SFE MN Steels

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Materials Science and Engineering A 387–389 (2004) 158–162

Correlations between the calculated stacking fault energy and the


plasticity mechanisms in Fe–Mn–C alloys
S. Allain a,∗ , J.-P. Chateau a , O. Bouaziz b , S. Migot a , N. Guelton b
a Laboratoire de Physique des Matériaux, Ecole des Mines, Parc de Saurupt, F-54000 Nancy, France
b IRSID, ARCELOR Group, Voie Romaine, BP 30320, F-57283 Maizière les Metz Cedex, France

Received 26 August 2003; received in revised form 1 December 2003

Abstract

A model is proposed for the evaluation of the stacking fault energy (SFE) in Fe–Mn–C austenitic alloys, at different temperatures. It
accounts for the variation of the Gibbs energy of each element during the austenite to ε martensite transformation, plus their interactions.
The Gibbs energy due to the antiferromagnetic to paramagnetic transition is also taken into account. The required data have been obtained
from the literature. The result shows a decrease of the SFE with temperature, with a saturation below the austenite Néel temperature. The
result agrees with the mechanical and thermal martensitic transformation limits proposed by Schumann. The plasticity mechanisms depend
on the SFE. The mechanical martensitic transformation occurs below 18 mJ/m2 , and twinning between 12 and 35 mJ/m2 , in agreement with
the tensile tests and the deformation microstructures observed in an Fe–22 wt.% Mn–0.6 wt.% C alloy at 77, 293 and 693 K.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Stacking fault energy; Gibbs energy; Néel temperature; Mechanical twinning; Martensitic transformation

1. Introduction been investigated after tensile tests at different temperatures


and correlated with the calculated value of the SFE.
The good mechanical properties of high manganese
austenitic steels are due to the competition between differ-
2. Calculation of the SFE
ent plasticity mechanisms. As reported in the literature [1],
their activations are triggered by the value of the stacking
2.1. Modeling of the SFE
fault energy (SFE) of the alloy. With a decreasing value of
the SFE, plasticity is achieved by: (i) partial and perfect
As first proposed by Hirth [2], we use a classical formal-
dislocations gliding, (ii) gliding and mechanical twinning,
ism (see for example Ferreira and Müllner [3] in the ternary
which increases the elongation to fracture and provides the
Fe–Ni–Cr system [3] or Lee and Choi [4] in the ternary
best hardening rate and (iii) gliding and ε or ␣ martensitic
Fe–Mn–Si system). We suppose that the SFE Γ is the Gibbs
transformations which give the best yield strength and a
energy required to create a platelet of ε martensite of a thick-
good hardening rate.
ness of only two atomic layers.
Since the SFE depends on temperature and on chemical
composition, it has a major influence on the mechanical Γ = 2ρ G␥→ε + 2σ (1)
properties of alloys.
with ρ the molar surface density along {1 1 1} planes,
First, this paper proposes a new assessment of the cal-
G␥→ε the molar Gibbs energy of the transformation ␥ →
culation of the SFE in the ternary austenitic Fe–Mn–C al-
ε and σ the surface energy of the interface ␥/ε. The value of
loys. The modeling is then used to calculate the evolution
the last term is generally taken between 5 and 15 mJ m−2
of the SFE with temperature for the austenitic Fe–22 wt.%
for transition metals [3–5]. The molar surface density is ge-
Mn–0.6 wt.% C steel. The deformation microstructures have
ometrically determined by introducing the lattice parameter
a of the alloy:
∗ 4 1
Corresponding author. Tel.: +33 3 83584143; fax: +33 3 83579794. ρ= √ 2 (2)
E-mail address: allain@mines.inpl-nancy.fr (S. Allain). 3a N

0921-5093/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2004.01.059
S. Allain et al. / Materials Science and Engineering A 387–389 (2004) 158–162 159

with N the Avogadro number. In the composition range stud- 2.3. Treatment of the ternary Fe–Mn–C system
ied, Ono et al. [6] have measured this parameter at room
temperature. According to their measurements, we retain a In the ternary Fe–Mn–C system, Eq. (7) becomes, by
constant value for the lattice parameter equal to 0.361 nm. neglecting the interactions terms of third order:
␥→ε ␥→ε ␥→ε
2.2. Modeling ∆Gγ→ε in the binary Fe–Mn system G␥→ε = xFe GFe + xMn GMn + xC GC
␥→ε ␥→ε ␥→ε
+ xFe xMn ΩFeMn + xFe xC ΩFeC + xMn xC ΩMnC
We first apply the methodology to the case of the binary ␥→ε
+ Gmg (8)
system Fe–Mn for the sake of simplicity. The contributions
of carbon are then introduced and discussed. The bone of contention in such a modeling is the choice
The Gibbs energy of a binary solid solution ϕ (in this case of the thermochemical and magnetic data sets. We will only
␥ or ε) is approached using the regular solution law. propose our choice and not discuss here the relevance of the
ϕ ϕ data sets. For more details see the interesting comparisons
Gϕ = xFe 0 GFe + xMn 0 GMn + RT(xFe ln(xFe ) in ref. [9].
ϕ ϕ
+ xMn ln(xMn )) + E Gm + Gmg (3)
2.4. Choices of the thermochemical and magnetic
with x and 0 Gφ the molar fractions and the molar Gibbs parameters
energy of the pure elements in the given paramagnetic phase
ϕ
state, T the temperature, R the gas constant, E Gm the excess The data used are mainly taken from studies of the Sci-
ϕ
molar Gibbs energy due to mixing and Gmg the molar Gibbs entific Group Thermodata Europe (SGTE). The values and
energy due to the magnetic state of the phase. functions are summed up in Table 1. Some free parameters
According to a subregular modeling, the excess term is have been reassessed in the present study to assure a better fit
expressed by [7]. with experimental measurements presented in the literature.
ϕ
E
Gm = xFe xMn (0 Lϕ + 1 Lϕ (xFe − xMn )) (4)
2.5. Discussion
with 0 Lϕ a linear function of the temperature but indepen-
dent of the alloy composition and 1 Lϕ a constant.
The data about the magnetic moment of the austenitic
Both the ␥ and the ε phases are antiferromagnetic, and thus
phase are purely empirical [9] and are fitted by introducing
have their own Néel temperatures (transition temperature be-
a free parameter. Cotes et al. [7] proposed to account for a
tween the paramagnetic and antiferromagnetic states). Ac-
second order term as follows:
cording to a law first proposed by Inden and Hillert [8], and
used by numerous authors in the case of the binary Fe–Mn β␥ = βFe xFe + βMn xMn − βFeMn xFe xMn (9)
[7,9,10], the magnetic contribution described in Eq. (3) can
with βFe and βMn the empirical values for the pure elements
be calculated:
in the ␥ state as proposed by Huang [14] and βFeMn the sec-
ϕ T
Gmg = RT ln(βϕ + 1)f ϕ (τ ϕ ) with τ ϕ = ϕ (5) ond order term which is used as a fit parameter. With their
TNéel data set, Cotes et al. find βFeMn = 1, but we choose βFeMn =
with βϕ the magnetic moment, a polynomial function of the 0.5 since this value provides a better fit with experimental
composition, and τ ϕ the scaled temperature besides the Néel results. The antiferromagnetic contribution of the ε phase is
ϕ ε
accounted by introducing the values TNéel and βε proposed
transition TNéel of phase ϕ, which only depends on the alloy
composition. The polynomial function f ϕ is defined in ref. by Huang [14]. Since the Néel temperature is significantly
[10] for instance. lower than the room temperature, many authors pay no heed
If we suppose that during the transformation γ → ε, the to this weak contribution. One should also notice that some
stoichiometry is kept constant without Suzuki effect (case former models (see [4] for example) do not explicitly intro-
of the martensitic transformations), duce a magnetic term. Nevertheless, the contribution is par-
␥ ␥
tially accounted for by the use of the initial Kaufman data
G␥→ε = xFe (0 GεFe − 0 GFe ) + xMn (0 GεMn − 0 GMn ) set [18,19].
␥ ␥
+ (E Gεm − E Gm ) + (Gεmg − Gmg ) (6) Too few data concerning the carbon interactions have been
found in the literature for both phase ␥ and ε. By using a
By using Eqs. (3) and (4) and a compact formalism, Eq. (6) value for Fe–C interaction term proposed by Adler et al. [15],
can be rewritten term by term as follows: Yang and Wan [12] have used an indirect method proposed
␥→ε ␥→ε ␥→ε by Ishida and Nishizawa [13]. This method has then been
G␥→ε = xFe GFe + xMn GMn + xFe xMn ΩFeMn
␥→ε successfully applied by using the SGTE data set and values
+ Gmg (7) proposed by Ishida. Since most of the authors only address
where the binary Fe–Mn system, few models integrate the effect of
␥→ε
carbon on the Néel temperature. We thus propose an empiric
ΩFeMn = (0 Lε − 0 L␥ ) + (1 Lε − 1 L␥ )(xFe − xMn ) function for a given composition range (15% < Mn (wt.%)
160 S. Allain et al. / Materials Science and Engineering A 387–389 (2004) 158–162

Table 1
Numerical values and functions used for the calculations (Eqs. (8) and (5))
Parameters Functions used (units) Reference
␥→ ε
GFe (I) −2243.38 + 4.309T (J mol−1 ) [11]
␥→ ε
GMn (I) −1000.00 + 1.123T (J mol−1 ) [11]
␥→ ε
GC (R) −22166 (J mol−1 ) [12,13]
␥→ε
ΩFeMn (I) 2180 + 532(xFe − xMn ) (J mol−1 ) [14]
␥→ε
ΩFeC (I) 42500 (J mol−1 ) [15]
␥→ε
ΩMnC (R) 26910 (J mol−1 ) [12,13]

TNéel (R) 669.27(1 − exp(−5.46xMn )) − 0.241xC (1 − xMn − xC ) − 109 (K) [1,16,17]
ε
TNéel (I) 580xMn (K) [11,14]
β␥ /µB (R) 0.7xFe + 0.62xMn − 0.5xFe xMn [9,14]
βε /µB (I) 0.62xMn [11,14]
σ (R) 0.009 (J m−2 ) [1,3,9]
(I) indicates the use of the values given in the literature and (R) for the values which are reassessed in this study (µB Bohr magneton).

< 45% and 0% < C (wt.%) < 1.1%). The function is fitted
to the data found in the papers quoted in Table 1 using a
least-squares method.
The room temperature iso-SFE lines are plotted in Fig. 1
in the carbon/manganese (wt.%) plane. On the same map,
the straight lines proposed by Schumann [20] to describe
the occurrence of the mechanical and thermal ε marten-
sitic transformation are plotted. Our iso-SFE lines follow
Schumann’s lines, which is in agreement with an activation
of the mechanical ε martensitic transformation controlled by
the SFE values. The mechanical transformation occurs if the
SFE is lower than about 18 mJ mol−1 as proposed by Rémy
[1].
In Fig. 1, the stability line at 973 K for the carbide pre- Fig. 2. Calculated evolution of the SFE with temperature for a Fe–22 wt.%
cipitation ␥ ↔ ␥ + M3 C (M = Fe or Mn) is plotted in ad- Mn–0.6 wt.% C steel. The arrows indicate the Néel temperatures for the
dition to fix the boundaries of the fully austenitic domain ␥ and ε phases. The chemical (dark grey) and the magnetic contributions
[21].To confirm the link between the SFE value and the ac- (light grey) are also plotted. For comparison, the experimental values
(cross) given by Rémy for a Fe–20 wt.% Mn–4 wt.% Cr–0.5 wt.% C steel
tivation of mechanical twinning and martensitic transforma- are also shown [1].

tion, we calculate the SFE versus temperature for different


alloys found in the literature [22–24]. For each of them, the
active deformation mechanisms are checked by its author
at given temperatures. This study confirms that mechanical

Fig. 1. The calculated iso-EFE lines in the carbon/manganese (wt.%) map


at 300 K. The straight lines proposed by Schumann [20] to describe the
regimes of mechanical and thermal martensitic transformations are also Fig. 3. Tensile strain–stress curves for the Fe–22 wt.% Mn–0.6 wt.% C
plotted. The dashed line indicates the border between austenite and the steel (grain size = 15 ␮m) at different temperatures. The stars on the
␥ + M3 C (M = Fe or Mn) field at 973 K [21]. curves refer to the TEM images shown in Fig. 4.
S. Allain et al. / Materials Science and Engineering A 387–389 (2004) 158–162 161

Fig. 4. (a) TEM dark field (DF) micrograph from a sample deformed at 673 K. A high dislocation density is observed with a cell substructure. No
evidence of twinning. (b) and (c) TEM DF micrograph from a sample deformed at 293 K. Twins are in contrast. (d) TEM DF micrograph from a sample
deformed at 77 K. Two ε martensitic variants appear in the grain.

ε martensitic transformation only occurs if the SFE is be- this alloy is lower than that of Fe–22 wt.% Mn–0.6 wt.% C
low 18 mJ mol−1 and mechanical twinning occurs if the SFE because chromium stabilizes the ␣ phase and thus reduces
is included in the 12–35 mJ mol−1 range for the studied G␥→ε .
grades. For the Fe–22 wt.% Mn–0.6 wt.% C steel, the SFE varies
significantly in the given temperature range. The deforma-
2.6. Results for a Fe–22Mn–0.6C (wt.%) steel tion mechanisms at different temperatures are predicted ac-
cording to the calculated SFE:
For the composition studied, the SFE values versus tem- T = 673 K → Γ = 80 mJ m−2 → dislocation gliding
perature are plotted in Fig. 2. We also plotted the chemi-
cal and the magnetic contributions as indicated. The Néel
temperatures of the ε and ␥ phases are indicated by the ar- T = 293 K → Γ = 19 mJ m−2
rows. Each magnetic transition gives an additional contri- → mechanical twinning + dislocation gliding
bution to G␥→ε and thus induces a change in the slope of
the SFE curve. The global behavior of the calculated SFE is T = 77 K → Γ = 10 mJ m−2
in agreement with the experimental results of Rémy [1] for
an Fe–20 wt.% Mn–4 wt.% Cr–0.5 wt.% C alloy and partic- → ε martensitic transformation

ularly the stabilization of the SFE below TNéel . The SFE of + dislocation gliding
162 S. Allain et al. / Materials Science and Engineering A 387–389 (2004) 158–162

3. Deformation mechanisms activated during Acknowledgements


tensile tests
The ARCELOR Company is gratefully acknowledged for
The experimental procedure for performing tensile tests the doctoral grant to one of us (A.S.) and for the financial
at different temperatures is described elsewhere [25]. Fig. 3 support.
shows the stress–strain behavior of the alloy at different tem-
peratures. The stars indicate when the samples were studied
by transmission electronic microscopy (TEM). At 673 K, the References
alloy has a low hardening rate leading to necking at 32%
strain. Dislocations are organized into cells at the end of de- [1] L. Rémy, thesis, Université d’Orsay, Paris VI, 1976.
formation as shown Fig. 4(a). No evidence of twinning or [2] J.P. Hirth, Metall. Trans. A 1A (1970) 2367–2374.
ε martensite was found. At room temperature, gliding and [3] P.J. Ferreira, P. Müllner, Acta Mater. 46 (1998) 4479–4484.
[4] Y.K. Lee, C.S. Choi, Metall. Mater. Trans. A 31A (2000) 355–
mechanical twinning (cf. Fig. 4(b) and (c)) enter in compe- 360.
tition, leading to a high hardening rate due to a dynamical [5] G.B. Olson, M. Cohen, Metall. Trans. A 7A (1976) 1897–
Hall-Petch effect. This hardening rate prevents necking, and 1904.
thus increases significantly the elongation to fracture. The [6] Y. Ono, T. Tsuchiyama, S. Takaki, Tetsu-To-Hagané 84 (1998)
organization of these microtwins is discussed in a separate 309.
[7] S. Cotes, M. Sade, A. Fernàndez Guillermet, Metall. Mater. Trans.
paper [26]. Neither X-ray diffraction nor TEM reveal the A 26A (1995) 1957–1969.
presence of ⑀ martensite. Finally, at 77 K, two deformation [8] M. Hillert, M. Jarl, Calphad 2 (1978) 227–238.
mechanisms are activated: dislocation gliding between the [9] A.P. Miodownik, Z. Metallkd. 89 (1998) 840–846.
ε martensite platelets, as shown Fig. 4(d). The micrograph [10] L. Li, T.Y. Hsu, Calphad 21 (1997) 443–448.
demonstrates an intense ε martensitic transformation, which [11] A.T. Dinsdale, Calphad 15 (1991) 317–425.
[12] W.S. Yang, C.M. Wan, J. Mater. Sci. 25 (1990) 1821–1823.
explains the high hardening rate but also the lower elonga- [13] K. Ishida, T. Nishizawa, Trans. Jpn Inst. Metall. 15 (1974) 225–231.
tion to fracture than at room temperature. [14] W. Huang, Calphad 13 (1989) 243–252.
[15] P.H. Adler, G.B. Olson, M. Cohen, Metall. Trans. 17 (1986) 1725–
1737.
[16] X. Wu, T.Y. Hsu, Mater. Charact. 45 (2000) 137–142.
4. Conclusion [17] Y.S. Zhang, X. Lu, X. Tian, Z. Qin, Mater. Sci. Eng. A 334 (2002)
19–27.
We have presented a model of the SFE that is based on a [18] G.L. Stepakoff, L. Kaufman, Acta Metall. 16 (1968) 13–22.
significant bibliographic study in the Fe–Mn–C system. The [19] J.F. Breedis, L. Kaufman, Metall. Trans. A 2A (1971) 2359–2371.
[20] V.H. Schumann, Neue Hütte 17 (1972) 605–609.
results are in good agreement with other authors’ works. This
[21] G.V. Raynor, V.G. Rivlin (Eds.), Phase Equilibria in Iron Ternary
modeling is then used to successfully predict the occurrence Alloys, Institute of Materials, London, UK, 1988.
of the deformation mechanisms in Fe–22 wt.% Mn–0.6 wt.% [22] Y. Tomota, M. Strum, J.W. Morris, Metall. Trans. A 17A (1986)
C austenitic steel. 537–547.
According to our measurements and data in the litera- [23] Y. Tomota, Y. Morioka, W. Nakagawara, Acta Mater. 46 (1998)
1419–1426.
ture, displacive transformations can occur during plasticity
[24] C.Y. Lim, Y.G. Kim, Mater. Sci. Eng. A 141 (1991) 67–72.
depending on the SFE; for Γ < 18 mJ mol−1 , the ε marten- [25] S. Allain, J.P. Chateau, O. Bouaziz, Steel Res. 73 (2002) 299–300.
sitic transformation occurs while for 12 mJ mol−1 < Γ < [26] S. Allain, J.P. Chateau, D. Dahmoun, O. Bouaziz, Mater. Sci. Eng.
35 mJ mol−1 mechanical twinning takes place. A, this volume.

You might also like