You are on page 1of 31

Version of Record: https://www.sciencedirect.

com/science/article/pii/S0921509318302053
Manuscript_af0cbe9a1325d407df780ee8b66ba558

The effect of annealing on the microstructural evolution and mechanical

properties in phase reversed 316LN austenitic stainless steel


D.M. Xu1, G.Q. Li1,2, X.L. Wan1,2*, R.D.K. Misra3, X.G. Zhang4, G. Xu1, K.M. Wu1

1
The State Key Laboratory of Refractories and Metallurgy, Wuhan University of
Science and Technology, Wuhan 430081, China
2
Key Laboratory for Ferrous Metallurgy and Resources Utilization of Ministry of
Education, Wuhan University of Science and Technology, Wuhan 430081, China
3
Laboratory for Excellence in Advanced Steel Research, Department of Metallurgical,
Materials, and Biomedical Engineering, University of Texas at El Paso, El Paso, TX
79968, USA
4
Institute for Materials Research, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai
980-8577, Japan
Corresponding author: X.L. Wan
Email: wanxiangliang@wust.edu.cn
Abstract
The present study aims to investigate the effect of annealing on microstructural
evolution and mechanical properties in phase reversion-induced ultrafine/fine-grained
316LN austenitic stainless steel. The commercial 316LN austenitic stainless steel was
cold rolled at room temperature to 90% thickness reduction and subsequently
annealed in the temperature range of 600-1000 °C for 1-100 minutes. Evolution of
phases in selected samples was identified and quantified by X-ray diffraction together
with the corresponding microstructural characterization through optical, scanning and
transmission electron microscopy, and electron backscattered diffraction. Mechanical
properties of selected samples were determined by the tensile test. The results
indicated that 46% α′-martensite and 54% deformed untransformed austenite were
obtained in 316LN austenitic stainless steel after 90% cold reduction. Ultrafine/fine
austenite grains nucleated at α′-martensite and deformed untransformed austenite via
nucleation and growth process on annealing. The average grain size increased
gradually with increased annealing temperature and time, with consequent decrease in
yield strength and increased elongation.
Keywords: Austenitic stainless steel; Phase reversion; Recrystallization; Grain size;
Mechanical property.

© 2018 published by Elsevier. This manuscript is made available under the Elsevier user license
https://www.elsevier.com/open-access/userlicense/1.0/
1. Introduction

The development of high strength steels characterized by high strength-high

ductility combination continues to be a major effort worldwide. The austenitic (γ)

stainless steels exhibit excellent ductility. However, the yield strength of austenitic

stainless steels is quite low because of coarse and soft γ-phase. Grain size refinement

is an effective method to increase the yield strength without significantly impairing

the ductility in austenitic stainless steels. The concept of phase reversion annealing

developed by Misra’s group in recent years [1-3], involving severe deformation (~ 45

- 75%) of metastable austenite to generate strain-induced martensite, followed by

annealing, was used to obtain nanograined/ultrafine-grained (NG/UFG) austenitic

stainless steels. Previous studies indicated that high yield strength (880 - 1000 MPa)

and high elongation (47 - 38%) combination was obtained in 301LN [3], 316LN [4],

and 204Cu [5] austenitic stainless steels, when the austenite grains were refined to

NG/UFG structure.

In this approach, different combination of strength and ductility can be achieved,

based on the grain size of austenitic stainless steels [3]. The grain size of reversed

austenite is a function of temperature-time annealing sequence. The influence of

annealing treatment on the formation of reversed austenite in austenitic stainless steels

with fully strain-induced martensite structure has been previously studied [2, 6-8].

However, only 46 - 67% of γ-phase in 316 austenitic stainless steels transformed into

strain-induced martensite after 80 - 90% cold-rolled (CR) reduction [4, 9-11]. The

phase reversion behavior in CR 316 steels should be complex due to the mixed

2
microstructures of deformed untransformed austenite and strain-induced martensite.

Furthermore, there is currently no reported study that explores microstructural

evolution during annealing of CR austenite.

Thus, the objective of the study described here is to study the effect of annealing

treatment on microstructural evolution and accompanying mechanical properties in

phase reversion annealed 316LN austenitic stainless steel containing strain-induced

martensite and deformed untransformed austenite.

2. Experimental procedure

The experimental material used in this study was a commercial 316LN austenitic

stainless steel of ~3 mm thickness and the chemical composition (in weight percent)

of steel is listed in Table 1. The strips were CR in a pilot plant to 90% reduction in

thickness at room temperature. Next, the strips were isothermally annealed in the

temperature range of 600 - 1000°C for 1 - 100 minutes in a tubular resistance furnace

under argon atmosphere, and then subsequently quenched in ice-water. The objective

of changing temperature-time annealing treatment was to study the microstructural

evolution and obtain samples with different grain size from nanograined (NG) to

coarse-grained (CG) regime. The development of strain-induced martensite in CR and

annealed specimens was evaluated using an optical microscope (OM, Olympus). The

samples for optical microscopy were etched in a solution, which consisted of 0.20 g

sodium-metabisulfate in 100 mL distilled water and 10 mL hydrochloric acid in 100

mL distilled water. The etching time was varied from 2 to 5 minutes, depending on

the fraction of martensite in each specimen, to enable martensite phase to be

3
adequately revealed as a dark phase. The annealed specimens for scanning electron

microscope (SEM, Nova400Nano, operated at 18 kV) studies were mechanically

polished and then electrochemically etched with 60% nitric acid solution. Some

selected annealed specimens were examined via TEM (JEM-2100) operated at 200 kV.

Thin foils were prepared by twin-jet electropolishing of 3 mm disks, punched from

the specimens, using a solution of 10% perchloric acid in acetic acid as electrolyte.

The mean grain size in annealed specimens was calculated from the measurements of

100 equiaxed grains with complete boundary. The size of each grain was measured by

the method based on average long and short axes of grains in SEM/TEM micrographs.

EBSD studies were carried out at a step size of 50 nm or 200 nm to evaluate the

microstructure of annealed samples. The samples were electrochemically etched with

20% perchloric acid alcohol solution operated at 25 °C with an applied potential of 15

V. The content of martensite and austenite were detected by the means of X-ray

diffraction (XRD, X’Pert Pro MPD). Spectra between 10 ° and 90 ° (2θ) were

recorded at scanning rate of 2 ° per minute using Cu Kα radiation (λ=0.1542 nm).

The annealed strips were machined to make tensile samples with a profile of 140

× 20 mm and 65 mm gage length. The uniaxial tensile tests were carried out at room

temperature at an engineering strain rate of 5×10-4 s-1. To study the deformed

microstructure as a function of strain, the tensile tests were interrupted at selected

engineering strain of 0.1. The area close to the highly stressed region was used for

TEM and EBSD observations. Meanwhile, in the sample with engineering strain less

than 0.1, the area close to the tensile fracture was selected for TEM observations.

4
3. Results

3.1. Microstructure and mechanical properties of commercial and CR 316LN steel

The optical micrographs illustrating the microstructure of commercial and CR

316LN steel are presented in Fig. 1a and 1b. Fig. 1a showed that the commercial

316LN steel had a coarse-grained austenite structure. The metastable γ-phase

transformed to strain-induced martensite during CR deformation. The microstructure

of CR 316LN steel with 90% reduction (Fig. 1b) consisted of deformed

untransformed austenite and strain-induced martensite. Figs. 1c and 1d suggested that

the strain-induced martensite was a combination of lath-type (Fig. 1c) and dislocation

cell-type (Fig. 1d) martensite, which are considered as effective sites for nucleation of

reversed austenite during annealing [6]. As previously described [4], the volume

fraction of α′-martensite in CR 316LN steel measured by XRD analysis was ~46%.

The engineering stress-engineering strain plots of samples are presented in Fig. 1e.

The figure shows that the commercial 316LN steel had low yield strength (~281 MPa)

and high elongation (~52%). The yield strength of CR 316LN steel was very high at

~1721 MPa, but the elongation was only ~1.3%.

3.2 Microstructural evolution of CR 316LN steel during annealing

A series of metallographic images illustrating the microstructural evolution of CR

samples annealed at 600°C for various times are presented in Fig. 2. The optical

micrograph of sample annealed for 2 minutes (Fig. 2a) showed that the microstructure

consisted of deformed untransformed austenite and α′-martensite. Some ultrafine dark

spots were present. These dark spots increased with increased annealing time to 10

5
minutes (indicated in the white circle in Fig. 2b). TEM observations revealed that the

reversed austenite grains nucleated at martensite. Fig. 2c showed that reversed

austenite with lath-type morphology existed near the ά-martensite structure, and the

reversed austenite grains with equiaxed morphology nucleated at the martensite

boundaries (Fig. 2d). The size of reversed austenite grains increased slightly when the

CR sample was annealed for 50 minutes, as shown in Fig. 2e. Furthermore, the

deformed untransformed prior austenite grains did not change on annealing for 100

minutes (Fig. 2f).

Similar to the sample annealed at 600°C, the microstructure of the samples

annealed at 700°C for 1 - 100 minutes are shown in Fig. 3. Some dark spots were

observed in Fig. 3a, which were caused by phase reversion transformation. TEM

micrographs revealed that the reversed austenite grain nucleated on martensite and

coarsened with increased annealing time to 5 minutes (Figs. 3b and 3c). Annealing

twins were also observed (Fig. 3c). With increased annealing time to 10 minutes,

some small grains formed in the interior of the deformed untransformed austenite,

because of recrystallization (Fig. 3d). When the time was increased, the size of newly

formed austenite grains increased to submicron-scale/micron-scale (Fig. 3e). Majority

of small austenite grains formed in the interior of deformed untransformed austenite

in Fig. 3f.

When the annealing temperature was increased to 800°C, nano-scale and

submicron-scale reversed austenite grains formed during the early stages (Fig. 4b).

With increased annealing time, a large number of grain boundaries were present in the

6
deformed untransformed austenite, which appeared as boundaries of small

recrystallized grains (Figs. 4c and 4d). The size of austenite grains associated with the

recrystallization of deformed untransformed austenite was at micrometer-scale. They

were obviously larger than NG/UFG phase reversed austenite grains formed from

ά-martensite. The austenite structure with bimodal grain size distribution was noted in

Figs. 4c and 4d. With increased annealing time to 50 and 100 minutes, the

microstructure was characterized by a large number of uniform defect-free equiaxed

austenite grains (Figs. 4e and 4f).

As shown in Fig. 5, the microstructure of sample annealed at 900°C was

significantly different from the samples annealed at lower temperatures. This apparent

difference is because the recrystallization (Fig. 5a) and reversion transformation (Fig.

5b) occurred during the early stage of annealing. Almost complete martensite

transformed into reversed austenite with annealing time up to 2 minutes and the

microstructure consisted of equiaxed austenite grains (Figs. 5c and 5d). The size of

austenite grains in samples annealed for 10 minutes were micrometer-scale (Fig. 5e)

and increased slightly with increased annealing time to 100 minutes (Fig. 5f).

Fig. 6 is the microstructure of samples annealed at 1000°C for varying times.

Martensite completely transformed into reversed austenite during the early stage of

annealing (Figs. 6a and 6b) and the microstructure consisted of predominantly

equiaxed austenite grains, whose size increased with increased annealing time (Figs.

6c-6f).

The crystallographic information of grain boundaries for selected samples, which

7
is important for studying the microstructural evolution during annealing, was

analyzed by EBSD. Figs. 7 and 8 are the EBSD analysis maps. The white and black

microstructure is austenite and martensite, respectively. Both martensite and austenite

were observed in sample annealed at 600°C for 5 minutes. Some austenite grains (as

shown in the red circle in Fig. 7a) were present in the interior of martensite, which

appeared to be reversed austenite. With increased annealing temperature and time, the

area fraction of martensite was decreased, and completely austenite grain structure

was obtained when the temperature was greater than 800°C (Fig. 7e). Abundant of

annealing twins were observed in austenite (marked by arrows in Fig. 8b and 8f).

Furthermore, the red, green and blue curves in Figs. 7 and 8 represent grain

boundaries with angle between 2 - 5°, 5 - 15° and 15 - 65° respectively, and the

percentage fraction and density of grain boundary with different misorientation range

of selected samples were measured and are presented in Table 2. It revealed that the

percentage fraction of high angle grain boundary increased in general and the number

density of grain boundary with different angle was decreased with increased annealing

temperature and time, respectively.

Fig. 9a shows the X-ray diffraction patterns for CR 316LN sample and samples

annealed at 900°C for 1 - 3 minutes. The CR 316LN sample had ά-martensite and

austenite peaks. As the annealing time was increased to 1 and 2 minutes, the austenite

peaks (γ(111), γ(200) and γ(220)) increased in intensity with respect to the martensite

peaks (α(110) and α (211)), implying an increase in the volume fraction of austenite.

When the annealing time was 3 minutes, martensite peaks disappeared, indicating that

8
100% austenite was obtained. The volume fraction of ά-martensite in samples with

annealing treatment is presented in Fig. 9b. It showed that 100% austenite was

obtained after 5, 3 and 1.5 minutes, when the annealing temperatures were 800°C,

900°C and 1000°C respectively. Figs. 9c and 9d are the mean grain size of samples

annealed at 600 - 1000°C for various time. The mean grain size increased with

increased annealing temperature and time.

3.3 Mechanical properties and deformation mechanisms of annealed 316LN steels

To study the influence of grain size on the mechanical property, the 316LN steel

annealed at 900°C and 1000°C for different times was subjected to tensile test. The

mean grain size and mechanical properties of samples as function of annealing

treatment are listed in Table 3. The mean grain size varied from NG structure to CG

structure. Part of α′-martensite continued to exist in the sample with NG structure

(Table 3). Table 3 also shows that various combination of yield strength and

elongation are obtained. The yield strength and elongation vs grain size plots in Fig.

10a revealed that the yield strength decreased from ~2155 MPa to ~265 MPa and

elongation increased from ~1.4 % to ~80.6 %, respectively when the grain size of

specimens gradually increased from the NG structure to the CG structure. Samples

with combination of very high yield strength-low elongation, high yield strength-high

elongation, and low yield strength-high elongation respectively were selected

(indicated in Fig. 10a) and the engineering stress-engineering strain curves plotted

from the tensile data (Fig. 10b). It is noted from Fig. 10b that the stress increased

rapidly with increased strain in NG steel (annealed at 900 °C for 1 minute) before

9
fracture. The curve of UFG steel (annealed at 900 °C for 2 minutes) with high yield

strength-high elongation combination in Fig. 10b revealed that the stress first

increased rapidly, followed by a distinct yield behavior and slow increase of stress,

with increasing strain. However, the tensile behavior of CG steel (annealed at 1000 °C

for 100 minutes) appeared to show a continuous yield and the stress increased slowly

with increasing strain until rupture, exhibiting low yield strength and high elongation.

The variation in the mechanical properties of 316LN steel is related to

microstructural evolution during tensile straining. The deformed microstructure of NG,

UFG and CG steels near the tensile fracture and with 0.1 strain are presented in Fig.

11, respectively. Fig. 11a shows the major deformation microstructure of NG steel in

the vicinity of fracture. The grain boundary of austenite was not clear and dislocations

were present. In some cases, α′-martensite with dislocations was observed and

confirmed by diffraction pattern in Fig. 11b. The deformed microstructure in UFG

316LN steel with 0.1 strain in Fig. 11c also revealed that dislocations were present in

the austenite grains. It was interesting that mechanical twins were observed in UFG

steel subjected to 0.1 strain (Figs. 11c and 11d). In contrast to UFG structure, shear

bands and strain-induced martensite laths were observed in CG steel with 0.1 strain,

as presented in Figs. 11e and 11f. The EBSD analysis was also used to observe the

characteristics of deformed microstructure for UFG and CG samples with 0.1 strain

and is presented in Fig. 12. Comparing the crystallographic characteristics in UFG

(Fig. 8b) and CG (Fig. 8f) samples before tensile straining, the EBSD maps of grain

angles between adjacent grains revealed a larger number of red and green curves were

10
obtained in both the samples with 0.1 strain (Fig. 12), implying numerous low angle

grain boundaries in UFG and CG samples during tensile straining. The percentage

fraction of low angle grain boundary in UFG and CG samples with 0.1 strain was

significantly greater than that before the tensile test, as shown in Table 4. Furthermore,

it also indicated that the number density of grain boundaries with low angle and high

angle in both the samples with 0.1 strain was more compared to prior to the tensile

test (Tables 2 and 4). The XRD analysis revealed that the content of martensite in the

CG sample increased to 9% with increase of strain to 0.1 (Tables 3 and 4), whereas,

the content of martensite in UFG samples with 0.1 strain was less than 5%, meaning

that the extent of strain-induced martensite transformation was negligible [12].

4. Discussion

4.1. Effect of austenite stability on strain-induced martensite formation in 316LN steel

Before discussing the effect of annealing on the microstructural evolution and

mechanical properties in austenitic stainless steel, it was important to evaluate the

microstructural characteristics of CR sample. The metastable γ-phase transformed to

strain-induced martensite during CR deformation in austenitic stainless steel below

Md temperature. Previous work indicated that almost complete austenite transformed

into strain-induced martensite in 301 and 304 austenitic stainless steels with 50 - 55%

CR reduction [2, 13]. However, only ~46% of metastable austenite in this 316LN

austenitic stainless steel transformed into ά-martensite even after ~90% CR reduction

in the present study. Comparing with 301 and 304 austenitic stainless steels, the

316LN austenitic stainless steel had high content of Ni, Cr and Mo, which effectively

11
increases the mechanical stability of austenite and inhibits the formation of

strain-induced martensite. The Md30 temperature (where 50% ά-martensite is present

after 30% tensile deformation) was used to represent the mechanical stability of

austenite in austenitic stainless steels. The Md30 temperature based on Nohara's

equation [14] was 26.6°C, 12.9°C and 7.6°C in commercial 301, 304, and 316LN

austenitic stainless steels, respectively [3, 4, 10, 15]. Thus, the strain-induced

martensite formation was inhibited and the volume fraction of ά-martensite was 46%

in CR 316LN steel.

4.2. Effect of annealing on the microstructural evolution in CR 316LN steel

It is generally accepted that ά-martensite in austenitic stainless steels would revert

back to austenite during the annealing process [8]. The new austenite grains primarily

nucleated at the martensite lath boundaries and grew though diffusional ά - γ

reversion mechanism, or ά-martensite transformed to austenite with lath-type

morphology by shear phase reversion mechanism [8]. The fraction of reversed

austenite gradually increased with increased annealing temperature and time [2].

Lastly, completely NG/UFG austenite structure was obtained in austenitic stainless

steels when ά-martensite was absent.

In our study, the microstructure of CR 316LN steel consisted partly of

ά-martensite and deformed untransformed austenite instead of fully ά-martensitic

structure (Fig. 1b). The ά-martensite in 316LN steel reverted back to austenite during

annealing at 600 - 1000°C, and when the annealing time exceeded to 5, 3 and 1.5

minutes at 800°C, 900°C and 1000°C respectively, 100% austenite was obtained (Fig.

12
9b). This implied that phase reversion was completed. Furthermore, the deformed

untransformed austenite also strongly affected the microstructural evolution during

the annealing process. The structure of deformed untransformed austenite did not

change even after 100 minutes, when the sample was annealed at 600°C (Fig. 2f). The

obvious variation in deformed untransformed austenite of the sample was observed at

700 - 800°C (Figs. 3d, 3f, 4c and 4d). Recrystallization occurred in the deformed

untransformed austenite, which was gradually replaced by defect-free austenite grains

(Figs. 4e and 4f). The size of defect-free austenite grains formed from the deformed

untransformed austenite by recrystallization and from ά-martensite by phase reversion

was micrometer-scale and nano-scale/submicron-scale, respectively (Figs. 4c and 4d).

Thus, austenite with bimodal grain size distribution was obtained via reversion

transformation at appropriate temperature-time annealing treatment [16]. NG/UFG

austenite grains with high grain boundary areal density and high interfacial energy are

thermodynamically unstable. Thus, austenite grains coarsened via grain boundary

mobility to reduce the grain boundary areal density. Meanwhile, the recrystallization

phenomenon occurred earlier and the austenite grains were uniform and equiaxed

when the sample was annealed at 900 - 1000°C (Figs. 5c and 6a). Then, austenite

grains continued to coarsen with increased annealing time (Figs. 9c and 9d).

4.3. Relationship of grain size and mechanical property of annealed 316LN steels

It is well-known that grain size is a major factor that determines mechanical

properties. As presented in Fig. 10a, the relationship of yield strength, elongation and

grain size indicated that the yield strength was decreased and elongation was

13
increased with increase in grain size from NG to CG structure. Thus, samples with

NG (annealed at 900 °C for 1 minute), UFG (annealed at 900 °C for 2 minutes) and

CG (annealed at 1000 °C for 100 minutes) structures were selected to analyze the

difference in mechanical properties.

The CG austenitic stainless steel exhibited a combination of low yield

strength-high elongation (Fig. 10b). It is well known that grain boundary has a

profound effect on the yield strength of steels leading to the well-known Hall-Petch

relationship, which describes the yield stress in terms of the grain size. The boundary

of annealing twin, similar to the grain boundary, is also considered as a barrier to

dislocation motion [17]. Thus, the low density of boundary of grain and twin didn’t

effectively restrict the movement of dislocations and caused low yield strength (~265

MPa) in CG sample. Numerous low and high angle grain boundaries formed in the

metastable austenite and some γ-phase transformed into strain-induced martensite

during tensile straining (Fig. 12b), which contributed to excellent elongation (TRIP

effect). Unlike the CG sample, the NG austenitic stainless steel exhibited a

combination of very high yield strength-low elongation. The microstructure of NG

steel consisted of nano-scale austenite grains and ~10% of ά-martensite lath structure.

The very high yield strength of NG sample is attributed to grain size effect governed

by Hall-Petch relationship. The lattice dislocation motion is blocked by high density

of grain boundaries in the NG sample (Table 2) [18]. Meanwhile, nano-scale

martensite laths played a role similar to that of grain boundaries and effectively

enhanced the yield strength in the NG sample. The elongation was low for the NG

14
sample. The nano-scale grains did not significantly work harden on deformation

because of very low dislocation storage capacity of the small grains [18].

The mechanical properties of UFG sample were intermediate between NG and CG

samples and were characterized by a combination of high yield strength-high

elongation (Fig. 10b). The TEM and EBSD observations revealed that the

microstructure was uniform with small equiaxed austenite grains and numerous high

angle grain boundaries (Figs. 5d and 8b). The UFG grains with high density of grain

boundaries exhibited high yield strength (~994 MPa), which was ~3.7 times higher

than the CG sample in accordance with the Hall-Petch relationship. The UFG sample

also exhibited high elongation (~40%), which is attributed to twinning-induced

plasticity (TWIP) effect (Figs. 11c and 11d). Both strain-induced martensite and

mechanical twinning are effective strain hardening mechanisms, which prevent strain

localization and thereby enhance elongation. The deformation mechanisms changed

from strain-induced martensite in the CG sample to deformation twinning in the UFG

sample because of differences in the stability of austenite. The stability of austenite

grain was not only affected by chemical composition but also grain size. The Md30

temperature based on Nohara's equation [14] indicative of mechanical stability of

austenite is presented in Table 3. It indicated that grain refinement decreased the Md30

temperature from ~13°C in CG sample to ~0°C in UFG steel, which inhibited

strain-induced martensite formation in metastable austenite. Thus, the lattice

displacement associated with strain was accommodated by twinning.

5. Conclusions

15
(1) The high mechanical stability of austenite in 316LN steels led to ~46% of

metastable austenite in 316LN steel transform into ά-martensite even at ~90% CR

reduction.

(2) The ultrafine/fine austenite grains nucleated at ά-martensite and deformed

untransformed austenite on phase reversion annealing. The density of grain

boundaries decreased and the mean grain size increased gradually with increased

annealing temperature and time, respectively.

(3) Annealed steels with combination of very high yield strength-low elongation,

high yield strength-high elongation, and low yield strength-high elongation were

obtained with NG, UFG and CG structures, respectively.

(4) The change in deformation mechanism from strain-induced martensite in CG

sample to mechanical twining in UFG sample is related to the increased stability of

austenite with decrease in grain size.

Acknowledgements

Authors gratefully acknowledge the financial support from National Natural

Science Foundation of China (No. 51501134). R.D.K. Misra gratefully acknowledges

support from the National Science Foundation, USA through grant number DMR

1458074.

References
[1] R.D.K. Misra, S. Nayak, S.A. Mali, J.S. Shah, M.C. Somani, L.P. Karjalainen. On
the significance of nature of strain-induced martensite on phase-reversion-
induced nanograined/ultrafine-grained austenitic stainless steel. Metall. Mater.
Trans. A 41A (2010) 3-12.
[2] R.D.K. Misra, J.S. Shah, S. Mali, P.K.C. Venkata Surya, M.C. Somani, L.P.
Karjalainen. Phase reversion induced nanograined austenitic stainless steels:
16
Microstructure, reversion and deformation mechanisms. Mater. Sci. Technol.
29(10) (2013) 1185-1192.
[3] V.S.A. Challa, X.L. Wan, M.C. Somani, L.P. Karjalainen, R.D.K. Misra. Strain
hardening behavior of phase reversion-induced nanograined/ultrafine-grained
(NG/UFG) austenitic stainless steel and relationship with grain size and
deformation mechanism. Mater. Sci. Eng. A 613 (2014) 60-70.
[4] D.M. Xu, G.Q. Li, X.L. Wan, R.L. Xiong, G. Xu, K.M. Wu, M.C. Somani, R.D.K.
Misra. Deformation behavior of high yield strength-high ductility ultrafine-
grained 316LN austenitic stainless steel. Mater. Sci. Eng. A 688 (2017) 407-415.
[5] A. Kisko, R.D.K. Misra, J. Talonen, L.P. Karjalainen. The influence of grain size
on the strain-induced martensite formation in tensile straining of anaustenitic
15Cr-9Mn-Ni-Cu stainless steel. Mater. Sci. Eng. A 578 (2013) 408-416.
[6] D.L. Johannsen, A. Kyrolainen, P.J. Ferreira. Influence of annealing treatment on
the formation of nano/submicron grain size AISI 301 austenitic stainless steels.
Metall. Mater. Trans. A 37A (2006) 2325-2338.
[7] P. Behjati, A. Kermanpur, A. Najafizadeh, H. S. Baghbadorani. Effect of annealing
temperature on nano/ultrafine grain of Ni-free austenitic stainless steel. Mater. Sci.
Eng. A 592 (2014) 77-82.
[8] M.C. Somani, P. Juntunen, L.P. Karjalainen, R.D.K. Misra, A. Kyröläinen.
Enhanced mechanical properties through reversion in metastable austenitic
stainless steels. Metall. Mater. Trans. A 40A (2009) 729-744.
[9] H. Wu, G. Niu, J. Cao, M. Yang. Annealing of strain-induced martensite to obtain
micro/nanometre grains in austenitic stainless. Mater. Sci. Technol. 33(4) (2017)
480-486.
[10] M. Eskandari, A. Najafizadeh, A. Kermanpur. Effect of strain-induced martensite
on the formation of nanocrystalline 316L stainless steel after cold rolling and
annealing. Mater. Sci. Eng. A 519 (2009) 46-50.
[11] Ï. Üçok, Teiiehi Ando, N. J. Grant. Property enhancement in Type 316L stainless
steel by spray forming. Mater. Sci. Eng. A 133 (1991) 284-287.
[12] R. Jenkins, R.L. Snyder. Diffraction theory, in introduction to X-ray powder
diffractometry, John Wiley & Sons, Inc., Hoboken, NJ, USA, 2012.
[13] S. Sabooni, F. Karimzadeh, M.H. Enayati, A.H.W. Ngan. The role of martensitic
transformation on bimodal grain structure in ultrafine grained AISI 304L stainless
steel. Mater. Sci. Eng. A 636 (2015) 221-230.
[14] K. Nohara, Y. Ono, N. Ohashi. Composition and grain size dependencies of
strain-induced martensitic transformation in metastable austenitic stainless steels.
J. Iron Steel Institute Japan 63 (1977) 772-782.
[15] A. Hedayati, A. Najafizadeh, A. Kermanpur, F. Forouzan. The effect of cold
rolling regime on microstructure and mechanical properties of AISI 304L stainless
steel. J. Mater. Process. Tech. 210 (2010) 1017-1022.
[16] H.B. Wu, F.J. Wu, S.W. Yang, D. Tang. The formation mechanism of austenite
structure with micro/sub-micrometer bimodal grain size distribution, Acta Metall.
Sin. 50(3) (2014) 269-274.
[17] C.S. Pande, B.B. Rath, M.A. Imam. Effect of annealing twins on Hall-Petch
17
relation in polycrystalline materials. Mater. Sci. Eng. A 367 (2004) 171-175.
[18] X.H. Chen, J. Lu, L. Lu, K. Lu. Tensile properties of a nanocrystalline 316L
austenitic stainless steel. Scripta Mater. 52 (2005) 1039-1044.

18
Fig. 1. Optical (a, b) and TEM (c, d) images of microstructure in (a) commercial and
(b-d) CR 316LN steels, and (e) Engineering strain - engineering stress curve of
commercial and CR 316LN steels, respectively.

1
Fig. 2. (a, b, f) Optical, (e) SEM and (c, d) TEM micrographs of specimens annealed
at 600°C for (a) 2 minute, (b-d) 10 minutes, (d) 50 minutes, and (e) 100 minutes,
respectively.

2
Fig. 3. (a, f) Optical, (d) SEM and (b, c, e) TEM micrographs of specimens annealed
at 700°C for (a) 1 minute, (b, c) 5 minutes, (d) 10 minutes, (e) 50 minutes, and (f) 100
minutes, respectively.

3
Fig. 4. (a) Optical, (b) SEM and (c-f) TEM micrographs of specimens annealed at
800°C for (a) 1 minute, (b) 2 minutes, (c) 5 minutes, (d) 10 minutes, (e) 50 minutes
and (f) 100 minutes, respectively.

4
Fig. 5. (a, c, e, f) SEM and (b, d) TEM micrographs of specimens annealed at 900°C
for (a, b) 1 minute, (c, d) 2 minutes, (e) 10 minutes and (f) 100 minutes, respectively.

5
Fig. 6. (a, c-f) SEM and (b) TEM micrographs of specimens annealed at 1000°C for
(a, b) 1 minute, (c) 5 minutes, (d) 10 minutes, (e) 50 minutes and (f) 100minutes,
respectively.

6
Fig. 7. EBSD analysis map of the grain angles between adjacent grains for CR
specimens annealed at (a) 600°C for 5 minutes, (b) 600°C for 100 minutes, (c) 700°C
for 5 minutes, (d) 700°C for 100 minutes, (e) 800°C for 5 minutes and (f) 800°C for
100 minutes, respectively. The red, green and blue curves represent grain boundaries
with angle between 2 - 5°, 5 - 15° and 15 - 65°, respectively.

7
Fig. 8. EBSD analysis map of the grain angles between adjacent grains for CR
specimens annealed at (a) 900°C for 1 minute, (b) 900°C for 2 minutes, (c) 900°C for
5 minutes, (d) 900°C for 100 minutes, (e) 1000°C for 5 minutes, and (f) 1000°C for
100 minutes, respectively. The red, green and blue curves represent grain boundaries
with angle between 2 - 5°, 5 - 15° and 15 - 65°, respectively.

8
Fig. 9. (a) XRD patterns for CR specimen and specimens annealed at 900 °C for 1 - 3
minutes, (b) Volume fraction of á-martensite and (c, d) austenite grain size vs time
plots for specimens annealed at 600 - 1000 °C for 1 - 100 minutes.

Fig. 10. (a) The plots of yield strength and elongation vs grain size of 316LN
annealed specimens, and (b) Engineering stress-engineering strain plots for specimens
with NG, UFG and CG structures. The selected samples with NG, UFG and CG
structure were annealed at 900 °C for 1 minute, 900 °C for 2 minutes and 1000 °C for
100 minutes, respectively.

9
Fig.11. TEM micrographs illustrating (a, b) deformed microstructure near fracture of
tensile sample with NG structure, and deformed microstructure of (c, d) UFG, and (e,
f) CG samples with 0.1 strain.

10
Fig. 12. EBSD analysis map of the grain angles between adjacent grains for (a) UFG
and (b) CG specimens with 0.1 strain, respectively. The red, green and blue curves
represent grain boundaries with angle between 2 - 5°, 5 - 15° and 15 - 65°,
respectively.

11
Table 1. Chemical compositions, grain size and SFE of the commercial 316LN steel
(wt.%).
C Si Mn N Cr Ni Mo SFE, mJ/m2 Mean grain size, μm Md30, °C

0.04 0.34 1.15 0.048 18.06 8.33 0.051 18.9 12.6 7.5

Table 2. Percentage fraction and density of grain boundaries with the different
misorientation range of steels with varying annealing treatment.
Percentage fraction, % Density, Number/μm2
Annealing temperature and time
2°~5° 5°~15° 15°~65° 2°~5° 5°~15° 15°~65°

600°C-5minutes 17.3 11.1 71.6 148 94 610

600°C-100minutes 10.0 10.8 79.1 79 85 317

700°C-5minutes 11.2 12.1 76.8 88 93 367

700°C-100minutes 2.3 6.6 91.1 3 8 105

800°C-5minutes 1.1 6.3 92.6 0.1 0.7 10

800°C-100minutes 1.2 6.2 92.6 0.1 0.6 8

900°C-1minutes 24.1 24.1 51.5 37 36 77

900°C-2minutes 1.7 5.3 93.0 0.5 1 25

900°C-5minutes 0.9 4.4 94.7 0.2 0.9 20

900°C-100minutes 1.9 6.3 91.8 0.1 0.4 7

1000°C-5minutes 1.2 6.3 92.5 0.04 0.08 0.9

1000°C-100minutes 1.2 5.6 93.2 0.007 0.03 0.5

1
Table 3. Grain size, fraction of martensite, tensile properties and Md30 of 316LN
steels with varying annealing treatment.
Annealing temperature Mean grain Volume fraction of Mean tensile Mean yield strength, Mean
Md30, °C
and time size, μm martensite, % strength, MPa MPa elongation, %

0 (90% CR) -- 46 1785 1721 1.3 --


900°C-1.0minute 0.35 10 2160 2155 1.4 -7.1
900°C-1.5minutes 1.2 3 1803 1665 3.8 -2.0
900°C-2.0minutes 2.2 0 1160 994 40.4 0.4
900°C-2.5minutes 2.4 0 1040 618 53.6 0.8
900°C-3.0minutes 2.6 0 1000 558 55.0 1.2
900°C-5.0minutes 3.0 0 1011 488 56.6 1.7
900°C-50minutes 4.5 0 966 375 61.2 3.4
900°C-100minutes 5.8 0 939 346 63.5 4.4
1000°C-5.0minutes 21 0 923 316 76.5 9.7
1000°C-50minutes 35 0 855 274 78.9 11.8
1000°C-100minutes 45 0 842 265 80.6 12.8

Table 4. Percentage fraction and density of grain boundaries with the different
misorientation range and fraction of martensite in steels with 0.1 strain.
Annealing temperature and Percentage fraction, % Density, Number/μm2 Fraction of
Sample
time 2°~5° 5°~15° 15°~65° 2°~5° 5°~15° 15°~65° martensite, %

UFG 900°C-2.0minutes 21.5 20.5 58.0 47 45 127 4

CG 1000°C-100minutes 47.1 8.6 44.3 2 0.4 2 9

You might also like