You are on page 1of 6

Scripta Materialia, Vol. 40, No. 7, pp.

831 836, 1999


Elsevier Science Ltd
Copyright 1999 Acta Metallurgica Inc.
Printed in the USA. All rights reserved.
1359-6462/99/$see front matter

Pergamon

PII S1359-6462(99)00025-1

MODELLING OF g/a TRANSFORMATION IN NIOBIUMCONTAINING MICROALLOYED STEELS


Kyung Jong Lee and Jae Kon Lee
Technical Research Laboratories, Pohang Iron and Steel Co., Ltd., P.O. Box 36,
Pohang, 790-785 Korea
(Received October 7, 1998)
(Accepted in revised form November 30, 1998)
Introduction
In Nb-containing steels where Nb is a strong carbide former throughout rolling and cooling, g/a
transformation is usually retarded. However, the formation of Nb precipitates formed during rolling
retards recrystallization of austenite, which would subsequently result in an acceleration of g/a
transformation due to reduction in the amount of solute Nb [1] as well as in the austenite grain size. The
effect of solute Nb on retardation of g/a transformation could be described as follows; (1) the activity
of carbon is reduced by solute Nb due to the strong interaction between carbon and niobium, and (2)
solute Nb is heavily segregated to g/a phase boundaries, and reduces the ferrite growth kinetics in the
intermediate transformation temperature range [2].
In Fe-C-M systems, where M is a substitutional alloying element, there is a transition from the
M-controlled slow-reaction regime with low-supersaturation to the carbon-controlled fast-reaction
regime with high-supersaturation. If M is a strong carbide forming element such as Nb, V, Ti, and Mo,
interface precipitation occurs at a proper migration rate of a g/a grain boundary, whereas matrix
precipitation in ferrite is prevailed when the growth rate is fast [2]. In the present study, the
transformation modes for ferrite, pearlite, bainite, and interface NbC precipitation were derived based
on the experimental and the theoretical approaches [3,4].
Modelling
Thermodynamic analysis of the Fe-C-Mn and Fe-C-Nb systems was conducted to understand the effect
of Mn and Nb on phase equilibria. The two-sublattice model, which was originally suggested by Hillert
and Staffansson [5], was used for the analysis. Thermodynamic data used in the present study were
obtained from the studies by Lee et al. [6] for the Fe-C-Mn system and by Ohtani et al. [7] for the
Fe-C-Nb system. Both local and para-equilibrium were applied to get phase equilibria. Phases appearing
in the Fe-rich side were classified as austenite (g), polygonal ferrite under local equilibrium (a, PFL),
polygonal ferrite under para-equilibrium (a, PFP), Widmanstatten ferrite under local equilibrium (a9,
WFL), Widmanstatten ferrite under para-equilibrium (a9, WFP), pearlite (PEP), bainite (BAI), and
alloy carbide (NbC). For Widmanstatten ferrite and bainite, additional shear energy (DGw, DGb) were
added to get the phase equilibria. The values of shear energy of 300 and 600 J/mol, which were
determined by Nanba et al. [8], were used for DGw and DGb, respectively.
831

832

g/a TRANSFORMATION

Vol. 40, No. 7

The nucleation rate (I) is expressed by the classical nucleation theory [9]:
I 5 ZbNexp(2DG*/kT)exp(2t/t)

(1)

where Z, b, DG*, and t are Zeldovich factor, frequency factor, critical driving force, and incubation
time, respectively. The shape of nuclei of ferrite, pearlite, and bainite are assumed to be pillbox. The
shape of nuclei of NbC carbide is assumed to be spherical.
It is necessary to determine whether nucleation/growth is controlled by the volume diffusion or the
interfacial diffusion of solute M or C. Nucleation of grain boundary ferrite is controlled by the
interfacial diffusion of Mn at high temperatures, while it is controlled by the volume diffusion of carbon
at low temperatures [9,10]. The solute-dragging effect due to segregation of Nb on a grain boundary is
considered in the diffusion of carbon by adding activation energy term (Qsd). The Qsd is formulated
from the solute-dragging effect of Nb on recrystallization, which was experimentally determined by
Sellars [11], multiplying by factor f. The factor f is selected as 5000 by adjusting the simulated results
with the experimental ones.
D 5 DCexp(2Qsd/RT) 5 DCexp(25000XNb(2750/T 2 1.85))

(2)

The diffusion-controlled growth rate can be solved for the flat interface and the curved interface [12]:
(C* 2 oC)/(Cp 2 C*) 5 lexp(l2)erfc(l)p0.5

(for flat interface)

(C* 2 oC)/(Cp 2 C*) 5 2l2[1 2 lexp(l2) erfc(l)p1/2]

(for curved interface)

(3)

where C, C*, CP, and dX*/dt are initial composition of solute at matrix, composition of solute in matrix
phase at interface, composition of solute in transformed phase at interface, and the moving rate of
interface, respectively. l can be solved with the degree of supersaturation of solute in order to get the
parabolic rate constant (a).

a 5 2lD0.5

(4)

For ellipsoid, the volume (V) transformed is expressed as:


V 5 4/3pabc 5 4/3p(aatn)(abtn)(actn)

(5)

where the superscript n is time exponent, and a, b, and c are thickness, length along grain face, and
length along grain edge, respectively. aa, ab, and ac are parabolic rate constants for each direction,
respectively. The growth rate along the edge direction is assumed to be 1/10 of that of the face direction.
The time exponent is formulated as a function of the amount of Nb. The overall transformation kinetics
are expressed as:
X/Xmax 5 1 2 exp[2SI(t)V(t 2 t)Dt]

(6)

where X is volume fraction transformed considering the effect of impingement, Xmax is maximum
volume fraction at a given temperature derived by the thermodynamic analysis, and V(t-t) is volume
of nuclei at time t which is formed at time t. Nucleation is assumed to stop when all the possible sites
are saturated by nucleation and growth. With the diffusional transformation, the remaining austenite is
enriched by carbon and Mn.

Vol. 40, No. 7

g/a TRANSFORMATION

833

Figure 1. Variation of phase equilibrium temperature with the Mn content at the fixed carbon content of 0.09%.

Experimental Procedure
Two steels, i.e., Nb-free steel (0.087C-1.46Mn (wt.%)) and Nb-containing steel (0.096C-1.4Mn0.022Nb), were used in this study. For convenience, the former is referred to as Steel-A, and the latter
Steel-B. Steels-A and -B were austenitized at 1000 and 1100C for 5 minutes, respectively, to eliminate
the austenite grain size effect on the g/a transformation. The measured austenite grain size of the two
steels is about 60 mm. They were quenched to temperatures between 780 to 640C by He gas, and the
isothermal transformation start (5%) and end (95%) times were measured by a transformation dilatometer. For the continuous cooling test, the specimens austenitized in the same way were quenched to
900C (cooling rate; 50C/sec) to eliminate the precipitation of carbonitrides, and were cooled to room
temperature with various cooling rates of 0.5, 1, 2, 3, 5, and 10C/sec. The continuous cooling
transformation start (5%) and end (95%) times were measured by a transformation dilatometer.
Results and Discussion
From the study of the Fe-Nb-C system, the phase boundaries determined below the Ae1 are quite similar
to the boundaries determined in the Fe-C system except the formation of Nb carbides when the Nb
concentration is lower than 0.05%. The volume Gibbs energy change for the formation of NbC can be
calculated assuming that Mn has no effect on the formation of carbides thermodynamically. From the
thermodynamic analysis of the Fe-Mn-C system, the equilibrium temperatures of Ae3 (ferrite start
temperature), Ae1 (pearlite start temperature), ACm (cementite start temperature), To (locus of G(a) 5
G(g)), TWF (Widmanstatten ferrite start temperature), and TBS (bainite start temperature) were derived
under local and para-equilibrium. Figure 1 shows the variation of the equilibrium temperatures with the
Mn content, and Figure 2 shows the isopleth section at 1.5% Mn where the ACm and the Ae3 are
extended below the Ae1. The determined equilibrium temperatures are decreased with increasing the
Mn content, and the equilibrium temperatures determined by para-equilibrium are always below the
temperatures of local equilibrium. The difference in equilibrium temperatures is increased with
increasing the Mn content.
By combining nucleation and growth kinetics, the calculated ferrite start time was compared with the
experimental dilatometric results of both steels. The experimentally determined isothermal transformation behavior was compared with the calculated one by comparing the transformation start (5%) and
end (95%) times, and the results are shown in Figures 3 and 4. Below the Ae3(para), austenite is
transformed to PFP without partitioning of Mn, and then is transformed to PFL with partitioning Mn

834

g/a TRANSFORMATION

Vol. 40, No. 7

Figure 2. Variation of phase equilibrium temperature with the carbon content at the fixed Mn content of 1.5%.

Figure 3. Comparison of the calculated TTT diagram with the experimental one in Steel-A.

Figure 4. Comparison of the calculated TTT diagram with the experimental one in Steel-B.

Vol. 40, No. 7

g/a TRANSFORMATION

835

Figure 5. Comparison of the calculated CCT diagram with the experimental one in Steel-A.

at the end of transformation. Below the TWF(para), austenite is transformed to WFP, and then
transformed to either PFL at relatively high temperatures or PFP at lower temperatures. Below the Ae1,
untransformed austenite is transformed to pearlite at the g/a interface. Below the TBS, austenite is
transformed to bainite.
At high temperatures, the interface precipitation occurs simultaneously as ferrite is formed. With
decreasing temperature, the mobility of g/a interface is larger than the nucleation rate of NbC, and the
precipitation of NbC is delayed, resulting in C-curve transformation of ferrite. As shown in Figures 3
and 4, with increasing the amount of Nb, the start of the g/a transformation is delayed. This indicates
that solute Nb has a strong effect on the start time of the g/a transformation. As Nb carbides are formed,
the amount of solute Nb is decreased, and the end time of the g/a transformation is similar, regardless
of the initial amount of Nb. In C-Mn steels, the transition of the transformation start of PFP to that of
WFP is quite continuous. However, in Nb-containing steels, a discontinuity is observed. It can be
explained by the fact that solute Nb has a stronger effect on the diffusional transformation of PF than
on the shear transformation of WFP, thereby resulting in more delay at high temperatures in Nbcontaining steels.
The experimentally determined continuous cooling transformation kinetics were compared with that
the calculated ones, and the results are shown in Figures 5 and 6. At a slow cooling rate, the present
model predicts that PFL will form first. With decreasing temperature, PFP, WFP, and PEP will
subsequently form. With increasing the cooling rate, PFP will be predominant, and the fraction of WFP
will increase. At a high cooling rate, g/a transformation will start with WFP, and will end with BAI.

Figure 6. Comparison of the calculated CCT diagram with the experimental one in Steel-B.

836

g/a TRANSFORMATION

Vol. 40, No. 7

At a slow cooling rate, when PF forms as a primary phase, NbC will form immediately after the
formation of PF. When the cooling rate is high enough to form WFP as a primary phase, the formation
of interface NbC will be relatively delayed. This is attributed to the high mobility of WFP.
Summary
The thermodynamic and kinetic behaviors of g/a transformation in Nb-free and Nb-containing steels
were modelled by classical nucleation and growth theories. By comparing with the experimental results,
the effect of Nb on the interfacial energy, diffusion coefficients, and growth rate of reactions were
determined. Using the model developed, the kinetics of g/a transformation and the interface NbC
precipitation were predicted. The predicted g/a transformation in the Nb-free steel was faster than that
in the Nb-containing steel, and was in reasonable agreement with the experimental results during both
isothermal and continuous cooling tests. It was found from the analyses that solute Nb restrained the
reaction of ferrite and pearlite which required long-range diffusion of solute more than the reaction of
Widmanstatten ferrite and bainite.
References
1.
2.
3.

S. Okaguchi, T. Hashimoto, and H. Ohtani, THERMEC 88, ed. I. Tamura, ISIJ, p. 330 (1988).
T. Abe, H. I. Aaronson, and G. J. Shiflet, Metall. Trans. A. 16A, 521 (1985).
K. J. Lee, K. B. Kang, J. K. Lee, O. Kwon, and R. W. Chang, in Proceedings of the International Symposium on
Mathematical Modelling of Hot Rolling of Steels, ed. S. Yue, Canadian Mining and Metallurgy, p. 435 (1990).
4. K. J. Lee, K. B. Kang, J. K. Lee, and O. Kwon, ISIJ Int. 32, 326 (1992).
5. M. Hillert and L. Staffansson, Acta Chem. Scand. 24, 3618 (1970).
6. B. J. Lee and D. N. Lee, CALPHAD. 13, 355 (1989).
7. H. Ohtani, M. Hasebe, and T. Nishizawa, CALPHAD. 13, 183 (1989).
8. S. Nanba, M. Katsumata, T. Inoue, S. Nakajima, G. Anan, A. Hiramatsu, A. Moriya, T. Watanabe, and M. Umemoto,
CAMP-ISIJ. 3, 871 (1990).
9. W. F. Lange, M. Enomoto, and H. I. Aaronson, Int. Mater. Rev. 34, 125 (1989).
10. M. Enomoto and H. I. Aaronson, Metall. Trans. A. 18A, 1547 (1987).
11. C. M. Sellars, in HSLA Steels Metallurgy and Application, ed. J. M. Gray et al., p. 73, ASM, Metals Park, OH (1986).
12. E. S. K. Menon, R. W. Hyland, and H. I. Aaronson, Scripta Metall. 18, 367 (1984).

You might also like