectioning, the samples were mounted in Bakelite and ground partial relaxation of these longitudinal stresses can occur within the
up to 4000 grit SiC paper using standard metallographic preparation samples, causing the curvature of the specimens. Large stresses parallel
procedures. To obtain a fine surface finish for electron backscatter to the scan direction have been reported in literature, and neutron
diffraction (EBSD), samples were polished using an aqueous oxidised diffraction experiments on similar specimen geometries (after support
porous silicon (OPS) suspension. Optical characterisation was per­ strut removal) have revealed possible evidence of stress redistribution,
formed using an Olympus-BXS1 optical microscope, while SEM imaging where considerable stresses were recorded for one side while stresses for
and EBSD mapping were carried out on a Zeiss Sigma 300 FEGSEM at the other side were closer to zero [39,51,52].
20 kV with a Bruker Quantax detector (step size between 1 and 2 µm The appearances of the samples after the various heat treatment
depending on grain size and acquisition time). EBSD data was processed conditions are shown in Fig. 4, where a clear variation in surface
and analysed using a combination of MTEX and ESPRIT 2 for orientation appearance can be seen. The change in the visual appearance can be
maps, grain analysis and misorientation maps. Surface cracks were prescribed to the changes in the oxide layer that forms on the specimen
recorded using optical microscopy and a crack density (number of cracks surface during heat treatment. In literature, the passivating Cr2O3 layer
per unit area) was calculated. A minimum threshold crack length of associated with stainless steels is broken down and replaced with an
50 µm was used to ensure consistency. Due to extensive branching oxide layer during heat treatment, consisting of an outer Fe2O3/Fe3O4
observed in almost all cracks, branched cracks were considered as single layer and an inner, Cr-rich spinel oxide layer (often regarded as a
cracks for crack count purposes. To quantify the degree of branching, a “barrier layer”). The thickness of this new outer oxide layer has been
crack-to-branching ratio was calculated, where the total number of reported to increase with higher temperatures/dwell times [53].
branches (threshold of 50 µm) were recorded per crack and averaged The morphology of the grains found in the 316L LPBF specimens
across each heat treatment. Hardness testing was performed, after after grinding and polishing, at the different heat treatment conditions,
grinding/polishing, on a sample corresponding to each of the different are shown using IPF Z coloured EBSD maps in Fig. 5 and Fig. 6, dis­
heat treatments and build orientations. A Vickers hardness indenter was playing the vertical and horizontal build orientations respectively. The
used with a load of 10 N for 10 s. Five surface measurements at random grains appear columnar, growing in parallel to the build direction due to
locations were recorded per sample and averaged. the associated high temperature gradients and share similarities with
the dendritic microstructure found in welds. However, unlike 316L
welds, the LPBF components are purely austenitic and do not have any
δ-ferrite phases, which is attributed to the rapid cooling rates [54]. The
columnar grain structure was present in all heat treatment conditions
except the 1250WQ treatment where full recrystallisation had been

4
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 5. EBSD maps in IPF Y colouring (parallel to BD) showing the morphology of the grains in the vertically printed samples after the following heat treatment
conditions: (a) AB; (b) 750FC; (c) 900FC; (d) 900WQ; (e) 1250WQ-recrystallised.

achieved. The recrystallised material (shown in Fig. 5e and Fig. 6e), between 45 and 50 µm in diameter. For all samples, the much larger
displayed equiaxed grain size with annealing twins. Both the vertically mean grain sizes when compared to the median reflect the large pro­
and the horizontally as-printed material are textured, as evident by the portion of small grains in the material as evident by the EBSD maps. This
pole figures in Fig. 7. is further shown by positive skew indicated by the interquartile ranges.
The majority of the grains are significantly larger than the typical In all the analysed regions, very large grains (above 10,000 µm2) were
equiaxed grains found in forged austenitic SS, although some very small observed with maximum grain sizes exceeding 30,000 µm2. The pro­
grains can still be observed. An average of the grain area and aspect portion of very large grains was between 6% and 8% for the as-built
ratios are displayed in Fig. 8 for each of the heat treatments. In general, material, which was decreased after the 750FC and 900FC heat treat­
the grains of the horizontally printed material are slightly larger than the ments to around 1–3%. The 900WQ treatment did not seem to affect
vertical orientations for the as-built and furnace cooled specimens. This proportion of very large grains significantly, remaining at between 6%
is likely due to larger scan vectors found in the horizontally printed and 7%. However, these observed trends may not be fully representative
material than the vertically built components leading to a more signif­ due to the relatively small area analysed by EBSD and the anisotropic
icant second thermal gradient [9]. There does not appear to be any grain nature of the printed material. Grain aspect ratios are relatively
growth during the furnace cooled heat treatments; instead both 750FC consistent across the different heat treatments and build orientations,
and 900FC conditions resulted in an apparent decrease of grain sizes averaging between 2.5 and 3.5, indicating a similar extent of elongation
compared to the as-built material for both print orientations. Despite the within the grains and suggesting that the reduction in grain size
900FC specimens being treated at a higher temperature, the average observed in the 750FC and 900FC specimens was not due to static
grain size is not significantly different from the 750FC specimens, likely recrystallisation of the deformed grains. The 1250WQ specimens, in
due to the shorter heating duration of the higher temperature treated both print directions, showed a lower aspect ratio of around 1.7 due to
samples. The 900WQ specimens showed similar grain sizes to the AB the nucleation and growth of equiaxed grains (presence of annealing
suggesting little if any grain scale change for this heat treatment con­ twins resulted in a slightly higher aspect ratio than 1).
dition, possibly due to the shorter heating time. The fully recrystallised The changes in the Vickers hardness of the specimens for the
material, 1250WQ, showed newly grown equiaxed grains that averaged different heat treatments are shown in Fig. 9. The as-built specimens

5
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 6. EBSD maps in IPF Y colouring (parallel to BD) showing the morphology of the grains in the horizontally printed samples after the following heat treatment
conditions: (a) AB; (b) 750FC; (c) 900FC; (d) 900WQ; (e) 1250WQ-recrystallised.

Fig. 7. Pole figures of the horizontally and vertically printed AB material showing texture. Scale is multiples of random texture.

6
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 8. Grain properties of 316L LPBF specimens exposed to different heat treatment: (a) Median and mean grain area, bars show the upper and lower quartiles; (b)
Mean grain aspect ratio, error bars show standard deviations.

Fig. 9. Hardness of 316L LPBF specimens exposed to different heat treatments. Error bars show standard deviation.

have a hardness of 190–195 HV that is considerably higher than that of annihilation/rearrangement. However, in this study, the horizontally
annealed 316L austenitic SS, that is typically around 150 HV [55]. The printed specimens displayed a consistently higher value of hardness
heat treatment at 750FC resulted in a fall in hardness by around 5–10 than the vertically built specimens for the heat treatments tested,
HV, and further decreased to approximately 175–185 HV for the 900FC whereas measurements by Kamariah et al. have suggested that speci­
treatment. The trend in hardness with increasing heat treatment tem­ mens of the vertical orientation exhibit a greater hardness [37]. This
perature are in good agreement with work by Kamariah et al. on simi­ discrepancy with literature is likely due to differences in the tempera­
larly printed 316L material [37]. However, whereas they have ture gradients present in the printed components. In both print orien­
attributed the decrease in hardness to a reduction in δ-ferrite at tem­ tations, the highest temperature gradient would be expected along the
peratures above 950 ◦ C, no δ-ferrite was observed in any of the materials build direction, but for the horizontal specimens produced in this study,
in this study. Here, it is more likely that the heat treatments at 900 ◦ C another considerable temperature gradient could be found along the
have allowed some degree of recovery to occur (reduction is dislocation scan direction (due to the long scan vectors employed resulting in heat
density), resulting in the lowered hardness [42,56]. The samples build-up) [9,39]. This would result in high tensile residual stresses in
exposed to the 900FC treatment possessed a lower Vickers hardness two directions that could result in the grains of the horizontally printed
value than the 900WQ specimens suggesting that a greater degree of material experiencing two significant strain components. The recrys­
recovery occurred in the furnace cooled specimens compared to the tallised specimens (1250WQ) exhibited the lowest hardness at approx­
water quenched equivalents. This is possibly due to the much slower imately 140 HV due to the nucleation and growth of strain-free grains.
cooling rate in the former, allowing more time for dislocation As a result of complete recrystallisation occurring in both vertically and

7
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 10. Final states of the different heat treated printed specimens after 20 days of corrosion with an applied tensile stress of 60 MPa.

8
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 11. Optical image of an SCC crack found in: (a) An as-built vertically built LPBF 316L component. (b) A 750FC heat treated horizontal LPBF 316L component.
Secondary electron SEM image showing: (c) highly branched cracking in an as-built vertically built LPBF 316L component. (d) part of a corrosion pit with an
associated SCC crack in a 750FC heat treated horizontal LPBF 316L component.

9
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 12. Surface crack densities of the different heat treated printed specimens after 20 days of corrosion with an applied tensile stress of 60 MPa. Error bars show
standard deviation.

horizontally printed material, there were no variations in hardness be­ where a much larger area of dark corrosion can be seen along the sup­
tween the two build orientations. port surface. Some speckled corrosion products can be seen within the
corroded areas possibly indicating the presence of pitting. The corrosion
3.2. Changes in corrosion/SCC behaviour with heat treatment for the vertically printed specimens also appears different: the surface
now contains red-brown corrosion products along with dark, speckled
Images showing the extent of corrosion on the surface of the various areas of localised corrosion likely to correspond to pits. The poorer
heat treated specimens (of both tested orientations) after 20 days of surface condition of the heat treated material when compared to the as-
corrosion testing with an applied tensile stress of 60 MPa are shown in built samples can be linked to the surface oxide. It has been reported in
Fig. 10. There are distinct differences in the corrosion patterns sug­ literature that the heat tint produced at higher treatment temperatures
gesting that the type of heat treatment has had significant effects on the result in reduced corrosion resistance, in particular pitting resistance
corrosion behaviour of the LPBF 316L material. The AB samples dis­ (through reduced pitting potentials) [59]. The two treatments at 900 ◦ C
played the least amount of surface corrosion, with most of the specimen appeared similar in terms of surface corrosion products. Various col­
unaffected by the chloride salt. When compared with forged 316L from oured corrosion products exist on the surface suggesting the presence of
an identical experimental setup, the LPBF 316L material shows superior multiple different compositions of iron oxides. Speckling can still be
general corrosion resistance; the entire surface of the forged 316L ma­ observed on the surface of specimens suggesting that pitting has still
terial had developed a layer of red-brown corrosion product within the occurred. Furthermore, the horizontal specimens still retain the prefer­
same timeframe [47]. In literature, the corrosion resistance of LPBF ence to corrode along the support edge, an indication that stresses have
316L compared to the wrought counterpart is highly disputed with ev­ not been fully relieved by either the 900FC or 900WQ heat treatments.
idence of both greater and poorer performance being presented [12,14, Again, the more heavily corroded surfaces for these heat treated samples
57]. This is likely due to the significant variations in properties that may when compared to the 750FC are likely due to the more deleterious
exist in LPBF material depending on the chosen printing parameters, oxide layer formed at higher temperature heat treatments [59]. The
such as: grain size, residual stress distribution, plastic strains, porosity recrystallised 316L material shows an absence of any speckled corrosion
and surface quality [58]. The superior corrosion resistance of LPBF 316L products and are covered by a uniform layer of red-brown corrosion
has been argued to be the result of the lack of detrimental MnS in­ products, indicating the lack of localised corrosion such as pitting. While
clusions (widely regarded to be detrimental to pitting resistance) and the reduced residual stresses and plastic strains reduce the likelihood of
formation of a more compact or thicker protective oxide layer [12,14, pitting in recrystallised forged material, this has not been generally
20,57,58]. However, for the as-built printed specimens in this study, observed for recrystallised additively manufactured austenitic SS
corrosion was typically observed in dark, circular patterns suggesting [60–64]. For LPBF 316L, lower pitting potential have been reported in
some form of localised corrosion such as pitting still occurs, possibly due literature, which has been attributed to an accelerated transition of
to localised large plastic strains present (see later Section 3.5). The build metastable pitting to steady-state pitting as a result of weak passive films
orientation is also an important factor to consider as while the corrosion formed at pores that have been enlarged during heat treatments [63]. It
in the vertical specimens appear randomly distributed; the horizontally is likely that the discrepancy with literature in this study is due to the
printed specimens tended to favour corrosion at a specific surface (the LPBF 316L material studied here not being heat treated in an inert at­
surface where the support structures were located). This has been sug­ mosphere. As a result of the oxide layer formed on the sample surface
gested that this is due to partial relaxation of longitudinal stresses during during the heat treatment/water quenching process, significant surface
removal of the support structures, resulting in higher tensile stresses corrosion can be observed on the recrystallised LPBF 316L. It is possible
distributed along the support surface [9]. that the increased surface corrosion may have masked the pitting
For the 750FC heat treatment, the preference for horizontally printed corrosion.
specimens to corrode at the support surface remains, indicating that After removing the surface corrosion products via grinding/polishing
higher residual stresses still exist close to the support surface. The SCC cracks were revealed in some of the samples. The morphology of the
corrosion appears significantly worse in the horizontal specimens, SCC cracks tended to be highly branched and propagated

10
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 13. Final states of the different heat treated printed specimens after 30 days of corrosion without an applied tensile stress. Note that specimens denoted with a *
were produced from a different print run than those used in the 20-day tests.

perpendicularly with respect to the build direction (even in the presence 900 ◦ C and 1250 ◦ C also appeared to be sufficient for preventing
of an applied stress in an opposing direction for the horizontally printed cracking in the horizontally printed specimens within the experimental
specimens). Examples of typical SCC cracks found in the printed mate­ timeframe. While heat tint has been shown in literature to be detri­
rial are shown in Fig. 11. The crack densities obtained from the polished mental to pitting corrosion and hence SCC, this trend is not observed
specimen surfaces are shown in Fig. 12 for the LPBF samples tested with here as the heat treatments at higher temperature did not result in
an applied stress of 60 MPa tested for 20 days. Crack densities were increased SCC susceptibility [59,65]. It is possible that other factors such
considerably lower than forged 304L, despite the presence of high ten­ as changes to the surface residual stress state through heat treatment
sile residual stresses, further showing the improved SCC resistance of may be more significant.
316L material, even in a non-ideal stress state [42]. The highest crack
density was found in the vertically printed AB specimens. After 750FC 3.3. Samples without an applied stress
heat treatment, the crack densities were reduced significantly for the
vertically printed material and at higher temperatures of 900 ◦ C and A second set of experiments was performed removing the effect of
1250 ◦ C no cracks were observed. Horizontal specimens displayed no applied stress and increasing the duration of the corrosion test to obtain
cracks in the as-built state; however, instances of cracking was detected more cracks for statistical analysis and to determine whether the re­
after the 750FC heat treatment. However, it is worth noting that the sidual tensile stresses present in the LPBF specimens were sufficient
standard deviation is larger than the mean measurement due to the fact alone for initiating SCC. Three sample conditions were chosen: as-built,
that only one of the tested samples showed any evidence of SCC. While it 750 ◦ C for 2 h then furnace cooled and 900 ◦ C for 0.5 h then water
is possible that the 750FC heat treatment could have had some detri­ quenched and prepared from a second batch of printed specimens. The
mental effect on the SCC susceptibility of the printed material, it is not surface corrosion characteristics displayed some similarities with the
conclusive as the SCC in the single sample could be due to a sample previous tests with applied stress (shown in Fig. 13), where as-built
defect. The second set of experiments (discussed later in Section 3.3) was specimens showed localised circles of corrosion and horizontal speci­
partially designed to address this, by increasing the test duration to mens displayed a preference for corrosion at the support surface.
allow for greater crack nucleation and growth. Heat treatments at However, the extent of corrosion was greater than 20 day loaded

11
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 14. Optical images showing the pitting/cracking in the specimens exposed to MgCl2 for 30-days without applied stress. Vertically printed: (a) AB*, (b) 750FC*,
(c) 900WQ*. Horizontally printed: (d) AB*, (e) 750FC*, (f) 900WQ*.

Fig. 15. Surface crack densities of the different heat treated printed specimens after 30 days of corrosion without an applied stress. Note that specimens denoted with
a * were produced from a different print run. Error bars show standard deviation.

12
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Table 4 horizontally printed AB* material displayed some cracking in the as-
Stress components measured using XRD for similar LPBF dogbone specimens. built state but still far lower than the vertically printed equivalent.
Data compiled from Williams et al. [39] and Vecchiato [9]. However, upon heat treating at 750 ◦ C, the number of observed cracks in
Build Orientation Heat Treatment Stress component Stress (MPa) the horizontal specimens increased significantly.
Horizontal As-built σBD 242 ± 20.0
σLD 102 ± 20.9 3.4. Influence of surface residual stresses
Horizontal 700 ◦ C (2 h) FC σBD 229 ± 44.4
σLD 127 ± 7.7 The various heat treatments are likely to result in different distri­
Horizontal 900 ◦ C (0.5 h) WQ σBD -71.8 ± 7.4
butions of residual stresses within the LPBF samples. Surface residual
σLD 26 ± 8.2
Vertical As-built σBD 349 ± 20.5 stresses obtained using X-ray diffraction from work by Williams et al.
σTD 40.1 ± 7.4 and Vecchiato on similar heat treatments of identical printed tensile
Vertical 700 ◦ C (2 h) FC σBD 119 ± 8.3 specimens are given in Table 4 [9,39]. Although the heat treatment at
0 ± 17.7
σTD
700 ◦ C (2 h) is slightly lower than the heat treatment of 750 ◦ C (2 h)
Vertical 900 ◦ C (0.5 h) WQ σBD -191.7 ± 19.5
σTD -135.4 ± 17.4
studied in this experiment, it is not significantly different and should still
provide a useful indication of the residual stresses to be expected within
the 750FC and 750FC* samples. The surface residual stress data from
specimens and corroded areas were larger in both sample print orien­ literature confirm that high tensile residual stresses are present in the
tations, possibly due to prolonged exposure to the corrosive environ­ as-built 316L specimens and reaffirm that the highest stress components
ment. For the 750FC* heat treatment, the horizontal specimens still are found parallel to the build direction. This is seen in the SCC orien­
initially corroded preferentially around the support surface but after the tations, where cracks mainly propagate perpendicularly to the build
duration of the extended test, nearly the entire gauge length had direction, as indicated in Fig. 14. A measure for the degree of crack
developed dark surface corrosion products, with heavy speckling. The branching is shown in Fig. 16 where more significant branching was
vertical specimens at the 750FC* treatment showed no preference in detected in the vertical AB* specimens compared to the horizontal. The
corrosion sites with dark areas developing across the surface along with higher crack density and greater degree of branching recorded in the
speckled corrosion. The corrosion appears more severe, in terms of both vertical AB* specimens could possibly be due to the as-built vertically
corroded area and speckling than the equivalent samples from the 20- printed samples possessing higher tensile residual stresses than the
day test with applied stress. The 900WQ* samples showed similar horizontally printed equivalents, along their respective build directions,
corrosion patterns to the 750FC* specimens for both print orientations. as seen in Table 4.
The morphology of the SCC cracks found in the AB* and 750FC* From Table 4, the 700 ◦ C furnace cooled heat treatment resulted in a
samples are shown in Fig. 14a, b, d and e. The majority of the cracks are significant reduction in surface tensile stresses for vertical build com­
associated with large corrosion pits and appear to propagate perpen­ ponents. For the 750FC and 750FC* vertically built specimens in this
dicular to the build direction. The 900WQ* treatment still appeared to study, a similar change in the stress state is to be expected and it is
be effective in reducing the risk of SCC, where even after 30-days of possible that this reduction in surface tensile stresses has resulted in the
corrosion testing, despite extensive pitting (shown in Fig. 14c and f), no lower crack density (Fig. 12 and Fig. 15) and crack branching (Fig. 16)
SCC were observed in samples of either build orientations. The crack observed from the corrosion tests than for AB specimens. However, the
densities obtained after 30 days are shown in Fig. 15. Although the crack horizontal 750FC LPBF components showed an increase in crack den­
densities appear to be much higher than that recorded for the 20-day sity, despite a minor decrease in the surface residual tensile stresses
applied stress test, similar trends could be observed. For vertically measured by Williams et al. on equivalent samples treated at 700 ◦ C
built specimens, crack densities were highest in the AB* condition and [39]. It is likely therefore that in the horizontally built specimens, other
decreased at the 750 ◦ C heat treatment (as for the 20-day tests). The material factors may make a greater contribution to the reduced SCC

Fig. 16. Average number of branches per crack for the different heat treated printed specimens after 30 days of corrosion without an applied stress. Note that
specimens denoted with a * were produced from a different print run. Error bars show standard deviation.

13
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 17. GROD maps representing the intragranular misorientation for the vertically printed LPBF 316L specimens after the different heat treatments: (a) AB; (b)
750FC; (c) 900FC; (d) 900WQ; (e) 1250WQ-recrystallised. Qualitative indicator of intragranular strain. Colour scale represents the deviation of the orientation with
respect to the grain average reference (maximum is red, minimum is dark blue).

Fig. 18. GROD maps representing the intragranular misorientation for the horizontally printed LPBF 316L specimens after the different heat treatments: (a) AB; (b)
750FC; (c) 900FC; (d) 900WQ; (e) 1250WQ-recrystallised. Qualitative indicator of intragranular strain. Colour scale represents the deviation of the orientation with
respect to the grain average reference (maximum is red, minimum is dark blue).

resistance, such as microstructural effects or the role of heat tint, as corrosion tests where neither 900WQ nor 900WQ* specimens displayed
discussed earlier. any SCC. Although, horizontally built specimens appeared to retain a
Surface X-ray diffraction measurements of equivalent water small tensile residual stress in the longitudinal direction (perpendicular
quenched specimens treated at 900 ◦ C by Vecchiato have shown that to build direction), this did not appear to be sufficient to induce SCC
considerable magnitudes of stresses can remain in the build direction within the duration of the tests. Furnace cooling after the heat treatment
after the heat treatment [9]. However, due to the water quenching, most at 900 ◦ C also appeared to be effective in reducing SCC as none of the
of the tensile residual stresses at the surface have inverted and become 900FC samples had developed cracks. Rather than a compressive stress
compressive due to cooling rate effects through thickness [39,66]. The state, it is likely that despite the short 30 min heat treatment; the higher
compressive stresses in the 900WQ/900WQ* specimens are generally temperature in combination with the slower cooling has allowed enough
considered to be beneficial for improving SCC resistance by inhibiting time for greater stress relief within the specimens compared to the water
the development of surface SCC cracks [67,68]. This is confirmed in the quenched process, and that these tensile stresses maybe low enough to

14
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 19. EBSD maps of a crack in the AB* specimens (without applied tensile stress). Vertically printed: (a) GROD, (b) Grain boundary map. Horizontally printed: (a)
GROD, (b) Grain boundary map.

inhibit SCC. samples likely explain why pitting corrosion is still prominent, despite
Despite a compressive surface residual stress state being desirable for the lack of SCC (due to compressive stresses along build direction). On
SCC resistance, considerable levels of plastic strain in the 900WQ/ the other hand, the recrystallised LPBF 316L (1250WQ), mostly con­
900WQ* specimens are still likely to exist, as this is not a recrystallisa­ sisting of strain-free grains, did not experience either pitting or SCC. In
tion heat treatment. This is indicated by the Vickers micro-hardness literature, recrystallised LPBF 316L (in an inert atmosphere) has been
values in Fig. 9 and the grain structures in Figs. 5c and 6c. Pitting re­ reported to have poorer pitting resistance despite the reduction in plastic
mains an important issue in highly strained material, since once the pit strain due to the effects of increasing pore size (due to temperature)
has grown deep enough, there is a possibility for SCC to develop below weakening the associated passive film [63]. It is likely that for the
the surface, albeit after some incubation time, as the core of the LPBF samples in this study that were heat treated in an open atmosphere,
specimen is now presumed to be under a tensile stress state (to reach surface oxidation and hence surface corrosion may have influenced the
equilibrium of forces) [9,66]. On the other hand, despite a tensile sur­ corrosion pitting observations.
face residual stress state being detrimental for SCC in furnace cooled The GROD maps in the vicinity of a typical crack found in the AB*
samples, any crack growth would be limited due to a decline in the specimens are shown in Fig. 19a and c, for horizontal and vertical print
magnitude of the tensile stresses towards the core. Neutron diffraction orientations respectively, with the corresponding grain boundary maps
measurements by Williams et al. have revealed that the core of as-built in Fig. 19b and d. The crack branching remains extensive in both build
and furnace cooled specimens to be under compressive stresses that orientations, although it appears that the vertically printed material
could assist in retarding or stopping SCC growth, preventing complete displays more branches, in agreement with the quantification of
component failure [39]. branching made by optical microscopy shown in Fig. 16. It is clear from
the grain boundary maps in Fig. 19b and d that the cracking observed in
both LPBF print orientations is, despite the branching, still trans­
3.5. Role of microstructure
granular, consistent with typical Cl- SCC in austenitic SS. Crack
branching predominantly occurs either at grain boundaries or intra­
Cold work/plastic strain is widely regarded as being detrimental to
the SCC and pitting resistance of austenitic SS. Grain reference orien­ granular misorientation bands (as indicated by the red arrows) shown by
the GROD map. It is likely that these bands of misorientation seen within
tation deviation (GROD) maps (displayed in Fig. 17 and Fig. 18 for
vertical and horizontal build orientations, respectively) can be used as a the grains represent deformation bands that have been suggested to be
energetically favourable for crack propagation. For intergranular SCC,
qualitative indicator of intragranular strain by comparing the misori­
entation of a pixel with the average grain misorientation [69]. When low angle grain boundaries (LAGBs) have sometimes been suggested to
be more resistant to SCC due to the lower associated grain boundary
compared to the fully recrystallised material, it is clear that the LPBF
material (regardless of heat treatment) still have significantly deformed energies and could similarly assist in delaying/inhibiting transgranular
crack propagation [70,71]. However, from the grain boundary maps in
grains present. The presence of high levels of plastic strain in the 900WQ

15
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 20. EBSD maps of a crack in the 750FC* specimens (without applied tensile stress). Vertically printed: (a) GROD, (b) Grain boundary map. Horizontally printed:
(a) GROD, (b) Grain boundary map.

Fig. 19, the transgranular SCC cracks observed here do not appear to be could have occurred during the fabrication process. At the cellular sub-
inhibited by the presence of LAGBs. grain scale it has been reported that structures exist due to redistribution
Grain boundary maps and GROD maps of typical cracks found in of elements during solidification [9]. Various studies in literature have
750FC* samples are shown in Fig. 20. Although the maps were taken at reported the segregation of Cr and Mo towards cell walls, elements
the same magnification, the degree of branching appears to be reduced critical to both corrosion and SCC resistance [7]. The elemental segre­
for both print orientations in comparison to the AB* specimens shown in gation could have both beneficial and detrimental effect on SCC resis­
Fig. 19. Bands of misorientation are still observed in the 750FC* maps tance, as the higher Cr and Mo content at cell walls could hinder crack
but these do not appear to cause branching of the SCC crack. The SCC propagation while the cell interior could become more susceptible to
cracks observed are still clearly transgranular and similarly to the AB* SCC/corrosion. The cellular structure has been reported to be influenced
material, LAGBs do not appear to inhibit SCC. by heat treatment after fabrication, where higher temperatures allow
“Special” coincidence site lattice (CSL) boundaries, in particular Σ3 diffusion of elements resulting in grain homogeneity [7,38]. Ronneberg
boundaries corresponding to twins, are also low energy grain boundaries et al. have suggested that the cellular network begins to disappear at a
and have been widely recognised to improve resistance to intergranular similar temperature range to the melt pool boundaries, above 700 ◦ C,
SCC in austenitic SS [72–74]. It is possible that they may have a similar and is fully removed at 1200 ◦ C [7]. The loss of the cell structure could
effect on transgranular SCC as more energy would be needed to cross be another reason why the 750 ◦ C heat treated horizontally printed
such grain boundaries, providing an inhibiting effect on crack nucle­ specimens performed worse, in terms of SCC resistance, than the as-built
ation and propagation [75]. Higher heat treatment temperatures have equivalents. It is important to note, however, that the vertically printed
been reported to have an effect of increasing the proportion of low ΣCSL specimens did not show a decrease in SCC resistance after the 750 ◦ C
grain boundaries, such as Σ3, in 316L [76]. Grain boundary analysis of heat treatment. It is possible likely surface tensile stresses have a pri­
the EBSD maps in Figs. 5 and 6 revealed that while the recrystallised mary role in SCC. After heat treatment, surface residual tensile stresses
316L specimens contain 45–50% Σ3 boundaries (by length compared to in the build direction only decreased from 242 MPa to 229 MPa for the
total boundary length), the other LPBF sample conditions all possessed horizontal specimens while the vertically printed samples fell dramati­
minimal amounts of Σ3 boundaries, below 4% by length. Even if Σ3 cally from 349 MPa to 119 MPa.
boundaries were to have a beneficial impact on transgranular SCC, the
very low proportion found in the LPBF materials renders their effect 3.6. Effect of encapsulation
negligible.
Chemical segregation is another concern when considering the SCC The surfaces of the heat treated encapsulated specimens generally
susceptibility of LPBF 316L. Although intergranular cracking was not appear very different to the equivalent samples heat treated in air
observed in any of the heat treated samples, suggesting that sensitisation (displayed in Fig. 21). The encapsulated 900FC specimens displayed no
during heat treatments was not an issue, other segregation of elements substantial oxide development on the surface aside from a few small

16
P. Dong et al. Additive Manufacturing 40 (2021) 101902

significantly worse corrosion present in the components with the oxide


scale, both encapsulated and non-encapsulated 900 ◦ C heat treatments
displayed no evidence of SCC, suggesting that the surface residual stress
state may be more significant.
The transgranular cracks observed in the encapsulated 750FC sam­
ples are similar to the non-encapsulated samples in appearance showing
a branched morphology with the direction of the crack propagation
parallel to the build direction as shown in Fig. 24 (even with 60 MPa
applied stress in a competing direction for the horizontally printed
material). As the residual stress state should be virtually identical be­
tween the encapsulated and non-encapsulated samples at 750 ◦ C and the
samples should have experienced similar changes in the cellular struc­
ture, the difference in crack density must be due to other material factors
related to the surface condition. From the images of the initial specimen
surfaces after heat treatment in Fig. 4 and Fig. 21, it is clear that there is
a significant difference in the generated surface oxide layer. In literature
it has been reported that the passivating oxide layer formed during LPBF
may be unstable and LPBF austenitic SS have poor re-passivation ability
attributed to porosity and precipitates [12,77,78]. For encapsulated
specimens at 750 ◦ C, the passivating layer may become damaged during
heating and the lack of oxygen could have limited the regeneration of
the passive layer resulting in the poorer corrosion and SCC resistance of
the encapsulated LPBF material. Due to the much shorter heating
duration of the 900 ◦ C encapsulated samples (only 0.5 h compared to
2 h), the surface still looks mostly unchanged, however, the black spots
forming along the specimen suggest that some oxidation is occurring
despite the inert atmosphere. If the heating was continued, the oxidation
could similarly extend across the entire specimen.
Fig. 21. Images of the LPBF components at each of the different encapsulated
heat treatments: (a) Vertically printed; (b) Horizontally printed. The face par­
allel to BD shown here was SCC tested. 4. Conclusions

In order to assess the effect of heat treatment on the SCC suscepti­


spots. The encapsulated 900WQ specimen showed oxidation of the
bility of LPBF 316L specimens, various heat treatment procedures were
entire steel surface after the quenching process. Interestingly, the
applied to printed dogbone specimens and exposed to MgCl2 salt-spray
encapsulated 750FC samples developed considerable surface oxidation
to simulate atmospheric salt deposition. Samples were tested at 75 ◦ C
products when heat treated in an inert atmosphere, that varied between
and 70% RH in an enclosed environment both with and without an
print orientations. For the vertically printed specimens, the entire sur­
applied tensile stress. The following conclusions were made:
face appeared oxidised, while in the horizontally printed specimens only
some regions were affected. The darkened regions of the horizontal
1. Horizontally printed specimens show preferential corrosion along
specimens are associated with the support surface, where higher tensile
the “support surface” where highest residual stresses were expected.
stresses and greater plastic strain are expected. High levels of plastic
This was not observed in the vertically printed specimens which
strain have been known to have a detrimental effect on the oxidation
showed localised corrosion circles, distributed randomly. The pref­
potential of austenitic SS in literature.
erence for corrosion of the support side in horizontally printed
Encapsulation in an inert argon atmosphere prior to the heat treat­
samples is retained in heat treatments up to 900 ◦ C but eliminated
ment process also appears to have had some influence on the surface
after recrystallisation at 1250 ◦ C.
corrosion behaviour and SCC behaviour of the LPBF 316L material when
2. As-built specimens show SCC in both vertical and horizontal print
compared with the non-encapsulated equivalents. The end states of
directions after 30 days of testing. Crack morphologies were highly
samples exposed to the three different heat treatment conditions with
branched and always aligned perpendicularly to the build direction
encapsulation are shown in Fig. 22 (samples were under a 60 MPa
(due to high tensile residual stresses), even in the presence of an
applied tensile stress and tested for 20 days). In comparison with the
applied tensile stress of 60 MPa in a perpendicular direction.
non-encapsulated 750FC samples, the encapsulated specimens showed a
3. Heat treatments at 750 ◦ C for 2 h followed by furnace cooling was
far greater extent of corrosion, in terms of area affected, especially for
insufficient to eliminate the risk of SCC for either print orientations
the horizontally built samples. Corrosion products appear darker and are
due to significant surface tensile residual stresses remaining along
speckled. For the horizontal specimens, there are clear signs that
the build direction. This particular heat treatment had a detrimental
corrosion still preferred the support edge, as seen in the non-
effect of increasing SCC crack density for horizontally printed spec­
encapsulated specimens. Crack density analysis in Fig. 23 show that
imens but a beneficial effect of reducing susceptibility to SCC for
the encapsulated samples at 750 ◦ C are much more susceptible to SCC
vertically printed samples.
than both the as-built and 750FC samples, suggesting a strong detri­
4. GROD maps show that highly plastic strained grains can exist for all
mental effect caused by encapsulation. However, the encapsulated
heat treated LPBF material, except the fully recrystallised material.
samples at 900 ◦ C (both FC and WQ) still showed no evidence of
In the as-built 316L components branching occurs either at grain
cracking due to the beneficial effects of stress relief/stress state inversion
boundaries or at the edge of misorientation bands.
although the surface corrosion patterns displayed a visible difference to
5. CSL boundaries, even if they are beneficial to transgranular SCC, are
the non-encapsulated samples. By comparing the final corroded states of
unlikely to be important for LPBF 316L material as their total
the 900ENFC specimens in Fig. 22 with the non-encapsulated equiva­
boundary fraction is negligible (less than 4% by length).
lents from Fig. 10, it is clear that the loss of the oxide scale has reduced
the extent of the surface corrosion considerably. However, despite the

17
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 22. Final states of the different encapsulated heat treated printed specimens after 20 days of corrosion with an applied tensile stress of 60 MPa.

18
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Fig. 23. Surface crack densities of the different encapsulated heat treated printed specimens after 20 days of corrosion with an applied tensile stress of 60 MPa
compared with AB and 750FC samples. Error bars show standard deviation.

Fig. 24. Optical images showing the pitting/cracking in: (a) Vertically printed 750ENFC (b) Horizontally printed 750ENFC. Applied tensile stress parallel to LD for
both specimens.

6. Encapsulation in argon atmosphere prior to 750 ◦ C heat treatment 8. 900 ◦ C heat treatments (both encapsulated/non-encapsulated and
for 2 h with furnace cooling resulted in a significant increase in both both furnace cooled/water quenched) were effective for improving
the corrosion and SCC susceptibility of the LPBF material. SCC resistance in LPBF 316 L components.
7. The difference in the observed corrosion patterns between the
encapsulated and non-encapsulated heat treated material highlight CRediT authorship contribution statement
the importance of the detrimental impact of oxide scales on corrosion
behaviour, in addition to microstructure and residual stress. P. Dong: Conceptualization, Methodology, Investigation, Formal

19
P. Dong et al. Additive Manufacturing 40 (2021) 101902

Analysis, Writing- Original draft preparation, Visualization. F. Vec­ [19] S. Lozano-Perez, J. Dohr, M. Meisnar, K. Kruska, SCC in PWRs: learning from a
bottom-up approach, Metall. Mater. Trans. E 1 (2) (2014) 194–210, https://doi.
chiato: Conceptualization, Investigation, Formal Analysis. Z. Yang:
org/10.1007/s40553-014-0020-y.
Methodology, Investigation, Formal Analysis. P.A. Hooper: Conceptu­ [20] X. Ni, D. Kong, Y. Wen, L. Zhang, W. Wu, B. He, Anisotropy in mechanical
alization, Resources, Writing - Review & Editing, Supervision, Funding properties and corrosion resistance of 316L stainless steel fabricated by selective
acquisition. M.R. Wenman: Conceptualization, Resources, Writing - laser melting Anisotropy in mechanical properties and corrosion resistance of 316L
stainless steel fabricated by selective laser melting,” no. June, 2019, doi: 10.100
Review & Editing, Supervision, Project Administration, Funding 7/s12613-019-1740-x.
acquisition. [21] T. Simson, A. Emmel, A. Dwars, J. Böhm, Residual stress measurements on AISI
316L samples manufactured by selective laser melting, Addit. Manuf. 17 (2017)
183–189, https://doi.org/10.1016/j.addma.2017.07.007.
[22] Y. Liu, Y. Yang, D. Wang, A study on the residual stress during selective laser
Declaration of Competing Interest melting ( SLM) of metallic powder, Int. J. Adv. Manuf. Technol. 87 (2016)
647–656, https://doi.org/10.1007/s00170-016-8466-y.
[23] T. Magnin, A. Chambreuil, B. Bayle, The corrosion-enhanced plasticity model for
The authors declare that they have no known competing financial stress corrosion cracking in ductile fcc alloys, Acta Mater. 44 (4) (1996)
interests or personal relationships that could have appeared to influence 1457–1470, https://doi.org/10.1016/1359-6454(95)00301-0.
the work reported in this paper. [24] D.T. Spencer, M.R. Edwards, M.R. Wenman, C. Tsitsios, G.G. Scatigno, P.R. Chard-
Tuckey, The initiation and propagation of chloride-induced transgranular stress-
corrosion cracking (TGSCC) of 304L austenitic stainless steel under atmospheric
Acknowledgements conditions, Corros. Sci. 88 (2014) 76–88, https://doi.org/10.1016/j.
corsci.2014.07.017.
[25] R.J. Williams, F. Vecchiato, J. Kelleher, M.R. Wenman, P.A. Hooper, C.M. Davies,
P. Dong and M.R. Wenman thank EDF and the EPSRC Centre for Effects of heat treatment on residual stresses in the laser powder bed fusion of 316L
Doctoral Training in Nuclear Energy (EP/L015900/1) for financial stainless steel: finite element predictions and neutron diffraction measurements,
support. F. Vecchiato, P.A. Hooper and M.R. Wenman acknowledge J. Manuf. Process. 57 (2020) 641–653, https://doi.org/10.1016/j.
jmapro.2020.07.023.
funding from AWE plc (contract 30338995) and EPSRC (EP/M507878/ [26] E. Otero, A. Pardo, M.V. Utrilla, E. Sáenz, F.J. Perez, Influence of microstructure on
1). The authors would also like to thank Richard J. Williams and Harry the corrosion resistance of AISI type 304L and type 316L sintered stainless steels
de Winton for providing the LPBF 316L dogbones tested in this study. exposed to ferric chloride solution, Mater. Charact. 35 (3) (1995) 145–151,
https://doi.org/10.1016/1044-5803(95)00099-2.
[27] D. Kong, C. Dong, Z. Zheng, F. Mao, A. Xu, X. Ni, C. Man, J. Yao, K. Xiao, X. Li,
References Surface monitoring for pitting evolution into uniform corrosion on Cu-Ni-Zn
ternary alloy in alkaline chloride solution: ex-situ LCM and in-situ SECM, Appl.
Surf. Sci. 440 (2018) 245–257, https://doi.org/10.1016/j.apsusc.2018.01.116.
[1] Y. Huang, M.C. Leu, J. Mazumder, A. Donmez, Additive manufacturing: current
[28] Y. Zuo, H. Wang, J. Xiong, The aspect ratio of surface grooves and metastable
state, future potential, gaps and needs, and recommendations, J. Manuf. Sci. Eng.
pitting of stainless steel, Corros. Sci. 44 (1) (2002) 25–35, https://doi.org/
137 (2015) 1–10, https://doi.org/10.1115/1.4028725.
10.1016/S0010-938X(01)00039-7.
[2] S. Ford, M. Despeisse, Additive manufacturing and sustainability: an exploratory
[29] E. Otero, A. Pardo, M.V. Utrilla, E. Sáenz, J.F. Álvarez, Corrosion behaviour of aisi
study of the advantages and challenges, J. Clean. Prod. 137 (2016) 1573–1587,
304l and 316l stainless steels prepared by powder metallurgy in the presence of
https://doi.org/10.1016/j.jclepro.2016.04.150.
sulphuric and phosphoric acid, Corros. Sci. 40 (8) (1998) 1421–1434, https://doi.
[3] Y.M. Wang, T. Voisin, J.T. McKeown, J. Ye, N.P. Calta, Z. Li, Z. Zeng, Y. Zhang,
org/10.1016/S0010-938X(98)00047-X.
W. Chen, T.T. Roehling, R.T. Ott, M.K. Santala, P. Depond, M.J. Matthews, A.
[30] R.F. Schaller, J.M. Taylor, J. Rodelas, E.J. Schindelholz, Corrosion properties of
V. Hamza, T. Zhu, Additively manufactured hierarchical stainless steels with high
powder bed fusion additively manufactured 17-4 pH stainless steel, Corrosion 73
strength and ductility, Nat. Mater. 17 (1) (2018) 63–71, https://doi.org/10.1038/
(7) (2017) 796–807, https://doi.org/10.5006/2365.
nmat5021.
[31] D. Kong, X. Ni, C. Dong, L. Zhang, C. Man, J. Yao, K. Xiao, X. Li, Heat treatment
[4] C. Qiu, M. Al Kindi, A.S. Aladawi, I. Al Hatmi, A comprehensive study on
effect on the microstructure and corrosion behavior of 316L stainless steel
microstructure and tensile behaviour of a selectively laser melted stainless steel,
fabricated by selective laser melting for proton exchange membrane fuel cells,
Sci. Rep. 8 (1) (2018) 7785, https://doi.org/10.1038/s41598-018-26136-7.
Electrochim. Acta 276 (2018) 293–303, https://doi.org/10.1016/j.
[5] S.M. Yusuf, Y. Chen, R. Boardman, S. Yang, N. Gao, Investigation on porosity and
electacta.2018.04.188.
microhardness of 316L stainless steel fabricated by selective laser melting, Metals 7
[32] Z. Duan, C. Man, C. Dong, Z. Cui, D. Kong, L. wang, X. Wang, Pitting behavior of
(2) (2017) 1–12, https://doi.org/10.3390/met7020064.
SLM 316L stainless steel exposed to chloride environments with different
[6] B. Vrancken, Study of Residual Stresses in Selective Laser Melting, Arenberg
aggressiveness: pitting mechanism induced by gas pores, Corros. Sci. 167 (2020),
Doctoral School, 2016.
108520, https://doi.org/10.1016/j.corsci.2020.108520.
[7] T. Ronneberg, C.M. Davies, P.A. Hooper, Revealing relationships between porosity,
[33] G. Sander, J. Tan, P. Balan, O. Gharbi, D.R. Feenstra, L. Singer, S. Thomas, R.
microstructure and mechanical properties of laser powder bed fusion 316L
G. Kelly, J.R. Scully, N. Birbilis, Corrosion of additively manufactured alloys: a
stainless steel through heat treatment, Mater. Des. 189 (2020), 108481, https://
review, Corrosion 74 (12) (2018) 1318–1350, https://doi.org/10.5006/2926.
doi.org/10.1016/j.matdes.2020.108481.
[34] X. Lou, M. Song, P.W. Emigh, M.A. Othon, P.L. Andresen, On the stress corrosion
[8] X. Ni, Dc Kong, Y. Wen, L. Zhang, Wh Wu, Bb He, L. Lu, Dx Zhu, Anisotropy in
crack growth behaviour in high temperature water of 316L stainless steel made by
mechanical properties and corrosion resistance of 316L stainless steel fabricated by
laser powder bed fusion additive manufacturing, Corros. Sci. 128 (2017) 140–153,
selective laser melting, Int. J. Miner. Metall. Mater. 26 (3) (2019) 319–328,
https://doi.org/10.1016/j.corsci.2017.09.017.
https://doi.org/10.1007/s12613-019-1740-x.
[35] X. Lou, P.L. Andresen, R.B. Rebak, Oxide inclusions in laser additive manufactured
[9] F. Vecchiato, Influence of Laser Parameters on the Deposition of 316L Stainless
stainless steel and their effects on impact toughness and stress corrosion cracking
Steel using Laser Powder Bed Fusion, Imperial College London, 2019.
behavior, J. Nucl. Mater. 499 (2018) 182–190, https://doi.org/10.1016/j.
[10] J.C. Lippold, D.J. Kotecki, Welding Metallurgy and Weldability of Stainless Steels,
jnucmat.2017.11.036.
John Wiley and Son, 2005.
[36] E. De Bruycker, M.L.M. Sistiaga, F. Thielemans, K. Vanmeensel, Corrosion testing
[11] M.P. Ryan, D.E. Williams, R.J. Chater, B.M. Hutton, D.S. McPhail, Why stainless
of a heat treated 316 L functional part produced by selective laser melting, Mater.
steel corrodes, Nature 415 (6873) (2002) 770–774, https://doi.org/10.1038/
Sci. Appl. 08 (2017) 223–233, https://doi.org/10.4236/msa.2017.83015.
415770a.
[37] M.S.I.N. Kamariah, W.S.W. Harun, N.Z. Khalil, F. Ahmad, M.H. Ismail, S. Sharif,
[12] D. Kong, C. Dong, X. Ni, X. Li, Corrosion of metallic materials fabricated by
Effect of heat treatment on mechanical properties and microstructure of selective
selective laser melting, npj Mater. Degrad. 3 (2019), https://doi.org/10.1038/
laser melting 316L stainless steel, IOP Conf. Ser. Mater. Sci. Eng. 257 (1) (2017),
s41529-019-0086-1 no. November 2018, 24.
012021, https://doi.org/10.1088/1757-899X/257/1/012021.
[13] A. Hemmasian Ettefagh, S. Guo, Electrochemical behavior of AISI316L stainless
[38] M.L. M. Sistiaga, S. Nardone, C. Hautfenne, J. Van Humbeeck, “Effect of heat
steel parts produced by laser-based powder bed fusion process and the effect of
treatment of 316L stainless steel produced by Selective Laser Melting ( SLM),” in
post annealing process, Addit. Manuf. 22 (2018) 153–156, https://doi.org/
Proceedings of the 27th Annual International Solid Freeform Fabrication 2016:
10.1016/j.addma.2018.05.014.
Proceedings of the 26th Annual International Solid Freeform Fabrication
[14] F. Andreatta, A. Lanzutti, E. Vaglio, G. Totis, M. Sortino, L. Fedrizzi, Corrosion
Symposium, 2016, no. November.
behaviour of 316L stainless steel manufactured by selective laser melting, Mater.
[39] R.J. Williams, F. Vecchiato, J. Kelleher, M.R. Wenman, P.A. Hooper, C.M. Davies,
Corros. 70 (9) (2019) 1633–1645, https://doi.org/10.1002/maco.201910792.
Effects of heat treatment on residual stresses in the laser powder bed fusion of 316L
[15] B.F. Brown, Stress Corrosion Cracking Control Measures, 1, U.S. government
stainless steel: finite element predictions and neutron diffraction measurements,
printing office, Washington, 1977. First.
J. Manuf. Process. 57 (2020) 641–653.
[16] R.C. Newman, 2001 W.R. Whitney Award Lecture: understanding the corrosion of
[40] P. Ganesh, R. Giri, R. Kaul, P. Ram Sankar, P. Tiwari, A. Atulkar, R.K. Porwal, R.
stainless steel, Corrosion 57 (12) (2001) 1030–1041.
K. Dayal, L.M. Kukreja, Studies on pitting corrosion and sensitization in laser rapid
[17] K. Sieradzki, R.C. Newman, Stress-corrosion cracking, J. Phys. Chem. Solids 48
manufactured specimens of type 316L stainless steel, Mater. Des. 39 (2012)
(11) (1987) 1101–1113, https://doi.org/10.1016/0022-3697(87)90120-X.
509–521, https://doi.org/10.1016/j.matdes.2012.03.011.
[18] R.C. Newman, R.P.M. Procter, Stress corrosion cracking: 1965–1990, Br. Corros. J.
25 (4) (1990) 259–270, https://doi.org/10.1179/000705990799156373.

20
P. Dong et al. Additive Manufacturing 40 (2021) 101902

[41] J. Hlinka, M. Kraus, J. Hajnys, M. Pagac, J. Petrů, Z. Brytan, T. Tański, Complex [60] K.R. Trethewey, M. Wenman, P. Chard-Tuckey, B. Roebuck, Correlation of meso-
corrosion properties of AISI 316L steel prepared by 3D printing technology for and micro-scale hardness measurements with the pitting of plastically-deformed
possible implant applications, Materials 13 (7) (2020) 1527. Type 304L stainless steel, Corros. Sci. 50 (4) (2008) 1132–1141, https://doi.org/
[42] G.G. Scatigno, M.P. Ryan, F. Giuliani, M.R. Wenman, The effect of prior cold work 10.1016/j.corsci.2007.11.026.
on the chloride stress corrosion cracking of 304L austenitic stainless steel under [61] G. Wu, P.M. Singh, Effect of elastic stresses on pitting behavior of stainless steel
atmospheric conditions, Mater. Sci. Eng. A 668 (2016) 20–29, https://doi.org/ 304, J. Electrochem. Soc. 166 (8) (2019) 0–7, https://doi.org/10.1149/
10.1016/j.msea.2016.05.037. 2.1381906jes.
[43] L. Caseres, T.S. Mintz, “Atmospheric Stress Corrosion Cracking Susceptibility of [62] N. Islami, S. Rashid, A.K. Ariffin, M.Z. Nuawi, Stress corrosion damage on
Welded and Unwelded 304, 304L and 316L Austenitic Stainless Steels Commonly austenitic stainless steel in sodium chloride, Int. J. Automot. Mech. Eng. 14 (1)
Use for Dry Cask Storage Containers Exposed to Marine Engironments,” U.S.NRC, (2017) 3824–3836.
2010. [63] D. Kong, C. Dong, X. Ni, L. Zhang, J. Yao, C. Man, X. Cheng, K. Xiao, X. Li,
[44] X. He, T. Mintz, D. Dunn, R. Pabalan, G. Oberson, T. Ahn, “Stress corrosion Mechanical properties and corrosion behavior of selective laser melted 316L
cracking of UNS S30400 stainless steel exposed to atmospheric ammonium nitrate stainless steel after different heat treatment processes, J. Mater. Sci. Technol. 35
and sodium chloride mixtures,” NACE - Int. Corros. Conf. Ser., no. 3885, pp. 1–14, (7) (2019) 1499–1507, https://doi.org/10.1016/j.jmst.2019.03.003.
2014. [64] Q. Chao, V. Cruz, S. Thomas, N. Birbilis, P. Collins, A. Taylor, P.D. Hodgson,
[45] C. Örnek, D.L. Engelberg, Toward understanding the effects of strain and chloride D. Fabijanic, On the enhanced corrosion resistance of a selective laser melted
deposition density on atmospheric chloride-induced stress corrosion cracking of austenitic stainless steel, Scr. Mater. 141 (2017) 94–98, https://doi.org/10.1016/j.
type 304 austenitic stainless steel under MgCl 2 and FeCl 3:MgCl 2 droplets, scriptamat.2017.07.037.
Corrosion 75 (2) (2019) 167–182, https://doi.org/10.5006/3026. [65] G. Hinds, L. Wickström, K. Mingard, A. Turnbull, Impact of surface condition on
[46] G.G. Scatigno, P. Dong, M.P. Ryan, M.R. Wenman, The effect of salt loading on sulphide stress corrosion cracking of 316L stainless steel, Corros. Sci. 71 (2013)
chloride-induced stress corrosion cracking of 304L austenitic stainless steel under 43–52, https://doi.org/10.1016/j.corsci.2013.02.002.
atmospheric conditions, Materialia (2019), https://doi.org/10.1016/j. [66] S. Hossain, M.R. Daymond, C.E. Truman, D.J. Smith, Prediction and measurement
mtla.2019.100509. of residual stresses in quenched stainless-steel spheres, Mater. Sci. Eng. A 373
[47] P. Dong, G.G. Scatigno, M.R. Wenman, Effect of salt composition and (2004) 339–349, https://doi.org/10.1016/j.msea.2004.02.014.
microstructure on stress corrosion cracking of 316L austenitic stainless steel for dry [67] F. Jiarui, W. Yan, The effect of shot-peening in SCC resistance of steam generator
storage canisters, J. Nucl. Mater. (2020), 152572, https://doi.org/10.1016/j. heat-transfer tube material, Nucl. Power Eng. 9 (5) (1988) 7–16.
jnucmat.2020.152572. [68] A. Turnbull, L. Crocker, S. Zhou, Do corrosion pits eliminate the benefit of shot-
[48] Renishaw, “Design for metal AM - a beginner’s guide,” 2017. peening? Int. J. Fatigue 116 (2018) 439–447, https://doi.org/10.1016/j.
[49] S. Ghosh, V. Kain, Effect of surface machining and cold working on the ambient ijfatigue.2018.07.004.
temperature chloride stress corrosion cracking susceptibility of AISI 304L stainless [69] M.N. Gussev, K.J. Leonard, In situ SEM-EBSD analysis of plastic deformation
steel, Mater. Sci. Eng. A 527 (3) (2010) 679–683, https://doi.org/10.1016/j. mechanisms in neutron-irradiated austenitic steel, J. Nucl. Mater. 517 (2019)
msea.2009.08.039. 45–56, https://doi.org/10.1016/j.jnucmat.2019.01.034.
[50] A. Turnbull, K. Mingard, J.D. Lord, B. Roebuck, D.R. Tice, K.J. Mottershead, N. [70] M.A. Arafin, J.A. Szpunar, A new understanding of intergranular stress corrosion
D. Fairweather, A.K. Bradbury, Sensitivity of stress corrosion cracking of stainless cracking resistance of pipeline steel through grain boundary character and
steel to surface machining and grinding procedure, Corros. Sci. 53 (10) (2011) crystallographic texture studies, Corros. Sci. 51 (1) (2009) 119–128, https://doi.
3398–3415, https://doi.org/10.1016/j.corsci.2011.06.020. org/10.1016/j.corsci.2008.10.006.
[51] H. Meier, C. Haberland, Experimental studies on selective laser melting of metallic [71] T. Fujii, T. Furumoto, K. Tohgo, Y. Shimamura, Crystallographic evaluation of
parts, Materwiss. Werksttech. 39 (9) (2008) 665–670, https://doi.org/10.1002/ susceptibility to intergranular corrosion in austenitic stainless steel with various
mawe.200800327. degrees of sensitization, Materials 13 (2020) 613.
[52] I. Yadroitsev, I. Yadroitsava, Evaluation of residual stress in stainless steel 316L [72] T. Liu, S. Xia, Q. Bai, B. Zhou, Y. Lu, T. Shoji, Evaluation of grain boundary network
and Ti6Al4V samples produced by selective laser melting, Virtual Phys. Prototyp. and improvement of intergranular cracking resistance in 316L stainless steel after
10 (2) (2015) 67–76, https://doi.org/10.1080/17452759.2015.1026045. grain boundary engineering, Materials (2019) 242, https://doi.org/10.3390/
[53] R.L. Higginson, C.P. Jackson, E.L. Murrell, P. Exworthy, R.J. Mortimer, D. ma12020242.
R. Worrall, G.D. Wilcox, Effect of thermally grown oxides on colour development of [73] M. Michiuchi, H. Kokawa, Z.J. Wang, Y.S. Sato, K. Sakai, Twin-induced grain
stainless steel, Mater. High Temp. 32 (1–2) (2015) 113–117, https://doi.org/ boundary engineering for 316 austenitic stainless steel, Acta Mater. 54 (19) (2006)
10.1179/0960340914Z.00000000083. 5179–5184, https://doi.org/10.1016/j.actamat.2006.06.030.
[54] J.W. Fu, Y.S. Yang, J.J. Guo, W.H. Tong, Effect of cooling rate on solidification [74] V.Y. Gertsman, S.M. Bruemmer, Study of grain boundary character along
microstructures in AISI 304 stainless steel, Mater. Sci. Technol. 24 (8) (2008) intergranular stress corrosion crack paths in austenitic alloys, Acta Mater. 49 (9)
941–944, https://doi.org/10.1179/174328408×295962. (2001) 1589–1598, https://doi.org/10.1016/S1359-6454(01)00064-7.
[55] J. Liu, Y. Song, C. Chen, X. Wang, H. Li, C. Zhou, J. Wang, K. Guo, J. Sun, Effect of [75] F. Eltaghoor, Transgranular Stress Corrosion Cracking of 316 L Stainless Steel in
scanning speed on the microstructure and mechanical behavior of 316L stainless Chloride Environment at 80oC, The University of Manchester, 2017.
steel fabricated by selective laser melting, Mater. Des. 186 (2020), 108355, [76] B. Wang, M.Y. Jiang, M. Xu, C.W. Cui, J. Wang, W. Cheng, Influence of heat
https://doi.org/10.1016/j.matdes.2019.108355. treatment on corrosion resistance of explosive welding 316L stainless steel
[56] G.G. Scatigno, Chloride-Induced Transgranular Stress Corrosion Cracking of composite plate, Mater. Res. Express 6 (10) (2019), 106575, https://doi.org/
Austenitic Stainless Steel 304L,” Imperial College London. 10.1088/2053-1591/ab358e.
[57] G. Sander, J. Electrochem, C. Soc, “On The Corrosion and Metastable Pitting [77] M. Laleh, A.E. Hughes, W. Xu, I. Gibson, M.Y. Tan, Unexpected erosion-corrosion
Characteristics of 316L Stainless Steel Produced by Selective Laser Melting On The behaviour of 316L stainless steel produced by selective laser melting, Corros. Sci.
Corrosion and Metastable Pitting Characteristics of 316L Stainless Steel Produced 155 (2019) 67–74, https://doi.org/10.1016/j.corsci.2019.04.028.
by Selective Laser Melting,” 2017, doi: 10.1149/2.0551706jes. [78] D. Kong, C. Dong, X. Ni, L. Zhang, H. Luo, R. Li, L. Wang, C. Man, X. Li, The
[58] J.R. Scully, N. Birbilis, “Corrosion of Additively Manufactured Alloys: A Review,” passivity of selective laser melted 316L stainless steel, Appl. Surf. Sci. 504 (2020),
vol. 9312, pp. 1318–1350, 2018. 144495, https://doi.org/10.1016/j.apsusc.2019.144495.
[59] F. Elshawesh, A. Elhoud, “Role of heat tint on pitting corrosion of 304 austenitic
stainless steel in chloride environment,” in EUROCORR 2004: long term prediction
and modeling of corrosion, 2004.

21

You might also like