Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Acta Materialia 242 (2023) 118481

Contents lists available at ScienceDirect

Acta Materialia
journal homepage: www.elsevier.com/locate/actamat

Full length article

Effects of initial 3D printed microstructures on subsequent


microstructural evolution in 316L stainless steel
Chunlei Zhang∗, Dorte Juul Jensen, Tianbo Yu∗
Department of Civil and Mechanical Engineering, Technical University of Denmark, Kgs. Lyngby DK-2800, Denmark

a r t i c l e i n f o a b s t r a c t

Article history: Plastic deformation of 3D printed components may occur when they are in use. Here we analyze the
Received 23 June 2022 effects of the initial 3D printed microstructure of 316 L stainless steel on the subsequent deformation be-
Revised 2 September 2022
havior, where we apply 10% and 30% thickness reductions by cold rolling. The microstructures are charac-
Accepted 24 October 2022
terized by electron backscatter diffraction (EBSD) and transmission electron microscopy (TEM). Compared
Available online 25 October 2022
to conventionally manufactured (solution treated) samples, deformation twinning is observed to occur at
Keywords: lower strains in 3D printed samples, and twins are observed to be thinner and intersecting inside some
3D printed 316L grains. These observations are linked to the pre-existing dislocation structure in the 3D printed samples,
Cold rolling where dislocations of various Burgers vectors facilitate deformation twinning. Furthermore, cells with an
Microstructure average size of 125 nm are observed to form inside the initial cellular structure. The strength calcu-
Deformation twinning lated based on microstructural parameters generally agrees with experimental results, showing a large
Mechanical properties
strengthening contribution from twin–matrix lamellae. The present study provides fundamental ideas for
microstructural engineering of 3D printed metals for even better mechanical properties.
© 2022 The Author(s). Published by Elsevier Ltd on behalf of Acta Materialia Inc.
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/)

1. Introduction Extensive recent research has shown that the printing parame-
ters, for example the scan strategy, laser power, and scan speed,
Metal 3D printing, a popular additive manufacturing (AM) pro- significantly affect the microstructure, porosity distribution, and
cess, has drawn enormous attention during the last few decades mechanical properties of printed samples, e.g., [13–15]. With care-
because it enables near net-shape manufacturing of components ful optimization of the printing parameters, almost fully dense
with complex shapes [1–3]. Laser powder bed fusion (L-PBF) is printed samples with reproducible microstructures can be obtained
in particular popular. This technology has been used to manufac- [16,17]. When a melt pool is cooling down, the <001> crystallo-
ture components of many alloys, e.g., Ti-6Al-4 V, TiAl-based al- graphic direction is favored for growth along the maximum tem-
loys, Inconel 718, AlSi10 Mg, and stainless steels [2,4–6]. Austenitic perature gradient (in metals of cubic crystal system), which is nor-
stainless steel 316 L is one of the most frequently used stainless mally perpendicular to the melt pool surface [18]. Therefore, the
steels, due to its good corrosion resistance and formability. Many texture of the printed components can be controlled by manip-
316 L components have complex geometries, for example pipeline ulating the printing parameters and thus the melt pool geome-
systems used in the nuclear industry [7], various tailor-made im- try during printing. For example, Sun et al. [18] reported that by
plants [8,9], and structural components used in the automotive and optimizing the laser power and scan strategy, the texture of the
aerospace industries [10], making conventional manufacturing pro- printed materials is highly dependent on the depth of the melt
cesses difficult and costly. Additionally, a promising result is that pool; <100> texture tends to form in shallow melt pools while
316 L samples produced by L-PBF may have even better mechanical <110> texture tends to form in deep melt pools. In addition, ma-
properties (higher strength and comparable ductility) than those of terials with <110> and <100> texture have been manufactured by
as the cast and even wrought counterparts [11,12]. Thus, there is utilizing the unidirectional scanning and 90° rotation between ad-
huge potential in producing 316 L components via AM techniques jacent layers, respectively [19,20]. The texture induced by printing
and a need for understanding the relationship between AM param- can affect the deformation behavior and thus also the strength and
eters, microstructure, and mechanical properties. ductility of the printed components; deformation twinning tends
to occur in domains with <011> texture under tensile loading [18].
For industrial applications, the scanning strategy of 67° laser ro-

Corresponding authors. tation is usually applied to avoid a strong texture and to make
E-mail addresses: chzhang@mek.dtu.dk (C. Zhang), tiyu@dtu.dk (T. Yu).

https://doi.org/10.1016/j.actamat.2022.118481
1359-6454/© 2022 The Author(s). Published by Elsevier Ltd on behalf of Acta Materialia Inc. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/)
C. Zhang, D. Juul Jensen and T. Yu Acta Materialia 242 (2023) 118481

the directions in the scanning plane (normal to building direction)


equivalent.
This improvement in mechanical properties of 3D printed
samples/components is generally attributed to the fine-scale mi-
crostructures resulting from the printing process. Wang et al.
[17] reported a complex heterogeneous cellular microstructure in
printed 316 L samples and suggested this as an explanation for the
improved properties. A recent study [21] suggests that it is the lo-
cal thermal expansion/shrinkage in a constrained medium which
gives rise to the formation of the cellular structure, rather than
other factors, e.g., dendritic micro-segregation and the resultant
constitutional stress, or misorientations between dendrites. Some- Fig. 1. Sketch of the rolling geometry of the as printed sample (P-0). The rolling
times also a cellular structure is observed. It is worth noting that direction (RD) is parallel to the building direction (BD); the normal direction (ND)
the cellular/columnar structure is actually cylindrical in 3D and the is parallel to the X direction; the transverse direction (TD) is parallel to the Y direc-
difference between seeing a cellular or a columnar structure is due tion. The small micrographs in the longitudinal section (LS) and the cross section
(CS) indicate the sections used for microstructural characterization.
to different viewing directions, i.e., it is a stereological effect [22].
Irrespective of how the pre-existing cellular/columnar structure is
formed, it is expected to affect the dislocation behavior, and there-
fore plays a vital role, in any subsequent deformation process. ish of the component, a final contour scanning, with input power
There has been extensive research, investigating the deforma- of 120 W and scanning speed of 0.22 m/s, was applied. The as
tion behavior of 3D printed 316 L, e.g., [17,23,24], mainly focus- printed microstructure was characterized by electron backscatter
ing on the analysis of the tensile behavior. It is suggested that diffraction (EBSD) and TEM. The results hereof are summarized in
the finely-spaced cellular/columnar boundaries together with ele- Supplementary materials S1.
mental segregation facilitate dislocation storage and deformation As sketched in Fig. 1, the 3D printed samples were unidirection-
twinning, contributing to the high strength and good ductility of ally rolled along BD at room temperature to thickness reductions
3D printed 316 L [17,25]. Moneghan et al. [26] suggested that of 10% and 30% (von Mises equivalent strains ε vm of 0.12 and 0.41,
the cellular/columnar structure can delocalize the strain and thus respectively). During cold rolling, the samples were lubricated by
increase the defect tolerance. In-situ transmission electron mi- mineral oil and turned upside down between each pass. The ratio
croscopy (TEM) observations revealed that the cellular/columnar of the contact length to the mean height l/h was kept at about one
boundaries are weak obstacles to both dislocation motion and de- to ensure through-thickness strain homogeneity. For convenience,
formation twinning [12]. The high temperature evolution of the the as printed sample, and the printed samples after 10% and 30%
cellular/columnar structure has also been analyzed [27]; the cel- thickness reductions are denoted as P-0, P-10%, and P-30%, respec-
lular/columnar structure appears to be stable up to 600 °C, below tively.
which the yield strength also appears to be unchanged, ensuring A Zeiss Supra 35 field emission gun scanning electron micro-
the properties of the printed components are maintained in a wide scope (FEG-SEM), equipped with an HKL EBSD system, was used
temperature range. to characterize the microstructure. EBSD samples were prepared by
These are all detailed investigations. However, to the best of au- grinding using SiC papers followed by mechanical polishing. A col-
thors’ knowledge, there is a lack of quantitative TEM characteriza- loidal silica suspension (0.04 μm) was used for the final polishing
tion of how the 3D printed microstructures (particularly the cel- step. TEM investigations were performed using a JEOL JEM-2100
lular/columnar structure) evolve during subsequent deformation. microscope at an acceleration voltage of 200 kV. Foils for the TEM
Such high fine-scale characterization, we consider essential for investigations were prepared by a modified window technique in
understanding the evolution of microstructures and properties of a perchloric acid based electrolyte [29]. EBSD and TEM characteri-
components when they are mechanically loaded while in-service, zations were carried out at the thickness center in both the cross
and for providing guidance for the practical application of printed section and the longitudinal section (see Fig. 1). EBSD maps were
components. The present study aims at providing such a detailed obtained using a step size of 500 nm. Simple noise reduction, e.g.,
microstructural characterization. Regarding the deformation mode, removing wild spikes and filling them with neighbor orientations,
we have here chosen cold rolling. It is of course clear that a printed was applied using Channel 5 software. MTEX was used for analyz-
component is not expected to be rolled for, or in, a practical ap- ing the EBSD data [30,31], e.g., the Schmid factor analysis. EBSD
plication. In spite of this we have chosen to study rolling of 3D orientation maps are in the present study colored according to the
printed 316 L, because the microstructural evolution of conven- normal direction to the observation surface in the inverse pole fig-
tionally manufactured (solution treated) 316 L during cold rolling ure. We have chosen 15° as the cutoff misorientation angle because
is well known [28]. By comparing the two sets of microstructures this is the value we used for the analysis of conventionally manu-
as a function of strain, it is thereby possible to directly quantify factured samples, and we define boundaries in EBSD maps accord-
the effects of the printing microstructures on the subsequent de- ing to their misorientation angles into low angle boundaries (LABs,
formation behavior—for example the balance between deformation 2–15°) and high angle boundaries (HABs, >15°). Twin boundaries
twinning and dislocation slip—and its effects on the evolution of are defined as boundaries which deviate less than 2° from the
mechanical properties. misorientation angle/axis of 60°/<111>. The Vickers microhardness
was measured using a Struers DuraScan hardness tester with a
2. Experimental load of 300 g. For all samples, the hardness was measured in the
longitudinal section, and each value was averaged over at least 12
Stainless steel 316 L samples, with sizes of 3 × 4 × 14 mm in measurements. Larger samples for tensile test with the dimensions
the scanning directions X, Y and the building direction (BD), were 130 × 12 × 3 mm were printed using the same printing parame-
produced by L-PBF. The laser power and scanning speed were set ters. Some of these larger samples were cold rolled to 30% thick-
to 370 W and 1.35 m/s, respectively. A 67° rotation of the scanning ness reduction before they were cut into dog-bone shaped speci-
direction between neighboring layers was applied, making the X mens. The tensile tests were done using a universal tensile testing
and Y scanning directions equivalent. To improve the surface fin- machine at a strain rate of 4 × 10−4 .

2
C. Zhang, D. Juul Jensen and T. Yu Acta Materialia 242 (2023) 118481

are observed to deform by twinning, compared to P-10% (14%). The


fraction of twin boundaries increases to 15% of the total bound-
ary length (see Supplementary materials S2). Twin boundaries are
at a tilt of ±45° to RD in the longitudinal section (Fig. 4(a)) and
parallel to TD in the cross section (Fig. 4(b)). These twin planes,
approximately ±45° rotated about TD from the rolling plane, are
close to the planes of the largest resolved shear stress. Although
single sets of deformation twins are still the most common obser-
vation, double sets are also frequently observed, as marked by the
white arrows in Fig. 4(a).
Fig. 2. Inverse pole figure coloring maps of P-10%. (a) Longitudinal section (TD col-
ored); (b) cross section (RD colored). The thick black lines mark twin boundaries; An example of a TEM observation of a grain with double sets of
gray lines and thin black lines mark LABs and other HABs, respectively. The black deformation twins is shown in Fig. 5(a1). The first set of twins is
arrows in (b) mark the grains with double sets of deformation twins. thinner and more narrowly spaced compared to the second set of
another twin variant, indicating that the first set of twins are fa-
vorable under the current stress state. The high magnification im-
3. Results age in Fig. 5(a2) shows very thin deformation twins with thick-
nesses down to a few nanometers. Microbands are commonly ob-
3.1. Microstructure after 10% thickness reduction (P-10%) served after 30% thickness reduction, and an example is shown in
Fig. 5(b1). Microbands form by different combinations of slip sys-
After 10% thickness reduction, deformation twins are observed tems or different partitioning of slip on the same slip systems in
in EBSD orientation maps as shown in Fig. 2. There is no obvious neighboring parts of the microstructure, and thus the orientations
change in the shape of grains (here domains delineated by HABs on the two sides are different. In the present sample, the misori-
are denoted as grains for simplicity although they appear different entation angles across the microbands are low, typically less than
from grains in conventionally manufactured materials) after defor- 5°, and microbands are observed to strongly interact with deforma-
mation compared to that of P-0 (see Supplementary materials S1). tion twins. As a result, twins are commonly observed to be sheared
In total, 14% of the grains (area fraction) are observed to have de- by microbands especially in the longitudinal section as shown in
formation twins at this strain. Most of the grains hosting twins ex- Fig. 5(b2).
hibit a single set of deformation twins, only two grains are ob- Fig. 5(c1) shows an example of a deformation twin penetrat-
served exhibiting double sets of deformation twins (marked by the ing through pre-existing CeDBs. The misorientation of the twin
black arrows in Fig. 2(b)). The fraction of twin boundaries is 5%, boundary is found to differ from the perfect 60°/<111> due to
while LABs amount to 67%, of the total boundary length (see Sup- subsequent deformation (the misorientation angle is observed to
plementary materials S2). be 48°). Although it appears to be rather ‘clean’ inside the twin
It is a common observation that deformation twins penetrate in the current view, dislocation boundaries are observed when the
and shear the columnar dislocation boundaries (CoDBs) after 10% sample is tilted to a proper position. A few stacking faults, aligned
cold rolling, as shown in Fig. 3(a). In the region delineated by parallel to the deformation twin, are also observed (see Fig. 5(c1)).
CoDBs, the dislocation density increases due to plastic deforma- Another example is shown in Fig. 5(c2), exhibiting double sets of
tion. In a few grains where two sets of deformation twins are twins intersecting with each other inside the cellular structure,
observed, the grains are subdivided into many parallelogrammatic e.g., at the CeDBs marked by the white dashed line. Fig. 5(c3)
regions (see Fig. 3(b)). A high density of dislocations is also ob- shows an example of extensively developed deformation twins in-
served in these parallelogrammatic regions, further subdividing the side the cellular structure, which is marked by the white dashed
microstructure by forming new dislocation boundaries. For some lines. Most deformation twins are observed to penetrate the CeDBs
CoDBs, e.g., the CoDB marked by the white dashed line in Fig. 3(c), while some are stopped by the CeDBs as marked by the arrows in
some deformation twins are stopped by the boundaries (see white Fig. 5(c3).
arrows in Fig. 3(c)) although others penetrate the boundaries (see Fig. 6 shows an example of the typical dislocation structure
red arrows in Fig. 3(c)), indicating the moderate effect of CoDBs in areas without deformation twins. There is a high density of
on deformation twinning. Twins are also seen in the cellular dis- dislocations between the pre-existing CoDBs/CeDBs. As shown in
location boundaries (CeDBs) structure. A magnified image, exhibit- Fig. 6(a), these dislocations start to organize into new cells of
ing double sets of deformation twins formed on two {111} planes, smaller size (rolling-induced cells) than that of the pre-existing
is shown in Fig. 3(d). The vertical set develops first and is thus dislocation structure. Fig. 6(b) shows interaction between CeDBs
sheared by the subsequent deformation twinning set. The latter is and dislocations activated during cold rolling, increasing the width
a twin bundle containing multiple thin deformation twins. A high of CeDBs as well as the overall dislocation density.
dislocation density inside the twin bundle is observed. Typical dis-
location structures observed in areas without deformation twins 3.3. Mechanical properties
are shown in Fig. 3(e, f). Some dislocations align on one of the
{111} planes as marked by the white dashed lines in Fig. 3(f). How- The hardness measured on the longitudinal section is shown
ever, a common observation is that rolling induced dislocations en- in Fig. 7(a), revealing a significant increase by cold rolling. Also
tangle and organize to form new dislocation boundaries between shown is the hardness variation of conventionally manufactured
the original CeDBs/CoDBs. 316 L samples during cold rolling (denoted as C-0, C-10%, C-30%
and C-50%) from a previous study [28] for comparison. The hard-
3.2. Microstructure after 30% thickness reduction (P-30%) ness of P-0 is considerably higher than that of C-0, indicating high
strength of 3D printed samples. The hardness of P-0 is comparable
After 30% thickness reduction, an elongated microstructural to that of C-10%, suggesting that the strengthening effect of the 3D
morphology is observed (compressed along ND) as shown in Fig. 4. printing process is similar to that of 10% cold rolling although the
The aspect ratio, i.e., the spacing of HABs along RD or TD to that two microstructures are significantly different. The hardness of P-
along ND, increases to 3 in the longitudinal section and to 1.5 in 30% is more than 350 Hv, which is comparable to that of C-50%.
the cross section (Fig. 4(a, b)). More grains (31% in area fraction) The yield strength of the printed sample is increased by 79% after

3
C. Zhang, D. Juul Jensen and T. Yu Acta Materialia 242 (2023) 118481

Fig. 3. TEM images showing the deformation microstructure of P-10%. (a) Longitudinal section view, showing deformation twins (DTs) penetrating CoDBs; (b) cross section
view, showing the interactions between double sets of deformation twins, dislocation boundaries, and loose dislocations; (c) cross section view, showing the interaction
between deformation twins and CoDBs (the white arrows mark twins that are stopped by the CoDB while the red arrows mark twins penetrating the CoDB); (d) cross section
view, showing double sets of deformation twins (inset is the diffraction pattern taken from an intersection area of the two sets of twins); (e, f) typical microstructures in
grains without deformation twins.

vided by CeDBs/CoDBs (sub-micron scale). This microstructure (see


Supplementary materials S1) is significantly different from those
in the initial state and cold rolled state of conventionally manufac-
tured 316 L and is expected to affect the subsequent deformation
behavior. It is therefore of relevance to analyze the deformation
behavior by comparing the microstructural evolution during defor-
mation in printed samples to that in conventionally manufactured
samples. In the present work, the focus is on deformation by cold
rolling. The microstructural evolution during cold rolling of con-
ventionally manufactured samples has been systematically char-
Fig. 4. Inverse pole figure coloring maps of P-30%. (a) Longitudinal section (TD col-
ored); (b) cross section (RD colored). Thick black lines mark twin boundaries; gray acterized previously [28], and can be summarized in brief as: At
lines and thin black lines mark LABs and other HABs, respectively. The white ar- small strain (10–30% cold rolling), a Taylor lattice structure forms
rows mark examples of double sets of twins. The grains are elongated along RD. due to planar slip of dislocations. Microbands start to form after
More deformation twins are detected compared to samples deformed to a lower 30% cold rolling, accommodating a high density of dislocations.
strain.
Only few deformation twins form at low strain (10% cold rolling),
while the volume fraction of twins increases with increasing strain
30% cold rolling (from 500 to 895 MPa), although the ductility is up to 50%, forming twin–matrix lamellae. As deformation twin-
severely reduced simultaneously (see the tensile curves in Fig. 7b). ning occurs later than microband formation, it is observed that
To estimate and compare the work hardening rate of the printed microbands are sheared by the subsequent twinning after 30% cold
and conventionally manufactured samples, the stress–strain curves rolling. It was furthermore found that deformation twinning occurs
are fitted by the standard Voce model [32] based on the hard- preferentially in grains with Copper orientation rather than Brass
ness data (see Supplementary materials S3 for details about the orientation and largely on planes with the highest Schmid factor
fitting), showing a vertical shift (Fig. 7c). The work hardening rate for twinning. These results from conventionally manufactured sam-
of the printed samples is higher than that of conventionally man- ples serve as a basis for the analysis of the results obtained in the
ufactured ones at the same stress level (see Fig. 7d). present study.

4.1.1. Dislocation activity


4. Discussion In contrast to the conventionally manufactured samples, Taylor
lattice structures are not observed to form during cold rolling in
4.1. Deformation of 3D printed 316L samples the 3D printed samples. In 316 L, planar slip of dislocations pre-
vails during cold rolling [28,33] due to its low stacking fault en-
After printing of 316 L samples, the microstructure consists of ergy. Therefore, at low strains, dislocations are activated and glide
grains of rather irregular shapes (micrometer scale) each subdi- on one or two favorable {111} slip planes, forming the Taylor lattice

4
C. Zhang, D. Juul Jensen and T. Yu Acta Materialia 242 (2023) 118481

Fig. 5. Bright field TEM images showing the morphology of deformation twins (DTs) in P-30%. Panel (a) shows double sets of deformation twins. (a1) Overview of double sets
of deformation twins inside a grain; (a2) a magnified image showing twin–twin interaction. Panel (b) shows the interaction between dislocations and deformation twins. (b1)
In a columnar structure; (b2) in a cellular structure. Panel (c) shows details of deformation twins in the CeDB structure. (c1) Cross section view, showing one deformation
twin developing in a cellular structure and some stacking faults (SF) parallel to the deformation twin; (c2) cross section view, showing double sets of twins intersect with
each other at the CeDBs, as marked by the white dashed line; (c3) longitudinal section view, showing finely spaced deformation twins in one cellular structure. The white
dashed lines mark the CeDBs.

Fig. 7. Mechanical properties. (a) Microhardness measured in the longitudinal sec-


tion (RD/ND plane) of 3D printed and conventionally manufactured samples as a
function of deformation strain. The error bars represent standard deviations. (b)
Engineering stress vs strain curves for P-0 and P-30%. (c) The stress–strain curves
fitted by the standard Voce model based on hardness data (see Supplementary ma-
Fig. 6. Microstructure of areas without deformation twins in P-30%. (a) Formation terials S3 for details about the fitting). Note the error bars here represent standard
of rolling-induced dislocation cells in the columnar structure. (b) Loose dislocations errors. (d) Work hardening rate vs stress based on the fitted stress–strain curves.
seen inside the cellular structure.

actions between dislocations. In areas without deformation twins,


structure in conventionally manufactured 316 L. However, in the small cells form within the pre-existing cellular/columnar structure
3D printed 316 L, the pre-existing dislocation structure with vari- to accommodate the increasing dislocation density, e.g., Fig. 6. Af-
ous Burgers vectors acts as both a source and an obstacle [17] to ter 30% cold rolling, the average size of these cells is about 125 nm
the dislocation motion (see Figs. 3 and 6), leading to more inter- (see Fig. 8), i.e., on average one quarter of the original CeDB/CoDB

5
C. Zhang, D. Juul Jensen and T. Yu Acta Materialia 242 (2023) 118481

Fig. 9. The twin thickness in 3D printed samples (P-10% and P-30%) and in conven-
tionally manufactured samples (C-10% and C-30%). Note that the scale of X-axis for
Fig. 8. The size distribution of dislocation cells in P-30%. The average cell size is
C-30% is different from others. The average twin thickness and standard deviations
about 125 nm, a quarter of the original CeDB spacing.
are given. The average twin thickness for the printed samples is smaller than that
for the conventionally manufactured samples.

spacing. It is worth noting that such new cells were not observed
in P-10% or in cold rolled conventionally manufactured 316 L [28]. Lomer–Cottrell locks:
a   a a
112̄ + [112] = [110]
4.1.2. Deformation twinning behavior 6 6 3
Compared to the conventionally manufactured samples, twin- In the printed samples, there are a large amount of partial dis-
ning starts from a lower strain and occurs more extensively in 3D locations with various Burgers vectors of the type a/6<112>, which
printed samples during deformation. In P-10%, deformation twin- would react with twinning partials to form such Lomer–Cottrell
ning can be detected by EBSD using a step size of 500 nm whereas locks with the Burgers vector a/3<110> to reduce the disloca-
it is not detectable in C-10% using the same step size. This trend tion energy. These dislocations do not easily glide in FCC metals
continues and after 30% cold rolling, deformation twinning occurs at room temperature due to a large Peierls–Nabarro stress [36].
more extensively in P-30% than in C-30% with the fraction of twin Therefore, twinning cannot continue in this case. Upon increas-
boundaries 72% vs 60% (P-30% vs C-30%) of the total high angle ing strain, the Lomer–Cottrell locks can cause stress concentration,
boundary length. It has been reported that deformation twinning thereby initiating another partial dislocation. If this partial disloca-
is affected by deformation temperature, strain rate, stacking fault tion inherits the initial Burgers vector, the twinning may continue.
energy, and grain size [34]. The first three factors cannot explain If a new partial is initiated, then a new twinning variant can form
the present results as they are similar in the two cases. Concern- provided the shear stress is large enough for the growth of the
ing the grain size effect, it is generally accepted that larger grains new twins, contributing to the frequently observed double sets of
favor deformation twinning [34], although this effect is small for twinning in 3D printed 316 L samples.
face-centered-cubic (FCC) metals [35]. The average grain size of the The present results suggest that the pre-existing dislocation
conventionally manufactured 316 L investigated is about 26 μm structure can also affect the morphology of deformation twins.
[28] while it is 25 μm (33 μm/17 μm along BD/X) in the present Thinner deformation twins (average thickness) are observed in 3D
printed sample, suggesting that the grain size cannot explain the printed 316 L than those in conventionally manufactured samples
difference in deformation twinning in the present case. after deformation (see Fig. 9). Furthermore, the median of twin
It follows that the pre-existing dislocation structure has a ma- thickness for printed samples is also smaller than that of conven-
jor effect on deformation twinning. The dislocations in the pre- tionally manufactured samples (see Supplementary materials S4
existing dislocation boundaries are mostly dissociated partial dislo- for details). Through two-sample t-Test, it is concluded that the dif-
cations with the Burgers vector 1/6<112> [12]. Experimental evi- ference between the twin thicknesses of the 3D printed and con-
dence is shown in Figure S2 in Supplementary materials, where the ventionally manufactured samples is significant within 95% con-
partial dislocations glide from one side of a cell to the other side, fidence. One may argue that the difference in the thickness and
forming a stacking fault in between during printing. Since partial density of deformation twins may be caused by the difference in
dislocations and stacking faults can serve as nucleation sites for stress level between printed samples and conventionally manufac-
twinning [34], it is thus reasonable to expect earlier and more ex- tured samples. However, the t-Test results (Fig. S4(c, d) in Sup-
tensive deformation twinning upon external loading as observed in plementary materials) for twin thicknesses of both printed sam-
the present study. However, the pre-existing dislocation structure ples (P-10% vs P-30%) and conventionally manufactured samples
could also act as a weak obstacle to the growth of the deformation (C-10% vs C-30%) show that there is no significant difference in
twins [12,17]. As mentioned above, some deformation twins are average twin thickness at strains from 10 to 30%, suggesting that
observed to penetrate the dislocation structures while others are the stress level, at least in this strain range, has no obvious effect
stopped (see Fig. 3). This is attributed to the interaction between on the twin thickness. It is therefore suggested that it is the cellu-
twinning partial dislocations and the pre-existing dislocation struc- lar structure, containing a high variety of partial dislocations, that
ture. When a twinning partial encounters a CeDB/CoDB, it may provides nucleation sites for twinning (ether by providing stack-
be repelled by the dislocation boundary and only breaks through ing faults or by providing twinning partial dislocations) and in-
upon further loading (larger stress). This has been observed by in- creases the twinning nucleation probability, resulting in more ef-
situ TEM [12]. Another possibility is that the twinning partial may ficient stress release, as well as thinner and denser deformation
be trapped by the dislocation boundary through the formation of twins. Liu et al. [12] presented evidence of deformation twins with

6
C. Zhang, D. Juul Jensen and T. Yu Acta Materialia 242 (2023) 118481

It is worth noting that there is large overlap between the distri-


bution of mt /ms for grains with twins and that for grains without
twins as shown in Fig. 10(f). This is due to the high symmetry of
FCC metals, resulting in many equivalent slip/twinning systems.
Another difference found is that all the deformation twinning
detected by EBSD occurs on the planes having the highest mt in
C-30% whereas only 13/19 (13 grains out of 19) exhibits the same
pattern in P-10% and this probability decreases to 31/100 in P-30%.
These phenomena suggest that in 3D printed samples, twinning se-
lection is a more complex process than that in conventionally man-
ufactured samples due to the effects of the pre-existing dislocation
structure and irregularly shaped grains.
For the case of double sets of deformation twins, one set of
twins develops on the plane with the highest mt while the other
set develops on a plane with very low mt (near zero). As a shear
stress is needed for the motion of partial dislocations on the twin-
ning plane for twinning propagation if the classic twinning the-
ory is correct [34], large internal stresses are expected in the 3D
printed samples with irregularly shaped grains and pre-existing
dislocation structures. Both the irregularly shaped grains and pre-
existing dislocation structures can cause a complex local stress
state and therefore provide the resolved shear stress required for
the development of deformation twins on planes other than the
one with the highest mt during the subsequent cold rolling. How-
ever, further understanding of this phenomenon requires more in-
formation on relations between the 3D microstructure and the lo-
cal stress, which may be obtained by advanced experimental stud-
ies using the 3DXRD technique [38].

4.3. Microstructure–property relationship


Fig. 10. Difference between grains with twins and without twins in P-30%. (a, c)
Orientation map and {111} pole figure of grains with twins in P-30% (Fig. 4(a)). (b,
d) For grains without twins. (e) The highest Schmid factor for twinning (mt ). (f) As has been analyzed above, the large variety of Burgers vectors
Ratio of the highest Schmid factor for twinning (mt ) to the highest Schmid factor of pre-existing dislocations leads to a more uniform deformation
for slip (ms ), i.e., (mt /ms ). The grains with twins show an orientation away from
and thinner twins in 3D printed samples than in conventionally
Copper orientation; the grains without deformation twins show an orientation near
Brass orientation. There is no obvious difference in the highest Schmid factor for
manufactured ones. Deformation twinning has been considered as
twinning between grains with twins and grains without twins, whereas the ratio of a mechanism to achieve a good strength–ductility combination for
Schmid factors (mt /ms ) is higher for grains with twins. metals, known as twinning induced plasticity (TWIP) [39,39–43].
This is because deformation twinning decreases the spacing of high
angle boundaries (HABs), resulting in a high work hardening rate
thicknesses ranging from 2 to 6 nm observed during in-situ ten-
due to the dynamic Hall–Petch effect and thus good ductility [44–
sile tests. This is thinner than those observed in the present study,
46]. Due to the strong twin–dislocation interaction (as has been
probably due to the difference in deformation condition.
extensively studied, e.g., [47–50]), twin boundaries of 60°/<111>
will gradually evolve into ordinary high angle boundaries with in-
4.2. Orientation dependence of deformation twinning
creasing deformation strain [51], as shown in Fig. 5(b1, b2). In ad-
dition, rolling-induced cells, a novel phenomenon observed in the
In conventionally manufactured and cold rolled samples, defor-
present work, are rarely reported in the literature for 3D printed
mation twinning is found to be orientation dependent [28]. In the
316 L after deformation, and their strengthening effect has not
3D printed samples, we have observed that grains with Brass ori-
been considered in the analysis of the mechanical properties. The
entation are not favorable for twinning (see Fig. 10), similar to the
twin–twin, twin–dislocation and dislocation-dislocation boundary
results for conventionally manufactured samples. However, many
interaction may affect the ductility and the work hardening rate of
grains with twins in 3D printed samples have an orientation away
the printed samples during deformation [45,52,53]. To understand
from the Copper orientation (see Fig. 10(c)), which is the favorable
the strengthening effect of the pre-existing cellular/columnar dis-
orientation for twinning in conventionally manufactured samples.
location structure and the subsequent deformation microstructure,
To understand this difference, a Schmid factor analysis is done. The
the microstructural parameters of both printed and conventionally
Schmid factor for rolling is defined [37] as:
manufactured samples (for comparison) are quantified, and their
(cosϕRD cosγRD − cosϕND cosγND ) strengthening contributions are analyzed and compared with ex-
m=
2 perimental observations.
where ϕ RD and ϕ ND are the angles between the slip/twinning On the assumption of additivity of the strength contribution
plane normal and RD and ND, respectively; γ RD and γ ND are the from dislocation structures (dislocations in low angle dislocation
angles between the slip/twinning direction and RD and ND, respec- boundaries plus loose dislocations, i.e., those not organized into
tively. The results show no apparent difference in Schmid factors low angle boundaries) and high angle boundaries [45], the flow
for twinning (mt ) between grains with twins and without twins stress σ at strain ε may be expressed by the following equation:
in the printed sample (see Fig. 10(e)). However, higher average ra-  √
tio of Schmid factors, i.e., mt /ms (ms denotes the Schmid factor for
σ (ε ) = σ0 + Mα G 1.5(SV θ )b + ρ0 b2 + K1 / D (1)
slip), is observed for the grains with twins (see Fig. 10(f)). Similar where σ0 is the frictional stress, including contributions from so-
results were obtained from conventionally manufactured samples. lutes and particles, independent of grain size; M is the Taylor fac-

7
C. Zhang, D. Juul Jensen and T. Yu Acta Materialia 242 (2023) 118481

Table 1
Constants used for the calculation of strength.

b α G K1 σ0 KT −M (= K1 )

0.254 nm 0.2 77 GPa 299 MPa·μm1/2 [55] 173 MPa [55] 9455 MPa·nm1/2

Table 2
Microstructural parameters and yield strength (both calculated and experimentally measured).

M SV θ D ρ0 ρ0T −M f T −M λ σcal σexp


(μm−1 ) (o ) (μm) (1014 m−2 ) (1014 m−2 ) (%) (nm) (MPa ) (MPa )
P-0 3.06 10 0.4 25 — — — — 486 500
P-30% 2.87a 14 0.8 22 8.8 15 31 22 1096 895
C-0 3.06 — — 26 — — — — 231 265 [56]
C-30% 2.82a 5 2 20 3.3 7.5 23 42 852 821 [55]
a
Taylor factors for P-30% and C-30% are estimated from corresponding EBSD maps. Considering the weak initial
textures, the Taylor factors for the undeformed samples are chosen to be 3.06. Details about the dislocation density
estimation can be found in the Supplementary materials S5.

sonable to assume the same strengthening effect for twin bound-


aries as that of grain boundaries for 316 L stainless steel. For P-30%,
the calculation overestimates the yield strength by 22%, which may
be partly caused by using the average spacing of the inhomoge-
neous T–M lamellae. There is a higher fraction of T–M lamellae of
smaller average HAB spacing in P-30% compared to C-30%, indicat-
ing a stronger effect of twinning in strengthening P-30% than in
C-30%. However, further investigations are needed to evaluate the
Fig. 11. Idealized sketches illustrating the deformed microstructure of printed sam- strengthening contribution of the inhomogeneous T–M lamellae in
ples and conventionally manufactured samples. (a) C-30%, showing the twin–matrix
3D printed samples more precisely.
(T–M) lamellar structure as well as the Taylor lattice structure consisting of domain
boundaries (DB) and microbands (MB); (b) P-30%, showing the T–M lamellar struc-
ture (double sets of twins) and the rolling–induced cells in the structure delineated 5. Conclusion
by cellular dislocation boundaries (CeDB).

The present study investigates the microstructural evolution of


tor; α is a constant; G is the shear modulus; SV is the area per unit 3D printed stainless steel 316 L during cold rolling and compares
volume of low angle dislocation boundaries (taken to be 2 times to that of conventionally manufactured counterparts. The strength-
the number of intersections per unit length PL , assuming random ening contributions of different microstructural features are ana-
distribution of low angle dislocation boundaries); θ is the aver- lyzed; the yield strength is calculated based on microstructural pa-
age misorientation angle of low angle dislocation boundaries; b is rameters and compared to experimental results. The main conclu-
the Burgers vector; ρ0 is the density of loose dislocations; K1 is sions are as follows:
the Hall-Petch constant; D is the grain size (spacing of high angle
(1) The pre-existing dislocation boundaries suppress long-range
boundaries). The constants for 316 L are shown in Table 1.
planar glide of dislocations, and therefore the Taylor lattice
For C-30% and P-30% containing a mixture of twin–matrix (T–
structure, which is commonly observed in conventionally
M) lamellar structures and dislocation structures, as shown in
manufactured 316 L, does not develop in 3D printed sam-
Fig. 11, the flow stress σ (ε ) is given by the rule of mixture [45,54]:
ples. Instead, rolling-induced dislocation cells of an average
 √ size of 125 nm form inside the cellular structure.
σ (ε ) = (1 − fT −M )(σ0 + Mα G 1.5SV θ b + ρ0 b2 + K1 / D ) (2) The pre-existing dislocation structure can pivot the deforma-
   tion twinning process in terms of the initiation and propaga-
KT −M
+ fT −M σ0 + √ + Mα Gb ρ0T −M (2) tion, leading to thinner and denser twins in 3D printed sam-
2λ ples. This is because the pre-existing dislocation structure
where fT −M is the volume fraction of T–M lamellae, ρ0T −M is the facilitates deformation twinning by providing twinning par-
dislocation density in T–M lamellae, λ is the average spacing of tial dislocations. Yet, this structure also acts as an obstacle—
T–M lamellar structure, and KT −M is a constant representing the either as a repelling force between dislocations or by form-
strengthening effect of HABs in T–M lamellar structure, taken to ing Lomer-Cottrell locks.
be the same as K1 , i.e., assuming the strengthening effect of twin (3) The 3D printed samples have a higher work hardening rate
boundaries is identical to that of grain boundaries. compared to the conventionally manufactured ones at the
In the above analysis, residual stresses are considered to not same stress level. Strength calculation suggests that the thin-
have a significant effect on the yield strength as the residual ner and denser deformation twins, whose formation is fa-
stresses are balanced on the sample scale. Table 2 shows the mi- cilitated by the pre-existing dislocation structure, contribute
crostructural parameters quantified from EBSD and TEM images, as significantly to the high strength of 3D printed samples after
well as the calculated and experimentally measured yield strength. deformation.
For C-0, only grain boundary strengthening is considered; for P-0,
both grain boundary and dislocation boundaries are considered but Declaration of Competing Interest
the strengthening contribution from loose dislocations is ignored.
The calculated yield strength σcal is almost identical to the experi- The authors declare that they have no known competing finan-
mental results σexp for the cases of P-0 and C-0. For C-30%, there is cial interests or personal relationships that could have appeared to
only 3% difference between σcal and σexp , indicating that it is rea- influence the work reported in this paper.

8
C. Zhang, D. Juul Jensen and T. Yu Acta Materialia 242 (2023) 118481

Acknowledgment [20] D. Kumar, G. Shankar, K.G. Prashanth, S. Suwas, Texture dependent strain hard-
ening in additively manufactured stainless steel 316L, Mater. Sci. Eng. A 820
(2021) 141483, doi:10.1016/J.MSEA.2021.141483.
The main part of this work was carried out within M4D project [21] K.M. Bertsch, G. Meric de Bellefon, B. Kuehl, D.J. Thoma, Origin of dislocation
supported by the European Research Council (ERC) under the Eu- structures in an additively manufactured austenitic stainless steel 316L, Acta
ropean Union’s Horizon 2020 research and innovation program Mater. 199 (2020) 19–33, doi:10.1016/j.actamat.2020.07.063.
[22] R.F. Joseph Aroh, D. Rowenhorst, J. Feng, A. Rollett, Crystallographic analysis
(Grant agreement No. 788567). The tensile testing part of the work, of cell structures in additively manufactured 316L [Conference presentation],
which was added based on comments from the reviewers, was car- in: Proceedings of the 19th International Conference on Textures of Materials
ried out in the RePowder project funded by the Danish Council (ICOTOM 19), 2021.
[23] D. Kong, C. Dong, S. Wei, X. Ni, L. Zhang, R. Li, L. Wang, C. Man, X. Li, About
for Independent Research (Grant No. 9041-00335B). CZ further ac-
metastable cellular structure in additively manufactured austenitic stainless
knowledges the financial support of the China Scholarship Council steels, Addit. Manuf. 38 (2021) 101804, doi:10.1016/J.ADDMA.2020.101804.
(CSC) (N0. 201706120026). [24] L. Zhang, X.Q. Ni, D.C. Kong, Y. Wen, W.H. Wu, B.B. He, L. Lu, D.X. Zhu,
Anisotropy in mechanical properties and corrosion resistance of 316L stain-
less steel fabricated by selective laser melting, Int. J. Miner. Metall. Mater. 26
Supplementary materials (2019), doi:10.1007/s12613- 019- 1740- x.
[25] K. Saeidi, X. Gao, Y. Zhong, Z.J. Shen, Hardened austenite steel with columnar
sub-grain structure formed by laser melting, Mater. Sci. Eng. A 625 (2015) 221–
Supplementary material associated with this article can be
229, doi:10.1016/J.MSEA.2014.12.018.
found, in the online version, at doi:10.1016/j.actamat.2022.118481. [26] M. Moneghan, C. Williams, R. Mirzaeifar, Deformation mechanisms and defect
tolerance in the microstructure of 3D-printed alloys, J. Mater. Res. 35 (2020)
References 1984–1997, doi:10.1557/JMR.2020.60.
[27] T. Voisin, J.B. Forien, A. Perron, S. Aubry, N. Bertin, A. Samanta, A. Baker,
[1] J.J. Lewandowski, M. Seifi, Metal additive manufacturing: a review of me- Y.M. Wang, New insights on cellular structures strengthening mechanisms and
chanical properties, Annu. Rev. Mater. Res. 46 (2016) 151–186, doi:10.1146/ thermal stability of an austenitic stainless steel fabricated by laser powder-
annurev- matsci- 070115- 032024. bed-fusion, Acta Mater. 203 (2021) 116476, doi:10.1016/J.ACTAMAT.2020.11.018.
[2] W.E. Frazier, Metal additive manufacturing: a review, J. Mater. Eng. Perform. 23 [28] C. Zhang, D. Juul Jensen, T. Yu, Microstructure and texture evolution during
(2014) 1917–1928, doi:10.1007/s11665- 014- 0958- z. cold rolling of 316L stainless steel, Metall. Mater. Trans. A (2021) 1–12, doi:10.
[3] E. MacDonald, R. Wicker, Multiprocess 3D printing for increasing component 1007/s11661- 021- 06367- 6.
functionality, Science 353 (80) (2016), doi:10.1126/science.aaf2093. [29] G. Christiansen, J.R. Bowen, J. Lindbo, Electrolytic preparation of metallic thin
[4] B.E. Carroll, T.A. Palmer, A.M. Beese, Anisotropic tensile behavior of Ti-6Al-4V foils with large electron-transparent regions, Mater. Charact. 49 (2002) 331–
components fabricated with directed energy deposition additive manufactur- 335, doi:10.1016/S1044-5803(03)0 0 032-9.
ing, Acta Mater. 87 (2015) 309–320, doi:10.1016/j.actamat.2014.12.054. [30] R. Hielscher, C.B. Silbermann, E. Schmidla, J. Ihlemannb, Denoising of crys-
[5] T.C. Dzogbewu, Additive manufacturing of TiAl-based alloys, Manuf. Rev. 7 tal orientation maps, J. Appl. Crystallogr. 52 (2019) 984–996, doi:10.1107/
(2020) 35, doi:10.1051/mfreview/2020032. S160 05767190 09075.
[6] K.N. Amato, S.M. Gaytan, L.E. Murr, E. Martinez, P.W. Shindo, J. Hernandez, [31] F. Bachmann, R. Hielscher, H. Schaeben, Texture analysis with MTEX- Free and
S. Collins, F. Medina, Microstructures and mechanical behavior of Inconel open source software toolbox, Solid State Phenom. 160 (2010) 63–68, doi:10.
718 fabricated by selective laser melting, Acta Mater. 60 (2012) 2229–2239, 4028/www.scientific.net/SSP.160.63.
doi:10.1016/j.actamat.2011.12.032. [32] X. Tian, Y. Zhang, Mathematical description for flow curves of some sta-
[7] Y. Zhong, L. Liu, S. Wikman, D. Cui, Z. Shen, Intragranular cellular segrega- ble austenitic steels, Mater. Sci. Eng. A 174 (1994) L1–L3, doi:10.1016/
tion network structure strengthening 316L stainless steel prepared by selective 0921- 5093(94)91120- 7.
laser melting, J. Nucl. Mater. 470 (2016) 170–178, doi:10.1016/j.jnucmat.2015. [33] D. Kuhlmann-Wilsdorf, Theory of plastic deformation: properties of low en-
12.034. ergy dislocation structures, Mater. Sci. Eng. A 113 (1989) 1–41, doi:10.1016/
[8] M.M. Dewidar, K.A. Khalil, J.K. Lim, Processing and mechanical properties of 0921- 5093(89)90290- 6.
porous 316L stainless steel for biomedical applications, Trans. Nonferrous Met. [34] J.W. Christian, S. Mahajan, Deformation twinning, Prog. Mater. Sci. 39 (1995)
Soc. China 17 (2007) 468–473, doi:10.1016/S1003- 6326(07)60117- 4. 1–157, doi:10.1016/0 079-6425(94)0 0 0 07-7.
[9] U.I. Thomann, P.J. Uggowitzer, Wear-corrosion behavior of biocompat- [35] Y.T. Zhu, X.Z. Liao, X.L. Wu, J. Narayan, Grain size effect on deformation
ible austenitic stainless steels, Wear 239 (20 0 0) 48–58, doi:10.1016/ twinning and detwinning, J. Mater. Sci. 48 (2013) 4467–4475, doi:10.1007/
S0 043-1648(99)0 0372-5. s10853- 013- 7140- 0.
[10] B. AlMangour, D. Grzesiak, J.M. Yang, In-situ formation of novel TiC-particle- [36] F.R.N. Nabarro, Dislocations in a simple cubic lattice, Proc. Phys. Soc. 59 (1947)
reinforced 316L stainless steel bulk-form composites by selective laser melting, 256–272, doi:10.1088/0959-5309/59/2/309.
J. Alloy. Compd. 706 (2017) 409–418, doi:10.1016/j.jallcom.2017.01.149. [37] J.R. Luo, A. Godfrey, W. Liu, Q. Liu, Twinning behavior of a strongly basal tex-
[11] G. Miranda, S. Faria, F. Bartolomeu, E. Pinto, S. Madeira, A. Mateus, P. Carreira, tured AZ31 Mg alloy during warm rolling, Acta Mater. 60 (2012) 1986–1998,
N. Alves, F.S. Silva, O. Carvalho, Predictive models for physical and mechanical doi:10.1016/j.actamat.2011.12.017.
properties of 316L stainless steel produced by selective laser melting, Mater. [38] H.F. Poulsen, S.F. Nielsen, E.M. Lauridsen, S. Schmidt, R.M. Suter, U. Lienert,
Sci. Eng. A 657 (2016) 43–56, doi:10.1016/j.msea.2016.01.028. L. Margulies, T. Lorentzen, D. Juul Jensen, Three-dimensional maps of grain
[12] L. Liu, Q. Ding, Y. Zhong, J. Zou, J. Wu, Y.L. Chiu, J. Li, Z. Zhang, Q. Yu, Z. Shen, boundaries and the stress state of individual grains in polycrystals and pow-
Dislocation network in additive manufactured steel breaks strength–ductility ders, J. Appl. Crystallogr. 34 (2001) 751–756, doi:10.1107/S0021889801014273.
trade-off, Mater. Today 21 (2018) 354–361, doi:10.1016/J.MATTOD.2017.11.004. [39] L. Bracke, K. Verbeken, L. Kestens, J. Penning, Microstructure and texture evo-
[13] B. Cheng, S. Shrestha, K. Chou, Stress and deformation evaluations of scan- lution during cold rolling and annealing of a high Mn TWIP steel, Acta Mater.
ning strategy effect in selective laser melting, Addit. Manuf. 12 (2016) 240–251, 57 (2009) 1512–1524, doi:10.1016/j.actamat.2008.11.036.
doi:10.1016/j.addma.2016.05.007. [40] S. Vercammen, B. Blanpain, B.C. De Cooman, P. Wollants, Cold rolling behavior
[14] D. Kong, X. Ni, C. Dong, X. Lei, L. Zhang, C. Man, J. Yao, X. Cheng, X. Li, Bio- of an austenitic Fe-30Mn-3Al-3Si TWIP-steel: the importance of deformation
functional and anti-corrosive 3D printing 316L stainless steel fabricated by se- twinning, Acta Mater. 52 (20 04) 20 05–2012, doi:10.1016/j.actamat.2003.12.040.
lective laser melting, Mater. Des. 152 (2018) 88–101, doi:10.1016/j.matdes.2018. [41] D. Barbier, N. Gey, S. Allain, N. Bozzolo, M. Humbert, Analysis of the tensile
04.058. behavior of a TWIP steel based on the texture and microstructure evolutions,
[15] H. Choo, K.L. Sham, J. Bohling, A. Ngo, X. Xiao, Y. Ren, P.J. Depond, Mater. Sci. Eng. A 500 (2009) 196–206, doi:10.1016/j.msea.2008.09.031.
M.J. Matthews, E. Garlea, Effect of laser power on defect, texture, and mi- [42] D. Molnár, X. Sun, S. Lu, W. Li, G. Engberg, L. Vitos, Effect of temperature on
crostructure of a laser powder bed fusion processed 316L stainless steel, Mater. the stacking fault energy and deformation behavior in 316L austenitic stainless
Des. 164 (2019) 107534, doi:10.1016/j.matdes.2018.12.006. steel, Mater. Sci. Eng. A 759 (2019) 490–497, doi:10.1016/j.msea.2019.05.079.
[16] Z. Sun, X. Tan, S.B. Tor, W.Y. Yeong, Selective laser melting of stainless steel [43] C. Haase, S.G. Chowdhury, L.A. Barrales-Mora, D.A. Molodov, G. Gottstein, On
316L with low porosity and high build rates, Mater. Des. 104 (2016) 197–204, the relation of microstructure and texture evolution in an austenitic Fe-28Mn-
doi:10.1016/j.matdes.2016.05.035. 0.28C TWIP steel during cold rolling, Metall. Mater. Trans. A Phys. Metall.
[17] Y.M. Wang, T. Voisin, J.T. McKeown, J. Ye, N.P. Calta, Z. Li, Z. Zeng, Y. Zhang, Mater. Sci. 44 (2013) 911–922, doi:10.1007/s11661- 012- 1543- 4.
W. Chen, T.T. Roehling, R.T. Ott, M.K. Santala, P.J. Depond, M.J. Matthews, [44] J.W. Christian, S. Mahajan, deformation twinning, Prog. Mater. Sci. 39 (1995)
A.V. Hamza, T. Zhu, Additively manufactured hierarchical stainless steels 1–157.
with high strength and ductility, Nat. Mater. 17 (2018) 63–70, doi:10.1038/ [45] N. Hansen, Hall–Petch relation and boundary strengthening, Scr. Mater. 51
NMAT5021. (2004) 801–806, doi:10.1016/J.SCRIPTAMAT.20 04.06.0 02.
[18] Z. Sun, X. Tan, S.B. Tor, C.K. Chua, Simultaneously enhanced strength and duc- [46] T. Zhu, J. Li, A. Samanta, H.G. Kim, S. Suresh, Interfacial plasticity governs strain
tility for 3D-printed stainless steel 316L by selective laser melting, NPG Asia rate sensitivity and ductility in nanostructured metals, Proc. Natl. Acad. Sci. U.
Mater. 10 (2018) 127–136, doi:10.1038/s41427- 018- 0018- 5. S. A. 104 (2007) 3031–3036, doi:10.1073/PNAS.0611097104.
[19] T. Ishimoto, K. Hagihara, K. Hisamoto, S.H. Sun, T. Nakano, Crystallographic tex- [47] J. Wang, H. Huang, Novel deformation mechanism of twinned nanowires, Appl.
ture control of beta-type Ti–15Mo–5Zr–3Al alloy by selective laser melting for Phys. Lett. 88 (2006) 203112, doi:10.1063/1.2204760.
the development of novel implants with a biocompatible low Young’s modu- [48] Y.T. Zhu, X.Z. Liao, X.L. Wu, Deformation twinning in nanocrystalline materials,
lus, Scr. Mater. 132 (2017) 34–38, doi:10.1016/J.SCRIPTAMAT.2016.12.038. Prog. Mater. Sci. 57 (2012) 1–62, doi:10.1016/J.PMATSCI.2011.05.001.

9
C. Zhang, D. Juul Jensen and T. Yu Acta Materialia 242 (2023) 118481

[49] S.L. Thomas, A.H. King, D.J. Srolovitz, When twins collide: twin junctions in [54] F.K. Yan, G.Z. Liu, N.R. Tao, K. Lu, Strength and ductility of 316L austenitic
nanocrystalline nickel, Acta Mater. 113 (2016) 301–310, doi:10.1016/J.ACTAMAT. stainless steel strengthened by nano-scale twin bundles, Acta Mater. 60 (2012)
2016.04.030. 1059–1071, doi:10.1016/J.ACTAMAT.2011.11.009.
[50] L. Lu, R. Schwaiger, Z.W. Shan, M. Dao, K. Lu, S. Suresh, Nano-sized twins in- [55] S. Kheiri, H. Mirzadeh, M. Naghizadeh, Tailoring the microstructure and me-
duce high rate sensitivity of flow stress in pure copper, Acta Mater. 53 (2005) chanical properties of AISI 316L austenitic stainless steel via cold rolling and
2169–2179, doi:10.1016/J.ACTAMAT.2005.01.031. reversion annealing, Mater. Sci. Eng. A 759 (2019) 90–96, doi:10.1016/J.MSEA.
[51] F. Heidelbach, H.R. Wenk, S.R. Chen, J. Pospiech, S.I. Wright, Orientation and 2019.05.028.
misorientation characteristics of annealed, rolled and recrystallized copper, [56] S. Tanhaei, K. Gheisari, S.R.A. Zaree, Effect of cold rolling on the mi-
Mater. Sci. Eng. A 215 (1996) 39–49, doi:10.1016/0921- 5093(96)10264- 1. crostructural, magnetic, mechanical, and corrosion properties of AISI 316L
[52] Q. Yu, J. Wang, Y. Jiang, R.J. McCabe, N. Li, C.N. Tomé, Twin–twin interactions in austenitic stainless steel, Int. J. Miner. Metall. Mater. 25 (2018), doi:10.1007/
magnesium, Acta Mater. 77 (2014) 28–42, doi:10.1016/J.ACTAMAT.2014.05.030. s12613- 018- 1610- y.
[53] J. Wang, X. Zhang, Twinning effects on strength and plasticity of metallic ma-
terials, MRS Bull. 41 (2016) 274–281, doi:10.1557/MRS.2016.67.

10

You might also like