Tensile Properties of 316L Stainless Steels Made by Laser Powder-Bed-Fusion Additive Manufacturing

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

LLNL-JRNL-706341

Tensile properties of 316L stainless


steels made by laser powder-bed-fusion
additive manufacturing

Z. Li, T. Voisin, J. T. McKeown, J. Ye, T. T. Roehling, T.


Braun, A. V. Hamza, Y. M. Wang

October 21, 2016

Acta materialia
Disclaimer

This document was prepared as an account of work sponsored by an agency of the United States
government. Neither the United States government nor Lawrence Livermore National Security, LLC,
nor any of their employees makes any warranty, expressed or implied, or assumes any legal liability or
responsibility for the accuracy, completeness, or usefulness of any information, apparatus, product, or
process disclosed, or represents that its use would not infringe privately owned rights. Reference herein
to any specific commercial product, process, or service by trade name, trademark, manufacturer, or
otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring by the
United States government or Lawrence Livermore National Security, LLC. The views and opinions of
authors expressed herein do not necessarily state or reflect those of the United States government or
Lawrence Livermore National Security, LLC, and shall not be used for advertising or product
endorsement purposes.
LLNL-JRNL-706341

Tensile properties, strain rate sensitivity, and activation volume of additively


manufactured 316L stainless steels
Zan Li1§, Thomas Voisin1§, Joseph T. McKeown1, Jianchao Ye1, Tom Braun1, Chandrika
Kamath2, Wayne E. King1, Y. Morris Wang1*
1
Physical and Life Sciences Directorate, 2Computation Directorate, Lawrence Livermore
National Laboratory, Livermore, CA 94550, USA

Abstract

The tensile properties of additively manufactured (AM) metals and alloys are among the most
important variables that impact the potential applications of these materials. Here we examine and
report on the tensile properties of AM 316L stainless steels fabricated by the laser powder-bed-
fusion (L-PBF) technique, via twelve sets of optimized laser processing parameters that produce
materials with density >98.8±0.10%. A heterogeneous microstructure is observed in all L-PBF
samples, including microscopic features such as dislocations, cellular walls, elemental
segregations, local misorientations, impurities, precipitates, and a large fraction of low-angle grain
boundaries (2-10o, ~40-60%). The derived average grain size defined by high-angle grain
boundaries (>10o) is ~30-50 µm. Tensile testing reveals a yield strength ranging from 552-635
MPa and a tensile-elongation-to-failure (TEF) of 0.09-0.42 for directly-printed samples, whereas
these values are 592-690 MPa and 0.29-0.50 for samples machined from the as-built rectangular
thin plates. In all samples, we observe a variation of tensile yield strength within ~15% but not the
TEF, suggesting marginal microstructural changes despite a wide range of laser processing
parameters. The large scatter of TEF in directly-printed samples originates from the sensitivity of
thin gauge geometry (~2 mm2 cross-section area) to the built-in flaws. We measured a substantially
higher strain rate sensitivity (~0.02-0.03) of L-PBF 316L compared to the coarse-grained
counterparts (~0.006), together with a small activation volume of ~20-30b3 (where b is the Burgers
vector of 316L). These deformation kinetics parameters suggest that the tensile plasticity of L-
PBF 316L is controlled by a much finer microstructural length scale than the measured grain size,
consistent with the high strength and juxtaposed nano- to macro-structures seen in these materials.
Strategies to optimize the tensile properties of AM materials are discussed.

Keywords: Additive manufacturing; 316L stainless steel; tensile properties; strain rate
sensitivity; activation volume
§
equal contribution to this work.

*Email: ymwang@llnl.gov.

1
LLNL-JRNL-706341

1. Introduction

Metals and alloys made by additive manufacturing (AM) (Das et al., 2016; Herzog et al., 2016;
Lewandowski and Seifi, 2016; Martin et al., 2017; Mower and Long, 2016; Murr et al., 2012;
Tancogne-Dejean et al., 2016; Wang et al., 2018) have attracted increasing interest in recent
scientific research, owing to the unique design freedom offered by AM technologies. AM parts are
directly built from stereolithography files without tooling (Gu et al., 2012b), making it a powerful
approach for technological utilities. Laser powder-bed-fusion (L-PBF), as one type of AM
technology, has gained popularity due to its versatility in fabricating various important engineering
alloys. In addition to common porosity and residual stress issues, it has been generally accepted
that the materials made by L-PBF have unique microstructure and mechanical properties that are
substantially different from those of conventional counterparts (Amato et al., 2012; Attar et al.,
2014; Gu et al., 2012a; Thijs et al., 2013a; Thijs et al., 2013b; Thijs et al., 2010; Vrancken et al.,
2014; Wang et al., 2018). This requires a thorough investigation of the structure-property
relationship of these AM materials.
The low carbon austenitic 316L stainless steel (316L SS, C<0.03 wt.%) is an important
engineering material that is broadly used in the oil and nuclear industries as well as the biomedical
field due to its high corrosion and oxidation resistance. Because of its broad applications and
excellent weldability, it has become a material-of-choice for many laser-based AM processing
studies (Fernandes de Lima and Sankare, 2014; Kamath et al., 2014; Saeidi et al., 2015b; Tolosa
et al., 2010; Wu et al., 2014; Zhang et al., 2014). As the L-PBF process involves layer-by-layer
melting and solidification of powder precursors, several characteristic microstructural features
inevitably exist in AM materials, including pores, columnar/elongated grains, layered structures,
and impurities (e.g., oxygen, nitrogen, and hydrogen). These complex microstructures are
influenced by a number of processing parameters such as laser beam size, laser power, scan speed,
scan strategy, hatch spacing (i.e., distance between two scan vectors), powder layer thickness, and
powder bed temperature, among which laser power, scan speed, and hatch spacing are considered
as important variables as these parameters define the input energy density to the fabricated
materials. Research to date has focused on producing near fully dense AM parts and materials, as
porosity is a pressing issue that plagues AM technologies. The existence of voids may come from
several sources, including balling or keyhole phenomena in L-PBF processes (King et al., 2014;
Zhong et al., 2016), lack of fusion defects, entrapment of reactive gas species from residual oxygen

2
LLNL-JRNL-706341

or moisture (Cherry et al., 2015), transfer of powder pore sources to the build parts (Cunningham
et al., 2017), voids between powder particles trapped in the melt pool (Voisin et al., 2018), and/or
the space left upon grain coalescence and growth during rapid solidification (McKeown et al.,
2014; Wang et al., 2012). Although the complete mechanisms of void formation remain to be
investigated and is difficult to prevent, voids could substantially deteriorate the mechanical
properties even when their volume fraction is very low (Sanders et al., 1997; Voisin et al., 2018;
Wang et al., 2012). In addition, other structural defects in AM materials such as weak layer welding
could undermine the tensile properties. Despite above issues, we recently reported (as well as a
few other groups) that high strength and high ductility are achievable in L-PBF 316L stainless
steels (Liu et al., 2018; Shamsujjoha et al., 2018; Sun et al., 2018; Wang et al., 2018). These
promising results strongly suggest that AM is a viable technique to produce materials with
excellent mechanical properties. However, there remain many scientific questions that have not
been well understood. For example, the deformation kinetics of these materials has not been
studied, which may shed light on the strain rate dependent deformation behavior of AM materials.
Due to the flaw-prone nature and anisotropic properties of many AM metals, uniaxial tensile
testing is considered as one of the most valuable tools to evaluate their mechanical properties.
Unlike other techniques such as uniaxial compression and nanoindentation, tensile testing is
especially sensitive to the trapped pores or weak welding in AM materials. In addition, the tensile
properties are strongly influenced by both intrinsic microstructures (e.g., grain size, impurities,
texture, residual stresses, and voids) and extrinsic factors (e.g., sample geometry, gauge length,
gauge surface roughness, and even sample cutting method) (Wang et al., 2012; Zhao et al., 2008),
making it challenging to directly compare the tensile properties of AM 316L SS in the literature.
Depending upon laser processing parameters, build orientation, and sample geometry, a wide range
of tensile yield strengths (~265-690 MPa) and total tensile-elongation-to-failure (TEF, ~0.01-0.94)
have been reported for 316L SS fabricated by L-PBF (Ahmadi et al., 2016; Benjamin and
Kirkpatrick, 1980; Carlton et al., 2016; Casati et al., 2016; Liu et al., 2018; Mertens et al., 2014;
Niendorf et al., 2013; Röttger et al., 2016; Shamsujjoha et al., 2018; Spierings et al., 2011; Sun et
al., 2018; Suryawanshi et al., 2017; Tolosa et al., 2010; Wang et al., 2018; Yadollahi et al., 2015;
Yadroitsev et al., 2009; Zhang et al., 2013; Zhang et al., 2014). This large variation of tensile
properties suggests that a systematic study of various influencing factors is needed. According to
the Hart’s criterion (Hart, 1967) for stable deformation of a strain rate sensitive material,

3
LLNL-JRNL-706341

1 𝜕𝜎
( 𝜕𝜀 ) + 𝑚 − 1 ≥ 0, (1)
𝜎 𝜀̇

where σ is the true flow stress, 𝜀 the true strain, 𝜀̇ the deformation strain rate, and m the strain rate
sensitivity. The strain rate sensitivity (m) and strain hardening behavior of materials (i.e., the first
term of eqn. (1)) are two intrinsic factors that control the tensile properties. However, these
parameters have not been studied for L-PBF 316L SS.
In this work, we systematically investigated the tensile properties of near fully dense 316L SS
made by L-PBF from a Concept M2 L-PBF machine. The materials were made using a set of laser
parameters with the laser power from 150-350 W and input volumetric energy densities from 53-
111 J/mm3, with a fixed build layer thickness of 30 µm. These parameters were selected such that
near fully dense samples could be fabricated. To minimize the sample porosity effect, all samples
tested in our work have densities near or better than ~99%. We chose the broadest possible range
of laser parameters to explore the microstructural effects (i.e., intrinsic factor). The effect of
extrinsic factors such as gauge length and build geometry was also investigated. It is noteworthy
that our samples are built in thin geometry (1.0-2.0 mm thick), the purpose of which is to test the
ultimate resolution and capability of L-PBF technique. For the first time, we investigate the strain
rate sensitivity and activation volume of these materials through strain rate jump test and
deformation at different strain rates. Overall, our results indicate that small and thin structures are
sensitive to built-in flaws that have strong influence on tensile ductility, albeit that the measured
density of these samples is better than 99%. In contrast, a small variation of tensile yield strength
(<15%) is observed despite a wide combination of laser power and speed being used. Strain
hardening, strain rate sensitivity, and strain rate-dependent tensile behavior suggest that the tensile
plasticity of L-PBF 316L SS is controlled by very fine microstructural length scales, consistent
with the deformation mechanisms and fracture behavior of these materials revealed by
transmission and scanning electron microscopy.

2. Experimental

2.1 Tensile sample fabrication and density measurements

The 316L SS AM tensile samples used in this study were fabricated by a commercial Concept
Laser M2 PBF machine equipped with a 400 W fiber laser in argon environment. To map out the
effects of laser processing parameters on tensile properties, twelve sets of optimized laser

4
LLNL-JRNL-706341

processing parameters were selected to build different samples, as shown in Table 1. The oxygen
content was maintained below ~0.1-0.2% during the AM process, and the build plate was 304
stainless steel. A plasma atomized 316L powder was used (Additive Metal Alloys, Ltd., USA),
which has particle sizes ranging from ~20-80 μm and a mean value of ~30 μm. All samples used
in this work were built via a continuous scanning strategy with a hatch angle of 45 degrees and a
layer thickness of 30 µm. Alternate layers were rotated by 90o. The normalized energy density (E*)
was calculated according to (Thomas et al., 2016; Wang et al., 2018)

𝐴𝑝
𝐸 ∗ = 2𝑣𝑙𝑟 , (2)
𝑏 𝜌𝐶𝑝 (𝑇𝑚 −𝑇0 )

where E* is the dimensionless normalized volumetric energy density required to melt the material,
𝐴 the surface absorptivity (~0.35) (Khairallah et al., 2016), 𝑝 the laser power (W), 𝑣 the laser
speed (m/s), 𝑙 the layer thickness (m), 𝑟𝑏 the beam radius (m), ρ the density of 316L SS, 𝐶𝑝 the
specific heat capacity (Jkg-1K-1), 𝑇𝑚 the melting temperature (1673K), and 𝑇0 the initial (or
powder bed) temperature. We choose E* over volumetric energy density as it is normalized by the
thermophysical properties of as-built materials and thus has certain advantages.
The build direction and geometry are known to affect the tensile properties of AM materials.
To this end, all tensile samples were built with the tensile axis vertical to the build direction (i.e.,
horizontally built, see Fig. 1a). In terms of geometry, we have built two types of tensile samples.
The first type is directly printed tensile samples with supporting structures to the gauge section
(hereinafter referred to as “directly-printed” samples). The second type are machined out from the
as-built rectangular plates (hereinafter referred to as “plate” samples). Two different gauge lengths
were fabricated for the directly-printed samples: 6.5 mm and 3.5 mm, with the nominal gauge
cross-sectional area of 2 mm (width) × 1.5 mm (thickness). The 6.5 mm gauge length sample is
designed in accordance with a subsized specimen specification following the ASTM E8/E8M-16a
standards ASTM Standard E8/E8M-16a: Standard Test Methods for Tension Test of Metallic
Materials, ASTM International, West Conshohocken, PA, 2016. As shown in Fig. 1a, in one single
build, twenty-four thin rectangular plates with a dimension of 40 mm (length) × 20 mm (width) ×
2 mm (thickness) and forty-eight directly-printed samples were built using twelve sets of laser
parameters. This allows for eight comparison samples for each set of laser parameters, including
four direct-printed and four plate samples. Two batches of samples were built. The plate tensile
samples were machined out from these plates with the gauge geometry of 6.5 mm (length) × 2 mm
5
LLNL-JRNL-706341

(width) × 2 mm (thickness). Note that the final thickness of both plate and directly-printed samples
is ~1.0 mm after polishing.
Density measurements were performed using the Archimedes’ principle with FluorinertTM
FC-40 (Sigma-Aldrich) as the immersion liquid due to its very low intermolecular forces and low
surface tension, which minimizes the effect of meniscus forces on the sample. Each sample was
weighed separately on a XP56 microbalance (±0.001 mg, Mettler Toledo, Switzerland) in air and
in the immersion liquid, which allowed for calculating the density of the samples as
(𝑚air ×𝜌liq )−(𝑚liquid ×𝜌air )
𝜌s = , (3)
𝑚air −𝑚liquid

where 𝜌s is the density of the sample, 𝑚air the weight of the sample in air, 𝑚liquid the weight of the
sample in the immersion liquid, ρair the density of air, and ρliq the density of the immersion liquid.
The density of the immersion liquid was determined using a density standard and confirmed with
a piece of diamond. All density measurements were repeated three times to obtain standard
measurement error bars. The theoretical density of 316L is taken as 7.98 g/cm3 (Kamath et al.,
2014).

2.2 X-ray diffraction (XRD)

X-ray diffraction (XRD) analysis was carried out in an x-ray diffractometer (D8 Advance,
Bruker). The XRD patterns were acquired using Cu Kα radiation at 40 kV and 40 mA and were
recorded in the range of 2θ = 35o-85o with a scan rate of 0.5 o/min. A total of seven directly-printed
and plate samples were characterized via XRD. Three tested samples were also examined by XRD
to identify the possible martensitic phase after deformation.

2.3 Electron backscatter diffraction (EBSD) imaging

EBSD images were acquired using an FEI Company (Hillsboro, OR) Quanta 200 scanning
electron microscope (SEM) equipped with a four-quadrant solid-state backscatter electron
detector. A TexSEM Laboratories (Provo, UT) OIM 4 system equipped with a Peltier cooled CCD
camera was used. Specimens were tilted to 70o in the SEM chamber, and an accelerating voltage
of 30 keV was used. A typical scan area of 1-4 mm2 was used for EBSD analysis with a step size
of 0.1-1 μm. The reported average grain size was obtained by using the planimetric method,
following ASTM standard E112-12.

6
LLNL-JRNL-706341

2.4 Transmission electron microscopy (TEM)


An FEI 80-300 Titan (Oregon, USA) TEM was used to characterize the microstructure of 316L
SS samples. It has a point-to-point resolution of ≤0.20 nm in high-resolution TEM mode and a
maximum resolution of 0.136 nm in high-angle annular dark-field (HAADF) in scanning TEM
mode. The TEM is also equipped with an EDAX system for energy-dispersive spectrometry (EDS)
and a Gatan Quantum 965 GIF for electron energy-loss spectrometry (EELS). A series of TEM
samples were prepared by electropolishing method at 5 oC with a solution of 20% perchloric and
80% acetic acid.

2.5 Focused-ion-beam (FIB) characterization

The cross-sections of as-built and deformed tensile samples were characterized via a FIB
method (FEI Nova 600 Dual-Beam FIB, Oregon, USA). Ion channeling images were acquired by
using a small ion current of 93 pA. Location specific TEM samples were prepared by a FIB liftout
technique with platinum as the surface protection layer. For deformed tensile samples, both
electropolishing and FIB liftout were used to prepare TEM samples.

2.6 Tensile tests

The tensile samples were polished by using sandpaper down to the metallurgical grit of 1200.
The side surfaces of gauge section were also hand-polished by 1200 grit SiC papers. Uniaxial
tensile tests were carried out in an Instron (Norwood, MA, USA) 4444 tensile machine at the strain
rate range of 5 ×10-5 s-1 to 1 ×10-1 s-1. Two load cells with a maximum capacity of 2.5 kN and 2.0
kN, respectively, were used. The tensile elongation was measured by an LE-01 laser extensometer
(Electronic Instrument Research, Irwin, PA, USA) with a displacement resolution of 1 μm. Two
silver tapes were attached to the sample gauge and act as reflective markers for the laser
extensometer. Different strain rate tests (strain rates: 5 ×10-5 s-1, 1 ×10-2 s-1 and 1 ×10-1 s-1) and
strain rate jump tests (strain rate range of 1 ×10-4 s-1 to 2 ×10-2 s-1) were conducted in order to
calculate the strain rate sensitivity (m) and activation volume (V) of tensile samples by using the
standard formula

∂lnσ
𝑚 = (∂ln ε̇ )T,ε (4)

and

7
LLNL-JRNL-706341

√3𝑘𝑇
𝑉= . (5)
𝑚𝜎

Here the definitions of σ, ε̇ , 𝜀 are the same as eqn. (1), T the deformation temperature, and k the
Boltzmann constant. At least two tensile tests were performed for each type of sample. Note that
a value of 3.06 (instead of √3 that is based on von Mises yield criterion) has been used in the
literature when assuming Taylor yield criterion (Blum, 2018; Conrad, 2003). The choice of pre-
factor however does not affect the overall conclusion of this work.

3 Results

3.1 Density, composition, phase, and microstructure of L-PBF 316L SS

As shown in Table 1, all the tensile samples fabricated in this work have densities greater than
98.8 ± 0.10%. The compositions of the tensile samples have been reported in previous work (Wang
et al., 2018). Compared to the standard AISI 316L SS, L-PBF 316L SS contains a trace amount of
oxygen and nitrogen. In addition, AM samples contain lower carbon and higher silicon contents
compared to those in the AISI standard. The as-fabricated AM 316L SS has mostly austenitic
phase, as indicated by XRD data shown in Fig. 1b. This is consistent with the literature using the
same L-PBF technique (Saeidi et al., 2015b). Note that a trace amount of martensitic phase was
reported in some L-PBF 316L SS (Tucho et al., 2018) but not detected in our conventional and
synchrotron XRD (Wang et al., 2018). XRD further reveals a shift of peak positions towards higher
values (compared to zero-stress peak positions) in our samples, suggesting that as-built samples
contain residual stresses.
To investigate the grain structures in as-built materials, we performed EBSD studies on several
plate and directly-printed samples. Fig. 2a shows a representative EBSD inverse pole figure (IPF)
of a directly-printed tensile sample. Grains appear rather irregular in shape. Inside a single grain,
however, we observe continuous evolution of misorientations, as indicated by the change of color
and the presence of numerous low angle grain boundaries (Fig. 2b). In fact, statistical results of
GB distribution in Fig. 2c reveal a large volume fraction of low-angle GBs (LAGBs, 2-10o, ~57%).
Our EBSD studies collectively indicate that the microstructure of as-built 316L SS is dominated
by irregular grain shapes, a large fraction of LAGBs, and local misorientations with individual
grains. The average grain size based upon HAGBs from EBSD is large for all samples we

8
LLNL-JRNL-706341

investigated (~30-50 µm). At the TEM scale, we found that it is difficult to discern difference
between sample to sample built by different processing conditions.

3.2 Tensile properties

3.2.1 Build geometry effect: directly-printed versus plate samples

Representative engineering stress-strain curves of samples cut from plates and directly-printed
samples are shown in Fig. 3a. In general, we observe a higher yield strength and higher TEF in
plate samples than those of directly-printed samples (see more details in the discussion). This is at
least true for the horizontal build direction reported in this work.

3.2.2 Gauge length effect (3.5 mm versus 6.5 mm)

As the measured TEF of materials depends upon gauge dimensions (especially gauge length),
in Fig. 3b we compare the TEF values of a series of tensile samples with two different gauge
lengths (i.e., 3.5 mm and 6.5 mm). Each sample was fabricated from one set of laser parameters
listed in Table 1. For directly-printed samples we observe a relatively narrow range of TEF values
(~0.30-0.35) when the gauge length is 3.5 mm, whereas a large scatter in data is observed for the
6.5 mm gauge length. This scatter in 6.5 mm gauge samples suggests that directly-printed samples
may contain more built-in flaws, making them more susceptible to flaw-induced fracture. Such
failure can be stochastic and does not represent the intrinsic ductility of flaw-free materials.
Interestingly, the gauge length seems to have little influence on the yield strength of AM samples,
as illustrated in Fig. 3c. The relatively small variation of tensile yield strengths in both gauge
lengths of directly-printed samples suggests negligible microstructure changes despite the wide
range of laser processing parameters investigated (see Table 1). At the same gauge length of 6.5
mm, we noticed that plate samples have slightly higher yield strength than those of directly-printed
samples.

3.2.3 Strain rate effects

The strain rate effect can shed light on the rate-controlling deformation mechanisms of AM
materials. We thus performed tensile tests at different strain rates, as well as strain rate jump tests
for several selected samples. Fig. 4 compares engineering stress-strain curves of a directly-printed
sample tested at three different strain rates. The yield strength increases with the increased strain

9
LLNL-JRNL-706341

rate. However, the TEF trend is less obvious and less reliable due to the existence of sample
defects. The estimated strain rate sensitivity (m) value is 0.0248 based on the yield strength at
different strain rates (inset of Fig. 4), which corresponds to an activation volume of 22b3 (where
b=0.258 nm is the Burgers vector of 316L SS).
The fracture surfaces of tensile samples tested at different strain rates are shown in Fig. 5a-c.
The average dimple size can be observed to decrease with the increasing strain rate, from 423.4 ±
177.3 nm at the strain rate of 5×10-5 s-1 to 226.3 ± 88.7 nm at 1×10-1 s-1. Note that the dimple size
measured from fracture surface is two orders of magnitude smaller than the average grain size
observed in our EBSD measurements. This suggests that the deformation and fracture behavior of
AM 316L SS is not controlled by the grain size. Nanosized inclusions, some of which are
highlighted by white circles in Fig. 5, are also observed on the fracture surface. At the slow strain
rate of 5×10-5 s-1 (Fig. 5a), the particles seem to preferentially locate in the center of the dimples,
suggesting that these particles may act as a source of stress concentration. Trapped large voids are
occasionally visible, one example of which is shown in the inset of Fig.5c (~8 µm in diameter).

3.3 Strain rate sensitivity and activation volume

Independently, the strain rate sensitivity (m) of a second directly-printed sample is further
measured by the strain rate jump test, Fig. 6. From the plot, we derive an average m of 0.022±0.002,
a value that is similar to the measurement above and is comparable to those of nanocrystalline face
centered cubic metals (e.g., nanocrystalline copper and nickel) (Wang et al., 2006). The measured
m is higher than that of coarse-grained 316L SS. For example, Khodabakhshia et al. reported an m
value of 0.0061 for a coarse-grained 316L SS (Khodabakhshia et al., 2019). The corresponding V
value is determined to be ~28±3b3 – a value that is again similar to those measured in
nanocrystalline metals and alloys (Rupert et al., 2011; Wang et al., 2006). A series of additional
nanoindentation jump tests have also been performed for other samples, which measured an m in
the range of 0.02-0.03, consistent with our tensile tests at different strain rates and jump tests. The
rather small activation volume observed in our materials suggests that there exist high-density
dislocation barriers for plastic flow during the deformation (Wang et al., 2006). This length scale
is substantially smaller than the average grain size and cellular sizes (~0.6 µm) reported in our
previous work (Wang et al., 2018). These obstacles include HAGBs, LAGBs, pre-existing
dislocations, nanoprecipitates, cellular walls, fusion boundaries, and local misorientations.

10
LLNL-JRNL-706341

3.4 Post mortem microstructure and surface morphology after fracture

316L SS is a low stacking fault energy material. In conventional 316L SS, twinning and
martensitic phase transformations have been commonly observed and contribute to the strain
hardening behavior. Our post mortem TEM studies in tensile samples deformed to low strain (true
strain ~0.06) revealed that the dominant deformation mechanism in AM 316L SS is dislocation
slip (see Fig. 7b-e). We observe that the deformed microstructure in AM 316L SS is highly
nonuniform, with both equiaxed and elongated cells. Some cellular walls are aligned crudely with
the tensile axis and decorated by a high density of dislocations. Some twinning activities were also
observed in a FIB ion-channeling image shown in Fig. 8a. As the strain increases (~0.23), however,
twinning becomes an important mechanism, Fig. 8b. Numerous deformation twins are observed.
Martensitic phase was not detected by our post mortem XRD studies. The results of FIB studies
are consistent with our previous TEM studies (Wang et al., 2018), where deformation twinning is
identified an important deformation mechanisms at large strains. A close examination of sample
surfaces after fracture reveals microcracks in some directly-printed samples (Fig. 8c), whereas the
plate samples are largely devoid of such defects (Fig. 8d).

4 Discussion

4.1 Factors influencing tensile strength and ductility

4.1.1 Sample density and impurities

Although 316L SS has been the subject of many AM processing studies, the tensile properties
can be undermined by the sample density issue. It has been argued that a sample density of 99%
or better is required in order to achieve good mechanical properties (Gong et al., 2015). To give a
better overview of the literature, Fig. 9a summarizes the tensile properties of our materials, as well
as those reported in the literature (Ahmadi et al., 2016; Benjamin and Kirkpatrick, 1980; Carlton
et al., 2016; Casati et al., 2016; Lavery et al., 2017; Liu et al., 2018; Mertens et al., 2014; Niendorf
et al., 2013; Pham et al., 2017; Röttger et al., 2016; Shamsujjoha et al., 2018; Spierings et al., 2011;
Sun et al., 2018; Suryawanshi et al., 2017; Tolosa et al., 2010; Wang et al., 2018; Yadollahi et al.,
2015; Yadroitsev et al., 2009; Zhang et al., 2013; Zhang et al., 2014; Zheng et al., 2017). To make
a meaningful comparison, here we only include 316L materials made by L-PBF technique. The

11
LLNL-JRNL-706341

large scattering in the literature data in both strength (~265-690 MPa) and TEF (0.01-0.93) is a
strong indication that the tensile properties of many materials could be affected by built-in flaws
(in addition to sample geometry effect). Overall, we observe that our plate samples (dashed blue
circle in the figure) have a higher strength, compared to those reported in the literature. This
suggests some unique benefits of our laser parameter optimization processes (Kamath et al., 2014).
These enhanced mechanical properties can at least partially be attributed to the near fully-dense
samples that we have fabricated. On the other hand, we observe a relatively large variation of TEF
(0.08-0.43) in long gauge (6.5 mm) directly-printed samples. This is a sign that sample flaws
remain an issue that influences the tensile properties (especially the tensile ductility), manifested
by occasionally visible large voids and microcracks (inset of Fig. 5c). The results further suggest
that the tensile ductility of small build structures is especially sensitive to built-in flaws, despite
the measured high density (i.e., better than 99%). This is consistent with some reports in the
literature (Salzbrenner et al., 2017). Our recent computed tomography results have in fact revealed
void coalescence mechanisms that could lead to stochastic failure behavior in AM materials
(Voisin et al., 2018).
To better understand the laser processing parameter effect on tensile ductility, in Fig. 9b we
examine the TEF as a function of normalized energy density (see eqn (2)). Despite the relatively
large scatter in the data, it is interesting that plate samples tend to have higher TEF values than the
directly-printed samples. This seems to be true in the entire range of laser processing parameters,
suggesting again the critical importance of pores/defects in small build structures. Recall that our
directly-printed samples have a small gauge section of 1.0 mm (thickness) × 2.0 mm (width).
Regarding the tensile yield strength (σ0.2%), in Fig. 9c we observe a narrow variation of σ0.2% as
the normalized energy density rises (except for the highest energy). For the directly-printed
samples, the σ0.2% value is between 552-635 MPa (~15% variation), whereas for the plate samples,
the σ0.2% falls between 592-690 MPa (~16% variation). These results suggest that the laser
processing parameters only induce marginal changes (if any) of microstructures in both types of
samples, provided that the laser beam size, build layer thickness and hatch spacing were fixed and
that all the samples have density better than ~99%. Under the same energy input conditions, we
observe that the plate samples exhibit slightly higher yield strengths. We hypothesize that different
residual stresses, microstructures, and/or pores might contribute to these measurable differences.

12
LLNL-JRNL-706341

In particularly, the effect of residual stress on tensile properties is a subject that is poorly
understood in the literature and deserves further investigations.
As reported earlier (Wang et al., 2018), our AM 316L SS contains 0.094±0.002% nitrogen and
0.032±0.002% oxygen (averaged from two AM samples). Nitrogen is a known solid solution
strengthening source for low-carbon steels (Byrnes et al., 1987; Karaman et al., 2001; Tong et al.,
2003). Furthermore, at 0.4 wt.% Karaman et al. (Karaman et al., 2001) reported that nitrogen tends
to suppress twinning activities in single crystal 316L SS. This appears not to be the case for AM
316L SS, where deformation twinning was observed rather earlier. For oxygen contamination, our
FIB and TEM studies in Fig. 7a,d,e suggest that oxygen reacts with silicon in AM 316L SS and
forms transition metal enriched silicate particles. Previously, these particles were reported by
Saeidi et al. as pure silicates (Saeidi et al., 2015a). Our previous analysis suggested the
strengthening effect of these particles is relatively small (Wang et al., 2018). However, the fracture
surface of our samples suggests that these silicate nanoparticles could act as the source of stress
concentration during tension, leading to particle-matrix interface debonding, and thus may have
influenced the ultimate tensile strength of these materials. Note that the as-built materials also
contain tens ppm level hydrogen, the effects of which on tensile properties are largely unknown
and deserves further investigation.
According to eqn. (1), the continuously increased strain hardening at high strains heralds a
large tensile elongation to failure in L-PBF 316L SS. However, this is not always the case for our
316L SS samples (for example, the directly-printed samples), suggesting that the fracture strength
of these materials could ultimately curb their overall tensile ductility. The microcracks observed
in Fig. 8c for directly-printed samples support our previous judgement that the tensile ductility of
small AM structures is often limited by their intrinsic flaws (Voisin et al., 2018).

4.1.2 Influence of GBs, cellular walls, and dislocations


The yield strength (σy) of most materials is known to follow the Hall-Petch relationship:
𝜎𝑦 = 𝜎0 + 𝐾𝑦 𝑑−1/2, (6)
where σ0 is the Peierls stress for dislocation motion, Ky the Hall-Petch slope, and d the average
grain size. By using (Chen et al., 2005)
𝜎𝑦 = 183.31 + 253.66/√𝑑 (MPa) (d in µm), (7)

13
LLNL-JRNL-706341

we estimate that the strength contribution for d~30-50 µm is ~219-230 MPa, which is far less than
the measured σy in our samples. This simple analysis suggests that the yield strength of AM 316L
SS is not controlled by the average grain size defined by HAGBs. Cellular structures, high-density
dislocation, and an excessive amount of LAGBs are expected to have strong influence on the yield
strength of these materials, as also suggested by our own modeling results (Wang et al., 2018).
Previously, the observed intragranular cellular spacing has been proposed to scale with the
strength of AM 316L SS (Zhong et al., 2016). However, others also reported that the size of these
cellular structures is nonuniform, ranging from 0.2-1.0 µm in average spacing in the same sample,
and thus a single size scaling law is difficult to establish (Trelewicz et al., 2016). Furthermore, a
softening effect has sometimes been reported in samples with smaller cellular sizes (Wang et al.,
2016). We also observed a nonuniform distribution of these cellular structures in our tensile
samples, the volume fraction of which seems to vary from sample to sample and may be correlated
with the laser parameters and build geometry. In addition, the cellular size is observed to vary from
region to region. Our statistical measurements based on cross-section etched surface of as-built
samples suggest that the plate samples have appreciably higher volume fraction of cellular
structures than the directly-printed samples. However, a substantially higher strengthening effect
was not observed in our plate samples. Importantly, our EBSD and inverse-pole figure orientation
mapping in TEM (Wang et al., 2018) did not reveal misorientations across most cellular
boundaries, suggesting that these cellular walls should be treated differently from traditional
interfaces such as HAGBs or dislocation walls. Recently, dislocation blockage has been observed
in thick cellular walls of AiSi10Mg alloys (Wu et al., 2016); but thin cellular walls can be
transmitted by dislocations. These results collectively argue that the strengthening efficiency of
these cellular structures is not well understood and remains to be investigated.

4.1.3 Strain hardening, strain rate sensitivity, and ductility

According to Hart’s criterion for strain rate sensitive material (see eqn. (1)) (Hart, 1967), the
uniform elongation can be maintained if a materials have a high m. The measured m (0.02-0.03)
in the AM 316L SS increases compared to that of coarse-grained 316L SS (Sun et al., 2014). This
three- to four-fold hike of m value can in principle help to diffuse and stabilize necking but is too
small to be effective at room temperature except at the very slow strain rates (Wang and Ma, 2004).
The large TEF observed in our AM 316L SS may primarily originate from its strong hardening

14
LLNL-JRNL-706341

capability that is inherently linked to the unique microstructure of AM materials. Assuming a


power law strain hardening behavior, we applied the Ludwig equation
𝜎 = 𝜎𝑦 + 𝐾𝑠 𝜀 𝑛 (8)
to calculate the strain hardening exponent (n) of our materials, where Ks is the strength coefficient.
We found that for all AM samples, a single power-law fit cannot be obtained. Instead, we extract
an n value of 0.06-0.07 for most samples at 0.01-0.02 strain, which increases to 0.16-0.19 at 0.10-
0.13 strain. This large increase in n value as a function of strain agrees with a shift of deformation
mechanisms from dislocation slips to twinning observed in Fig. 8 and our in situ synchrotron X-
ray diffraction experiments (Wang et al., 2018). In addition, we argue that the hierarchical
microstructure and multiple length scale distributions in AM 316L SS may have a strong influence
on hardening behavior. The effectiveness of heterogeneous and hierarchical microstructures in
enhancing the work hardening behavior has been broadly reported in pure metals (Fang et al.,
2011; Wang et al., 2002), steels (Chen et al., 2016; Wei et al., 2014), and aluminum alloys
(Liddicoat et al., 2010).

4.2 Strategies to optimize the tensile properties of AM metals

Our findings above suggest that AM structures with low dimensions (such as thin walls,
narrow/thin bars, and lattice structures) are sensitive to trapped flaws, leading to low tensile
ductility. Yet, the tensile yield strength remains high when the materials are in near full density.
This suggests that further elimination of pores in as-built materials (especially small structures)
remain to be critical to further improve the materials properties. In addition to pore sources from
powder, computer simulations (Khairallah et al., 2016; Voisin et al., 2018) suggest that the pore
formation in L-PBF is a complex event and can be significantly affected by recoil pressure and
Marangoni convection phenomena. Spatter, denudation, and melt pool depression can all generate
voids at the edge of a scan track, at the bottom of melt pool, and at the end of a melt track. For thin
structures, one type of defects/pores is likely linked to void formation at the beginning or end of a
melt track due to the laser power ramp down. To mitigate this type of defect, slow ramp down of
laser power is desirable to allow the melt enough time to smooth out the surface tension. One
possible solution to this problem is thus to use a lower laser power (and/or laser speed) provided
that near fully dense samples can be fabricated. This is consistent with our observation in Fig. 9b,
where higher energy density tends to cause low tensile ductility. Our previous studies (Kamath et

15
LLNL-JRNL-706341

al., 2014) have, however, also indicated that there exists a relatively narrow window at low laser
power in order for samples to reach full density. As such, a tradeoff might be needed among laser
power, laser speed, and build time; for example, by applying medium laser power and speed. The
medium laser power is also likely to help minimize the keyhole-related defects in thin structures
as the laser turns around and the power ramps up.
The intrinsic heterogeneity with multiple length scales (from nano- to micro-meters) in AM
metals is considered beneficial to the tensile ductility due to the enhanced strain hardening
capability of inhomogeneous or gradient microstructures (Castelluccio and McDowell, 2017; Fang
et al., 2011; Liddicoat et al., 2010; Wang et al., 2002; Wei et al., 2014). The rising experimental
evidence has indeed demonstrated that high strength and high ductility are achievable in as-built
316L SS (Liu et al., 2018; Shamsujjoha et al., 2018; Sun et al., 2018; Wang et al., 2018). However,
it is unclear whether these as-built microstructures have reached their full potential in terms of
mechanical performance. Heat treatment has been found to be another effective strategy to mitigate
the residual stresses, strain incompatibility, and elemental segregation, and thus could help to
enhance the tensile ductility. Note that traditional heat treatment methods typically lead to a
tradeoff between strength and ductility. A few exceptions have been reported in nanostructured
materials where low temperature and/or short annealing time could simultaneously enhance the
tensile strength and ductility (Huang et al., 2006; Wang et al., 2002; Wang et al., 2004). This
phenomenon has not been explored in AM metals. A carefully conducted research towards this
direction is desirable for 316L SS presented here and many other AM metals and alloys reported
in the literature.

5 Conclusion

In summary, we have systematically investigated the tensile properties, strain rate sensitivity
and activation volume of near fully dense 316L SS fabricated by L-PBF technique. The main
findings of our work are summarized as follows:
(1) L-PBF 316L SS has complex and heterogeneous microstructures, both structurally and
chemically. Compared to conventional materials, some unique microstructural features in AM
metals include fusion boundaries, cellular walls, precipitates, impurities, elemental
segregations (along both HAGBs and cellular walls), local lattice misorientations, and a large
fraction of LAGBs (as high as ~57% of total boundaries).

16
LLNL-JRNL-706341

(2) The tensile elongation-to-failure (TEF) of AM materials is sensitive to the sample geometry
and built-in flaws. This is especially true for small geometry structures where the fracture
strength can ultimately limit their tensile ductility despite a high intrinsic strain hardening
capability of AM 316L SS. This suggests that sample density is a poor indicator for tensile
properties.
(3) When the samples reach near full density, the tensile yield strength varies within ~15% despite
a broad range of laser speed and power were used. This suggests that other processing variables
such as build layer thickness are important to further alter the microstructure of AM metals.
(4) Nanoinclusions in L-PBF 316L SS appear to have little influence on the yield strength (Wang
et al., 2018) but may have affected the fracture strength and thus tensile ductility of these
materials.
(5) A substantially higher strain rate sensitivity (m~0.02-0.03) is observed in L-PBF 316L SS
compared to the conventional coarse-grained counterparts that have an m value of 0.0061
(Khodabakhshia et al., 2019). A small activation volume (~20-30b3) is also measured in L-
PBF 316L SS. These deformation kinetics parameters suggest that the controlling length scale
(tens of nanometers) for plastic deformation in AM 316L SS is nearly three orders of
magnitude smaller than the measured average grain size.

Acknowledgement

We thank N. Teslich, A.M. Rubenchik, N. Calta, T.T. Roehling, S. Burke, and E. Sedillo for their
helpful discussion and experimental assistance. The work was performed under the auspices of the
US Department of Energy by LLNL under contract No. DE-AC52-07NA27344.

17
LLNL-JRNL-706341

References
Ahmadi, A., Mirzaeifar, R., Moghaddama, N.S., Turabi, A.S., Karaca, H.E., Elahinia, M., 2016. Effect of
manufacturing parameters on mechanical properties of 316L stainless steel parts fabricated by selective
laser melting: A computational framework. Mater. Des. 112, 328-338.
Amato, K.N., Gaytan, S.M., Murr, L.E., Martinez, E., Shindo, P.W., Hernandez, J., Collins, S., Medina, F.,
2012. Microstructures and mechanical behavior of Inconel 718 fabricated by selective laser melting. Acta
Mater. 60, 2229-2239.
Attar, H., Boenisch, M., Calin, M., Zhang, L.-C., Scudino, S., Eckert, J., 2014. Selective laser melting of
in situ titanium-titanium boride composites: Processing, microstructure and mechanical properties. Acta
Mater. 76, 13-22.
Benjamin, D., Kirkpatrick, C.W., 1980. Properties and selection: stainless steels, tool materials and special
purpose metals, 9th ed. American Society for Metals, OH.
Blum, W., 2018. Discussion: Activation volumes of plastic deformation of crystals. Scr. Mater. 146, 27-
30.
Byrnes, M.L.G., Grujicic, M., Owen, W.S., 1987. NITROGEN STRENGTHENING OF A STABLE
AUSTENITIC STAINLESS-STEEL. Acta Metall. 35, 1853-1862.
Carlton, H.D., Haboub, A., Gallegos, G.F., Parkinson, D.Y., MacDowell, A.A., 2016. Damage evolution
and failure mechanisms in additively manufactured stainless steel. Mater. Sci. Eng. A 651, 406-414.
Casati, R., Lemke, J., Vedani, M., 2016. Microstructure and Fracture Behavior of 316L Austenitic Stainless
Steel Produced by Selective Laser Melting. J. Mater. Sci. Technol. 32, 738-744.
Castelluccio, G.M., McDowell, D.L., 2017. Mesoscale cyclic crystal plasticity with dislocation
substructures. Int. J. Plast. 98, 1-26.
Chen, A., Liu, J., Wang, H., Lu, J., Wang, Y.M., 2016. Gradient twinned 304 stainless steels for high
strength and high ductility. Mater. Sci. Eng. A 667, 179-188.
Chen, X.H., Lu, J., Lu, L., Lu, K., 2005. Tensile properties of a nanocrystalline 316L austenitic stainless
steel. Scr. Mater. 52, 1039-1044.
Cherry, J.A., Davies, H.M., Mehmood, S., Lavery, N.P., Brown, S.G.R., Sienz, J., 2015. Investigation into
the effect of process parameters on microstructural and physical properties of 316L stainless steel parts by
selective laser melting. Int. J. Adv. Manuf. Technol. 76, 869-879.
Conrad, H., 2003. Grain size dependence of the plastic deformation kinetics in Cu. Mater. Sci. Eng. A 341,
216-228.
Cunningham, R., Narra, S.P., Montgomery, C., Beuth, J., Rollett, A.D., 2017. Synchrotron-based X-ray
microtomography characterization of the effect of processing variables on porosity formation in laser
powder-bed additive manufacturing of Ti-6Al-4V. JOM 69, 479-484.
Das, S., Bourell, D.L., Babu, S.S., 2016. Metallic materials for 3D printing. MRS Bull. 41, 729-741.
Fang, T.H., Li, W.L., Tao, N.R., Lu, K., 2011. Revealing Extraordinary Intrinsic Tensile Plasticity in
Gradient Nano-Grained Copper. Science 331, 1587-1590.
Fernandes de Lima, M.S., Sankare, S., 2014. Microstructure and mechanical behavior of laser additive
manufactured AISI 316 stainless steel stringers. Mater. Des. 55, 526-532.
Gong, H., Rafi, K., Gu, H., Janaki Ram, G.D., Starr, T., Stucker, B., 2015. Influence of defects on
mechanical properties of Ti–6Al–4 V components produced by selective laser melting and electron beam
melting. Mater. Des. 86, 545-554.
Gu, D., Hagedorn, Y.-C., Meiners, W., Meng, G., Batista, R.J.S., Wissenbach, K., Poprawe, R., 2012a.
Densification behavior, microstructure evolution, and wear performance of selective laser melting
processed commercially pure titanium. Acta Mater. 60, 3849-3860.
Gu, D.D., Meiners, W., Wissenbach, K., Poprawe, R., 2012b. Laser additive manufacturing of metallic
components: materials, processes and mechanisms. Int. Mater. Rev. 57, 133-164.
Hart, E.W., 1967. Theory of tensile test. Acta Metall. 15, 351-355.
Herzog, D., Seyda, V., Wycisk, E., Emmelmann, C., 2016. Additive manufacturing of metals. Acta Mater.
117, 371-392.

18
LLNL-JRNL-706341

Huang, X.X., Hansen, N., Tsuji, N., 2006. Hardening by annealing and softening by deformation in
nanostructured metals. Science 312, 249-251.
Kamath, C., El-dasher, B., Gallegos, G.F., King, W.E., Sisto, A., 2014. Density of additively-manufactured,
316L SS parts using laser powder-bed fusion at powers up to 400 W. Int. J. Adv. Manuf. Technol. 74, 65-
78.
Karaman, I., Sehitoglu, H., Maier, H.J., Chumlyakov, Y.I., 2001. Competing mechanisms and modeling of
deformation in austenitic stainless steel single crystals with and without nitrogen. Acta Mater. 49, 3919-
3933.
Khairallah, S.A., Anderson, A.T., Rubenchik, A., King, W.E., 2016. Laser powder-bed fusion additive
manufacturing: Physics of complex melt flow and formation mechanisms of pores, spatter, and denudation
zones. Acta Mater. 108, 36-45.
Khodabakhshia, F., Farshidianfarb, M.H., Gerlichc, A.P., Noskod, M., Trembošovád, V., Khajepour, A.,
2019. Microstructure, strain-rate sensitivity, work hardening, and fracture behavior of laser additive
manufactured austenitic and martensitic stainless steel structures. Mater. Sci. Eng. A 756, 745-561.
King, W.E., Barth, H.D., Castillo, V.M., Gallegos, G.F., Gibbs, J.W., Hahn, D.E., Kamath, C., Rubenchik,
A.M., 2014. Observation of keyhole-mode laser melting in laser powder-bed fusion additive manufacturing.
J. Mater. Process. Technol. 214, 2915-2925.
Lavery, N.P., Cherry, J., Mehmood, S., Davies, H., Girling, B., Sackett, E., Brown, S.G.R., Sienz, J., 2017.
Effects of hot isostatic pressing on the elastic modulus and tensile properties of 316L parts made by powder
bed laser fusion. Mater. Sci. Eng. A 693, 186-213.
Lewandowski, J.J., Seifi, M., 2016. Metal additive manufacturing: a review of mechanical properties.
Annual review of materials research 46, 151-186.
Liddicoat, P.V., Liao, X.Z., Zhao, Y.H., Zhu, Y.T., Murashkin, M.Y., Lavernia, E.J., Valiev, R.Z., Ringer,
S.P., 2010. Nanostructural hierarchy increases the strength of aluminium alloys. Nat. Commun. 1, 63.
Liu, L., Ding, Q., Zhong, Y., Zou, J., Wu, J., Chiu, Y.-L., Li, J., Zhang, Z., Yu, Q., Shen, Z., 2018.
Dislocation network in additive manufactured steel breaks strength-ductility trade-off. Mater. Today 21,
354-361.
Martin, J.H., Yahata, B.D., Hundley, J.M., Mayer, J.A., Schaedler, T.A., Pollock, T.M., 2017. 3D printing
of high-strength aluminium alloys. Nature 549, 365-369.
McKeown, J.T., Kulovits, A.K., Liu, C., Zweiacker, K., Reed, B.W., LaGrange, T., Wiezorek, J.M.K.,
Campbell, G.H., 2014. In situ transmission electron microscopy of crystal growth-mode transitions during
rapid solidification of a hypoeutectic Al-Cu alloy. Acta Mater. 65, 56-68.
Mertens, A., Reginster, S., Paydas, H., Contrepois, Q., Dormal, T., Lemaire, O., Lecomte-Beckers, J., 2014.
Mechanical properties of alloy Ti–6Al–4V and of stainless steel 316L processed by selective laser melting:
influence of out-of-equilibrium microstructures. Powder Metall. 57, 184-189.
Mower, T.M., Long, M.J., 2016. Mechanical behavior of additive manufactured, powder-bed laser-fused
materials. Mater. Sci. Eng. A 651, 198-213.
Murr, L.E., Gaytan, S.M., Ramirez, D.A., Martinez, E., Hernandez, J., Amato, K.N., Shindo, P.W., Medina,
F.R., Wicker, R.B., 2012. Metal Fabrication by Additive Manufacturing Using Laser and Electron Beam
Melting Technologies. J. Mater. Sci. Technol. 28, 1-14.
Niendorf, T., Leuders, S., Riemer, A., Richard, H.A., Tröster, T., Schwarze, D., 2013. Highly anisotropic
steel processed by selective laser melting. Metall. Mater. Trans. B 44, 794-796.
Pham, M.S., Dovgyy, B., Hooper, P.A., 2017. Twinning induced plasticity in austenitic stainless steel 316L
made by additive manufacturing. Mater. Sci. Eng. A 704, 102-111.
Röttger, A., Geenen, K., Windmann, M., Binner, F., Theisen, W., 2016. Comparison of microstructure and
mechanical properties of 316 L austenitic steel processed by selective laser melting with hot-isostatic
pressed and cast material. Mater. Sci. Eng. A 678, 365-376.
Rupert, T.J., Trenkle, J.C., Schuh, C.A., 2011. Enhanced solid solution effects on the strength of
nanocrystalline alloys. Acta Mater. 59, 1619-1631.

19
LLNL-JRNL-706341

Saeidi, K., Gao, X., Lofaj, F., Kvetkova, L., Shen, Z.J., 2015a. Transformation of austenite to duplex
austenite-ferrite assembly in annealed stainless steel 316L consolidated by laser melting. J. Alloys Compd.
633, 463-469.
Saeidi, K., Gao, X., Zhong, Y., Shen, Z.J., 2015b. Hardened austenite steel with columnar sub-grain
structure formed by laser melting. Mater. Sci. Eng. A 625, 221-229.
Salzbrenner, B.C., Rodelas, J.M., Madison, J.D., Jared, B.H., Swiler, L.P., Shen, Y.L., Boyce, B.L., 2017.
High-throughput stochastic tensile performance of additively manufactured stainless steel. J. Mater.
Process. Technol. 241, 1-12.
Sanders, P.G., Youngdahl, C.J., Weertman, J.R., 1997. The strength of nanocrystalline metals with and
without flaws. Mater. Sci. Eng. A 234, 77-82.
Shamsujjoha, M., Agnew, S.R., Fitz-Gerald, J.M., Moore, W.R., Newman, T.A., 2018. High Strength and
Ductility of Additively Manufactured 316L Stainless Steel Explained. Metall. Mater. Trans. A 49A, 3011-
3027.
Spierings, A., Herres, N., Levy, G., 2011. Influence of the particle size distribution on surface quality and
mechanical properties in AM steel parts. Rapid Prototyping J. 17, 195-202.
Sun, C., Ma, J., Yang, Y., Hartwig, K.T., Maloy, S.A., Wang, H., Zhang, X., 2014. Temperature and grain
size dependent plastic instability and strain rate sensitivity of ultrafine grained austenitic Fe-14Cr-16Ni
alloy. Mater. Sci. Eng. A 597, 415-421.
Sun, Z., Tan, X., Tor, S.B., Chua, C.K., 2018. Simultaneously enhanced strength and ductility for 3D-
printed stainless steel 316L by selective laser melting. Npg Asia Mater. 10, 127-136.
Suryawanshi, J., Prashanth, K.G., Ramamurty, U., 2017. Mechanical behavior of selective laser melted
316L stainless steel. Mater. Sci. Eng. A 696, 113-121.
Tancogne-Dejean, T., Roth, C.C., Woy, U., Mohr, D., 2016. Probabilistic fracture of Ti-6Al-4V made
through additive layer manufacturing. Int. J. Plast. 78, 145-172.
Thijs, L., Kempen, K., Kruth, J.-P., Van Humbeeck, J., 2013a. Fine-structured aluminium products with
controllable texture by selective laser melting of pre-alloyed AlSi10Mg powder. Acta Mater. 61, 1809-
1819.
Thijs, L., Sistiaga, M.L.M., Wauthle, R., Xie, Q., Kruth, J.-P., Van Humbeeck, J., 2013b. Strong
morphological and crystallographic texture and resulting yield strength anisotropy in selective laser melted
tantalum. Acta Mater. 61, 4657-4668.
Thijs, L., Verhaeghe, F., Craeghs, T., Van Humbeeck, J., Kruth, J.P., 2010. A study of the micro structural
evolution during selective laser melting of Ti-6Al-4V. Acta Mater. 58, 3303-3312.
Thomas, M., Baxter, G.J., Todd, I., 2016. Normalised model-based processing diagrams for additive layer
manufacture of engineering alloys. Acta Mater. 108, 26-35.
Tolosa, I., Garciandia, F., Zubiri, F., Zapirain, F., Esnaola, A., 2010. Study of mechanical properties of
AISI 316 stainless steel processed by "selective laser melting", following different manufacturing
strategies. Int. J. Adv. Manuf. Technol. 51, 639-647.
Tong, W.P., Tao, N.R., Wang, Z.B., Lu, J., Lu, K., 2003. Nitriding iron at lower temperatures. Science 299,
686-688.
Trelewicz, J.R., Halada, G.P., Donaldson, O.K., Manogharan, G., 2016. Microstructure and Corrosion
Resistance of Laser Additively Manufactured 316L Stainless Steel. JOM 68, 850-859.
Tucho, W.M., Lysne, V.H., Austbo, H., Sjolyst-Kverneland, A., Hansen, V., 2018. Investigation of effects
of process parameters on microstructure and hardness of SLM manufactured SS316L. J. Alloys Compd.
740, 910-925.
Voisin, T., Calta, N.P., Khairallah, S.A., Forien, J.-B., Balogh, L., Cunningham, R.W., Rollett, A.D., Wang,
Y.M., 2018. Defects-dictated tensile properties of selective laser melted Ti-6Al-4V. Mater. Des. 158, 113-
126.
Vrancken, B., Thijs, L., Kruth, J.P., Van Humbeeck, J., 2014. Microstructure and mechanical properties of
a novel beta titanium metallic composite by selective laser melting. Acta Mater. 68, 150-158.

20
LLNL-JRNL-706341

Wang, D., Song, C.H., Yang, Y.Q., Bai, Y.C., 2016. Investigation of crystal growth mechanism during
selective laser melting and mechanical property charcterization of 316L stainless steel parts. Mater. Des.
100, 291-299.
Wang, Y.M., Chen, M.W., Zhou, F.H., Ma, E., 2002. High tensile ductility in a nanostructured metal.
Nature 419, 912-915.
Wang, Y.M., Cheng, S., Wei, Q.M., Ma, E., Nieh, T.G., Hamza, A., 2004. Effects of annealing and
impurities on tensile properties of electrodeposited nanocrystalline Ni. Scr. Mater. 51, 1023-1028.
Wang, Y.M., Hamza, A.V., Ma, E., 2006. Temperature-dependent strain rate sensitivity and activation
volume of nanocrystalline Ni. Acta Mater. 54, 2715-2726.
Wang, Y.M., Ma, E., 2004. Three strategies to achieve uniform tensile deformation in a nanostructured
metal. Acta Mater. 52, 1699-1709.
Wang, Y.M., Ott, R.T., van Buuren, T., Willey, T.M., Biener, M.M., Hamza, A.V., 2012. Controlling
factors in tensile deformation of nanocrystalline cobalt and nickel. Phys. Rev. B 85, 014101.
Wang, Y.M., Voisin, T., McKeown, J.T., Ye, J., Calta, N.P., Li, Z., Zeng, Z., Zhang, Y., Chen, W.,
Roehling, T.T., Ott, R.T., Santala, M.K., Depond, P.J., Matthews, M.J., Hamza, A.V., Zhu, T., 2018.
Additively manufactured hierarchical stainless steels with high strength and ductility. Nat. Mater. 17, 63-
71.
Wei, Y., Li, Y., Zhu, L., Liu, Y., Lei, X., Wang, G., Wu, Y., Mi, Z., Liu, J., Wang, H., Gao, H., 2014.
Evading the strength- ductility trade-off dilemma in steel through gradient hierarchical nanotwins. Nat.
Commun. 5, 3580.
Wu, A.S., Brown, D.W., Kumar, M., Gallegos, G.F., King, W.E., 2014. An Experimental Investigation into
Additive Manufacturing-Induced Residual Stresses in 316L Stainless Steel. Metall. Mater. Trans. A 45A,
6260-6270.
Wu, J., Wang, X.Q., Wang, W., Attallah, M.M., Loretto, M.H., 2016. Microstructure and strength of
selectively laser melted AlSi10Mg. Acta Mater. 117, 311-320.
Yadollahi, A., Shamsaei, N., Thompson, S.M., Seely, D.W., 2015. Effects of process time interval and heat
treatment on the mechanical and microstructural properties of direct laser deposited 316L stainless steel.
Mater. Sci. Eng. A 644, 171-183.
Yadroitsev, I., Pavlov, M., Bertrand, P., Smurov, I., 2009. Mechanical properties of samples fabricated by
selective laser melting. 14èmes Assises Européennes du Prototypages & Fabrication Rapide, Paris.
Zhang, B., Dembinski, L., Coddet, C., 2013. The study of the laser parameters and environment variables
effect on mechanical properties of high compact parts elaborated by selective laser melting 316L powder.
Mater. Sci. Eng. A 584, 21-31.
Zhang, K., Wang, S., Liu, W., Shang, X., 2014. Characterization of stainless steel parts by Laser Metal
Deposition Shaping. Mater. Des. 55, 104-119.
Zhao, Y.H., Guo, Y.Z., Wei, Q., Dangelewiez, A.M., Zhu, Y.T., Langdon, T.G., Zhou, Y.Z., Lavernia, E.J.,
Xu, C., 2008. Influence of specimen dimensions on the tensile behavior of ultrafine-grained Cu. Scr. Mater.
59, 627-630.
Zheng, Z., Wang, L.F., Yan, B., 2017. Effects of laser power on the microstructure and mechanical
properties of 316L stainless steel prepared by selective laser melting. Int. J. Mod. Phys. B 31, 1744015.
Zhong, Y., Liu, L.F., Wikman, S., Cui, D.Q., Shen, Z.J., 2016. Intragranular cellular segregation network
structure strengthening 316L stainless steel prepared by selective laser melting. J. Nucl. Mater. 470, 170-
178.

21
LLNL-JRNL-706341

Table 1. Laser processing parameters used in this work to fabricate tensile samples. Four directly-
printed and four plate samples were fabricated for each set of laser parameters. The definition of
A1 parameter can be seen in reference (Kamath et al., 2014). The density measurements were
obtained for directly-printed samples only.

Set# Laser Laser speed Trace width Density in


power (mm/s) (μm) percentage
(W) (%)
1 150 700 150 μm, A1=50% 99.5 ± 0.06
2 150 700 150 um, A1=70% 99.4 ± 0.09
3 150 900 150 um, A1=50% 99.4 ± 0.13
4 150 900 150um, A1=70% 99.3 ± 0.09
5 250 1000 150 um, A1=50% 99.4 ± 0.10
6 250 1000 150 um, A1=70% 99.3 ± 0.10
7 250 1300 150 um, A1=50% 99.4 ± 0.06
8 250 1300 150 um, A1=70% 99.4 ± 0.10
9 350 1400 150 um, A1=50% 99.0 ± 0.09
10 350 1400 150 um, A1=70% 99.5 ± 0.09
11 350 1700 150 um, A1=50% 99.2 ± 0.12
12 350 1700 150 um, A1=70% 98.8 ± 0.10

22
LLNL-JRNL-706341

Figure 1. Sample geometry, phase, and microstructure of L-PBF 316L stainless steels (SS). (a) A
photograph of as-built rectangular plates and directly-printed tensile samples. The build orientation
is parallel to the z-axis in the image. (b) X-ray diffraction (XRD) pattern of an as-built 316L SS,
indicative of austenitic phase.

23
LLNL-JRNL-706341

Figure 2. Electron backscatter diffraction (EBSD) of a directly-printed 316L SS tensile sample


built with the laser parameter set#3. The scan step size is 1 µm. High-angle grain boundaries
(HAGBs) with misorientations >10o are colored with dark lines. (a) EBSD inverse pole figure
(IPF). The build direction is labelled with an arrow. (b) The same IPF where only HAGBs and
low-angle grain boundaries (LAGBs, 2-10o, colored in red) are shown. (c) Grain boundary
misorientation distribution extracted from (a) and (b). The fraction of LAGBs is determined to be
57%.

24
LLNL-JRNL-706341

Figure 3. Representative tensile engineering stress-strain curves of directly-printed and plate


samples and the dependence of tensile behavior on gauge length. (a) A comparison of engineering
tensile stress-strain curves for directly-printed (red curve) and plate samples (blue curve). (b) The
measured engineering tensile-elongation-to-failure (TEF), and (c) the yield strength as a function
of gauge length for selected samples that span the entire laser power and speed range. The readers
are referred to the main text for detailed discussion.

25
LLNL-JRNL-706341

Figure 4. Tensile engineering stress-strain curves of a directly-printed samples (laser parameter


Set#6) at the strain rate range of 5×10-5s-1 — 1×10-1s-1. The inset is a ln(σ) – ln(𝜀̇) plot to extract
the strain rate sensitivity (m) of AM 316L, where σ is the yield strength and 𝜀̇ the strain rate. Two
sets of experiments were performed to confirm the strain rate trend.

26
LLNL-JRNL-706341

Figure 5. The fracture surface of an AM 316L SS tested at a strain rate of (a) 5×10-5 s-1, (b) 1×10-
3 -1
s , and (c) 1×10-1 s-1, respectively. Some of the nanoparticles in (a) are accentuated inside white
circles. Similar nanoinclusions can also been seen in (b) and (c). Note in (a) that most nanoparticles
are found to locate in the center of a dimple. The inset in the lower left corner of (c) shows a large
void uncovered on the fracture surface.

27
LLNL-JRNL-706341

Figure 6. Strain rate jump test performed on an AM 316L SS, with a strain rate range of 1×10-4 s-
1
to 2×10-2 s-1. The derived strain rate sensitivity (m) is 0.022±0.002 (inset). Note that the m
reported here represents an average value in 5% strain.

28
LLNL-JRNL-706341

Figure 7. Microstructure and composition analysis of nanoprecipitates in AM 316L SS. (a) A


cross-sectional focused-ion-image (FIB) image of as-built 316L SS. Dark particles are visible
(some are highlighted with solid white circles). (b) A bright-field transmission electron micrograph
of as-deformed 316L SS (strain level ~0.06). (c) A HAADF image of straight cellular boundaries,
with the segregation of nanoparticles along the boundaries clearly revealed. The segregation of Cr,
Mo, and Si are also observed. (d) A HAADF image of a high-angle grain boundary (HAGB),
indicating the segregation of nanoparticles along the boundary. (e) Energy dispersive x-ray
spectrum taken from the nanoparticles, suggesting that the nanoprecipitates are silicon oxides.

29
LLNL-JRNL-706341

Figure 8. Deformation microstructure and surface morphology of tensile samples after fracture.
Focused-ion-beam (FIB) ion-channeling (ion current, 93pA) images of tensile deformed samples
after (a) 0.06 and (b) 0.23 true strains, respectively. The surface morphologies of (c) directly-
printed and (d) plate samples after fracture. Note the microcracks on the surface of directly-printed
sample. The build direction and tensile axis are the same for all the samples and marked in (b).
Some deformation twins in (a) are indicated with white arrows.

30
LLNL-JRNL-706341

31
LLNL-JRNL-706341

Figure 9. A summary of tensile yield strength (σ0.2%) versus tensile-elongation-to-failure (TEF) in


our AM 316L SS, in comparison with values collected in the literature. (a) Tensile σ0.2% versus
engineering TEF plot. The data from our plate and directly-printed samples are accentuated inside
the dashed blue and red circles, respectively. For reference, the data of a wrought 316L SS is also
included (Carlton et al., 2016). Note that TEF values depend strongly on the sample gauge length,
which is not considered in the plot. (b) Engineering TEF in various AM 316L SS as a function of
normalized energy density. Note that one of the directly-printed sample reaches 43% TEF. (c)
Tensile σ0.2% is plotted against normalized energy density for all our samples. Note that TEF of the
plate samples is generally higher than that of the directly printed samples.

32

You might also like