You are on page 1of 7

Corrosion Science 161 (2019) 108189

Contents lists available at ScienceDirect

Corrosion Science
journal homepage: www.elsevier.com/locate/corsci

Short Communication

On the unusual intergranular corrosion resistance of 316L stainless steel T


additively manufactured by selective laser melting
Majid Laleha,b, Anthony E. Hughesb,c, Wei Xua, Nima Haghdadib, Ke Wangb, Pavel Cizekb,
Ian Gibsona, Mike Yongjun Tana,b,

a
Deakin University, School of Engineering, Faculty of Science and Technology, Geelong Waurn Ponds Campus, Waurn Ponds, Victoria, 3216, Australia
b
Deakin University, Institute for Frontier Materials, Geelong Waurn Ponds Campus, Waurn Ponds, Victoria, 3216, Australia
c
Commonwealth Scientific and Industrial Research Organisation (CSIRO), Mineral Resources, Private Bag 10, Clayton South, Victoria, 3169, Australia

ARTICLE INFO ABSTRACT

Keywords: The intergranular corrosion (IGC) resistance of 316L stainless steel (316L SS) produced by selective laser melting
Intergranular corrosion (SLM) was investigated using microscopy analysis and electrochemical measurements. The IGC resistance of
316L stainless steel SLM-produced 316L SS, determined using a double-loop electrochemical potentiokinetic reactivation test, was
Additive manufacturing found to be substantially higher than that of conventional 316L SS. This unusual behaviour was explained by the
Grain boundary
fact that no Cr-rich precipitates were detected for SLM-produced specimens after long-term sensitisation heat-
Precipitation
treatment and those SLM-produced specimens exhibited a high frequency of twin boundaries and low-angle
grain boundaries along with fine grains, leading to the avoidance of localised Cr depletion.

1. Introduction properties such as weaker erosion-corrosion resistance of SLM-pro-


duced 316L stainless steel (hereafter 316L SS) [11].
Austenitic stainless steels are widely used in engineering structures It is well-known that an important microstructural characteristic of
operating at elevated temperatures such as in steam generating plants a crystalline material that could be highly influenced by the thermal
as piping and superheating tube materials. Under such harsh environ- history during the production process is the grain boundary (GB)
mental conditions, intergranular corrosion (IGC) of conventional character. Such change in GB character could lead to changes in the IGC
stainless steels is frequently observed, especially in sections joined by behaviour of SLM-produced stainless steels. Research on IGC resistance
welding at temperatures between 500 and 850 °C. Additive manu- over the years has shown that the GB character has a crucial influence
facturing (AM), an emerging net-shape layer-wise fabrication process, is on precipitation behaviour and IGC susceptibility [12–16]. Measure-
being adopted for producing structural components used at elevated ments of GB character, with special emphasis on coincidence site lattice
temperatures because of its ability to produce integrated complex parts (CSL) boundaries, have been the focus of GB engineering in improving
in a single step without the need for joining like welding. Selective laser corrosion resistance. In spite of extensive studies on the microstructural
melting (SLM) is a powder-bed AM technique, in which a component is evolution and properties of SLM-produced 316L SS [17–22], the IGC
produced by selectively melting consecutive layers of powder on top of behaviour of SLM-produced stainless steels has yet to be studied in any
each other using a high-energy laser beam [1–3]. Upon irradiation, the detail. A recent report has found that the SLM-produced 316L SS ex-
powder is melted and forms a tiny melt pool. This generates extremely hibited a more rapid sensitization of GBs upon exposure to elevated
high temperatures up to 105 °C and rapid cooling up to 106-108 °C/s temperature compared to the wrought counterpart, however, the pre-
within the melt pool [4,5]. The thermal history that materials experi- cipitation behaviour and its subsequent influence on the IGC resistance
ence during SLM processing is very different from that established by are not well understood [23].
conventional manufacturing techniques [6,7]. The rapid melting and In the present study, an attempt was made to provide insight into
solidification in combination with cyclic heating and cooling upon the how the nonconventional thermal history associated with SLM pro-
deposition of subsequent layers result in a microstructure that differs cessing can influence the GB character, precipitation behaviour and IGC
from traditionally-produced parts [8–10]. This has resulted in unusual susceptibility with the support of microscopy analysis and


Corresponding author at: Deakin University, School of Engineering, Faculty of Science and Technology, Geelong Waurn Ponds Campus, Waurn Ponds, Victoria,
3216, Australia.
E-mail address: mike.tan@deakin.edu.au (M.Y. Tan).

https://doi.org/10.1016/j.corsci.2019.108189
Received 24 May 2019; Received in revised form 14 August 2019; Accepted 28 August 2019
Available online 30 August 2019
0010-938X/ © 2019 Elsevier Ltd. All rights reserved.
M. Laleh, et al. Corrosion Science 161 (2019) 108189

electrochemical measurements. (DL-EPR) test in 0.5 M H2SO4 + 0.01 M KSCN solution at room tem-
perature. For DL-EPR testing, firstly the potential was scanned in the
2. Experimental procedure anodic direction from open circuit potential (OCP) to +0.35 V vs. the
reference electrode at a scanning rate of 100 mV/min and then im-
2.1. Materials and heat-treatment mediately reversed to the starting point at the same scanning rate. DOS
was reported as the ratio of maximum current density in the reactiva-
Commercially available gas-atomized spherical 316L SS powder, tion loop (Ir) to current density in the activation loop (Ia). The DL-EPR
with a particle size range between 5 and 40 μm, was used to produce experiments were repeated at least three times and the average values
specimens. Gas-atomisation was performed under argon gas atmo- of DOS are reported. After DL-EPR testing, the specimens were ex-
sphere. An SLM®-125HL machine was used to produce cubic specimens amined using optical microscopy (Olympus GX41) to analyse the in-
with dimensions of 1.0 × 1.0 × 1.0 cm3. Prior to the SLM fabrication, tergranular attack.
the build plate was pre-heated to a temperature of 200 °C and the build
chamber was purged with purified argon until the oxygen level was 3. Results and discussion
reduced to below 100 ppm. The main processing parameters were laser
power of 150 W, scanning speed of 600 mm/s, hatch spacing of 80 μm In general, it is believed that the loss of intergranular corrosion
and layer thickness of 30 μm. The scanning pattern was bidirectional resistance at the GBs in stainless steels is due to the formation of GB
with 67° angle between successive layers. The density of the SLM-pro- chromium carbides that cause localised Cr depletion in the region ad-
duced specimens was 99.5 ± 0.1%, measured based on the jacent to these precipitates [25–27]. The intergranular attack is be-
Archimedes method, using an electronic Densimeter (Model SD-200 L) lieved to be accelerated by the increasing potential difference between
with 0.0001 g/cm3 resolution. The commercial (wrought) 316L SS was grain interiors (cathode) and GBs (anode). The DL-EPR test is a common
received as 3 mm thick plate and then was solution annealed at 1100 °C method for assessing such behaviour based on the assumption that only
for 1 h followed by water quenching, prior to sensitisation studies [24]. the sensitised GBs become active, while grain interiors are unsensitised
Chemical composition of the SLM-produced specimen is presented in [28,29]. Therefore, gaining further insight into the characteristics of
Table 1, showing a comparable chemical composition with that of the GBs and grain interiors as well as their correlation with the DL-EPR
commercial counterpart. For the sensitisation experiments, the speci- behaviour is essential for a critical understanding of the IGC behaviour
mens were heat-treated at 700 °C for 60 h and then immediately water of SLM-produced 316L SS.
quenched. As shown in Fig. 1, the EBSD analyses of the microstructures in the
commercial and SLM-produced 316L SS specimens reveals an obvious
2.2. Microstructural characterisation difference in grain size and shape as well as the GB character. The
commercial 316L SS mostly comprised equiaxed grains (Fig. 1a).
For microscopy analysis, the specimens were mechanically polished However, grains in the SLM-produced specimen showed a fairly elon-
to a mirror finish with the last stage being a 0.3 μm oxide polishing gated morphology on the build plane (parallel to the build direction), as
suspension (OPS). A scanning electron microscope (SEM ZEISS SUPRA presented in Fig. 1b. The same specimen showed elongated grains se-
55 V FEG) equipped with an energy dispersive spectroscopy (EDS) de- parated with melt pool boundaries when viewed from the transverse
tector and operated at 20 kV was used for imaging and EDS analysis. plane (perpendicular to the build direction), as shown in Fig. 1c. The
Transmission electron microscopy (TEM) was employed for a detailed mean grain size was measured to be 19.2 ± 0.7 μm, 8.4 ± 0.8 μm and
analysis of precipitates, using a JEOL JEM 2100 F microscope equipped 10.4 ± 0.9 μm for commercial, and SLM-produced specimens at
with an EDS detector and operated at 200 kV. EDS mapping was per- transverse and build planes, respectively, indicating a smaller grain size
formed in a scanning-transmission TEM (STEM) mode using a spot size for the SLM-produced specimens compared to their commercial coun-
of 1 nm. For TEM analysis, thin slices with a thickness of approximately terpart. The relative fraction and length of the GBs, measured from the
500 μm were cut from the specimens and then mechanically ground to a EBSD data, are summarised in Table 2 for all three microstructures.
thickness of around 100 μm. Discs with a diameter of 3 mm were sub- Interestingly, these microstructures comprised a significant fraction of
sequently punched from the slices and further ground using 4000 ∑3 boundaries and a relatively low fraction of ∑9 boundaries. A detailed
abrasive paper until their thickness was reduced to 60–70 μm. Final analysis of angle/axis pair associated with ∑3 boundaries (Brandon's
thinning was carried out by twin-jet electro-polishing using a Struers criterion [30]) revealed that almost 35%, 25% and 31% of commercial,
Tenupol 5 device. Electron backscatter diffraction (EBSD) acquisition SLM-produced at transverse and build plane microstructures were of ∑3
was performed using a ZEISS LEO 1530 FEG SEM, with a scan step size character. As can be seen in the misorientation angle distribution plots
of 0.3 μm at an accelerating voltage of 20 kV. Oxford Instruments Aztec for the surfaces subjected to EBDS, the ∑3 peak at 60° is much sharper
software was used for acquisition and indexing the EBSD patterns. for the commercial specimen. However, a relatively lower fraction of
Further EBSD data post-processing was performed using HKL Channel 5 the 60° boundaries is of a < 111 > misorientation axis in the com-
software. An area of 500 × 500 μm2 was used for EBSD analysis. A mercial microstructure compared to those in the SLM-produced speci-
ZEISS FEI QUANTA focused ion beam (FIB) SEM was used for milling of mens. The same trend was observed for the misorientation axis peak
cross-sections at grain boundaries. at < 110 > for the misorientation angle of 39°. This was not surprising
as a ∑9 boundary is dictated by the geometrical constraints of two in-
2.3. Electrochemical tests tersecting ∑3 boundaries [31].
It is important to note that an analysis of the misorientation angle
The sensitized specimens were evaluated for degree of sensitization alone can be misleading in characterising the GB character [32]. Due to
(DOS) using a double loop electrochemical potentiokinetic reactivation the difference in the grain size, the length of different boundaries per

Table 1
Chemical composition of the commercial and as-built SLM-produced 316L SS.
Element (wt%) C Cr Ni Mn Mo Si P S Nb Ti Cu Fe

Commercial 0.02 16.9 10.30 1.77 2.11 0.50 0.04 0.01 0.03 < 0.01 0.39 Bal.
As-built SLM 0.02 16.6 11.60 1.11 2.70 0.56 0.01 0.01 0.01 < 0.01 0.08 Bal.

2
M. Laleh, et al. Corrosion Science 161 (2019) 108189

Fig. 1. EBSD IPF maps (a,b,c) and band contrast maps (d,e,f) of the commercial specimen (a,d), SLM-produced specimen in the build plane (b,e), and in the
transverse plane (c,f). Z shows the build direction. The red and blue lines in (d–f) are ∑3 and ∑9 CSL boundaries, respectively. Misorientation angle/axis distribution
for the austenite/austenite boundaries for commercial specimen (g), SLM-produced specimen in the build plane (h), and in the transverse plane (i). The scale bars in
(a–f) are the same. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

Table 2 that of the commercial specimen, respectively. Hence, from the EBSD
Grain boundary characteristics for all three microstructures measured from the analysis, it is clear that the grain size and shape, as well as the GB
EBSD results presented in Fig. 1. character of the SLM-produced specimens are quite different from those
0.7° < θ < 2° 2° < θ < 15° 15° < θ ∑3 of their commercial counterpart, which is expected to influence the
precipitation behaviour during the sensitisation heat-treatment and
Commercial Fraction 1.0% 6.0% 92.5% 35.1% subsequent IGC resistance.
Length 394.80 μm 2.34 mm 3.58 cm 1.33 cm
Fig. 2 shows micrographs of the specimens after the sensitisation
SLM-produced Fraction 9.4% 4.7% 85.0% 24.8% heat-treatment, indicating the formation of precipitates along GBs. The
(transverse Length 8.23 mm 4.14 mm 7.45 cm 2.09 cm
dark spherical spots in Fig. 2b, c are inclusions, which are inherent in
plane)
austenitic stainless steels produced by SLM [11,18], and are not due to
SLM-produced Fraction 10.7% 4.0% 84.6% 31.3% the sensitisation heat-treatment. For the commercial 316L SS, coarse
(build plane) Length 8.39 mm 3.14 mm 6.63 cm 2.37 cm
precipitates were formed along GBs (Fig. 2a, d). STEM-EDS analysis
showed that these GB precipitates in commercial 316L SS are enriched
in Cr and C (Fig. 2d), indicating the formation of chromium carbides.
given area should also be considered in evaluating the GB character. It For the SLM-produced specimens, however, relatively fine precipitates
was revealed in Table 2 that the relative length of low-angle GBs are visible (Fig. 2b, c). Interestingly, the STEM-EDS analysis shown in
(2° < θ < 15°), which are of lower energy compared to their high-angle Fig. 2e reveals that these precipitates are indeed enriched in Mo and Si
counterparts, is approximately 75% and 35% higher in the micro- instead of Cr (Table 3), which can be considered as Laves phase. At first
structure of the SLM-produced specimens at the transverse and build glance, one may consider these precipitates in the SLM-produced 316L
plane surfaces, respectively, relative to the commercial specimen. The SS as sigma and chi phases, which are commonly found in duplex
lengths of the ∑3 boundaries measured at the transverse and build plane stainless steels when subjected to sensitisation heat-treatment [32–35].
surfaces of the SLM-produced specimen were 57% and 73% higher than In duplex stainless steels, these precipitates (sigma/chi) can quickly

3
M. Laleh, et al. Corrosion Science 161 (2019) 108189

Fig. 2. SEM micrographs showing the precipitation along grain boundaries for commercial (a) and SLM-produced 316L SS at the transverse plane (b) and build plane
(c). The black spherical spots in (b,c) are the inclusions, which are inherent in stainless steels produced by SLM, and are not due to the sensitisation heat-treatment.
STEM images of the precipitates for the commercial (d) and SLM-produced (e) 316L SS specimens with their EDS elemental map analysis. The dark spot shown in the
dashed red line in (e) is the inclusion and is not due to the sensitisation heat-treatment. (For interpretation of the references to colour in this figure legend, the reader
is referred to the web version of this article.)

Table 3 obviously higher DOS (25.7) was recorded for the commercial 316L SS
STEM/EDS analysis of the precipitates in SLM-produced specimens after heat- compared to those recorded for the SLM-produced specimens, i.e. 0.08
treatment at 700 °C for 60 h from Fig. 2. and 0.06 for transverse and build planes, respectively, as depicted in
Element (wt%) Fe Cr Mo Si Fig. 3b. The lack of reactivation peak for the SLM-produced specimens
indicates a minimal density of Cr-rich precipitates, in agreement with
Laves phase 54.41 ± 0.32 16.15 ± 0.45 26.78 ± 0.65 2.67 ± 0.17 the STEM-EDS compositional analysis (Fig. 2e). The high DOS in the
commercial 316L SS is consistent with the presence of Cr-rich pre-
cipitates shown in SEM and STEM images in Fig. 2. It seems that the IGC
form when the Cr content is between 20 and 30%. Furthermore, the resistance is mostly controlled by the Cr-rich (chromium carbides)
sigma/chi precipitation tend to occur preferentially in the delta ferrite precipitates rather than Mo-rich (Laves) ones [41]. Optical microscopic
islands [36]. But, based on the EBSD results, no delta ferrite was de- examination after DL-EPR measurements revealed a ‘ditch’ structure for
tected for the SLM-produced specimens in our study, and the material is commercial 316L SS, where all the GBs were deeply corroded (Fig. 3c
a single-phase austenite. In addition, chemical compositional analysis and d). For the SLM-produced 316L SS specimens at both transverse
by STEM/EDS (Table 3) shows the enrichment of the Si in the pre- (Fig. 3e, f) and build (Fig. 3g, h) planes, the extent of attack at GB
cipitates in the SLM-produced specimens, which confirms these pre- regions is much less than that of their commercial counterpart.
cipitates are Laves phase. This is in agreement with the literature, To get a further insight into the corrosion attack along GBs, the
where the presence of Si has been reported in the Laves phase [37–39]. cross-section of the intergranular corrosion was examined using FIB-
More importantly, no chromium carbides were detected for the SLM- SEM. Fig. 4(a–c) shows FIB images for the commercial specimen in-
produced specimens. So, the precipitation behaviour for commercial dicating a substantial GB attack extending into the matrix (after the DL-
and SLM-produced 316L SS is different under exposure to long-term EPR test). By contrast, the GB attack for the SLM-produced specimens is
sensitisation heat-treatment. much less than that of their commercial counterpart. So, the extent of
For the additively manufactured 316L SS, the segregation of Mo in the attack along the GBs correlates well with the DOS values estimated
the cell walls has already been shown in the literature [18]. One pos- from the DL-EPR curves; a higher DOS corresponds to a deeper attack
sible explanation for the formation of Laves precipitates in the SLM- along GBs.
produced 316L SS might be the diffusion of Mo from the cell walls to The results reported here show the substantially higher IGC re-
the GBs during the long-term sensitisation heat-treatment, which needs sistance of the SLM-produced specimens. When subjected to sensitisa-
further clarification in the future work. A similar explanation has been tion between 500–850 °C, the formation of GB Cr-rich precipitates in
made by Isik et al. [40], showing that segregation of Mo to grain conjunction with the presence of Cr-depleted regions near GBs tends to
boundaries in steel can promote the formation of Laves phase during cause IGC in stainless steels [25–27]. Previous studies on the relation-
high-temperature exposure. ship between GB characteristics and IGC resistance in stainless steels
To find the influence of precipitation on the IGC resistance, DL-EPR have shown that only low-angle GBs and twin boundaries were immune
tests were carried out, revealing a definite reactivation peak for the from being sensitised [15,42–46]. In the present study, the substantial
commercial specimen, as shown in Fig. 3a. In contrast, DL-EPR analysis decrease in the degree of sensitisation of SLM-produced specimens can
of the SLM-produced specimens revealed no reactivation peaks. An

4
M. Laleh, et al. Corrosion Science 161 (2019) 108189

Fig. 3. (a) DL-EPR curves for the commercial and SLM-produced 316L SS specimens recorded in 0.5 M H2SO4 + 0.01 M KSCN solution at room temperature; (b) DOS
values measured from the peak current densities of the forward and reverse scans in DL-EPR curves; optical images after DL-EPR tests for (c, d) commercial 316L SS
showing that all grain boundaries got severely attacked, and SLM-produced 316L SS at transverse (e, f) and build (g, h) surfaces, respectively, indicating a negligible
attack along grain boundaries.

Fig. 4. FIB cross-sections made on typical corroded grain boundaries to measure the depth of attack for (a–c) commercial and (d–g) SLM-produced 316L stainless
steel indicating an aggressive attack along the grain boundaries of the commercial specimen, and a slight attack for the SLM-produced 316L SS. The white rectangles
in (a, d, e) show the regions where the high magnification images in (b, c, f, g) were taken.

5
M. Laleh, et al. Corrosion Science 161 (2019) 108189

be manifested in two ways viz. the smaller grain size and larger length melting: materials and applications, Appl. Phys. Rev. 2 (4) (2015) 041101.
of CSL boundaries. It is believed that the degree of sensitisation has an [3] D. Herzog, V. Seyda, E. Wycisk, C. Emmelmann, Additive manufacturing of metals,
Acta Mater. 117 (2016) 371–392.
inverse relationship with the grain size [13,47]. The smaller the grain [4] D. Gu, H. Wang, G. Zhang, Selective laser melting additive manufacturing of Ti-
size, the larger the total GB surface area, which in turn provides a larger based nanocomposites: the role of nanopowder, Metall. Mater. Trans. A 45 (1)
number of sites for precipitates to nucleate. However, the precipitates (2014) 464–476.
[5] M. Das, V.K. Balla, D. Basu, S. Bose, A. Bandyopadhyay, Laser processing of SiC-
need to attain a critical size to cause sensitisation. In the case of fine- particle-reinforced coating on titanium, Scr. Mater. 63 (4) (2010) 438–441.
grained materials, such as the SLM-produced material in the present [6] E.A. Jägle, P.-P. Choi, J. Van Humbeeck, D. Raabe, Precipitation and austenite re-
study, the carbon availability for every single nucleus is restricted be- version behavior of a maraging steel produced by selective laser melting, J. Mater.
Res. 29 (17) (2014) 2072–2079.
cause of the higher number of nuclei and therefore chromium carbides [7] S. Kelly, S. Kampe, Microstructural evolution in laser-deposited multilayer Ti-6Al-
can’t precipitate, which is consistent with the observations in SEM and 4V builds: part II. Thermal modeling, Metall. Mater. Trans. A 35 (6) (2004)
STEM images in Fig. 2, where no trace of chromium carbides was de- 1869–1879.
[8] Z. Wang, T.A. Palmer, A.M. Beese, Effect of processing parameters on micro-
tected for SLM-produced specimens. Furthermore, the intergranular
structure and tensile properties of austenitic stainless steel 304L made by directed
corrosion was suppressed due to the uniform distribution of highly energy deposition additive manufacturing, Acta Mater. 110 (2016) 226–235.
coherent CSL boundaries, minimising the extent of intergranular cor- [9] R. Pokharel, L. Balogh, D. Brown, B. Clausen, G. Gray, V. Livescu, S. Vogel,
rosion along the GBs [48], as indicated in the FIB sections in Fig. 4. S. Takajo, Signatures of the unique microstructure of additively manufactured steel
observed via diffraction, Scr. Mater. 155 (2018) 16–20.
After subjected to the long-term sensitisation heat-treatment, the [10] M. Pham, B. Dovgyy, P. Hooper, Twinning induced plasticity in austenitic stainless
SLM-produced specimens have substantially lower IGC susceptibility steel 316L made by additive manufacturing, Mater. Sci. Eng. A 704 (2017)
compared to their conventionally-produced counterpart. Thermo-me- 102–111.
[11] M. Laleh, A.E. Hughes, W. Xu, I. Gibson, M.Y. Tan, Unexpected erosion-corrosion
chanical treatments have been usually applied to develop a maximum behaviour of 316L stainless steel produced by selective laser melting, Corros. Sci.
proportion of low-angle GBs to improve the IGC resistance of austenitic 155 (2019) 67–74.
stainless steels, which is known as grain boundary engineering (GBE). [12] N. Haghdadi, D. Abou-Ras, P. Cizek, P. Hodgson, A. Rollett, H. Beladi, Austenite-
ferrite interface crystallography dependence of sigma phase precipitation using the
The results presented here show the potential of AM techniques in the five-parameter characterization approach, Mater. Lett. 196 (2017) 264–268.
fabrication of IGC immune microstructures with finer grain size and [13] M. Laleh, F. Kargar, Suppression of chromium depletion and sensitization in aus-
abundant CSL boundaries, without resorting to any post-processing. tenitic stainless steel by surface mechanical attrition treatment, Mater. Lett. 65 (12)
(2011) 1935–1937.
Although GBE and its influence on the IGC resistance have been widely [14] A. Chen, W. Hu, D. Wang, Y. Zhu, P. Wang, J. Yang, X. Wang, J. Gu, J. Lu,
studied for conventionally-produced stainless steels, tailoring GB Improving the intergranular corrosion resistance of austenitic stainless steel by high
characteristics in situ via manipulation of the SLM processing conditions density twinned structure, Scr. Mater. 130 (2017) 264–268.
[15] M. Shimada, H. Kokawa, Z. Wang, Y. Sato, I. Karibe, Optimization of grain
opens a new avenue for the direct SLM manufacture of IGC resistant
boundary character distribution for intergranular corrosion resistant 304 stainless
316L SS components. steel by twin-induced grain boundary engineering, Acta Mater. 50 (9) (2002)
2331–2341.
4. Summary [16] T. Watanabe, An approach to grain-boundary design for strong and ductile poly-
crystals, Res. Mech. 11 (1) (1984) 47–84.
[17] J.J. Lewandowski, M. Seifi, Metal additive manufacturing: a review of mechanical
The IGC resistance of an additively manufactured 316L SS was properties, Annu. Rev. Mater. Res. 46 (2016) 151–186.
studied using a combination of microscopy analysis and electro- [18] Y.M. Wang, T. Voisin, J.T. McKeown, J. Ye, N.P. Calta, Z. Li, Z. Zeng, Y. Zhang,
W. Chen, T.T. Roehling, Additively manufactured hierarchical stainless steels with
chemical measurements. The relationship between the IGC and GB high strength and ductility, Nat. Mater. 17 (1) (2018) 63.
character of the SLM-produced 316L SS was established for the first [19] K. Saeidi, X. Gao, Y. Zhong, Z.J. Shen, Hardened austenite steel with columnar sub-
time. No Cr-rich precipitates were detected for SLM-produced speci- grain structure formed by laser melting, Mater. Sci. Eng. A 625 (2015) 221–229.
[20] G. Sander, S. Thomas, V. Cruz, M. Jurg, N. Birbilis, X. Gao, M. Brameld,
mens after a long-term sensitisation heat-treatment. Subsequently, DL- C. Hutchinson, On the corrosion and metastable pitting characteristics of 316L
EPR tests showed substantially lower DOS values for the SLM produced stainless steel produced by selective laser melting, J. Electrochem. Soc. 164 (6)
316L SS compared to its commercial counterpart. Post-corrosion mi- (2017) C250–C257.
[21] Q. Chao, V. Cruz, S. Thomas, N. Birbilis, P. Collins, A. Taylor, P.D. Hodgson,
croscopy analysis revealed a continuous network of ditched GBs with
D. Fabijanic, On the enhanced corrosion resistance of a selective laser melted
grooves extending deeply into the bulk in the case of commercial 316L austenitic stainless steel, Scr. Mater. 141 (2017) 94–98.
SS; while for the SLM-produced specimens, the attack along GBs was [22] S.-H. Sun, T. Ishimoto, K. Hagihara, Y. Tsutsumi, T. Hanawa, T. Nakano, Excellent
mechanical and corrosion properties of austenitic stainless steel with a unique
less pronounced, indicating higher resistance of SLM-produced 316L SS
crystallographic lamellar microstructure via selective laser melting, Scr. Mater. 159
to IGC. This behaviour was found to be related to the GB character, (2019) 89–93.
where the SLM-produced specimens exhibited a high frequency of twin [23] D. Macatangay, S. Thomas, N. Birbilis, R. Kelly, Unexpected interface corrosion and
boundaries and low-angle GBs along with fine grains. sensitization susceptibility in additively manufactured austenitic stainless steel,
Corrosion 74 (2) (2017) 153–157.
[24] X. Yu, S. Chen, Y. Liu, F. Ren, A study of intergranular corrosion of austenitic
Data availability statement stainless steel by electrochemical potentiodynamic reactivation, electron back-
scattering diffraction and cellular automaton, Corros. Sci. 52 (6) (2010) 1939–1947.
[25] B. Hopkinson, K. Carroll, Chromium distribution around grain boundary carbides
The raw/processed data required to reproduce these findings cannot found in austenitic stainless steel, Nature 184 (4697) (1959) 1479.
be shared at this time as the data also forms part of an ongoing study. [26] H. Kokawa, M. Shimada, Y.S. Sato, Grain-boundary structure and precipitation in
sensitized austenitic stainless steel, Jom 52 (7) (2000) 34–37.
[27] V. Ĉíhal, I. Kašová, Relation between carbide precipitation and intercrystalline
Acknowledgement corrosion of stainless steels, Corros. Sci. 10 (12) (1970) 875–881.
[28] A.P. Majidi, M.A. Streicher, Potentiodynamic reactivation method for detecting
Financial support from Deakin University Postgraduate Research sensitization in AISI 304 and 304L stainless steels, Corrosion 40 (8) (1984)
393–408.
Scholarship (DUPRS) is greatly appreciated. Deakin University’s
[29] G. Aydoğdu, M. Aydinol, Determination of susceptibility to intergranular corrosion
Advanced Characterisation Facility is acknowledged for use of the mi- and electrochemical reactivation behaviour of AISI 316L type stainless steel, Corros.
croscopy instruments and assistance from Dr Adam Taylor and Dr Mark Sci. 48 (11) (2006) 3565–3583.
[30] D. Brandon, The structure of high-angle grain boundaries, Acta Metall. 14 (11)
Nave.
(1966) 1479–1484.
[31] G.S. Rohrer, V. Randle, C.-S. Kim, Y. Hu, Changes in the five-parameter grain
References boundary character distribution in α-brass brought about by iterative thermo-
mechanical processing, Acta Mater. 54 (17) (2006) 4489–4502.
[32] N. Haghdadi, M. Laleh, A. Kosari, M.H. Moayed, P. Cizek, P.D. Hodgson, H. Beladi,
[1] I. Gibson, D.W. Rosen, B. Stucker, Additive Manufacturing Technologies, Springer, The effect of phase transformation route on the intergranular corrosion suscept-
2014. ibility of 2205 duplex stainless steel, Mater. Lett. 238 (2019) 26–30.
[2] C. Yap, C. Chua, Z. Dong, Z. Liu, D. Zhang, L. Loh, S. Sing, Review of selective laser [33] N. Llorca-Isern, H. López-Luque, I. López-Jiménez, M.V. Biezma, Identification of

6
M. Laleh, et al. Corrosion Science 161 (2019) 108189

sigma and chi phases in duplex stainless steels, Mater. Charact. 112 (2016) 20–29. [41] A.B. Rhouma, T. Amadou, H. Sidhom, C. Braham, Correlation between micro-
[34] C.-C. Hsieh, W. Wu, Overview of intermetallic sigma (σ) phase precipitation in structure and intergranular corrosion behavior of low delta-ferrite content AISI
stainless steels, ISRN Metall. 2012 (2012). 316L aged in the range 550–700 °C, J. Alloys Compd. 708 (2017) 871–886.
[35] L. Tan, Y. Yang, In situ phase transformation of Laves phase from Chi-phase in Mo- [42] T. Fujii, K. Tohgo, Y. Mori, Y. Shimamura, Crystallography of intergranular corro-
containing Fe–Cr–Ni alloys, Mater. Lett. 158 (2015) 233–236. sion in sensitized austenitic stainless steel, Mater. Charact. 144 (2018) 219–226.
[36] D.E. Villanueva, F. Junior, R. Plaut, A. Padilha, Comparative study on sigma phase [43] R. Jones, V. Randle, G. Owen, Carbide precipitation and grain boundary plane se-
precipitation of three types of stainless steels: austenitic, superferritic and duplex, lection in overaged type 316 austenitic stainless steel, Mater. Sci. Eng. A 496 (1–2)
Mater. Sci. Technol. 22 (9) (2006) 1098–1104. (2008) 256–261.
[37] A. Aghajani, C. Somsen, G. Eggeler, On the effect of long-term creep on the mi- [44] E. Trillo, L. Murr, Effects of carbon content, deformation, and interfacial energetics
crostructure of a 12% chromium tempered martensite ferritic steel, Acta Mater. 57 on carbide precipitation and corrosion sensitization in 304 stainless steel, Acta
(17) (2009) 5093–5106. Mater. 47 (1) (1998) 235–245.
[38] A. Aghajani, F. Richter, C. Somsen, S. Fries, I. Steinbach, G. Eggeler, On the for- [45] C. Hu, S. Xia, H. Li, T. Liu, B. Zhou, W. Chen, N. Wang, Improving the intergranular
mation and growth of Mo-rich Laves phase particles during long-term creep of a corrosion resistance of 304 stainless steel by grain boundary network control,
12% chromium tempered martensite ferritic steel, Scr. Mater. 61 (11) (2009) Corros. Sci. 53 (5) (2011) 1880–1886.
1068–1071. [46] R. Jones, V. Randle, Sensitisation behaviour of grain boundary engineered auste-
[39] M.I. Isik, A. Kostka, V.A. Yardley, K.G. Pradeep, M.J. Duarte, P.-P. Choi, D. Raabe, nitic stainless steel, Mater. Sci. Eng. A 527 (16–17) (2010) 4275–4280.
G. Eggeler, The nucleation of Mo-rich Laves phase particles adjacent to M23C6 [47] R. Singh, S.G. Chowdhury, B.R. Kumar, S.K. Das, P. De, I. Chattoraj, The importance
micrograin boundary carbides in 12% Cr tempered martensite ferritic steels, Acta of grain size relative to grain boundary character on the sensitization of metastable
Mater. 90 (2015) 94–104. austenitic stainless steel, Scr. Mater. 57 (3) (2007) 185–188.
[40] M. Isik, A. Kostka, G. Eggeler, On the nucleation of Laves phase particles during [48] M. Michiuchi, H. Kokawa, Z. Wang, Y. Sato, K. Sakai, Twin-induced grain boundary
high-temperature exposure and creep of tempered martensite ferritic steels, Acta engineering for 316 austenitic stainless steel, Acta Mater. 54 (19) (2006)
Mater. 81 (2014) 230–240. 5179–5184.

You might also like