You are on page 1of 29

Author’s Accepted Manuscript

Impact of intercritical annealing temperature and


strain state on mechanical stability of retained
austenite in medium Mn steel

Yong-Gang Yang, Zhen-Li Mi, Mei Xu, Qi Xiu,


Jun Li, Hai-Tao Jiang
www.elsevier.com/locate/msea

PII: S0921-5093(18)30541-0
DOI: https://doi.org/10.1016/j.msea.2018.04.041
Reference: MSA36361
To appear in: Materials Science & Engineering A
Received date: 22 February 2018
Revised date: 2 April 2018
Accepted date: 10 April 2018
Cite this article as: Yong-Gang Yang, Zhen-Li Mi, Mei Xu, Qi Xiu, Jun Li and
Hai-Tao Jiang, Impact of intercritical annealing temperature and strain state on
mechanical stability of retained austenite in medium Mn steel, Materials Science
& Engineering A, https://doi.org/10.1016/j.msea.2018.04.041
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Impact of intercritical annealing temperature and strain
state on mechanical stability of retained austenite in medium
Mn steel

Yong-Gang Yang1, Zhen-Li Mi1*, Mei Xu1, Qi Xiu1, Jun Li2, Hai-Tao Jiang1

1
Institute of Engineering Technology, University of Science and Technology
Beijing, Beijing 100083, China
2
Materials department of Automotive Engineering Research Institute, Chery
Automobile Co., Ltd, Wuhu, Anhui 241009, China

yeungyg@163.com
mizhen_li@126.com
xumei619@163.com
sepxiu2011@126.com
yang.y.g@outlook.com
jianght_ustb@163.com

*
Corresponding author. Tel.: 86-10-62336603; fax: 86-10-62332947

Abstract
Proper mechanical stability in retained austenite is a crucial factor for medium Mn

steel to attain high strength and excellent ductility. In this paper, intercritical

annealing and quenching and partitioning (IA & Q&P) processes are applied to

medium Mn steel, to obtain retained austenite with proper mechanical stability. The

effect of intercritical annealing temperature on mechanical stability of retained


1
austenite is studied using interrupted tests. Moreover, from the perspective of usage of

medium Mn steel, the mechanical stability of retained austenite at uniaxial tension

and plane strain states is further investigated, to understand the transformation of

retained austenite. Results indicate that high intercritical annealing temperature

(680°C) caused the austenite to transform immediately, due to the limited resistance

of austenite to martensite transformation, caused by low carbon (C) content. This low

C content also led to low stacking-fault energy, which may be another way to promote

the transformation of austenite. Additionally, due to the high Schmid factor, caused

by enhanced grain rotation and strain state, the retained austenite has relatively low

stability at the early stage of the plane strain state. With increasing martensite

surrounding the austenite, the retained austenite achieves high mechanical stability at

the later stage of the plane strain state.

Keywords: medium Mn steel, quenching and partitioning, mechanical stability,

retained austenite, intercritical annealing temperature, strain state

1. Introduction
Due to its remarkable performance, medium Mn steel containing retained austenite

(RA) microstructure is the focus of research in new generation automotive steel [1-2].

Medium Mn steel exhibits high strength and ductility, which are crucial to achieve the

goal of reducing the weight of automobiles, while meeting safety requirements. The

good mechanical properties are primarily due to the enhanced transformation-induced

2
plasticity (TRIP) effect resulting from the transformation of austenite-to-martensite

[3].

The resistance of RA to transform into martensite when it is subjected to

deformation, which is also known as mechanical stability, is a significant indicator in

the use of medium Mn steel in manufacturing vehicle bodies. RA with extremely high

stability cannot transform into martensite, thus resulting in poor performance of the

medium steel [4]. RA with extremely low stability transforms immediately when

subjected to a small strain, due to which, the medium steel cannot achieve sustainable

work hardening rate. Moreover, medium steel is in complex strain state, such as plane

strain, multiaxial tension and shear stress state, when it is being used. The

transformation of RA under complex strain state may influence the safety level of the

automobile. Therefore, mechanical stability of RA is a crucial question that should be

considered with respect to the properties of medium Mn steel, especially if it is

subjected to processes such as stamping, bending, drawing, or mechanical cycling

during usage. Understanding the elements impacting mechanical stability of RA is

considerably important in using medium Mn steel in automobiles.

It is suggested that the mechanical stability of RA is affected by grain size [5-8],

morphology [5,9], and carbon and manganese content [4,10-11] in RA, as well as the

crystallographic orientation [12,13], and the constraining influence of the surrounding

phase [10,11,14-18]. In medium Mn steel, austenite-reverted transformation annealing

(ART-annealing) heat treatment is usually required to form a large fraction of

austenite [2]. However, recent investigation revealed that the long annealing duration

3
in ART-annealing led to excessive C and Mn enrichment in austenite, which rendered

austenite too stable and deteriorated the TRIP effect [5]. Thus, different heat

treatments should be used to explore the stability of retained austenite in order to

obtain optimal mechanical properties of medium Mn steel. Speer et al. [19] originally

proposed the quenching and partitioning (Q&P) process to stabilize the retained

austenite during heat treatments, achieving an excellent combination of strength and

ductility. Following Speer, Moor et al. [20] obtained good combinations of tensile

strength (1,000-1,200 MPa) and total elongation (14-20%), using intercritical

annealing and quenching & partitioning (IA & Q&P) treatment. Therefore, the IA &

Q&P process may be a good way to obtain RA with proper mechanical stability for

medium Mn steel.

The aim of the present work focuses on the mechanical stability of RA in medium

Mn steel subjected to IA & Q&P process. The effect of intercritical annealing

temperature on the mechanical stability of RA is analyzed. Furthermore, the

mechanical stability of RA at uniaxial tension and plane strain states is investigated, in

order to understand the transformation of RA during usage. The mechanical stability

is studied by conducting interrupted tensile and plane strain tests and by determining

the remaining austenite fraction at different levels of strain. Factors linking the effect

of intercritical annealing temperature and strain state on the mechanical stability of

RA are proposed and discussed.

2. Experimental Procedure

4
2.1 Heat treatments processes of the investigated steel

The chemical composition of the investigated steel was Fe-0.22C-1.53Si-6.83Mn

(wt.%). The investigated steel was melted using a vacuum induction furnace. The

ingots were forged to produce 35-mm-thick billets, followed by hot rolling. After hot

rolling, the investigated steel was cold rolled to a thickness of 1.2 mm. Cylindrical

dilatometry samples, displayed in Figure 1a, were machined from the cold rolled sheet

steel. Using the cylindrical specimens, the critical transformation temperatures were

determined with a DIL805 dilatometer. The results showed that the austenite-starting

temperature (Ac1) and austenite-finishing temperature (Ac3) were 582°C and 745°C,

respectively. The martensite-starting temperature (Ms) was 175°C. With the help of

the critical transformation temperatures, three different IA & Q&P processes, shown

in Figure 1b, were selected to create the retained austenite with different degrees of

stability. The steel sheets subjected to different IA & Q&P processes are labelled as: (a)

Specimen IA640, which was heated to 640°C for 600 s, quenched to 75°C and

reheated to 450°C for 60 s; (b) Specimen IA660, which was heated to 660°C for 600 s,

quenched to 75°C and reheated to 450°C for 60 s; and (c) Specimen IA680, which

was heated to 680°C for 600 s, quenched to 75°C and reheated to 450°C for 60 s.

5
Fig 1 (a) Dimensions of the cylindrical specimens. (b) Heat treatment process of the
investigated steel.

2.2 Microstructural analyses

Microstructures of IA640, IA660, and IA680 specimens were observed by scanning

electron microscopy (SEM) and transmission electron microscopy (TEM). Specimens

for SEM and TEM analysis were cut from the IA640, IA660 and IA680 specimens

before and after deformation. After mechanical grinding, the SEM specimens were

mechanically polished and etched with 4% nitric acid for 10 s. The TEM specimens

were mechanically grinded to a thickness of 50 µm and then electro-polished with a

twin-jet electropolisher at -20°C using a solution containing 85 vol% C2H5OH and 15

vol% HClO4. The volume fraction of retained austenite was measured using X-ray

diffraction (XRD) and calculated based on the integrated intensity of (200)α, (211)α

ferrite/martensite peaks and (200)γ, (220)γ, and (311)γ austenite peaks. The texture

components and lattice parameter were measured using XRD. The measurements

were conducted in the 40°-130° 2θ, with a step size of 0.042°. The carbon

concentration in the RA was calculated based on its lattice parameter αγ [21]:

6
 =3.566+0.0453c +0.00095 Mn  0.0056 Al (1)

where αγ is in Å, χc, χMn, χAl are the contents of C, Mn and Al, in wt%.

2.3 Uniaxial tensile tests and plane strain tests

Dog-bone tensile test samples (gauge length: 25 mm; width: 6 mm; thickness: 1.2

mm) were prepared with gauge length along the rolling direction, using a wire cutting

machine. Uniaxial tensile tests were conducted with a deformation rate of 1 mm/min

on a SANS CMT 5105 testing machine. Rectangular plane strain samples (length: 25

mm; width: 6 mm; thickness: 1.2 mm) were cut with length along the rolling direction.

Plane strain tests were performed with a deformation rate of 1 mm/min on a

DSH-50-10 reversing cold mill.

3. Results

3.1 Strain-induced transformation of retained austenite

Retained austenite transforms into martensite when it is subject to strain, which can

delay the onset of necking and improve the ductility of steel [22]. The stress-strain

curves after IA & Q&P treatment at 640℃, 660℃, and 680℃ are displayed in Figure

2(a). Both the stress-strain curves of the IA640 and the IA660 specimens exhibit clear

yielding plateaus, followed by significant serrated flow before fracture, except for the

IA680 specimen in which no plateaus could be observed. Retained austenite in every

specimen at different strains, illustrated by the magenta-colored shapes, were

measured using XRD, to obtain the transformation of retained austenite. The

relationships between strain and transformation of retained austenite in IA640, IA660

and IA680 specimens are shown in Figure 2(b). The following equation is used to
7
describe the austenite consumption with the increase in strain [23].

f  f 0 exp  k  (2)

log f  log f0  K e (3)

Where f0 is the volume fraction of retained austenite (RA) before deformation and f is

the volume fraction of RA at strain ε. Eq. (3) is the logarithmic form of Eq. (2) and

K  k log e . A lower value of K corresponds to a higher driving force for

transformation and lower austenite stability.

Based on the fitting curve in Figure 1(b), the K values of IA640, IA660 and IA680

specimens are -0.027, -0.035 and -0.044, respectively. It indicates that the stability of

retained austenite in this steel decreased with the increasing intercritical temperature.

In other words, the majority of retained austenite in the IA680 specimen transformed

immediately when subjected to a small strain, while that in IA640 and IA660

specimens remained slightly stable when subjected to a large strain. It was

advantageous for the IA640 and the IA660 specimens to possess a good combination

of strength and ductility, as retained austenite could progressively transform into

martensite during the entire deformation process. This result was also confirmed by

the relatively higher yield strength and elongation of the IA640 and the IA660

specimens (Figure 1a).

8
Fig 2 Strain-induced transformation of retained austenite after IA& Q&P treatment at 640℃,
660℃, and 680℃. (a) Stress-strain curves of the specimens (b) Relationship between the volume
fraction of retained austenite and strain

Figure 3 provides the strain-induced transformation of retained austenite at the

uniaxial tension and the plane strain state. The specimens subjected to different strain

states were treated with IA & Q&P process at an intercritical annealing temperature of

660℃. The XRD patterns of the specimens at different plane strain levels are

displayed in Figure 3(a). Based on these patterns, volume fractions of retained

austenite at plane strain state have been calculated and plotted against the equivalent

strain in Figure 3(b). Moreover, the volume fractions of retained austenite at different

tensile strain levels are also presented in Figure 3(b). The volume fraction of retained

austenite progressively decreased with the increasing strain in both strain states. The

stability of retained austenite in both strain states was also calculated using Eq. (2)

and Eq. (3). The K value of IA660 specimen at the uniaxial tension is -0.035.

However, the K value at the plane strain state differed across two different

transformation stages, in order to better fit the strain-induced transformation curve,

and are marked as Kp1 and Kp2. The values of Kp1 and Kp2 are -0.0509 and -0.0028,

respectively, which represent the transformation rates of retained austenite. Here, the

larger the value of K, the higher was the stability of RA. Consequently, the stability of
9
RA was lower at the first stage of the plane strain state than at the uniaxial tension

state; however, the stability of RA greatly increased at the second stage of the plane

strain state.

Fig 3 Strain-induced transformation of retained austenite at different strain states. (a) XRD
patterns of the specimens at different plane strain levels (b) Relationship between the volume
fraction of retained austenite and strain at different strain states.

3.2 Microstructural examination

Fig 4 SEM images of the microstructures of the investigated steel IA & Q&P treatment at 640℃
(a), 660℃ (b), and 680℃ (c). ‘γL’ and ‘γE’ represent the lamellar and equiaxed austenite grains;
‘M’ and ‘α’ represent the martensite and ferrite grains; ‘θ’ denotes cementite.

Figure 4 depicts the microstructures of the investigated steel after IA & Q&P

treatment at 640℃, 660℃, and 680℃. The microstructures of the IA & Q&P process

samples are mainly composed of ferrite, martensite and retained austenite grains. The

austenite grains have two types of morphologies – lamellar and equiaxed. Moreover,

after intercritical annealing at 640℃ and 660℃, the martensite and austenite grains

are slightly finer than those annealed at 680℃. This phenomenon is reported by Cho

10
et al. [24] in 4Mn steel; high annealing temperature results in the coarsening of the

earlier austenite grain size and, in consequence, leads to coarsening of austenite and

ferrite grains at room temperature. In addition, Figure 4a displays a considerable

number of cementite particles precipitated during the IA & Q&P process at 640℃.

However, the cementite particles were not observed in the IA660 and IA680

specimens (Figure 4b and 4c). The results indicate that the temperature for the

cementite phases to completely disappear was less than 660℃.

Fig 5 Carbon content and the volume fraction of retained austenite in the IA640, IA660 and IA680
specimens measured by XRD. ‘RA’ represents the retained austenite.

The carbon content and the volume fraction of retained austenite in the IA640,

IA660 and IA680 specimens are displayed in Figure 5. The C content of RA was

measured to be 1.028, 1.005 and 0.906 wt. %, respectively, while the C content of the

matrix was 0.02 wt. %, indicating that C atoms were partitioned to RA. The RA

fractions of IA640, IA660 and IA680 were 28.4, 32.0 and 41.7%, respectively. The

change in RA fraction with the increasing annealing temperature implied that more

RA can be obtained under optimal annealing temperature. It is also seen that the C

content of RA decreased when the IA temperature increased from 640℃ to 680℃,

which suggests that there is insufficient carbon to diffuse into austenite at 660℃ and
11
680℃. This decreasing C content in RA is probably attributable to the increasing RA

fraction. The partitioned carbon atoms from matrix to RA are limited during the IA &

Q&P process [10, 11]. Thus, the C content in RA decreases with the increasing RA

fraction.

3.3 Texture characterization

Figure 6 displays the crystallographic textures of ferrite and retained austenite for

the IA660 specimen before deformation. The textures of ferrite were rather weak,

with maxima of 3.0 mrd (multiples of a random distribution) occurring for the E {111}

<110> and B {011} <112> components. The initial preferred orientation of the

retained austenite is the weak B {011} <211>, G {110} <001> and C {112} <111>

components, with maxima of 3.0, 2.5 and 1.5 mrd, respectively. The crystallographic

textures for the IA660 specimen at 27% tensile strain are displayed in Figure 7. The E

{111} <110> component was also measured using XRD. Compared to the ferrite

textures of specimens prior to deformation (Figure 6a), the mrd of E {111} <110>

component increased significantly, indicating that the {111} <110> component of the

ferrite is further enhanced by applying mechanical load. Moreover, G {110} <001>

component also occurred in the IA660 specimen at 27% tensile strain due to the

rotation of grains. However, B {011} <112> component could not be measured, which

may also be caused by the rotation of grains under mechanical load. The retained

austenite textures in specimens at 27% tensile strain (Figure 7b) have similar

components to that in Figure 6b. Furthermore, a slightly increasing mrd of C {112}

<111> components can be seen in Figure 7b due to the orientation rotation.


12
Fig 6 Orientation distribution functions (ODFs) in φ2 = 45°section for unstrained IA660
specimen. (a) ODFs for ferrite (b) ODFs for austenite. ‘E’ and ‘B’ represent the {111}<110>
component and the {011}<112> brass component, ‘C’ and ‘G’ denote the {112}<111> copper
component and {110}<001> Goss component, respectively.

Fig 7 Orientation distribution functions (ODFs) in φ2 = 45°section for IA660 specimen at 27%
tensile strain. (a) ODFs for ferrite (b) ODFs for austenite. ‘E’ and ‘B’ represent the {111}<110>
component and the {011}<112> brass component, ‘C’ and ‘G’ denote the {112}<111> copper
component and {110}<001> Goss component, respectively.

Figure 8 displays the crystallographic textures of ferrite and retained austenite for

the IA660 specimen at 21% plane strain. The E {111} <110> and B {110} <112>

components of the ferrite at this strain level had a higher mrd than that in specimens

prior to deformation. It can be seen from Figure 8b that a maxima of 3.0 and 2.5 mrd

occurred for the G {110} <001> and C {112} <111> components of retained austenite.

This result illustrated that the textures of retained austenite increased when compared

to the textures in the undeformed samples. Furthermore, can be seen that the G {110}

<001> and C {112} <111> components of RA in plane strain (Figure 8b) are higher

than that in tensile strain (Figure 7b), suggesting that plane strain has a stronger effect

13
on facilitating rotation of retained austenite grains. In addition, the ferrite and retained

austenite for the IA660 specimen at 56% plane strain presented similar

crystallographic textures, irrespective of the degree of specimen strain.

Fig 8 Orientation distribution functions (ODFs) in φ2 = 45°section for IA660 specimen at 21%
plane strain. (a) ODFs for ferrite (b) ODFs for austenite. ‘E’ and ‘B’ represent the {111}<110>
component and the {011}<112> brass component, ‘C’ and ‘G’ denote the {112}<111> copper
component and {110}<001> Goss component, respectively.

3.4 Transformation characteristics of retained austenite

Figure 9 displays the TEM examined microstructures for the unstrained IA660

specimen. The lamellar and equiaxed austenite grains can still be observed in Figure

9a, which is confirmed by indexing the diffraction patterns of the selected area. The

lamellar austenite grains had an ultrafine width of 150-300 nm and they were

transformed from martensite laths with ultrafine size during the IA & Q&P treatment

[25]. The equiaxed austenite grains had a coarse size of 400-600 nm. In addition, the

stacking fault was also found in equiaxed austenite grains (Figure 9b). The occurrence

of stacking fault is probably attributable to the low stacking fault energy [26]. When

the austenite grains nucleated during the IA process, there was shuffling of atomic

shells due to the low stacking fault energy, resulting in the occurrence of stacking

fault.

14
Fig 9 TEM micrographs showing morphology of retained austenite and stacking fault for
unstrained IA660 specimen. ‘γL’ and ‘γE’ represent the lamellar and equiaxed austenite grains.

Figure 10 displays the morphology of stacking fault and twin martensite for the

IA660 specimen at different plane strain levels. When the sample was strained to 6%,

plastic deformation occurred in the ferrite grains, contributing to an increase in

dislocation density within ferrite grains. At the same time, the stacking fault could

also be observed in the austenite grain. Compared with the unstrained sample in

Figure 9b, the amount of stacking fault increased significantly, indicating that the

stacking fault was enhanced by applying mechanical load. With further increasing

strain (9%), there was even greater stacking fault. Furthermore, twinned martensite

formed in the intersection zones of stacking faults (Figure 10b), which is confirmed

by the diffraction patterns of the selected area. Martensitic transformation from

stacking faults has been found in stainless steel by Venables [27] and Whelan [28].

They reported that the stacking faults facilitated the product of intermediate ε phase,

resulting in the occurrence of martensitic transformation. Twin martensite could be

obviously seen at 21% and 56% strain (Figure 10c-d). According to Song [29], the

lamellar retained austenite could transform into twin martensite during deformation.

Therefore, the twin martensite in this study, probably transformed from retained

austenite. Nevertheless, a mass of dislocations took place at 21% and 56% strain due
15
to a large degree of deformation (Figure 10c-d).

Fig 10 TEM micrographs showing morphology of stacking fault and twin martensite for IA660
specimen at different plane strain levels.(a) 6% (b) 9% (c) 21% (d) 56%

4. Discussion

4.1 Austenite mechanical stability as a function of intercritical annealing


temperature

The results of Section 3.1 show that the stability of RA decreased with the

increasing intercritical annealing (IA) temperature. Specifically, C concentration in

RA was the highest in IA640, followed by IA660 and IA680, as shown in Figure 5.

This sequence coincided with the sequence of decreasing RA stability after different

IA & Q&P processes. Therefore, the decrease in RA mechanical stability is mainly

attributable to the low C concentration. In fact, the dependence of RA mechanical

stability on carbon content has been reported in previous studies [10-11]. Higher

solute C content resulted in higher strength of RA. Moreover, with higher strength

16
(YS) of RA, there is more resistance of the volume change accompanying the

austenite to martensite transformation, increasing the stability of austenite. In addition,

it can be seen that the RA morphology and crystallographic orientation were similar in

IA640, IA660 and IA680 specimens due to the similar IA & Q&P process, which

contributed equivalent austenite mechanical stability. Although the grain size of RA

slightly increases with the increasing IA temperature, the austenite grain size does not

appear to be an independent factor that affects the stability of the austenite, and the

influence of grain size may also be attributed to the C content [10]. Furthermore,

whereas the RA could be stabilized by Mn partitioning, steel with a higher C content

and a slightly lower Mn content could also lead to austenite grains larger than 500 nm

that are still stable at room temperature [10,30]. This strongly suggests that the C

content has a greater influence on the stabilization of the austenite than the Mn

content. Therefore, in this study, the low C concentration is probably the main factor

for the decrease in RA mechanical stability at high IA temperature.

Apart from the C content in RA, the intercritical annealing temperature can

influence austenite stability in other ways. During IA & Q&P process, C atoms

diffused from the supersaturated martensite into RA. In the case of the IA680

specimen, the higher C diffusivity at 680 ℃ contributed to complete C

homogenization through austenite. However, the relatively lower IA temperature (640℃

and 660℃) may be insufficient to complete C homogenization. Thus, the C atoms

gather at the martensite-austenite interface, which was confirmed by

three-dimensional atom-probe tomography [31]. As the C accumulated at the

17
martensite-austenite interface, it results in higher C diffusivity to RA during the

partitioning process. Furthermore, the higher diffusivity caused C homogenization

through austenite in the IA640 and IA 660 specimens, while the lower diffusivity

created the heterogeneous C content through austenite in the IA680 specimen. Thus,

after the partitioning process, the martensite-austenite interface accumulated high C

content in the IA680 specimen, while the interior of the austenite had lower C content.

According to the discussion in the previous paragraph, lower solute C content resulted

in the lower stability of RA. Therefore, the transformation preferentially took place in

the interior of the austenite with lower C content, and then transformation of the

interface occurred with the increasing strain. Due to the aforementioned process, the

austenite with heterogeneous C content transformed early. This behavior is further

confirmed by Song [29]. In their research, the martensitic transformation took place in

the interior of the austenite, while the interface still remained. This research agrees

with the results of the present study and emphasizes the importance of C

homogenization during the IA & Q&P process, on the mechanical stability of RA.

It is well known that stacking fault energy (SFE) governs the transformation of RA

during deformation [32]. For example, a SFE of 12-35 mJ/m2 favors the formation of

mechanical twins, while a SFE below 18 mJ/m2 favors martensite formation [33-36].

Thus, the stacking fault energy (SFE), impacted by intercritical annealing temperature,

may in other ways influence austenite stability. The range of SFEs of the retained

austenite grains in the IA640, IA660 and IA680 specimens were estimated using the

following equation [33, 34]:

18
SFE  2G   2 (4)

where σ is the {111} interface energy between γ and ε, and is considered to be 5

mJ/m2 for the investigated steel [34, 36]. ρ is the molar surface density of atoms along

the {111} planes (as calculated by Equation (5)). It is calculated from the lattice

parameter (αγ) of austenite. ∆Gγ→ε is the molar Gibbs energy of the γ → ε

transformation, as a function of the composition. In the ternary Fe-Mn-Si-C system,

∆Gγ→ε is calculated by Equation (6) (neglecting the third order interaction terms).

4 1
 (5)
3  N
2

where N is the Avogadro constant. αγ represents the average lattice parameter of

austenite, which is calculated as 0.359 nm based on the Rietveld refinement of the

XRD data.
       
G    Fe GFe   Mn GMn   Si GSi  C GC   Fe  Mn FeMn
 
(6)
  Fe  c Fec
 
  Fe  Si FeSi
 
  Mn  Si MnSi
 
  Mn  C MnC
   
 Gmg

where χM and ∆GMγ→ε represent the molar fractions and the molar Gibbs energy of the

pure elements (Fe, Mn, Si, and C), ΩMmγ→ε is the excess free energy coefficient for the

binary systems, ∆Gmgγ→ε is the molar Gibbs energy due to the magnetic state of the

phase. The Si content of austenite grains is around 1.43 wt.%. The C content of the

austenite grains is 1.028, 1.005 and 0.906 wt. %, respectively, as shown in Figure 5.

The Mn content of the austenite grains is estimated to be in the range of 8-11 wt.%

[31,33]. ∆GMγ→ε andΩMmγ→ε at different temperatures were calculated using equations

in Table 1. (in which XC is atomic percentage of C, ∆Gmgγ→ε magnetic contribution,

TNφ Néel temperature, R gas constant, T absolute temperature, βφ magnetic moment,

19
ΩMmγ→ε excess free energies, μB Bohr magneton, p fraction of magnetic enthalpy

absorbed above TNφ).

Table 1 Thermodynamic parameters and equations to calculate ∆GMγ→ε andΩMmγ→ε


Parameter Function Ref.

∆GFeγ→ε -2243.38+4.309T (J/mol) [34]


∆GMnγ→ε -1000+1.123T (J/mol) [34]
∆GSiγ→ε -150-8T (J/mol) [35]
∆GCγ→ε -22166 (J/mol) [35]
ΩFeMnγ→ε 2873-717(XFe- XMn) (J/mol) [36]
ΩFeSiγ→ε 2850+3520(XFe- XSi (J/mol) [35]
ΩFeCγ→ε 42500 (J/mol) [34]
ΩMnCγ→ε 26910 (J/mol) [34]
ΩMnSiγ→ε 1780 (J/mol) [35]
γ→ε γ→ε ε γ
∆Gmg ∆Gmg = ∆Gmg - ∆Gmg --
∆Gmg φ
f(T/ TNφ)RTln(1+βφ/μB), φ=γ, ε [34]

79 1 474  1   3  9  15 
 1  
140 P 497  P   6 135 600 
1 ,
D
φ
when τ≤1 (τ= T/ TN );
f(T/ TNφ) [35]
 
5
 15 25

 10  315  1500 
  , whenτ>1
D
For any φ, P =0.28, D=2.34
βγ/μB 0.7 XFe+0.62 XMn-0.64 XFe XMn-4 XC (J/mol) [35]
βε/μB 0.62 XMn - 4 XC (J/mol) [35]
TNγ 251.71+681 XFe -1575 XFe -1740 XFe(K) [35]
TNε 580 XMn(K) [35]

The calculated results of SFEs of retained austenite grains are 18.3, 17.1, and

12.2mJ/m2 for IA640, IA660, and IA680 specimens, respectively. The SFEs are

sufficient for austenite grains to transform when specimens are subjected to strain.

20
Moreover, the sequence of SFEs caused by different IA temperatures also coincided

with the sequence of decreasing RA stability. Therefore, mechanical stability of RA

correlates with the previous conditions of the intercritical annealing temperature.

4.2 Mechanical stability of retained austenite at different strain states

There are significant differences in the mechanical stability of RA when the RA is

strained at different strain states, as shown in Figure 3. Due to the same process of IA

& Q&P at 660℃, the retained austenite possessed the same grain size, morphology, C

concentration, C homogenization and SFEs, which in principle should lead to

equivalent austenite mechanical stability. However, as deduced from the XRD results

in Figure 3, the mechanical stability of RA was lowest in the first stage of the plane

strain state, followed by the uniaxial tension state and the second stage of the plane

strain state. These differences in the mechanical stability of RA are a result of

differences in the crystallographic textures, which are discussed in this section.

Texture results show that the G {110} <001> and C {112} <111> components of

RA in 21% plane strain are higher than that in tensile strain. Moreover, from Figure

3b, the RA texture in 21% plane strain represents the grain rotation result during the

first stage of the plane strain state. Thus, the higher G {110} <001> and C {112} <111>

components in 21% plane strain indicates the higher grain rotation of RA during the

first stage of the plane strain state. According to Hu et al. [37], grain rotation promotes

favorable orientation with the highest resolved shear stress. Moreover, Blondé et al.

[38] reported that the mechanical stability of RA is affected by grain orientation.

Based on the orientation results measured by high-energy X-ray diffraction, it is


21
concluded that the austenite grains oriented along the loading direction are

transformed preferentially, due to the highest resolved shear stress. In other words, the

RA with the highest resolved shear stress has the lowest mechanical stability during

deformation. Therefore, higher grain rotation, in this study, may be one of the reasons

why RA obtained the lowest mechanical stability in the first stage of the plane strain

state.

The relationship between crystallographic textures and mechanical stability of RA

is described by the Schmid factor m, which relates to the maximum resolved shear

stress. The Schmid factor mu of RA at uniaxial tension can be estimated by equation

(7) [38, 39]:

mu  cos    cos   (7)

where λ represents the angle between the tensile direction and the <110> slip direction,

and Φ is the angle between the tensile direction and the {111} slip plane normal

direction. λ and Φ are shown in Figure 11a. The Schmid factor mp of RA at planar

strain state can be estimated by the equation (8) [39]:

mp  cos   cos     cos   cos   (8)

where α represents the angle between the tensile direction and the <110> slip

direction, β is the angle between the tensile direction and the {111} slip plane

normal direction, γ is the angle between the compression direction and the <110>

slip direction, andδ is the angle between the compression direction and the {111}

slip plane normal direction. α, β, γ, and δ are shown in Figure 11b.

22
Fig 11 Schematic diagrams describing the Schmid factor parameters at different strain states. (a)
the uniaxial tension state (b) the planar strain state

The calculated results of the Schmid factor are displayed in Table 2. The RA

oriented with the C {112} <111>, obtained equivalent Schmid factor at both strain

states. However, the Schmid factor of G {110} <001> and B {110} <112> at planar

strain state is much higher than that at uniaxial tension state. According to the in situ

high-energy X-ray diffraction results [38], the high Schmid factor corresponds to the

low mechanical stability of RA. Thus, the RA oriented with the G {110} <001> and B

{110} <112>, obtained low mechanical stability at the plane strain state. In the first

stage of the plane strain state, the C {112} <111>, G {110} <001> and B {110} <112>

components are the main crystallographic textures of RA (as indicated by Figure 8).

Therefore, the RA has low mechanical stability due to the low mechanical stability

components (G {110} <001> and B {110} <112>) in the first stage of the plane strain

state. Furthermore, TEM results show that several extended stacking faults occur

during the first stage of the plane strain state (Figure 10). However, only a few

stacking faults are observed at the uniaxial tension state. According to Weidner [40],

stacking faults belonging to slip systems with high Schmid factors can be largely

extended by the applied external stress. In this study, the main texture components of

RA obtain high Schmid factors in the planar strain state (Table 2). Thus, the
23
occurrence of several extended stacking faults is probably promoted by the high

Schmid factors. Furthermore, Shimizu [41] illustrated that martensite platelets are

found to be associated with austenite stacking faults, which facilitate the

transformation of RA during the strain-induced transformation process. Therefore, the

occurrence of several extended stacking faults, promoted by high Schmid factors, may

be another way for textures to influence austenite stability.

Table 2 Calculated results of the Schmid factor at the uniaxial tension and the planar strain state.

Texture components mu mp
C {112} <111> 0.272 0.272
G {110} <001> 0.408 0.816
B {110} <112> 0.408 0.680

In the case of the second stage of the plane strain state, despite the higher Schmid

factor caused by textures, the martensite phase surrounding the austenite may increase

mechanical stability of the RA. After the first stage of the plane strain state, several

martensite formations occurred due to the strain-induced transformation of RA (as

displayed in Figure 3 and Figure 10). Furthermore, with more volume fraction of

martensite surrounding the austenite, there is more resistance by the martensite in

accommodating the volume change accompanying the austenite to martensite

transformation [3, 10, 29]. Thus, RA has high mechanical stability due to the

increasing martensite surrounding the austenite, in the second stage of the plane strain

state.

Conclusion

24
In this paper, the intercritical annealing and quenching and partitioning (IA & Q&P)

process is applied to medium Mn steel, in order to obtain retained austenite with

proper mechanical stability. The effect of intercritical annealing temperature and

strain state on the mechanical stability of RA is investigated. The conclusions can be

summarized as follows:

(1) The medium Mn steel could obtain retained austenite with various mechanical

stability values by changing intercritical annealing temperature of IA & Q&P process.

Moreover, the stability of RA differed across two stages at plane strain state, while

showing uniformity in the single stage at uniaxial tension state.

(2) The mechanical stability of retained austenite decreased with increasing

intercritical annealing temperature of IA & Q&P process. The low mechanical

stability of retained austenite after IA & Q&P process at 680℃ is due to the limited

resistance of the austenite to martensite transformation, caused by the low C content

and the heterogeneities of carbon concentration in austenite. Furthermore, low C

content also leads to low stacking fault energy, promoting the transformation of

austenite during deformation.

(3) The mechanical stability of RA was lowest in the first stage of the plane strain

state, followed by the uniaxial tension state and the second stage of the plane strain

state. The retained austenite has relatively low stability in the first stage of the plane

strain state due to the high Schmid factor, caused by enhanced grain rotation and

strain state. Nevertheless, the retained austenite has high mechanical stability at the

25
later stage of the plane strain state, due to the increase in martensite surrounding the

austenite.

Acknowledgements
Financial supports from the State Key Research and Development Program of China
(2016YFB0101605, 2017YFB0304404 ) are greatly acknowledged.

References:
[1]D.K. Matlock, J.G. Speer, Third generation of AHSS: microstructure design concepts, in: A.
Haldar, S. Suwas, D. Bhattacharjee (Eds.), Microstructure and Texture in Steels, Springer, London,
2009, pp. 185–205.
[2]B. Hu, H. Luo, F. Yang, H. Dong. Recent progress in medium-Mn steels made with new
designing strategies, a review, J. Mater. Sci. Technol., 33(2017) 1457-1464.
[3]J. Hidalgo, K.O. Findley, M.J. Santofimia. Thermal and mechanical stability of retained
austenite surrounded by martensite with different degrees of tempering. Mat. Sci. Eng. A,
690(2017) 337-347.
[4]Z.C. Li, H. Ding, R.D.K. Misra, Z.H. Cai. Microstructure-mechanical property relationship and
austenite stability in medium-Mn TRIP steels: The effect of austenite-reverted transformation and
quenching-tempering treatments. Mat. Sci. Eng. A, 2017, 682:211-219.
[5]D. De Knijf, C. Föjer, L.A.I. Kestens, R. Petrov, Factors influencing the austenite stability
during tensile testing of Quenching and Partitioning steel determined via in-situ Electron
Backscatter Diffraction, Mat. Sci. Eng. A, 638 (2015) 219227
[6]M.M. Wang, C.C. Tasan, D. Ponge, A. Kostka, D. Raabe, Smaller is less stable: size effects on
twinning vs. transformation of reverted austenite in TRIP-maraging steels, Acta Mater. 79 (2014)
268–281.
[7]V.S.A. Challa, R.D.K. Misra, M.C. Somani, Z.D. Wang, Influence of grain structure on the
deformation mechanism in martensitic shear reversion-induced Fe-16Cr-10Ni model austenitic
alloy with low interstitial content: coarse-grained versus nano-grained/ultrafine-grained structure,
Mat. Sci. Eng.: A 661 (2016) 51-60.
[8]R.D.K. Misra, V.S.A. Challa, P.K.C. Venkatsurya, Y.F. Shen, M.C. Somani, L.P. Karjalainen,
Interplay between grain structure, deformation mechanisms and austenite stability in
phase-reversion-induced nanograined/ultrafine-grained austenitic ferrous alloy, Acta Mater. 84
(2015) 339-348.
[9]J. Min, L.G. Hector Jr, L. Zhang, J. Lin, J.E. Carsley, L. Sun, Elevated-temperature mechanical
stability and transformation behavior of retained austenite in a quenching and partitioning steel,
Mat. Sci. Eng. A, 673 (2016) 423-429.
[10]H. Luo Comments on "austenite stability of ultrafine-grained transformationinduced plasticity
steel with Mn partitioning" by S. Lee, S.J. Lee and B.C. de Cooman, Scripta Materialia 65,

26
225-228, Scr. Mater. 66 (2012) 829-831, 2011.
[11]R.M. Wu, W. Li, C.L. Wang, Y. Xiao, L. Wang, X.J. Jin, Stability of retained austenite through
a combined intercritical annealing and quenching and partitioning (IAQP) treatment, Acta Metall.
Sin. (Engl. Lett.). 28 (2015) 386-393.
[12]P. Hilkhuijsen, H.J.M. Geijselaers, T.C. Bor, E.S. Perdahcıoğlu, A.H. vd Boogaard, R.
Akkerman. Strain direction dependency of martensitic transformation in austenitic stainless steels:
The effect of γ-texture γ - texture mathContainer Loading Mathjax, Mat. Sci. Eng. A 573 (2013)
100-105.
[13]S.O. Kruijver, L. Zhao, J. Sietsma, S.E. Offerman, N.H. van Dijk, E.M. Lauridsen, L.
Margulies, S. Grigull, H.F. Poulsen, S. van der Zwaag. In situ observations on the mechanical
stability of austenite in TRIP-steel J. Phys. IV (2003) 499-502.
[14]C. Wang, H. Ding, M. Cai, B. Rolfe, Characterization of microstructures and tensile properties
of TRIP-aided steels with different matrix microstructure, Mat. Sci. Eng. A, 610 (2014) 65-75.
[15]B. Fultz, J.I. Kim, Y.H. Kim, H.J. Kim, G.O. Fior, J.W. Morris, The stability of precipitated
austenite and the toughness of 9Ni steel, Metall. Trans. A 16 (1985) 2237-2249.
[16]P.J. Jacques, F. Delannay, J. Ladrière, On the influence of interactions between phases on the
mechanical stability of retained austenite in transformation-induced plasticity multiphase steels,
Metall. Mater. Trans. A 32 (2001) 2759-2768.
[17]W.S. Li, H.Y. Gao, H. Nakashima, S. Hata, W.H. Tian, In-situ study of the
deformation-induced rotation and transformation of retained austenite in a lowcarbon steel treated
by the quenching and partitioning process, Mat. Sci. Eng. A 649 (2016) 417-425.
[18]B. Fultz, J.W. Morris, The mechanical stability of precipitated austenite in 9Ni steel, Metall.
Trans. A 16 (1985) 2251–2256.
[19]J. Speer, D.K. Matlock, B.C. De Cooman, J.G. Schroth. Carbon partitioning into austenite
after martensite transformation, Acta Materialia 51 (2003) 2611-2622.
[20]E.D. Moor, J.G. Speer, D.K. Matlock, J.H. Kwak, S.B. Lee, Quenching and Partitioning of
CMnSi Steels Containing Elevated Manganese Levels, Steel Res. Int. 83 (2012) 322-327.
[21]N.H. Van Dijk, A.M. Butt, L. Zhao, J. Sietsma, S.E. Offerman, J.P. Wright, S. Van Der Zwaag,
Thermal stability of retained austenite in TRIP steels studied by synchrotron X-ray diffraction
during cooling, Acta Mater. 53 (2005) 5439-5447.
[22]Noriyuki Tsuchida, Y. Tomota, A micromechanic modeling for transformation induced
plasticity in steels, Mater. Sci. Eng. A 285 (1–2) (2000) 346-352.
[23]Wen Shi, Lin Li, Chun-Xia Yang, Ren-Yu Fu, Li Wang, Partick Wollants, Straininduced
transformation of retained austenite in low-carbon low-silicon TRIP steel containing aluminum
and vanadium, Mater. Sci. & Eng. A 429 (1-2) (2006) 247-251.
[24]Lawrence Cho, Eun Jung Seo, Bruno C. De Cooman. Near-Ac3 austenitized ultra-fine-grained
quenching and partitioning (Q&P) steel, Scripta Materialia 123 (2016) 69-72.
[25]Z.Z. Zhao, J.H. Liang, A.M. Zhao, J.T. Liang, D. Tang, Y.P. Gao. Effects of the austenitizing
temperature on the mechanical properties of cold-rolled medium-Mn steel system, J. alloy. &
compd. 691 (2017) 51-59.
[26] I. Karaman, H. Sehitoglu, K. Gall, Y. I. Chumlyakov, and H. J. Maier, Deformation of single
crystal Hadfield steel by twinning and slip, Acta Mater. 48 (2000) 1345-1359.
[27]J. A. Venables. The martensite transformation in stainless steel, The Philosophical Magazine:
A, 7:73,1962, 35-44,

27
[28]M. J. Whelan, P. B. Hirschand, R. W. Horne. Dislocations and stacking faults in stainless steel,
Proceedings of the royal society A, 1957, 7, 524-538.
[29]C. Song, H. Yu, L. Li, T. Zhou, J. Lu, X. Liu. The stability of retained austenite at different
locations during straining of I&Q&P steel. Mater. Sci. Eng. A 670 (2016) 326-334.
[30]H. Dong, Paper presented at the fourth Symposium on Heat-resistant Steels and Alloys, April
11-13, Beijing, China, 2011.
[31] Z.J. Xie, C.J. Shang, S.V. Subramanian, X.P. Ma, R.D.K. Misra. Atom probe tomography and
numerical study of austenite stabilization in a low carbon low alloy steel processed by two-step
intercritical heat treatment, Scripta Materialia 137 (2017) 36-40.
[32]S. Martin, S. Wolf, U. Martin, L. Krüger, D. Rafaja, Deformation Mechanisms in Austenitic
TRIP/TWIP Steel as a Function of Temperature, Metall. Mater. Trans. A 47 (2016) 49-58.
[33]B. Hu, H. Luo. A strong and ductile 7Mn steel manufactured by warm rolling and exhibiting
both transformation and twinning induced plasticity, J. alloy. & compd.725 (2017) 684-693.
[34]S. Allain, J.-P. Chateau, O. Bouaziz, S. Migot, N. Guelton. Correlations between the
calculated stacking fault energy and the plasticity mechanisms in Fe-Mn-C alloys, Materials
Science and Engineering A 387-389 (2004) 158-162.
[35]Curtze S, Kuokkala V T, Oikari A, et al. Thermodynamic modeling of the stacking fault
energy of austenitic steels. Acta Materialia, 2011, 59(3): 1068-1076.
[36]A. Dumay, J.P. Chateau, S. Allain, S. Migot, O. Bouaziz. Influence of addition elements on the
stacking-fault energy and mechanical properties of an austenitic Fe-Mn-C steel, Mater. Sci. Eng. A
483 (2008) 184-187.
[37]G. Hu, X. Cai, Y. Rong, Fundamentals of material science, second ed., Shanghai Jiao Tong
University Press, Shanghai, China 2010.
[38]R. Blondé, E. Jimenez-Melero, L. Zhao, J.P. Wright, E. Brück, S. van der Zwaag, N.H. van
Dijk, High-energy X-ray diffraction study on the temperature-dependent mechanical stability of
retained austenite in low-alloyed TRIP steels, Acta Mater. 60 (2012) 565-577.
[39]P. Yang. Electron backscatter diffraction technology and its application, second ed.,
Metallurgical Industry Press, Beijing, 2007.
[40]A. Weidner, S. Martin, V. Klemm, U. Martin, H. Biermann, Stacking faults in high-alloyed
metastable austenitic cast steel observed by electron channelling contrast imaging. Scripta
Materialia, 64 (2011) 513-516.
[41]K. Shimizu, M. Oka, C.M. Wayman. The association of martensite platelets with austenite
stacking faults in a Fe-8Cr-1C alloy. Acta Metallurgica, 18(1970) 1005-1011.

28

You might also like