Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Physical Properties

Towards an austenite decomposition model for TRIP steels


Fateh Fazeli and Matthias Militzer
The current status of developing a fundamental model for describing the overall austenite decomposition kinetics to ferrite and carbidefree bainite in low carbon TRIP steels alloyed with Mn and Si is reviewed. For ferrite growth, a model is proposed where both interface and carbon diffusion-controlled ferrite formation are considered in a mixed-mode approach. The kinetic model is coupled with Thermocalc to obtain necessary thermodynamic information. Spherical geometry with an outer ferrite shell is assumed to capture in a simple way the topological conditions for growth. The mixed-mode modelling philosophy has been identified to permit a rigorous incorporation of the solute drag effect of substitutional alloying elements, in particular Mn. The Purdy-Brechet solute drag theory is adopted to characterize the interaction of Mn with the moving austenite-ferrite interface. The challenges of quantifying the required solute drag parameters are discussed with an emphasis on a potential solute drag interaction of Mn and Si. The model is extended to non-isothermal processing paths to account for continuous and stepped cooling occurring on the run-out table of a hot strip mill or on a continuous annealing line. The transformation start temperature during cooling is predicted with a model combining nucleation and early growth which had previously been validated for conventional low carbon steels. The overall model is evaluated by comparing the predictions with experimental data for the ferrite growth kinetics during continuous cooling of a classical TRIP steel with mass contents of 0.19 % C, 1.49 % Mn and 1.95 % Si. Extension of the model to include bainite formation remains a challenge. Both diffusional and displacive model approaches are discussed for the formation of carbide-free bainite. It is suggested to develop a combined nucleation and growth model which would enable to capture a potential transition from a diffusional to a displacive transformation mode with decreasing temperature.

Microstructure engineering is increasingly gaining attention in the steel industry to quantitatively link the operational parameters of a process (e.g. hot strip rolling, continuous annealing) with the properties of the produced steel. So far, these models have emphasized hot strip rolling of plain carbon and microalloyed steels with a ferritepearlite microstructure [1...4]. A crucial part of hot strip mill models is the run-out table component with the austenite decomposition sub-model. The significance of process modelling will be magnified for TRIP and other multiphase steels because of the tight processing windows to generate the desired complex microstructure which may include bainite, martensite and retained austenite in addition to ferrite. Traditionally, semi-empirical approaches utilizing the Johnson-Mehl-Avrami-Kolmogorov (JMAK) equation in conjunction with the additivity principle have been employed in process models which were applied to capture the ferrite and pearlite formation kinetics under industrial processing conditions. Jones and Bhadeshia [5] proposed an extended JMAK model for simultaneous transformation to several transformation products. In addition to polygonal ferrite and pearlite, they also included Widmansttten ferrite into an austenite decomposition model. However, incorporation of bainite and martensite into an overall model applicable to the industrial process is still a challenge. Further, it is crucial for TRIP steels to quantitatively account for the effects of substitutional alloying elements on the austenite decomposition including ferrite formation. Studies on laboratory alloys have suggested a retarding effect due to solute drag of alloying elements such as Mn, Si, Mo etc. [6]. More fundamental modelling approaches appear to be required to clarify detail of these effects. In binary Fe-C alloys, the kinetics of ferrite formation is

governed by carbon diffusion in remaining austenite; a variety of diffusion models have been developed [7; 8]. In binary Fe-X alloys, where X represents a substitutional alloying element, the reaction is interface-controlled for sufficiently high amounts of undercooling such that the transformation can occur without redistribution of the alloying element. Thus, in addition to diffusional approaches, mixed-mode models have been proposed which combine both carbon diffusion and interface reaction to describe the growth kinetics of ferrite in Fe-C-X alloys and in lean alloyed, low to medium carbon steels [9]. In pure diffusion models, local equilibrium concentration of carbon is assumed at the / interface, with a variety of thermodynamic conditions being proposed, i.e. orthoequilibrium, paraequilibrium and no partition local equilibrium, to describe the role of substitutional alloying elements. No partition local equilibrium has been suggested to reasonably reflect experimental growth rates [10]. For a better quantitative description of experimental results observed for continuous cooling transformation, the introduction of an effective segregation factor for alloying elements has been proposed assuming orthoequilibrium conditions at the interface [11]. Alternatively, Kop et al. [12] predicted ferrite growth in low carbon steels assuming interface control. In this approach, the interface velocity was employed as a fit parameter to describe experimental observations resulting in cooling rate dependent mobility terms. This is in contradiction to the physical definition of mobility and may mask the potential effects of alloying elements on the transformation kinetics. Mixed-mode models permit a more accurate consideration of the underlying physics since both diffusion and interface reaction are explicitly taken into account thereby allowing the incorporation of solute drag in a rigorous manner. In the present study the mixed-mode approach has been adopted to incorporate solute drag explicitly. The proposed model is applied to describe the ferrite formation kinetics in a TRIP steel. Further, approaches are delineated to de271

Fateh Fazeli, Graduate Student; Dr. Matthias Militzer, Associate Professor, The Centre for Metallurgical Process Engineering, The University of British Columbia, Vancouver, Canada

Int. Conf. on TRIP-Aided High Strength Ferrous Alloys

Physical Properties

velop an overall austenite decomposition model for TRIP steels by including the bainite portion of the reaction.

Mixed-mode model for ferrite growth


Austenite decomposition to ferrite involves the reconstruction of BCC from FCC crystal that is accomplished by the individual jumping of substitutional atoms across the interface. Simultaneously, carbon partitioning has to take place, usually at a rate much higher than the interface reaction. However, the intrinsic interfacial friction may become appreciable, in particular when substitutional alloying elements are present. According to the theory of thermally activated growth [13], the interface velocity, v, is proportional to the driving pressure for interface migration, as given by:

= M G

(1)

where M is the interface mobility and G is the Gibbs free energy difference per mole across the interface. The driving pressure, i.e. G, can be calculated based on the local chemistry at the interface. In contrast to the local equilibrium condition assumed for the diffusion-controlled models, the interfacial carbon concentration is not constant [9]. Carbon builds up at the interface as growth proceeds, it changes from the initial to the equilibrium carbon concentration of austenite, i.e. from
co to ceq . This change, c, of interfacial carbon content is

associated with a non-zero carbon net flux, Jnet, across the interface and can be expressed by:

c =

J net 1 c = v ci nt ceq + DC x int v v

(2)

is the carbon equilibrium concentration in ferwhere ceq rite, cint the carbon concentration at the austenite side of the / interface and DC denotes the diffusivity of carbon

quilibrium and paraequilibrium, may be established at the / interface. In orthoequilibrium both carbon and substitutional elements redistribute between ferrite and austenite whereas in paraequilibrium carbon redistributes without any partitioning of substitutional elements. Different growth rates of ferrite will be predicted by using each of these conditions. Thermocalc software version N with the Fe2000 database has been incorporated to perform the required thermodynamic calculations. Figure 1 schematically shows how thermodynamic assumptions affect the predictions of ferrite growth kinetics in a typical Fe-C-Mn alloy. The calculations assume planar growth geometry, the carbon diffusivity of gren [15] and an interface mobility as proposed by Krielaart et al. [9]. Planar growth geometry is fulfilled when thickening of a ferrite plate is considered. For the general case of the overall austenite decomposition kinetics, the growth geometry is governed by detail of ferrite nucleation. Transforming from undeformed austenite microstructure, ferrite nucleation preferentially occurs on the three potential grain boundary sites, i.e. grain corners, edges and surfaces. For low undercooling or sufficiently small cooling rates, corner nucleation dominates resulting in a growth geometry where a spherical ferrite particle is surrounded by the remaining austenite; the dimension of the diffusion field is determined by the final ferrite grain size. Large undercoolings or sufficiently high cooling rates lead to nucleation at the grain boundary surface sites such that the prior austenite grain boundaries are covered by ferrite plates. Planar geometry can describe the early growth stage in this situation whereas for later growth stages a spherical growth geometry is more appropriate where an exterior shell of ferrite surrounds the interior remaining austenite. Here, the diffusion field length is simply determined by the austenite grain size. Using the same parameters as for the calculations shown in figure 1 and assuming that ferrite and austenite grain sizes are equal, the growth kinetics predicted by these three assumed geometries are quite different as shown in figure 2. Spherical ferrite growing from the grain corners represents the slowest initial transforma-

in austenite. Carbon volume diffusion and interface reaction are coupled by the above equations to give the momentary interfacial carbon concentration, i.e. the growth kinetics is mixed-mode controlled. Evaluation of carbon profile in the remaining austenite is accomplished by considering both the flux due to interface migration and the flux arising from the presence of a concentration gradient, i.e.:
2 dc c dx c = + DC dt x dt x 2

(3)

To handle the problem of a moving / boundary for a finite medium, an implicit finite difference variable-grid method [14] was employed to solve the above equation where dx/dt represents the velocity of each individual node. Depending on the partitioning mode of substitutional elements, different thermodynamic situations, e.g. orthoe272

Figure 1. Effect of thermodynamic boundary condition on the growth predictions for ferrite in Fe-0.2 % C-1.5 % Mn at 730 C assuming an austenite grain size of 80 m and planar growth geometry

Int. Conf. on TRIP-Aided High Strength Ferrous Alloys

Physical Properties

concentration of solute in the boundary region, cso is the bulk composition of solute, and v is the velocity of the moving boundary. Due to an asymmetrical solute distribution, solute exerts a drag pressure, PSD, on the moving interface, which strongly depends on interface velocity and the parameters Db/ , E and Eo. The chemical potential difference of Mn in austenite and ferrite at the interface, 2 E, is a function of interfacial carbon concentration and follows from the assumed thermodynamic condition. The solute drag pressure is converted to the free energy dissipated per mole of substitutional atoms and can be calculated by [18]:

Figure 2. Effect of geometry on the growth kinetics of ferrite in Fe0.2 % C-1.5 % Mn at 730 C

Gdrag = PSDVm = cSO (C 1)


1

E dX X

(5)

tion rate while its final growth rate surpasses that of the other geometries. This comparison clarifies the significance of appropriate geometry selection in modelling the overall kinetics of austenite decomposition. Similarly, the significance of geometry assumptions had been pointed out by van Leeuwen et al. [16] for massive transformations which are interface controlled.

where Vm is the molar volume, C and X are dimensionless parameters defined by C = cs/cso and X = x/, respectively. The amount of dissipated free energy, Gdrag, is a function of the normalized velocity, V = v /Db. The free energy dissipated by drag reduces the available driving pressure at the / interface. Hence the rate of interface migration is evaluated based on the effective driving pressure; i.e.: v = M Gint Gdrag

Incorporation of solute drag effect


The retarding effect of slow diffusing elements, e.g. Mn, on the moving / interface has been attributed to solute drag [6], which reflects the apparent discrepancies observed between predicted and experimental growth kinetics. Several attempts have been made to quantify the effect of solute interaction with the / interface [11; 17]. Here, the approach of Purdy and Brechet [17] is adopted. The assumed potential well of solute inside the / interphase boundary is shown in figure 3 where 2 is the interface thickness, 2 E is the chemical potential difference of Mn in austenite and ferrite, i.e. - , and Eo is the depth of the potential well. At steady state, the net flux of solute through the boundary is zero. The velocity-dependent concentration profile of solute in the boundary region is defined from: Db cS Db cS E + + v ( cS cSO ) = 0 x RT x (4)

(6)

thereby incorporating the retarding effect of solute drag explicitly into the mixed-mode model.

Results and discussion


Modelling of ferrite. The predictions of the mixedmode model for the growth rate constant of ferrite with and without solute drag effect for orthoequilibrium and paraequilibrium are represented in figure 4 for an iron al-

where Db is solute diffusivity across the interface, cs is the

Figure 3. Chemical potential of austenite-stabilizing solute in the interphase boundary as proposed by Purdy and Brechet [17]

Figure 4. Experimental growth rate constant for Fe-0.21 % C-1.52 % Mn [10] and model predictions with and without solute drag effect (SDE) for orthoequilibrium (OE) and paraequilibrium (PE)

Int. Conf. on TRIP-Aided High Strength Ferrous Alloys

273

Physical Properties

loy with a mass contents of 0.21 % C and 1.52 % Mn where thickening rates of ferrite plates had been quantified experimentally [10]. The interfacial parameters Db/ and Eo are not well known. For the present calculations,
where DMn Db = 10 DMn is the bulk diffusivity of Mn in

ceq co

c* co

TN

1 / 2 d

ceq co dT DC ceq ceq TS

(9)

austenite [19], =1 nm and Eo = kT [17] are adopted. Based on these assumptions, paraequilibrium conditions and solute drag are in best agreement with the experimental results in terms of growth rate constant as a function of temperature. Further, the capability of the proposed model is evaluated for describing the kinetics of ferrite formation during the processing of hot-rolled TRIP steel with a mass contents of 0.19 % C, 1.49 % Mn and 1.95 % Si. The experimental data were taken from Park et al. [20], who observed a predominant ferrite microstructure for cooling rates of 1 and 5 C/s and an austenite grain diameter of 20 m. The model has been extended to continuous cooling by integrating along the time-temperature path. The calculations are performed assuming nucleation site saturation at the austenite grain boundary such that the growth geometry is represented by a spherical ferrite shell outlining the austenite grain. This situation is achieved when the fraction transformed is approximately 5 %. Thus, the transformation start temperature was taken to be the temperature where in the experiment a fraction transformed of 0.05 was recorded. A separate model is required to predict the transformation start by considering detail of the nucleation process. A model which combines nucleation and early growth had been proposed earlier to predict the transformation start temperature, TS, during continuous cooling of conventional low carbon steels [11; 21]. The transformation start; i.e. 5 % transformed, is assumed to reflect nucleation site, saturation conditions at austenite grain boundaries. For the prediction of TS, carbon diffusion controlled early growth of corner ferrite nucleated at TN is adopted such that:
dRf dT ceq co 1 = DC R dT dt ceq ceq f

Adopting the above approach and assuming orthoequilibrium conditions, TN = 810 C, and a value of 3 for the c*/co ratio, the predicted transformation start temperatures are in good agreement with the experimental measurements in the TRIP steel, as illustrated in figure 5. A good description of the experimental data can also be obtained when paraequilibrium is assumed. This confirms the finding of Tanaka et al. [22] who showed that both orthoequilibrium and paraequilibrium can be used to adequately describe ferrite nucleation in Fe-C-Mn-X alloys. Further, the employed nucleation temperature is associated with an undercooling of 62 C below the Ae3 temperature, TAe3, of 872 C. This is consistent with the findings for conventional low carbon steels and confirms the tendency that the required undercooling increases from approximately 40 to 60 C with the addition of alloying elements [21]. Similarly, the c*/co ratios observed for conventional low carbon steels increase with alloying additions, in particular Mn; i.e. from 1.3 to 2.2 when the Mn mass content is raised from approximately 0.5 % to 1.35 % [21]. The c* value determined for the TRIP steel is significantly larger than these values thereby confirming the trend with alloying elements. Interestingly, the significant increase of c* for the TRIP steel appears to be mainly attributable to the addition of Si. An increased c* is associated with a higher undercooling being required for the transformation to start. This can be interpreted in terms of a solute drag effect of these alloying additions. The current results suggest synergistic effects of Mn and Si upon nucleation and growth kinetics, which is consistent with the findings of Tanaka et al. [22]. As shown, the transformation start model is a useful tool for continuous cooling conditions but does not give any detail for the early transformation kinetics; i.e. before reaching the 5 % level of fraction transformed.

(7)

where Rf is the radius of the ferrite particle, co is the aver and ceq are the age carbon bulk concentration, and ceq equilibrium carbon concentration in ferrite and austenite, respectively, based on the employed thermodynamic condition. A limiting carbon concentration, c*, is introduced above which ferrite nucleation is inhibited. Then, the condition of nucleation site saturation is attained when:

Rf

c* co d 2 co ceq

(8)
Figure 5. Comparison of the undercooling predicted for transformation start with experimental data obtained by Park et al. [20] for a TRIP steel (Fe-0.19 % C-1.49 % Mn-1.95 % Si) with an Ae3 temperature of 872 C

where d denotes the austenite grain size. Considering a constant cooling rate and combining equations (7) and (8), the following expression is obtained:
274

Int. Conf. on TRIP-Aided High Strength Ferrous Alloys

Physical Properties

For the subsequent ferrite growth, which is described with the mixed-mode model, the solute drag parameters Db/ and Eo are employed as adjustable parameters; all other parameters are taken as aforementioned. For the solute diffusivity across the interface an Arrhenius relationship is adopted; such that the normalizing factor for the interface velocity is given by: Db Do = exp ( Q / RT ) (10)

are clearly defined in terms of their physics and, while in detail to be inferred from experimental data, the order of magnitude of their values is prescribed; Eo is of order kT and Db is a few orders of magnitude larger than the bulk diffusivity of the substitutional alloying element under consideration. Modelling of bainite. In TRIP-assisted multiphase steels the stabilization of retained austenite is mainly accomplished by carbon enrichment of austenite during its partial transformation to carbide-free bainite. The terminology carbide-free bainite is controversial as recently discussed by Hillert and Purdy [23] because the term bainite was originally introduced for transformation products with distinct morphologies which contain carbides. In TRIP steels similar morphologies are observed, however, without carbide formation. Thus, the term carbide-free bainite is frequently employed in the TRIP literature and will be used here as well. So far no significant attempt in modelling the bainite reaction for this class of steels has been reported. The first challenge in modelling the kinetics of bainite formation arises from understanding its underlying mechanisms, which have been a matter of debate for a long time. Displacive [24] and diffusional ledgewise growth [25] of bainite are the two proposed growth mechanisms. In addition to these principle questions surrounding bainite, it has to be considered that in TRIP steels bainite forms in small sized austenite grains embedded in the ferrite matrix. Transformation behaviour of these austenite grains is expected to be quite different from that of conventional austenite microstructures which usually serve as starting point in laboratory studies of bainite formation. Further, the transition temperature at which austenite decomposes to bainite rather than ferrite has to be predicted accurately for a given cooling path. The criterion for this transition depends on the bainite formation mechanism. For example, a critical driving force for displacive transformation could be assumed if a displacive approach were adopted. In case of diffusional mechanisms, the criterion of a critical interface velocity may be useful similar to that proposed for initiation of pearlite formation [26]. Minote et al. [27] discussed bainite formation for a classical TRIP steel with mass contents of 0.21 % C, 1.53 % Mn and 1.54 % Si. They suggested that the bainite kinetics can be described at higher temperatures (>350 C) adopting a diffusional approach whereas at lower temperature (<350 C) a displacive model seems to agree with the experimental data. The formal mathematics of both models is quite similar. However, the employed parameters are describing growth only in case of the diffusional model and nucleation only in case of the displacive model. The proposed transition from a growth controlled model to a nucleation controlled model as temperature is decreased suggests the development of a combined nucleation and growth model to capture the bainite kinetics for industrial processing paths. This transition would also exclude the semi-empirical JMAK approach in conjunction with the additivity principle as a generally viable model approach. Thus, it is proposed to adopt a fundamental
275

Here an activation energy Q of 132 kJ/mol is assumed which is half the value of the activation energy of Mn bulk diffusion in austenite [19] and the pre-exponential factor Do/ is taken to be 7 cm/s. For =1 nm, the ratio of diffusion across the interface to Mn bulk diffusion in austenite increases then from 50 at 800 C to 1000 at 620 C. For the potential well depth Eo (in eV), a linear temperature dependence is assumed as follows: Eo = 2.34kT 0.11 (11)

which is, as suggested by Purdy and Brechet [17], of order kT in the transformation temperature range decreasing from 1.22 kT at 800 C to 0.99 kT at 620 C. It is noteworthy that the model predictions are rather sensitive to the selection of the parameter Eo. As illustrated in figure 6, paraequilibrium together with this choice of the solute drag parameters permits an accurate description of ferrite formation for the cooling rates of 1 and 5 C/s, respectively. Without incorporating solute drag, the growth rates would be drastically overpredicted similar to what is shown in figure 4 for the Fe-C-Mn alloy. Similar to the more empirical JMAK approach, the proposed fundamental ferrite growth model requires three adjustable parameters. These parameters are associated with the solute-interface interaction; i.e. the preexponential factor in the term associated with the diffusivity across the interface and two parameters to capture the temperature dependence of the potential well depth. The advantage of this approach is then that these parameters

Figure 6. Comparison of model prediction and experimental data [20] for ferrite growth in a TRIP steel (Fe-0.19 % C-1.49 % Mn1.95 % Si) during continuous cooling at rates of 1 and 5 C/s

Int. Conf. on TRIP-Aided High Strength Ferrous Alloys

Physical Properties

modelling philosophy similar to that proposed for ferrite formation. The anticipated model will combine autocatalytic nucleation with the mixed-mode growth of bainite laths in two dimensions. Two different values for mobility to evaluate thickening and lengthening have to be assumed. A solute drag effect may also have to be incorporated. The development of such a model requires a number of careful laboratory studies characterizing the bainite formation conditions of TRIP steels.

Conclusions
The effect of substitutional alloying elements on the ferrite growth kinetics can be described by incorporating solute drag into a fundamental mixed-mode growth model. The model has three adjustable parameters which characterize detail of the solute drag phenomenon. Applying this model to continuous cooling transformation of a TRIP steel replicates consistently the experimental data with a set of physically meaningful solute drag parameters. To develop an overall austenite decomposition model for TRIP steel, a model for the bainite portion of the reaction has yet to be developed. Given the complexity of the bainite formation mechanisms, this is a challenging task. Current knowledge suggests to follow the fundamental modelling philosophy proposed for ferrite and develop a bainite model which combines nucleation and growth. This approach appears to be suitable to also provide information on detail of carbon distribution in remaining austenite which is crucial for a reliable prediction of the fraction and distribution of retained austenite in TRIP microstructures.

Acknowledgements
The authors would like to thank the Natural Sciences and Engineering Research Council of Canada for financial support. One of the authors (F. Fazeli) extends his thanks to Stelco Inc. for providing a Fellowship. The discussions with Dr. W.J. Poole are greatly appreciated.

References
[1] Hodgson, P.D.; Gibbs, R.K.: ISIJ Intern. 32 (1992), p. 1329/38. [2] Senuma, T.; Suehiro, M.; Yada, H.: ISIJ Intern. 32 (1992), p. 423/32.

[3] Samarasekera, I.V.; Jin, D.Q.; Brimacombe, J.K.: The Application of Microstructure Engineering to the Hot Rolling of Steel, Proc. 38th Mechanical Working and Steel Processing Conf., 1997, ISS, Warrendale, PA, Vol. XXXIV, p. 313/27. [4] Andorfer, J.; Auzinger, D.; Buchmayr, B.; Giselbrecht, W.; Hribernig, G.; Hubmer, G.; Luger, A.; Samoilov, A.: Prediction of the as Hot Rolled Properties of Plain Carbon Steels and HSLA Steels, Proc. Thermec97 Intern. Conf. on Thermomechanical Processing of Steels and Other Materials, 1997, TMS, Warrendale, PA, Vol. II, p. 2069/75. [5] Jones, S.J.; Bhadeshia, H.K.D.H.: Acta Mater. 45 (1997), p. 2911/20. [6] Bradley, J.R.; Aaronson, H.I.: Metall. Trans. 12A (1981), p. 1729/41. [7] Atkinson, C.; Aaron, H.B.; Kinsman, K.R.; Aaronson, H.I.: Metall. Trans. 4 (1973), p. 783/92. [8] Kamat, R.G.; Hawbolt, E.B.; Brown, L.C.; Brimacombe, J.K.: Metall. Trans. 23A (1992), p. 2469/80. [9] Krielaart, G.P.; Sietsma, J.; van der Zwaag, S.: Mater. Sci. Eng. 237A (1997), p. 216/33. [10] Purdy, G.R.; Weichert, D.H.; Kirkaldy, J.S.: Trans. TMS-AIME 230 (1964), p. 1025/34. [11] Militzer, M.; Pandi, R.; Hawbolt, E.B.: Metall. Trans. 22A (1996), p. 1547/56. [12] Kop, T.A.; van Leeuwen, Y.; Sietsma, J.; van der Zwaag, S.: ISIJ Intern. 40 (2000), p. 713/18. [13] Christian, J.W.: The Theory of Transformation in Metals and Alloys, 2nd edn., Pergamon Press, Oxford, 1982. [14] Murray, W.D.; Landis, F.: Trans. AIME 81 (1959), p. 106/12. [15] gren, J.: Scr. Metall. 20 (1986), p. 1507/11. [16] van Leeuwen, Y.; Vooijs, S.I.; Sietsma, J.; van der Zwaag, S.: Metall. Mater. Trans. 29A (1998), p. 2925/31. [17] Purdy, G.R.; Brechet, Y.J.M.: Acta Mater. 43 (1995), p. 3763/774. [18] Cahn, J.W.: Acta Met. 10 (1962), p. 789/98. [19] Oikawa, H.: Tetsu-to-Hagan 68 (1982), p. 1489/97. [20] Park, S.H.; Han, H.N.; Lee, J.K.; Lee, K.J.: Microstructural Evolution of Hot Rolled TRIP Steels During Cooling Control, Proc. 40th Mechanical Working and Steel Processing Conf., 1998, ISS, Warrendale, PA, Vol. XXXVI, p. 283/91. [21] Militzer, M.; Hawbolt, E.B.; Meadowcroft, T.R.: Metall. Mater. Trans. 31A (2000), p. 1247/59. [22] Tanaka, T.; Aaronson, H.I.; Enomoto, M.: Metall. Trans. 26A (1995), p. 547/59. [23] Hillert, M.; Purdy, G.R.: Scr. Mater. 43 (2000), p. 831/33. [24] Bhadeshia, H.K.D.H.: Bainite in Steels, 1st edn., The Institute of Materials, London, 1992. [25] Aaronson, H. I.; Reynolds, W. T. Jr.; Shiflet, G.J.; Spanos, G.: Metall. Trans. 21A (1990), p. 1343/80. [26] Militzer, M.; Pandi, R.; Hawbolt, E.B.; Meadowcroft, T.R.: Modelling the Phase Transformation Kinetics in Low-Carbon Steels, Proc. Intern. Symp. on Hot Workability of Steels and Light AlloysComposites, 1996, CIM, Montreal, p. 373/80. [27] Minote, T.; Torizuka, S.; Ogawa, A.; Niikura, M.: ISIJ Intern. 36 (1996), p. 201/07.

276

Int. Conf. on TRIP-Aided High Strength Ferrous Alloys

You might also like