You are on page 1of 10

Materials & Design 210 (2021) 110057

Contents lists available at ScienceDirect

Materials & Design


journal homepage: www.elsevier.com/locate/matdes

Deformation heterogeneity induced coarse grain refinement of the


mixed-grain structure of 316LN steel through limited deformation
condition
Yangqi Li a, Ping Shen b, Haiming Zhang a,⇑, Kai Dong c, Yang Deng a, Xianwang Chen a, Zhenshan Cui a,⇑
a
Institute of Forming Technology and Equipment, School of Materials Science and Engineering, Shanghai Jiao Tong University, Shanghai 200030, China
b
Center for Advanced Solidification Technology (CAST), School of Materials Science and Engineering, Shanghai University, Shanghai 200444, China
c
Shanghai Electric SHMP Casting & Forging Co., Ltd., Shanghai 200245, China

h i g h l i g h t s g r a p h i c a l a b s t r a c t

 MCG grain size plays a significant role


in strain partitioning.
 MCGs can be refined by limited cold
deformation followed by rapidly
heating process.
 Distinct dislocation piles up in MCG
GBs under small deformation and
initiates SRX.
 SRX occurs at both GBs and interior of
the MCG when deformation is
relatively large.

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

Article history: The defect of mixed-grain structure with millimeter-grade coarse grains (MCGs) is frequently found in
Received 13 July 2021 heavy forgings, such as nuclear main pipes manufactured from 316LN steel. Renovation of such kind forg-
Revised 16 August 2021 ings can only be conducted through recrystallization. However, the allowed deformation is limited
Accepted 19 August 2021
because the geometric size of forgings should be maintained during the renovation. For this purpose,
Available online 20 August 2021
the deformation heterogeneity of 316LN steel with mixed-grain structures were studied experimentally,
which demonstrated that the MCGs were more inclined to deform than fine grains even if MCGs are with
Keywords:
hard orientations. Therefore, a two-stage method was studied to refine MCGs, which takes the advantage
316LN austenitic stainless steel
Mixed-grain structure
of this deformation heterogeneity. The first stage was to impose a limited deformation to the steel at
Deformation heterogeneity room temperature, suppressing the dynamic recovery and leading the preferentially deformed MCGs
Grain refinement to accumulate enough deformation energy storage. The second stage was rapidly heating the deformed
In-situ SRX observation materials to realize static recrystallization (SRX). The results showed that the MCGs were completely
refined by SRX even if the reduction rate is only 8%. The SRX mechanisms and refining assessments of
MCGs under different deformation conditions were discussed. This work, thus, provides an effective
method to salvage the heavy structures in engineering.
Ó 2021 Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://
creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction

⇑ Corresponding authors. Austenitic stainless steel ASTM-TP316LN (316LN) is widely


E-mail addresses: hm.zhang@sjtu.edu.cn (H. Zhang), cuizs@sjtu.edu.cn (Z. Cui). used in the nuclear power station equipment, e.g., the main pipe

https://doi.org/10.1016/j.matdes.2021.110057
0264-1275/Ó 2021 Published by Elsevier Ltd.
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
Y. Li, P. Shen, H. Zhang et al. Materials & Design 210 (2021) 110057

[1,2], due to its excellent combination of high strength, good corro- mission electron microscopy (TEM) was also employed to observe
sion resistance, and adequate weldability [3,4]. However, the the dislocation distribution and to elucidate the SRX mechanism of
mixed-grain structures were frequently detected in the 316LN samples under different deformation conditions. The effects of
steel forgings, in which coarse grains of sub-millimeter or even MCG refinement were compared and evaluated.
millimeter size (MCGs) arise due to the unavoidable temperature
gradient and inhomogeneous deformation during multi-pass hot
2. Materials and methods
forging process [5]. The existence of the mixed-structure with
abnormally large grains will have catastrophic effects on the
2.1. Specimen preparation
mechanical performances of materials [6,7], so that the forgings
should be scrapped if the large grains cannot be refined. Heavy
The materials investigated in this study are from as-forged
forgings like the main pipes often cost million dollars or even more,
316LN stainless steels, which is provided by Shanghai Electric
therefore, refining the mixed-grain structures of the steel is crucial
SHMP Casting & Forging Co., Ltd. The chemical composition of
to avoid the forgings abandoned. However, the renovation of the
the studied material is exhibited in Table 1. The as-received mate-
mixed-grain forgings is limited by the allowable deformation:
rials with MCG mixed-grain structure were shaped into cuboid
large deformation will cause a forging distorted in shape and also
compression samples by wire cut. The dimension of the sample
leads it to be scrapped. Therefore, for the austenitic stainless steel
was designed as 3.5 mm (length)  2.5 mm (width)  5 mm
whose grain cannot be refined by heat treatment, a suitable grain
(height, compression direction) for the convenience of in-situ
refinement method under the limitation of deformation is crucially
observations. Before and after the mechanical testing, all the sam-
required.
ples underwent the standard mechanical grinding followed with
In order to prepare metallic materials with good mechanical
the electrochemical polishing as well as the EBSD measurements
performances, the formation mechanisms [8,9] and refining meth-
using a SEM-EBSD system from the FEI Nova 230/IAC SEM. The
ods [10,11] of the mixed-grain structure have attracted extensive
electrolyte for electrochemical polishing is 10% oxalic acid solution,
research interest over the past few decades. Two common pro-
the voltage is 2 V, and the electrolytic time is 5 min at a room tem-
cesses to obtain fine grains without changing the chemical compo-
perature. The electrochemical polishing is a suitable method for
sition are the phase transformation and recrystallization,
austenitic stainless steel to remove the layers strained by mechan-
respectively [12,13]. However, the 316LN austenitic stainless steel
ical polishing and grinding. The results show that, for specimens of
is a single-phase material and no phase transformation takes place
316LN steel, the resolution rates of EBSD tests can reach more than
during the conventionally thermo-mechanical treatment [14].
95% when the surface layer of 15 lm in thickness was removed by
Consequently, the recrystallization is the only way for the grain
the electropolishing process.
refinement of this steel in the practical engineering. In our previous
work, the refinement of the mixed-grain structure with MCGs was
studied through isothermal compression experiments, i.e., through 2.2. Experimental methods
dynamic recrystallization (DRX) [15]. It was concluded that the
MCGs can be completely fragmented and refined under the defor- 2.2.1. Compression experiments at room temperature
mation condition of temperature of 1200 °C and reduction rate of The compression experiments were conducted at room temper-
50%. However, this method is not suitable for the renovation of ature (RT) and strain rate of 0.001 s 1 on a SANS CMT5305 (MTS)
forgings with coarse grains because of the limitation of allowed universal testing machine. To minimize the effect of the friction
deformation. between the die and the sample during compression tests, the sur-
In our recent work [16], it is observed in the room temperature faces of upper and lower die were polished. Fig. 1a shows the con-
condition that MCGs were preferentially deformed than fine grains figuration of compression tests, and Fig. 1b shows the typical
even if some of them are with hard orientations, and local shear mixed-grain structure with MCGs observed by an optical micro-
bands were found to be intensively formed inside MCGs. Since scopy (OM; Zeiss Axio Imager M2m). The reduction rates (RRs)
many literatures have reported that shear bands are privileged are designed as 25%, 15%, 10%, 8% and 6.5%. The RRs of compression
sites for the nucleation of new grains and provides driving force experiments were manually guaranteed by the comparison of
for recrystallization [17–19], it is inferred that the MCGs are possi- height measurements before and after the sample deformation.
bly refined if they receive more dislocation density than fine grains
due to the deformation heterogeneity, even in small deformation 2.2.2. In-situ SRX experiments
conditions. However, if the deformation happened in high temper- In order to investigate the SRX mechanism of 316LN steel with
ature environment, large part of dislocation can be annihilated due mixed-grain structure, in-situ experiments of the samples
to dynamic recovery. Therefore, we assume that the refinement of obtained by compression tests at RT were conducted on the
the MCGs may be achieved through a two-stage method: the first HTLSCM setup (VL2000DX-SVF17SP, Yonekura MFG Co., Japan).
stage is to deform the materials by a limited deformation at room Fig. 2 shows the schematic illustration of temperature history
temperature or a low temperature that can suppress the dynamic and the HTLSCM setup. The principle of using the thermal etching
recovery [20,21]; the second stage is rapidly heating the deformed method to display the grain boundary (GB) at high temperature of
materials to realize static recrystallization (SRX). This idea is aimed HTLSCM can be found in detail in Yogo et al. [22]. The reliability of
to take full advantages of the deformation energy to prompt the this method to reveal GBs has been proved by literature [23]. First,
recrystallization. the samples were heated to 200 °C at a rate of 50 °C/min, which is
To realize this idea, this study takes the nuclear grade 316LN equivalent to a drying treatment to the sample. Then, the samples
austenitic stainless steel with the MCG mixed-grain structure as were directly heated to the testing temperature 1250 °C at the fast-
the research object, and the SRX processing conditions were imple- est rate of 1000 °C/min, with the holding time for 10 min. This tem-
mented to refine the coarse grains. In order to find the minimum perature is determined for the convenience of observation because
deformation amount that is effective to refine the MCGs, tests with it was observed that the materials turned black when the holding
different deformation conditions were carried out. To probe the temperature is between 900 °C and 1200 °C. Also, the austenite
characteristics of microstructure evolution of mixed-grain struc- phase is stable at 1250 °C [14]. After holding, the samples were
ture, the SRX processes were in-situ observed by applying high quenched by nitrogen air to maintain the newly formed
temperature laser scanning confocal microscopy (HTLSCM). Trans- microstructure. During the holding time, the sample was pho-
2
Y. Li, P. Shen, H. Zhang et al. Materials & Design 210 (2021) 110057

Table 1
The chemical composition (wt. %) of the studied material.

Element C Cr Ni Mo Mn N Si P S Fe
wt. % 0.02 17.46 11.33 2.11 1.62 0.148 0.47 0.02 0.002 66.82

tographed at frequency of 1 Hz to record the microstructure evolu- even if some of them are with hard orientations, i.e., the grain size
tion. After in-situ SRX observations, the tested samples were exam- plays a more important role in strain partitioning than the crystal-
ined by EBSD measurements to evaluate the refinement of MCGs. lographic orientation when the difference of grain size is big. When
the deformation increased to some extent (e.g., 10% RR), obvious
2.2.3. TEM observations slip traces began to appear on the surfaces of FGs.
In order to explore the relationships between deformation con- It is known that annealing twins usually occur when heat treat-
dition and the SRX process, TEM by using JEM2100 transmission ments are conducted on strained metals with low stacking fault
electron microscope at the voltage of 200 kV were conducted to energies (e.g., the face-centered cubic metallic material). The role
observe the dislocation distribution. For the pre-deformed cuboid of the annealing twin playing in plastic deformation is usually
samples with MCG mixed-grain structures in the center, new cylin- thought to resemble that of a new grain, and the twin boundary
drical specimens with diameter of 3 mm were prepared by wire acts as a barrier against the dislocation motion [24]. An interesting
cutting, in which the center of each cylinder specimen coincides phenomenon is shown in Fig. 3e that the MCG was penetrated by
with the center of the corresponding sample. Thin slices were then an annealing twin band. However, the slip traces on the MCG sur-
prepared from the cylindrical specimens. The thin slices for TEM face were not blocked by the twin band, on the contrary, if regard-
observations were first mechanically polished down to 80 lm, fol- ing the twin band as a boundary, the slip traces on the upper and
lowed by twin-jet electropolishing using an electrolyte (main- lower surfaces of MCG have good consistencies both in number
tained at 35 °C) of 4% perchloric acid in ethanol, and finally ion and width, indicating that the internal annealing twin band has lit-
polished by Gatan PIPS II. tle effect on the strain partitioning of MCG.

3. Results 3.2. MCG refinements during in-situ SRX experiments with 25%, 15%
and 10% RRs
3.1. Surface morphologies after compression experiments at RT
After grinding and electrochemical polishing of each
After compression experiments at RT, the surface morphologies deformed sample, in-situ SRX observations were carried out.
of samples under different RRs are shown in Fig. 3, in which the The microstructure evolutions of thermal etched samples with
slip traces are clearly exhibited on the surfaces. It can be found that compression of 25%, 15% and 10% RRs are shown in video 7, 6
MCGs exhibits much more dense slip traces than the fine grains and 5, respectively, in which the samples exhibited surface slip
(FGs), inferring that the deformation is very inhomogeneous and traces in both MCGs and FGs under these deformation conditions
the MCGs receives much more strain than the FGs. When deforma- (see Fig. 3). Fig. 5 is a graphical illustration of these videos. Gen-
tion is small (e.g., 8% RR shown in Fig. 3d and 6.5% RR shown in erally, the SRX incubation time of samples was prolonged with
Fig. 3e), some FGs even have no obvious slip trace on their surfaces. the decrease of the RR. In the sample with 25% RR, it was
Fig. 4 shows a local morphology and the corresponding Schmid fac- observed that new GBs appeared in the interior of MCGs before
tor (SF) distribution of the sample underwent 8% compression. It heated to 1250 °C (see Fig. 5a1), while in the sample with 10%
can be found from Fig. 4a that significant surface slip traces were RR, new GBs appeared after being heated to 1250 °C for 60 s
observed in the MCG 1 and 2, while no slip trace in the clusters (see Fig. 5c2). Nevertheless, the SRX finish time of each sample
of fine grains 3 and 4 (see Fig. 4c and d). Note that the SFs of the was basically consistent. It indicates that the amount of plastic
MCG 1 and 2 are 0.43 and 0.39, respectively, both are lower than deformation plays a significant role in the incubation of SRX
the SFs of the clusters of fine grains 3 and 4. It indicates that at a but has less effect on the completion of SRX. Besides, the intra-
small RR, the plastic deformation is mainly undertaken by MCGs granular nucleation was observed in the sample with 25% RR as

Fig. 1. (a) Configuration of compression tests at room temperature, (b) typical


morphology of mixed-grain structure with MCGs in 316LN austenitic stainless steel. Fig. 2. Schematic illustration of the SRX process and the HTLSCM setup.

3
Y. Li, P. Shen, H. Zhang et al. Materials & Design 210 (2021) 110057

Fig. 3. Surface morphologies of each sample after compression tests under different RRs: (a) 25%, (b) 15%, (c) 10%, (d) 8%, (e) 6.5%.

shown in Fig. 5a1, while the nucleation at GBs was observed in eventually forming new GBs (see the microstructure after air
the sample with 10% RR as shown in Fig. 5c2. The GB migrations quenching in Fig. 6b1 and the corresponding all Euler map
were observed in recrystallized grains during the temperature obtained by EBSD measurements in Fig. 6c1). However, in video
holding, as shown in Fig. 5a3, b4 and c4. 2, it was found that the bulging motions from the MCG GBs
The observed different SRX nucleation mechanisms indicate joined forming a ‘collision front line’ and new GBs were formed
different SRX incubation processes. Therefore, the SRX incubation eventually (see the microstructure after air quenching in Fig. 6b2
of samples with 25% and 10% RRs were recorded in detail, and and the corresponding all Euler map obtained by EBSD measure-
the results are shown in video 1 (25% RR) and video 2 (10% ments in Fig. 6c2). These experimental phenomena together with
RR). Fig. 6 is a graphical illustration of the two videos. From the nucleation mechanisms in Fig. 5 reveal that the new grains
video 1, it was observed that the SRX process started when nucleate both at the MCG GBs and inside the MCG when the
the temperature rose to 1100 °C. The recrystallized area sample received a relatively large deformation (e.g., 25% RR),
expanded like a flood tide from the MCG GBs as well as from and grow up in the manner of GB bulging. However, if the sam-
the interior of MCG. This movement is actually driven by the ple received a relatively small deformation (e.g., 10% RR), the
GB bulging. If focusing only on the site marked in Fig. 6a1 and new grains nucleate only at the MCG GBs, and also grow up in
then examining video 1, two types of flood tide movements the manner of GB bulging. Additionally, some original MCG
(i.e., GB bulging) can be found: one coming from the GBs on GBs in the sample with 10% RR before heating were observed
the upper right and the other coming from an inner-MCG site. as remained after the SRX process; while the original MCG GBs
These two movements joined forming a ‘collision front line’, is not remained in the sample with 25% RR. It indicates that

Fig. 4. The role of grain size in mixed-grain structure playing in the strain partitioning. (a) Surface morphology of the sample under 8% RR, (b) the corresponding Schmid
factor distribution along the compression direction of the undeformed sample, and (c), (d) are zoomed-in observations in (a).

4
Y. Li, P. Shen, H. Zhang et al. Materials & Design 210 (2021) 110057

Fig. 5. Thermal etched microstructure evolution of samples with deformation of (a) 25%, (b) 15% and (c) 10% RRs at the holding temperature of 1250 °C for 600 s. HT
represents the holding time.

the deformation energy at MCG GBs of the sample with 10% RR 3.3. MCG refinements during in-situ SRX experiments with 8% and 6.5%
is nonuniformly distributed. The MCG GBs with high deforma- RRs
tion energy cannot be remained due to the recrystallization
nucleation and growth after heating, while the MCG GBs with The microstructural evolutions of thermal etched samples with
low deformation energy will be remained because they are in compression of 8% and 6.5% RRs are then shown in in video 4 and 3,
a relatively stable state. respectively. Fig. 7 is a graphical illustration of the two videos.

Fig. 6. Graphical illustrations of video 1 and 2: (a) SRX incubations, (b) microstructures after air quenching and (c) all Euler maps (marked by original MCG GBs of samples
before SRX) of two air quenched samples obtained by EBSD measurements. White dash lines in Fig. 6c2 are the identified original MCG GBs. The sample with 25% RR: 1, The
sample with 10% RR: 2.

5
Y. Li, P. Shen, H. Zhang et al. Materials & Design 210 (2021) 110057

Fig. 7. Thermal etched microstructure evolution of samples with small deformation of (a) 8% RR and (b) 6.5% RR at the holding temperature of 1250 °C for 600 s. HT
represents the holding time. (a4) and (b4) are all Euler maps (marked by original MCG GBs of samples before SRX) of two air quenched samples obtained by EBSD
measurements. White dash lines in Fig. 7a4 and b4 are the identified original MCG GBs.

Under these small deformation conditions, many surface slip traces tion partitioning among MCGs. As shown in Fig. 8a2, white lines
appeared in the MCGs but not in surrounding FGs. As shown in represent the low-angle grain boundaries (LAGBs). Generally, the
Fig. 7a2 and b2, new GBs began to appear in the sample with 8% LAGBs stands for the misorientation angle between 2 and 15 deg,
RR after it was kept at 1250 °C for 70 s, while in sample with but for a clear display, the LAGBs shown by the EBSD result were
6.5% RR, new GBs appeared after 80 s at the same temperature. defined as misorientation between 3 and 10 deg. It can be seen that
One can that even under 8% RR deformation condition, the MCGs the LAGBs among MCGs of the sample under 6.5% RR was unevenly
were significantly refined, as shown in Fig. 7a3 and a4. This obser- distributed before heating. Due to the close relationship between
vation is valuable since many reports consistently showed that the the LAGBs and local deformation [26], the deformation partitioning
grain refinement of 316LN austenitic stainless steel by DRX among MCGs was then identified inhomogeneous. It can be found
requires a large amount of strain [15,25]. Here, due to its large that the deformation partitioning of this incompletely refined MCG
grain size, the preferentially deformed MCGs were refined by SRX was lower, which also means that less deformation energy storage
method under a small deformation condition, thus providing a is available for SRX.
simple and feasible method to refine the mixed-grain structure
in engineering. Additionally, many original MCG GBs were
remained as shown in Fig. 7a4 and b4. 4. Discussion
It is noted that an MCG still existed in the sample under 6.5% RR
after heating, as shown in Fig. 7b3. This ‘new’ MCG is inside the 4.1. Refining assessments of MCGs under different deformation
‘old’ MCG. This phenomenon prompted us to further compare the conditions
EBSD measurements in this region before and after heating, and
the results are presented in Fig. 8. The ‘new’ MCG has the same In this work, three indexes are used to evaluate the MCG refine-
Euler angles as the ‘old’ MCG, i.e., (153, 37, 45 /deg), revealing that ment in material with the mixed-grain structure: the average grain
this MCG is actually a portion of the original MCG and the original size dm, the maximum grain size dmax and the grain size uneven
MCG was partially refined. It indicates that the refinement of the factor Z0. Note that to reduce the fluctuation of dmax among differ-
MCG is not significant under 6.5% RR. The incomplete refinement ent samples, its value is defined as an average of three MCGs with
can be interpreted by the difference in the amount of the deforma- largest grain sizes in the visual field. The average grain size dm is

Fig. 8. Optical observations and all Euler angle maps obtained by EBSD measurements of the sample with 6.5% RR (a) before and (b) after SRX process.

6
Y. Li, P. Shen, H. Zhang et al. Materials & Design 210 (2021) 110057

determined by the linear intercept method [27]. The grain size These slip bands are parallel to each other, indicating that MCGs
uneven factor Z0 is defined as the ratio of dmax to d0, where d0 is undertake the deformation in a primary slip pattern. Different from
the grain size with the highest occurrence rate, and its value is the samples under larger deformation conditions, the sample with
determined by the EBSD measurements. The results of refining 6.5% RR presented only the slip band morphology without disloca-
assessments are presented in Fig. 9. Fig. 9 also presents the corre- tion tangle or even scattered dislocation lines. For the sample with
sponding reference for each index of the uniform FG sample which 8% RR shown in Fig. 10b, an obvious slip band morphology was also
is from a well-fined studied material. It can be seen that the observed, and it is expectable to find that the slip bands are both
indexes dm and dmax all increase with the decrease of deformation, larger in number and width than those of the sample with 6.5%
revealing that the larger the deformation, the better the refinement RR. Moreover, scattered dislocation lines between the parallel slip
of the mixed-grain structure in SRX process. However, except for bands were observed in the sample under 8% RR. It indicates that
the sample under 6.5% RR, the grain size uneven factors Z0 of other the increase of deformation not only leads to the increase of the
samples have little difference, and they are all close to that of the overall dislocation density, but also enriches the dislocation struc-
uniform FG material, indicating that the discussed deformation tures. For the sample under 10% RR shown in Fig. 10c, the number
conditions except 6.5% RR can effectively eliminate the mixed- and width of parallel slip bands became larger. Apart from the scat-
grain structure and improve the homogeneity of grain size of the tered dislocation lines between parallel slip bands, the dislocation
material. One can find that d0 does not decrease but increase a little tangles were also observed. Fig. 11 presents the TEM image in the
after the recrystallization process. Our previous work studied grain GB region of the sample with 10% RR. The selected-area electron
growth behavior of 316LN steel, and it was concluded that the diffraction (SAED) spot patterns (see Fig. 11a and b) corresponding
grain size becomes coarser severely with the increase of tempera- to the regions in the circles in Fig. 11c reveal the different lattice
ture from 1150 to 1250 °C [15]. Therefore, it is speculated that the structures of the upper and the lower parts, confirming the exis-
increase of d0 is due to the grain growth because the heating tem- tence of GB. One can find the striking dislocation pile-ups in the
perature is high enough (1250 °C) even the holding time is 10 min. MCG GB region. In addition, the slip band moving to the GB also
became the privileged site for dislocation pile-ups. For the sample
4.2. Relationships between dislocation distributions and SRX with 15% RR as shown in Fig. 10d, slip band tangles were observed,
mechanisms indicating the activation of multiple slip systems. When the defor-
mation condition came to 25% RR, a new dislocation structure, i.e.,
Since the holding temperature of the SRX process is the same, dislocation cells, were distinctly observed as shown in Fig. 10e and
the different SRX behaviors as well as the different refining results f. These cell structures with the size range of 0.5 lm to 1.4 lm are
of the mixed-grain structures can be attributed to the dislocation formed based on the tangled slip bands. The phenomena, that only
distributions under different deformation conditions. Fig. 10 shows parallel slip bands were observed in MCGs under small deforma-
the TEM images inside the MCG. As shown in Fig. 10a, the typical tion conditions while slip band tangles and dislocation cells in
slip band morphology was observed in the sample with 6.5% RR. MCGs under large deformation conditions, may be connected with

Fig. 9. Refining assessments of MCGs under different deformation conditions as well as the corresponding reference for each index of the material with the uniform fine
grains (FGs): (a) dm, (b) dmax and (c) Z0. (d) The table lists the grain size with the highest occurrence rate d0 of each sample determined by the EBSD measurements.

7
Y. Li, P. Shen, H. Zhang et al. Materials & Design 210 (2021) 110057

Fig. 10. TEM images of MCGs in the deformed samples with (a) 6.5%, (b) 8%, (c) 10%, (d) 15% and (e, f) 25% RRs.

the nitrogen element. As presented in Table 1, the studied material bands easily move to the MCG GBs, contributing to dislocation
is 316LN austenitic stainless steel with 14 N (0.14 wt%). The addi- pile-ups at GBs. Although the dislocation density at the local GBs
tion of nitrogen is to enhance the creep resistance, corrosion resis- has met the nucleation condition for SRX under a small deforma-
tance, tensile strength, ductility and fatigue strength [3,4,28]. tion condition (e.g., 6.5% RR), the number of recrystallization nucle-
According to the work of Masumura et al. [29] and Ojima et al. ation at MCG GBs is limited, and the driving force for the growth of
[30], the nitrogen also plays an important role in dislocation nucleated grains is insufficient, which leads to the partial refine-
behavior. Dislocations are aligned on the same slip plane, because ment of MCG, i.e., the incomplete SRX during heating process.
cross slipping is remarkably suppressed, especially at low strains. It As deformation progresses, the MCG continues to undertake the
explains why the slip band tangles and the formation of dislocation main deformation, and the dislocation density at MCG GBs
cells were delayed to higher strains. increases gradually. Under a suitable small deformation condition
The dislocation distributions and various structures can con- (e.g., 8% RR or 10% RR), the dislocation density at MCG GBs satisfies
tribute to discuss the SRX mechanisms of samples under different the entire refinement condition for MCG. At this case (see video 2),
deformation conditions. As shown in Fig. 12, the SRX mechanism the number of the recrystallization nucleation at MCG GBs and the
with a limited deformation condition can be described as follows. driving force for the grain growth are adequate. The nucleated
The plastic deformation partitioning in the MCG mixed-grain grains grow up through GB bulging, and the SRX is basically com-
structure is dominated by grain size. At the early deformation pleted when the GB bulging is in a collision with the other. The GBs
stage, due to its abnormally large grain size, the MCGs in the will migrate locally during heating process. And eventually a stable
mixed-grain structures are more inclined to deform than the sur- fine microstructure is formed.
rounding fine grains. The MCG undertakes the plastic deformation Under a relatively large deformation condition, the MCG under-
in a primary slip pattern, leading to the formation of parallel slip takes the plastic deformation in a multi-slip pattern. Apart from
bands. Due to the absence of slip band tangles, these parallel slip the increase of the dislocation density at MCG GBs, a new disloca-

Fig. 11. (a, b) SAED spot patterns corresponding to the region in the circles in (c). (c) TEM image in the GB region of the sample with 10% RR.

8
Y. Li, P. Shen, H. Zhang et al. Materials & Design 210 (2021) 110057

Fig. 12. Schematic of the fragment and refinement of the MCG in mixed-grain structures based on the relationships between dislocation distributions and SRX mechanisms.

tion structure, i.e., the dislocation cell-structure, is formed in the (3) Under a relatively large deformation condition, slip band
interior of the MCG. These dislocation cells based on the slip band tangles were observed inside MCGs, indicating that the MCGs
tangles or dislocation walls [31], are excellent sites for recrystal- undertook the deformation in a multi-slip pattern. The tangled slip
lization nucleation [32,33]. Both the nucleated grains at MCG GBs bands in MCGs contributed to the formation of intragranular dislo-
and the intragranular-nucleated grains grow up in the manner of cation cells. Therefore, the nucleation of SRX occurred both at the
GB bulging, leading to a better refinement of MCG. It should be MCG GBs and in the interior of MCGs, and the nucleated grain also
noted that although we use a description of a relatively large defor- grew up in the manner of GB bulging. The MCGs were refined as
mation condition, the compression deformation, however, is only well at this deformation condition.
25% RR. Many studies have reported that the deformation condi-
tion of the DRX incubation for austenitic stainless steels is in the Declaration of Competing Interest
true strain range of 15% to 20% (i.e., 14% to 18% RR of compression
deformation) or even higher [25,34]. Therefore, the advantage to The authors declare that they have no known competing finan-
refine the mixed-grain structure with MCGs using the two-stage cial interests or personal relationships that could have appeared
method is emphasized here. More importantly, the refinement pro- to influence the work reported in this paper.
cess based on the limited deformation condition is easier to be
applied in engineering. Acknowledgements

The authors gratefully acknowledge the funding support from


the National Natural Science Foundation of China with the project
5. Conclusions of Nos. 51675335 and 52075329, as well as the support from
Shanghai Rising-Star Program (20QA1405300).
A two-stage grain refinement method, i.e., a limited cold defor-
mation followed by rapidly heating process, was applied to refine
the mixed-grain structure emerged in the nuclear grade 316LN Appendix A. Supplementary data
austenitic stainless steel. The SRX behaviors of samples under dif-
ferent cold deformation conditions were studied by in-situ obser- Supplementary data to this article can be found online at
vation technique. The conclusions can be summarized as follows: https://doi.org/10.1016/j.matdes.2021.110057.
(1) The MCGs in the mixed-grain structure were preferentially
deformed than fine grains even if MCGs are with hard orientations, References
i.e., the grain size plays a more significant role in strain partitioning
[1] X.W. Duan, J.S. Liu, Research on damage evolution and damage model of 316LN
than the crystallographic orientation. Since the MCGs preferen- steel during forging, Mater. Sci. Eng. A 588 (2013) 265–271.
tially and continuously undertook the main deformation, the accu- [2] H.C. Wu, B. Yang, S.L. Wang, M.X. Zhang, Effect of oxidation behavior on the
mulated dislocation density in MCGs can meet the deformation corrosion fatigue crack initiation and propagation of 316LN austenitic stainless
steel in high temperature water, Mater. Sci. Eng. A 633 (2015) 176–183.
energy requirement of the SRX process at a limited deformation [3] J.E. Pawel, A.F. Rowcliffe, D.J. Alexander, M.L. Grossbeck, K. Shiba, Effects of low
condition. It was proved that the compression with only 8% reduc- temperature neutron irradiation on deformation behavior of austenitic
tion rate followed by rapidly heating process (1250 °C  360 s) can stainless steels, J. Nucl. Mater. 233-237 (1996) 202–206.
[4] G.V. Prasad Reddy, R. Sandhya, S. Sankaran, P. Parameswaran, K. Laha, Creep–
completely refine the MCGs in the mixed-grain structure. fatigue interaction behavior of 316LN austenitic stainless steel with varying
(2) Under a small deformation condition, the observed parallel nitrogen content, Mater. Des. 88 (2015) 972–982.
slip bands inside MCGs indicates that the MCGs undertook the [5] S. Wang, B. Yang, M. Zhang, H. Wu, J. Peng, Y. Gao, Numerical simulation and
experimental verification of microstructure evolution in large forged pipe used
plastic deformation in a primary slip pattern. The slip bands easily
for AP1000 nuclear power plants, Ann. Nucl. Energy 87 (2016) 176–185.
moved to the MCG GBs due to the lack of slip band tangles, leading [6] X. Shang, H. Zhang, Z. Cui, M.W. Fu, J. Shao, A multiscale investigation into the
to the striking dislocation pile-ups at GBs. Therefore, the nucle- effect of grain size on void evolution and ductile fracture: Experiments and
ation of SRX preferentially occurred at the MCG GBs, and the nucle- crystal plasticity modeling, Int. J. Plast. 125 (2020) 133–149.
[7] M.B. Lezaack, A. Simar, Avoiding abnormal grain growth in thick 7XXX
ated grains grew up in the manner of GB bulging, finally leading to aluminium alloy friction stir welds during T6 post heat treatments, Mater. Sci.
the refinement of MCGs. Eng. A 807 (2021) 140901, https://doi.org/10.1016/j.msea.2021.140901.

9
Y. Li, P. Shen, H. Zhang et al. Materials & Design 210 (2021) 110057

[8] S.G. Kim, Y.B. Park, Grain boundary segregation, solute drag and abnormal [21] K. Huang, R.E. Logé, A review of dynamic recrystallization phenomena in
grain growth, Acta Mater. 56 (15) (2008) 3739–3753. metallic materials, Mater. Des. 111 (2016) 548–574.
[9] X. Wang, Z. Huang, B. Cai, N. Zhou, O. Magdysyuk, Y. Gao, S. Srivatsa, L. Tan, L. [22] Y. Yogo, K. Tanaka, K. Nakanishi, In-situ observation of grain growth of steel at
Jiang, Formation mechanism of abnormally large grains in a polycrystalline high temperature, Mater. Trans. 50 (2009) 280–285.
nickel-based superalloy during heat treatment processing, Acta Mater. 168 [23] C. Garcı́a de Andrés, F.G. Caballero, C. Capdevila, D. San Martı́n, San Martı´n,
(2019) 287–298. Revealing austenite grain boundaries by thermal etching: advantages and
[10] M.-S. Chen, Z.-H. Zou, Y.C. Lin, H.-B. Li, W.-Q. Yuan, Effects of annealing disadvantages, Mater. Charac. 49 (2) (2002) 121–127.
parameters on microstructural evolution of a typical nickel-based superalloy [24] S. Miura, Y. Saeki, Plastic deformation of copper crystals containing a twin
during annealing treatment, Mater. Charac. 141 (2018) 212–222. band, Transactions of the Japan Institute of Metals 17 (1976) 253–260.
[11] G.-Q. Wang, M.-S. Chen, H.-B. Li, Y.C. Lin, W.-D. Zeng, Y.-Y. Ma, Methods and [25] F. Chen, K.e. Qi, Z. Cui, X. Lai, Modeling the dynamic recrystallization in
mechanisms for uniformly refining deformed mixed and coarse grains inside a austenitic stainless steel using cellular automaton method, Comput. Mater. Sci.
solution-treated Ni-based superalloy by two-stage heat treatment, J. Mater. 83 (2014) 331–340.
Sci. Technol. 77 (2021) 47–57. [26] D. Zhu, M. Zhang, Y. Wang, Electron Backscattered Diffraction Study of
[12] R.D. Doherty, D.A. Hughes, F.J. Humphreys, J.J. Jonas, D.J. Jensen, M.E. Kassner, Microstructural Evolution During Isothermal Deformation of High-N
W.E. King, T.R. McNelley, H.J. McQueen, A.D. Rollett, Current issues in Mn18Cr18 Alloy, Metall. Mater. Trans. B 50 (4) (2019) 1662–1673.
recrystallization: a review, Mater. Sci. Eng. A 238 (2) (1997) 219–274. [27] H. Engqvist, B. Uhrenius, Determination of the average grain size of cemented
[13] A.N. Bucsek, L. Casalena, D.C. Pagan, P.P. Paul, Y. Chumlyakov, M.J. Mills, A.P. carbides, International Journal of Refractory Metals and Hard Materials 21 (1-
Stebner, Three-dimensional in situ characterization of phase transformation 2) (2003) 31–35.
induced austenite grain refinement in nickel-titanium, Scr. Mater. 162 (2019) [28] J. Ganesh Kumar, M. Chowdary, V. Ganesan, R.K. Paretkar, K. Bhanu Sankara
361–366. Rao, M.D. Mathew, High temperature design curves for high nitrogen grades of
[14] S. Wang, M. Zhang, H. Wu, B. Yang, Study on the dynamic recrystallization 316LN stainless steel, Nucl. Eng. Des. 240 (6) (2010) 1363–1370.
model and mechanism of nuclear grade 316LN austenitic stainless steel, [29] T. Masumura, Y. Seto, T. Tsuchiyama, K. Kimura, Work-Hardening Mechanism
Mater. Charac. 118 (2016) 92–101. in High-Nitrogen Austenitic Stainless Steel, Mater. Trans. 61 (4) (2020) 678–
[15] D.-S. Sui, H.-M. Zhang, H.-Y. Zhu, Z. Zhu, Z.-S. Cui, Effects of process parameters 684.
on fragment and refinement of millimeter-grade coarse grains for 316LN steel [30] M. Ojima, Y. Adachi, Y. Tomota, K. Ikeda, T. Kamiyama, Y. Katada, Work
during hot cogging, J. Iron Steel Res. Int. 24 (5) (2017) 529–535. hardening mechanism in high nitrogen austenitic steel studied by in situ
[16] Y.Q. Li, H.M. Zhang, M.X. Liu, Z.S. Cui, Study of the Influence of the Mixed-Grain neutron diffraction and in situ electron backscattering diffraction, Mater. Sci.
Structure on the Microscopic Deformation Behavior of 316LN Steel, in: Glenn Eng. A 527 (1-2) (2009) 16–24.
Daehn et al. (Eds.), Forming the Future:Proceedings of the 13th International [31] H. Mughrabi, Dislocation wall and cell structures and long-range internal
Conference on the Technology of Plasticity, Springer, 2021, pp. 1757–1766. stresses in deformed metal crystals, Acta Metall. 31 (9) (1983) 1367–1379.
[17] F. Ebrahimi, Q. Zhai, D. Kong, Deformation and fracture of electrodeposited [32] R.K. Ray, W.B. Hutchinson, B.J. Duggan, A study of the nucleation of
copper, Scr. Mater. 39 (3) (1998) 315–321. recrystallization using HVEM, Acta Metall. 23 (7) (1975) 831–840.
[18] A.T. Lim, M. Haataja, W. Cai, D.J. Srolovitz, Stress-driven migration of simple [33] B. Bay, N. Hansen, Initial stages of recrystallization in aluminum of commercial
low-angle mixed grain boundaries, Acta Mater. 60 (3) (2012) 1395–1407. purity, Metall. Mater. Trans. A 10 (3) (1979) 279–288.
[19] Y. Gu, Y. Xiang, D.J. Srolovitz, J.A. El-Awady, Self-healing of low angle grain [34] K.A. Babu, S. Mandal, C.N. Athreya, B. Shakthipriya, V.S. Sarma, Hot
boundaries by vacancy diffusion and dislocation climb, Scr. Mater. 155 (2018) deformation characteristics and processing map of a phosphorous modified
155–159. super austenitic stainless steel, Mater. Des. 115 (2017) 262–275.
[20] F. Humphrey, M. Hatherly, Recrystallisation and Related Annealing
Phenomena, Elsevier Ltd, Oxford, 2004.

10

You might also like