MATERIALS TRANSACTIONS
Online ISSN : 1347-5320
Print ISSN : 1345-9678
ISSN-L : 1345-9678
Microstructure of Materials
Influence of Morphology of Cementite on Kinetics of Austenitization in the Binary Fe–C System
Toshinobu NishibataKoutaro HayashiTakayuki SaitoManabu FukumotoMasanori Kajihara
著者情報
ジャーナル フリー HTML

2020 年 61 巻 9 号 p. 1740-1749

詳細
Abstract

When a binary Fe–C alloy with the ferrite (α) and cementite (θ) two-phase microstructure is isothermally annealed at a certain high temperature for the single-phase region of the austenite (γ) phase, the γ phase is produced at the α/θ interface by the reactive diffusion between the α and θ phases. Usually, this phenomenon is called austenitization. Owing to austenitization, the θ phase will completely dissolve into the γ phase at sufficiently long annealing times. For the flat plate of the γ phase produced between the α and θ lamellae, the one-dimensional diffusion of C occurs along the direction normal to the α/γ and γ/θ interfaces. In contrast, for the spherical particle of the θ phase distributed in the matrix of the α phase, the θ phase particle is covered with a spherical shell of the γ phase. In such a case, the three-dimensional diffusion of C in the spherical coordinate system occurs along the radial direction. The kinetics of the C diffusion is different from each other between the one-dimensional and three-dimensional coordinate systems. Consequently, the morphology of the θ phase will influence the growth behavior of the γ phase. To examine such influence, the dissolution of the θ phase was theoretically analyzed using kinetic models under various assumptions. On the basis of the analysis, the time-temperature-dissolution (TTD) diagram was constructed for each shape of the θ phase. This diagram provides quantitative information on the relationship between the dissolution time and the annealing temperature. According to the TTD diagram, the dissolution of the θ phase into the γ phase takes place much faster for the spherical morphology than for the flat one.

Fig. 14 The ratio td-S/td-P (Φ) versus the temperature T shown as dotted, dashed and solid curves for r0 = 0.15, 0.25 and 0.50 µm, respectively.

1. Introduction

Binary Fe–C alloy with eutectoid composition shows single-phase microstructure of the austenite (γ) phase at temperatures of T > Te and indicates two-phase microstructure consisting of the ferrite (α) and cementite (θ) phases at temperatures of T < Te. Here, T is the temperature, Te is the eutectoid temperature of 1000 K (727°C), and the concentration of C for the eutectoid composition is 0.76 mass% C or 3.46 at% C.1) If the eutectoid Fe–C alloy with the α + θ two-phase microstructure is isothermally annealed at a certain temperature of T > Te, the γ phase will form at the α/θ interface owing to the reactive diffusion2) between the α and θ phases. Such a phenomenon is usually called austenitization. During austenitization, the α/γ interface moves towards the α phase, and the γ/θ interface migrates towards the θ phase. At sufficiently long annealing times, the α and θ phases completely dissolve into the γ phase, and thus the γ single-phase microstructure is realized for the eutectoid alloy.

The kinetics of austenitization was investigated by many researchers.316) For instance, the dissolution of the θ phase into the γ phase in the ternary Fe–Cr–C system was experimentally observed by Liu et al.14) In their experiment, a ternary Fe–2.06 at% Cr–3.91 at% C alloy with the γ + θ two-phase microstructure was isothermally annealed at 1008 K (735°C) for various times up to 1.08 Ms (300 h). During annealing, the θ phase gradually dissolves into the γ phase. Using a kinetic model, they also quantitatively analyzed the observation.14) The quantitative analysis could satisfactorily reproduce the kinetic behavior of the θ phase dissolution. For such austenitization, the single-phase microstructure of the γ phase is realized by the dissolution of the θ phase into the γ phase. Thus, in their kinetic model,14) only the migration of the θ/γ interface was taken into consideration.

Various kinetic models for the single-interface migration10,11) were conveniently used to analyze experimental observations of austenitization.39) If the shrinkage of the α phase contributes to austenitization, however, such kinetic models cannot be reliably applicable to the quantitative analysis. The reliability of the quantitative analysis will be improved by kinetic models for the double-interface migration.4,12,13,16) When a flat plate of the γ phase forms between lamellae of the α and θ phases, the one-dimensional diffusion of C occurs along the direction perpendicular to the α/γ and γ/θ interfaces. On the other hand, spherical particles of the θ phase may be distributed in the matrix of the α phase for the eutectoid alloy with spheroidizing annealing. In such a case, each particle of the θ phase is covered with a spherical shell of the γ phase, and thus the α/γ and γ/θ interfaces possess curvature. Consequently, the three-dimensional diffusion of C in the spherical coordinate system takes place. Since the kinetics of the C diffusion is different from each other between the one-dimensional and three-dimensional coordinate systems, the morphology of the θ phase will influence the growth behavior of the γ phase. Kinetic models for the one-dimensional and three-dimensional coordinate systems were proposed by Akbay et al.16) In their study, however, attention was focused on the θ dissolution and austenitization in the one-dimensional coordinate system. Therefore, the influence of the morphology of the θ phase on the growth behavior of the γ phase is not understood well yet. In the present study, the dissolution behavior of the θ phase with planar and spherical shapes was quantitatively analyzed using different kinetic models. An analytical solution of the diffusion equation was used for the planar shape, and a numerical technique was adopted for the spherical shape. The influence of the morphology on the kinetics of austenitization was discussed on the basis of the analysis. Through the discussion, an analytical equation to describe accurately the dissolution behavior of the θ phase with the spherical shape was successfully obtained from the numerical result.

2. Methods for Analysis

2.1 Diffusion equations

Let us consider a binary Fe–C alloy with the α + θ two-phase microstructure. An example of microstructure for such an alloy is schematically depicted in Fig. 1. In this figure, a spherical particle of the θ phase with the initial radius of r = r0 is located in the matrix of the α phase. If the binary Fe–C alloy is isothermally annealed at a temperature of T > Te for a certain period of t > 0, a thin layer of the γ phase will form at the original α/θ interface. Here, r is the radius of the spherical θ phase, t is the annealing time, T is the annealing temperature, and Te is the eutectoid temperature. Owing to isothermal annealing, the γ phase grows and hence the θ phase shrinks. A phase diagram in the binary Fe–C system is schematically drawn in Fig. 2(a), and the concentration profile of C across the γ phase is shown in Fig. 2(b). In the present study, the following assumptions were adopted to simplify the analysis: (1) the local equilibrium is realized at the θ/γ and γ/α interfaces; (2) the concentrations of C in the θ and α phases are uniform and coincide with that in the θ phase at the θ/γ interface and that in the α phase at the γ/α interface, respectively; (3) the diffusion coefficient of C in the γ phase is independent of the concentration of C in the γ phase; and (4) the molar volume of each phase is constant independently of the chemical composition and the temperature and equivalent to one another among the θ, γ and α phases; and (5) the capillarity effect due to the curvature of the interface is negligible. Owing to Assumption 3, Fick’s second law is expressed by the following equation.17)   

\begin{equation} \frac{\partial c^{\gamma}}{\partial t} = \frac{D^{\gamma}}{r^{n}}\frac{\partial}{\partial r}\left(r^{n}\frac{\partial c^{\gamma}}{\partial r} \right) \end{equation} (1)
Here, cγ is the concentration of C in the γ phase, Dγ is the diffusion coefficient of C in the γ phase, r is the distance, t is the annealing time, n is the exponent of the power function rn. The dimensions of r, t, cγ and Dγ are m, s, mol/m3 and m2/s, respectively, and n is dimensionless. Furthermore, n = 0 for the diffusion of C in the one-dimensional coordinate system, n = 1 for that along the radial direction in the two-dimensional polar coordinate system, and n = 2 for that along the radial direction in the three-dimensional spherical coordinate system.17) In the present study, however, attention is focused on the cases with n = 0 and 2. The origin of r is the center of a flat sheet of the θ phase for n = 0 and that of a spherical particle of the θ phase for n = 2. We consider the diffusion of C along the direction perpendicular to the flat sheet for n = 0 and that of C along the radial direction for n = 2. Thus, r is the distance for n = 0 but the radius for n = 2. The flux balance at the moving interface is described as follows.2,1820)   
\begin{equation} (c^{\theta \gamma} - c^{\gamma \theta})\frac{\mathrm{d}r^{\theta \gamma}}{\mathrm{d}t} = - J^{\gamma \theta} = D^{\gamma}\left(\frac{\partial c^{\gamma}}{\partial r} \right)_{r = r^{\theta \gamma}} \end{equation} (2a)
  
\begin{equation} (c^{\gamma \alpha} - c^{\alpha \gamma})\frac{\mathrm{d}r^{\gamma \alpha}}{\mathrm{d}t} = J^{\gamma \alpha} = - D^{\gamma}\left(\frac{\partial c^{\gamma}}{\partial r} \right)_{r = r^{\gamma \alpha}} \end{equation} (2b)
Here, rθγ and rγα are the locations of the θ/γ and γ/α interfaces, respectively, on the r axis; cθγ and cγθ are the concentrations of C in the θ and γ phases, respectively, at the θ/γ interface with r = rθγ; cγα and cαγ are those in the γ and α phases, respectively, at the γ/α interface with r = rγα; and Jγθ and Jγα are the diffusional fluxes of C in the γ phase at the θ/γ and γ/α interfaces, respectively. Due to Assumption 2, cθ = cθγ and cα = cαγ, and thus Jθγ = 0 and Jαγ = 0. Here, cθ and cα are the concentrations of C in the θ and α phases, respectively, Jθγ is the diffusional flux of C in the θ phase at the θ/γ interface, and Jαγ is that in the α phase at the γ/α interfaces. Therefore, as shown in eqs. (2a) and (2b), the migration rate vθγ = drθγ/dt of the θ/γ interface is determined only by Jγθ, and that vγα = drγα/dt of the γ/α interface is governed merely by Jγα. The initial and boundary conditions for Fig. 2 are expressed as follows.   
\begin{equation} c^{\theta}\ (0 \leq r < r_{0},\ t = 0) = c^{\theta\gamma} \end{equation} (3a)
  
\begin{equation} c^{\alpha}\ (r_{0} < r \leq \infty,\ t = 0) = c^{\alpha\gamma} \end{equation} (3b)
  
\begin{equation} r^{\theta\gamma}\ (t = 0) = r_{0} \end{equation} (3c)
  
\begin{equation} r^{\gamma\alpha}\ (t = 0) = r_{0} \end{equation} (3d)
  
\begin{equation} c^{\theta}\ (0 \leq r < r^{\theta\gamma},\ t > 0) = c^{\theta\gamma} \end{equation} (4a)
  
\begin{equation} c^{\gamma}\ (r = r^{\theta\gamma},\ t > 0) = c^{\gamma\theta} \end{equation} (4b)
  
\begin{equation} c^{\gamma}\ (r = r^{\gamma\alpha}, t > 0) = c^{\gamma\alpha} \end{equation} (4c)
  
\begin{equation} c^{\alpha}\ (r^{\gamma\alpha} < r \leq \infty,\ t > 0) = c^{\alpha\gamma} \end{equation} (4d)
As mentioned earlier, the origin of r is the center of a flat sheet of the θ phase for n = 0 and that of a spherical particle of the θ phase for n = 2. Hence, r0 is half of the initial thickness of the flat sheet for n = 0 or the initial radius of the spherical particle for n = 2. Under the initial and boundary conditions of eqs. (3a)(3d) and (4a)(4d), respectively, eq. (1) will be solved for n = 0 and 2 by appropriate techniques in the present study.

Fig. 1

Schematic process of austenitization for spherical particle of cementite (θ).

Fig. 2

(a) Schematic phase diagram in the binary Fe–C system, and (b).

2.2 Analytical solution for one-dimensional diffusion

For one-dimensional diffusion with n = 0, an analytical solution of eq. (1) for n = 0 is obtained as follows under the initial and boundary conditions of eqs. (3a)(3d) and (4a)(4d), respectively.2,1820)   

\begin{equation} c^{\gamma} = c^{\gamma \theta} - \frac{c^{\gamma \alpha} - c^{\gamma \theta}}{\text{erf}(\lambda^{\gamma \alpha}) - \text{erf}(\lambda^{\theta \gamma})}\left\{\text{erf}(\lambda^{\theta \gamma}) - \text{erf}\left(\frac{r - r_{0}}{\sqrt{4D^{\gamma}t}} \right) \right\} \end{equation} (5)
Here, λθγ and λγα are the dimensionless proportionality coefficients for the θ/γ and γ/α interfaces, respectively. Using these parameters, rθγ and rγα are expressed as functions of t by the following equations.2,1820)   
\begin{equation} r^{\theta \gamma} = r_{0} + \lambda^{\theta \gamma}\sqrt{4D^{\gamma}t} \end{equation} (6a)
  
\begin{equation} r^{\gamma \alpha} = r_{0} + \lambda^{\gamma \alpha}\sqrt{4D^{\gamma}t} \end{equation} (6b)
As previously mentioned, r0 is half of the initial thickness of the flat sheet of the θ phase for n = 0. Inserting eqs. (5), (6a) and (6b) into eqs. (2a) and (2b), we obtain the following equations.   
\begin{equation} \omega^{\theta \gamma} = \sqrt{\pi} \lambda^{\theta \gamma}\exp \{(\lambda^{\theta \gamma})^{2}\} \{\text{erf}(\lambda^{\theta \gamma}) - \text{erf}(\lambda^{\gamma \alpha})\} \end{equation} (7a)
  
\begin{equation} \omega^{\gamma \alpha} = \sqrt{\pi} \lambda^{\gamma \alpha}\exp \{(\lambda^{\gamma \alpha})^{2}\} \{\text{erf}(\lambda^{\gamma \alpha}) - \text{erf}(\lambda^{\theta \gamma})\} \end{equation} (7b)
Here, ωθγ and ωγα are the supersaturations for the θ/γ and γ/α interfaces, respectively, defined by the following equations.   
\begin{equation} \omega^{\theta \gamma} \equiv \frac{c^{\gamma \theta} - c^{\gamma \alpha}}{c^{\theta \gamma} - c^{\gamma \theta}} \end{equation} (8a)
  
\begin{equation} \omega^{\gamma \alpha} \equiv \frac{c^{\gamma \theta} - c^{\gamma \alpha}}{c^{\gamma \alpha} - c^{\alpha \gamma}} \end{equation} (8b)
According to Assumption 1, the C concentration cφψ (φ, ψ = θ, γ, α) at the φ/ψ interface coincides with the corresponding C concentration of the two-phase φ-ψ tie-line at the annealing temperature T in the phase diagram of the binary Fe–C system. As a result, at a given value of T, ωθγ and ωγα in eqs. (7a) and (7b) are readily determined from eqs. (8a) and (8b), respectively. In such a case, λθγ and λγα are only unknown parameters in eqs. (7a) and (7b). Consequently, two unknown parameters λθγ and λγα can be evaluated from two simultaneous equations, eqs. (7a) and (7b). Since eqs. (7a) and (7b) are implicit functions of λθγ and λγα, however, the evaluation should be conducted by an appropriate numerical technique. For the numerical evaluation, the Newton-Raphson method21) was used in the present study. Inserting the evaluated values of λθγ and λγα into eqs. (6a) and (6b), we can quantitatively estimate the values of rθγ and rγα as functions of t. The migration distance lφψ of the φ/ψ interface is defined as follows.   
\begin{equation} l^{\theta \gamma} \equiv r_{0} - r^{\theta \gamma} = - \lambda^{\theta \gamma}\sqrt{4D^{\gamma}t} \end{equation} (9a)
  
\begin{equation} l^{\gamma \alpha} \equiv r^{\gamma \alpha} - r_{0} = \lambda^{\gamma \alpha}\sqrt{4D^{\gamma}t} \end{equation} (9b)
According to the concentration profile of C in Fig. 2(b) and the flux balance in eqs. (2a) and (2b), we may readily expect that the θ/γ interface migrates towards the θ phase and the γ/α interface moves into the α phase. Consequently, in eqs. (9a) and (9b), rγα > r0 > rθγ, lθγ > 0 and lγα > 0, and hence λθγ < 0 and λγα > 0. Equations (9a) and (9b) can be rewritten as follows.   
\begin{equation} (l^{\theta \gamma})^{2} = 4(\lambda^{\theta \gamma})^{2}D^{\gamma}t = K^{\theta \gamma}t \end{equation} (10a)
  
\begin{equation} (l^{\gamma \alpha})^{2} = 4(\lambda^{\gamma \alpha})^{2}D^{\gamma}t = K^{\gamma \alpha}t \end{equation} (10b)
Equations (10a) and (10b) are usually called a parabolic relationship, where Kθγ and Kγα are the parabolic coefficients. Since λθγ and λγα are dimensionless, the dimension of Kθγ and Kγα coincides with that of Dγ. On the other hand, the migration rate vφψ = drφψ/dt of the φ/ψ interface is obtained from eqs. (6a) and (6b) as follows.   
\begin{equation} v^{\theta \gamma} = \frac{\mathrm{d}r^{\theta \gamma}}{\mathrm{d}t} = \lambda^{\theta \gamma}\sqrt{\frac{D^{\gamma}}{t}} \end{equation} (11a)
  
\begin{equation} v^{\gamma \alpha} = \frac{\mathrm{d}r^{\gamma \alpha}}{\mathrm{d}t} = \lambda^{\gamma \alpha}\sqrt{\frac{D^{\gamma}}{t}} \end{equation} (11b)
Since λθγ < 0 and λγα > 0, vθγ < 0 and vγα > 0 in eqs. (11a) and (11b), respectively. Inserting rθγ = 0 into eq. (6a), we obtain the annealing time td for the complete dissolution of the θ phase in the α phase as follows.   
\begin{equation} t_{\text{d}} = \frac{1}{4D^{\gamma}}\left(\frac{r_{0}}{\lambda^{\theta \gamma}} \right)^{2} \end{equation} (12)
Hereafter, td is called the dissolution time. From eq. (12), it is obvious that td is proportional to the square of r0 for given values of λθγ and Dγ.

2.3 Approximate solution for three-dimensional diffusion

Analytical solutions of eq. (1) for three-dimensional diffusion with n = 2 under various initial and boundary conditions are extensively discussed by Carslaw and Jaeger.22) On the basis of the analytical solution for precipitation of a spherical particle in a matrix with infinite size, three different approximations were compared with the analytical solution by Aaron et al.23) They are stationary-interface, stationary-field and linear-gradient approximations. Various approximations were also proposed by many researchers.10,11,16,24,25) On the other hand, unlike the precipitation, no analytical solution of eq. (1) is known for dissolution of a spherical particle with finite size in a matrix with infinite size. Thus, stationary-field and stationary-interface approximations were used to evaluate the kinetics of the dissolution by Aaron et al.23) The following relationship is assumed for the stationary-field approximation.16,23,25)   

\begin{equation} \frac{\partial c^{\gamma}}{\partial t} = 0 \end{equation} (13)
From eqs. (1) and (13), we obtain the following equation.16)   
\begin{equation} c^{\gamma} = A\int_{r^{\gamma\theta}}^{r}\frac{\mathrm{d}r}{r^{n}} + B \end{equation} (14)
The integration constants A and B in eq. (14) for n = 2 are calculated as follows under the initial and boundary conditions of eqs. (3a)(3d) and (4a)(4d).16)   
\begin{equation} A = \cfrac{c^{\gamma \theta} - c^{\gamma \alpha}}{\cfrac{1}{r^{\theta \gamma}} - \cfrac{1}{r^{\gamma \alpha}}} \end{equation} (15a)
  
\begin{equation} B = c^{\gamma \theta} \end{equation} (15b)
From eqs. (2a)(2b), (14), (15a) and (15b), we obtain the following equations for n = 2.16)   
\begin{equation} (r^{\theta \gamma})^{3} = (1 + \varOmega)(r_{0})^{3} - \varOmega (r^{\gamma \alpha})^{3} \end{equation} (16a)
  
\begin{align} t &= \frac{1}{2D^{\gamma}\omega^{\gamma \alpha}\varOmega}\{(1 + \varOmega)(r_{0})^{2} - \varOmega (r^{\gamma \alpha})^{2} \\ &\quad - [(1 + \varOmega)(r_{0})^{3} - \varOmega (r^{\gamma \alpha})^{3}]^{\frac{2}{3}}\} \end{align} (16b)
  
\begin{equation} \frac{\mathrm{d}r^{\theta \gamma}}{\mathrm{d}t} = \frac{D^{\gamma}\omega^{\theta \gamma}r^{\gamma \alpha}}{r^{\theta \gamma}(r^{\theta \gamma} - r^{\gamma \alpha})} \end{equation} (16c)
  
\begin{equation} \frac{\mathrm{d}r^{\gamma \alpha}}{\mathrm{d}t} = \frac{D^{\gamma}\omega^{\gamma \alpha}r^{\theta \gamma}}{r^{\gamma \alpha}(r^{\gamma \alpha} - r^{\theta \gamma})} \end{equation} (16d)
Here, Ω is defined as16)   
\begin{equation} \varOmega \equiv \frac{\omega^{\theta \gamma}}{\omega^{\gamma \alpha}} = \frac{c^{\gamma \theta} - c^{\gamma \alpha}}{c^{\theta \gamma} - c^{\gamma \theta}}\frac{c^{\gamma \alpha} - c^{\alpha \gamma}}{c^{\gamma \theta} - c^{\gamma \alpha}} = \frac{c^{\gamma \alpha} - c^{\alpha \gamma}}{c^{\theta \gamma} - c^{\gamma \theta}}. \end{equation} (17)
Furthermore, like eqs. (9a) and (9b), lθγ and lγα are defined as follows.   
\begin{equation} l^{\theta \gamma} \equiv r_{0} - r^{\theta \gamma} \end{equation} (18a)
  
\begin{equation} l^{\gamma \alpha} \equiv r^{\gamma \alpha} - r_{0} \end{equation} (18b)
As mentioned earlier, r0 is the initial radius of a spherical particle of the θ phase for n = 2. In contrast, the dissolution time td for n = 2 corresponding to eq. (12) is estimated by the equation16)   
\begin{equation} t_{\text{d}} = \frac{(r_{0})^{2}}{2D^{\gamma}\omega^{\gamma \alpha}}\left[\frac{1 + \varOmega}{\varOmega} - \left(\frac{1 + \varOmega}{\varOmega} \right)^{\frac{2}{3}} \right]. \end{equation} (19)
From eq. (19), it is evident that td is proportional to the square of r0 for given values of Ω, ωγα and Dγ.

2.4 Numerical technique for three-dimensional diffusion

For n = 2, eq. (1) is rewritten as follows.   

\begin{equation} \frac{\partial c^{\gamma}}{\partial t} = \frac{D^{\gamma}}{r^{2}}\frac{\partial}{\partial r}\left(r^{2}\frac{\partial c^{\gamma}}{\partial r} \right) = D^{\gamma}\frac{\partial^{2}c^{\gamma}}{\partial r^{2}} + \frac{2D^{\gamma}}{r}\frac{\partial c^{\gamma}}{\partial r} \end{equation} (20)
Under the initial and boundary conditions of eqs. (3a)(3d) and (4a)(4d), respectively, eq. (20) can be numerically calculated using a technique reported in previous studies.2628) For the numerical calculation, a Crank-Nicolson implicit method29) was combined with a finite-difference technique.30) The calculation procedure was explained in detail elsewhere,31) where the kinetics of the reactive diffusion in various coordinate systems was theoretically analyzed by a numerical technique.

3. Results and Discussion

3.1 Temperature dependence of supersaturation

As previously mentioned, the dimension of cφ (φ = θ, γ, α) is mol/m3. Furthermore, cφ is related with the mol fraction xφ of C in the φ phase as follows.   

\begin{equation} c^{\varphi} = \frac{x^{\varphi}}{V_{\text{m}}^{\varphi}} \end{equation} (21)
Here, $V_{\text{m}}^{\varphi }$ is the molar volume of the φ phase with the dimension of m3/mol, and xφ is dimensionless. According to Assumption 4, $V_{\text{m}}^{\theta }$, $V_{\text{m}}^{\gamma }$ and $V_{\text{m}}^{\alpha }$ are constant independently of cφ and T, and $V_{\text{m}}^{\theta } = V_{\text{m}}^{\gamma } = V_{\text{m}}^{\alpha }$. Thus, cφ is automatically replaced with xφ in all the relevant equations mentioned above.

The mol fractions xφψ and xψφ for the φ/ψ tie-line in the binary Fe–C system were calculated with Thermo-Calc Ver. S using SSOL4 database in the temperature range of T = 1000–1173 K. The calculations are shown as solid curves in Fig. 3. As reported in a previous study,28) the temperature dependence of xφψ is reliably described by the following equation in the temperature range of T = 1000–1173 K.   

\begin{equation} x^{\varphi \psi} = a_{0} + a_{1}T + a_{2}T^{2} + a_{3}T^{3} \end{equation} (22)
The values of the coefficient ai (i = 0, 1, 2, 3) for xαγ, xγα, xγθ and xθγ are listed in Table 1. Utilizing the values of xαγ, xγα, xγθ and xθγ shown as the solid curves in Fig. 3, ωθγ and ωγα were calculated from eqs. (8a) and (8b), respectively. The calculated values of ωθγ and ωγα are plotted as dashed and solid curves, respectively, against the temperature T in Fig. 4. As can be seen, both ωθγ and ωγα increase with increasing temperature T. As T increases from 1000 K to 1173 K, ωθγ merely increases from 0 to 0.3. On the other hand, ωγα increases from 0 to 10 with increasing value of T from 1000 K to 1140 K. Hence, the dependence of ωγα on T is much more remarkable than that of ωθγ on T.

Fig. 3

The mol fractions xαγ, xγα and xγθ of C versus the temperature T calculated with Thermo-Calc Ver. S using SSOL4 database.

Table 1 The values of the coefficient ai (i = 0, 1, 2, 3) in eq. (22) for xαγ, xγα, xγθ and xθγ in Fig. 3.
Fig. 4

The supersaturations ωθγ and ωγα versus the temperature T shown as dashed and solid curves, respectively.

3.2 Analysis for one-dimensional diffusion

According to Assumption 3, Dγ is independent of cφ or xφ. Thus, Dγ is only a function of T. In such a case, the dependence of Dγ on T is usually expressed as follows.   

\begin{equation} D^{\gamma} = D_{0}^{\gamma}\exp \left(-\frac{Q^{\gamma}}{RT} \right) \end{equation} (23)
Here, $D_{0}^{\gamma }$ is the pre-exponential factor, Qγ is the activation enthalpy, and R is the gas constant. The values of $D_{0}^{\gamma } = 1.50 \times 10^{ - 5}$ m2/s and Qγ = 135 kJ/mol are reported by Bhadeshia and Honeycombe.32) Their parameters were adopted to estimate the value of Dγ. Using the result in Fig. 4, the dependencies of λθγ and λγα on T were calculated from eqs. (7a) and (7b), respectively. The calculations of λθγ and λγα are shown as dashed and solid curves, respectively, in Fig. 5. As T increases from 1000 K to 1173 K, λθγ decreases from 0 to −0.13. In contrast, λγα increases from 0 to 1.5 with increasing value of T from 1000 K to 1165 K. Therefore, the dependence of λγα on T is much more remarkable than that of λθγ on T. Such a tendency is attributed to the dependencies of ωθγ and ωγα on T in Fig. 4. The positive value of λγα shows the migration of the γ/α interface towards the α phase, and the negative one of λθγ indicates that of the θ/γ interface towards the θ phase.

Fig. 5

The dimensionless proportionality coefficients λθγ and λγα versus the temperature T shown as dashed and solid curves, respectively.

According to eq. (12), td is proportional to the square of r0 for given values of λθγ and Dγ. Thus, from eq. (12), td was calculated to be 1.9 s for r0 = 0.5 µm at T = 1073 K. The value r0 = 0.5 µm for n = 0 corresponds to the initial thickness of 1 µm for the flat sheet of the θ phase. Utilizing the result in Fig. 5, rθγ and rγα were calculated from eqs. (6a) and (6b), respectively, as functions of t within t = 0–1.9 s for r0 = 0.5 µm at T = 1073 K. The calculations of rθγ and rγα are indicated as solid and dashed curves, respectively, in Fig. 6. From eqs. (18a) and (18b), the thickness l for the flat layer of the γ phase is calculated as follows.   

\begin{equation} l = l^{\gamma \alpha} + l^{\theta \gamma} = r^{\gamma \alpha} - r^{\theta \gamma} \end{equation} (24)
The dependence of l on t is also represented as a dotted line in Fig. 6. In this figure, the ordinate shows the logarithms of rθγ, rγα and l, and the abscissa indicates the logarithm of t. As shown in Fig. 5, λγα is positive, but λθγ is negative. Thus, as t increases from 0 to 1.9 s, rγα increases from 0.5 to 4.5 µm, and rθγ decreases from 0.5 to 0 µm. Consequently, after isothermal annealing for t = 1.9 s at T = 1073 K, the flat sheet of the θ phase with the initial thickness of 1 µm completely disappears, and the flat layer of the γ phase grows up to 4.5 µm in thickness on both sides of the θ phase.

Fig. 6

The distances rθγ and rγα and the thickness l of the γ phase for n = 0 versus the annealing time t shown as solid, dashed and dotted curves, respectively.

As expressed by eqs. (10a) and (10b), the growth of the γ phase and the disappearance of the θ phase obey the parabolic relationship. From these equations, we obtain the following equations.   

\begin{equation} K^{\theta \gamma} = 4(\lambda^{\theta \gamma})^{2}D^{\gamma} \end{equation} (25a)
  
\begin{equation} K^{\gamma \alpha} = 4(\lambda^{\gamma \alpha})^{2}D^{\gamma} \end{equation} (25b)
Using the result in Fig. 5, Kθγ and Kγα were calculated as functions of T from eqs. (25a) and (25b), respectively, at T = 1000–1173 K. The calculations of Kθγ and Kγα are shown as dashed and solid curves, respectively, in Fig. 7. As can bee seen, both Kθγ and Kγα monotonically increase with increasing value of T. Since the absolute value is greater for λγα than for λθγ in Fig. 5, Kγα is greater than Kθγ in Fig. 7. The dependencies of Kθγ and Kγα on T are indicated as dashed and solid curves, respectively, also in Fig. 8. In this figure, the ordinate shows the logarithms of Kθγ and Kγα, and the abscissa represents the reciprocal of T. It may be anticipated that the dependencies of Kθγ and Kγα on T are expressed by the following equation of the same formula as eq. (23).   
\begin{equation} K^{\varphi \phi} = K_{0}^{\varphi \phi}\exp \left(- \frac{Q_{K}^{\varphi \phi}}{RT} \right) \end{equation} (26)
Here, $K_{0}^{\varphi \phi }$ is the pre-exponential factor, and $Q_{K}^{\varphi \phi }$ is the activation enthalpy. According to the result in Fig. 8, however, eq. (26) does not necessarily hold in the whole temperature range of T = 1000–1173 K. Such irregular temperature dependence of the parabolic coefficient is reported also for numerical analysis of carburization in the binary Fe–C system.28)

Fig. 7

The parabolic coefficients Kθγ and Kγα versus the temperature T shown as dashed and solid curves, respectively.

Fig. 8

The logarithms of Kθγ and Kγα versus the reciprocal of T shown as dashed and solid curves, respectively.

According to eqs. (25a) and (25b), the dependency of Kφψ on T is attributed to those of λφψ and Dγ on T. Thus, in the same manner as Fig. 8, the dependencies of |λθγ| and λγα on T in Fig. 5 are represented as dashed and solid curves, respectively, in Fig. 9. In this figure, the ordinate indicates the logarithms of |λθγ| and λγα, and the abscissa shows the reciprocal of T. Although both |λθγ| and λγα are monotone increasing functions of T in Fig. 5, they increase in irregular manners with increasing value of T in Fig. 9. However, the temperature dependencies of |λθγ| and λγα may be approximately expressed by the following equations in a certain temperature range.   

\begin{equation} |\lambda^{\theta \gamma}| = |\lambda_{0}^{\theta \gamma}|\exp \left(- \frac{Q_{\lambda}^{\theta \gamma}}{RT} \right) \end{equation} (27a)
  
\begin{equation} \lambda^{\gamma \alpha} = \lambda_{0}^{\gamma \alpha}\exp \left(- \frac{Q_{\lambda}^{\gamma \alpha}}{RT} \right) \end{equation} (27b)
Here, $\lambda _{0}^{\theta \gamma }$ and $\lambda _{0}^{\gamma \alpha }$ are the pre-exponential factors, and $Q_{\lambda }^{\theta \gamma }$ and $Q_{\lambda }^{\gamma \alpha }$ are the activation enthalpies. In the temperature range of T = 1063–1173 K, the pre-exponential factor and the activation enthalpy were estimated from the calculations in Fig. 9 by the least-squares method. The estimation provides $|\lambda _{0}^{\theta \gamma }| = 10.3$ and $Q_{\lambda }^{\theta \gamma } = 42$ kJ/mol for |λθγ| and $\lambda _{0}^{\gamma \alpha } = 6.40 \times 10^{3}$ and $Q_{\lambda }^{\gamma \alpha } = 81$ kJ/mol for λγα. On the other hand, the corresponding estimation at T = 1063–1173 K from the calculations in Fig. 8 gives $K_{0}^{\theta \gamma } = 6.76 \times 10^{ - 3}$ m2/s and $Q_{K}^{\theta \gamma } = 220$ kJ/mol for Kθγ and $K_{0}^{\gamma \alpha } = 2.95 \times 10^{3}$ m2/s and $Q_{K}^{\gamma \alpha } = 299$ kJ/mol for Kγα. Inserting eqs. (23), (26), (27a) and (27b) into eqs. (25a) and (25b), we obtain   
\begin{align} K^{\theta \gamma} &= 4|\lambda_{0}^{\theta \gamma}|^{2}D_{0}^{\gamma}\exp \left(- \frac{Q^{\gamma} + 2Q_{\lambda}^{\theta \gamma}}{RT} \right) \\ &= K_{0}^{\theta \gamma}\exp \left(- \frac{Q_{K}^{\theta \gamma}}{RT} \right) \end{align} (28a)
  
\begin{align} K^{\gamma \alpha} &= 4(\lambda_{0}^{\gamma \alpha})^{2}D_{0}^{\gamma}\exp \left(- \frac{Q^{\gamma} + 2Q_{\lambda}^{\gamma \alpha}}{RT} \right) \\ &= K_{0}^{\gamma \alpha}\exp \left(- \frac{Q_{K}^{\gamma \alpha}}{RT} \right), \end{align} (28b)
and thus   
\begin{equation} Q_{K}^{\theta \gamma} = Q^{\gamma} + 2Q_{\lambda}^{\theta \gamma} \end{equation} (29a)
  
\begin{equation} Q_{K}^{\gamma \alpha} = Q^{\gamma} + 2Q_{\lambda}^{\gamma \alpha}. \end{equation} (29b)
As mentioned earlier, eqs. (26), (27a) and (27b) hold merely approximately in a certain temperature range. Nevertheless, at T = 1063–1173 K, $Q^{\gamma } + 2Q_{\lambda }^{\theta \gamma } = 135 + 2 \times 42 = 219$ kJ/mol against $Q_{K}^{\theta \gamma } = 220$ kJ/mol in eq. (29a) and $Q^{\gamma } + 2Q_{\lambda }^{\theta \gamma } = 135 + 2 \times 81 = 297$ kJ/mol against $Q_{K}^{\gamma \alpha } = 299$ kJ/mol in eq. (29b). Thus, it is concluded that irregular temperature dependencies of Kθγ and Kγα in Fig. 8 are attributed to those of λθγ and λγα, respectively, in Fig. 9.

Fig. 9

The logarithms of |λθγ| and λγα versus the reciprocal of T shown as dashed and solid curves, respectively.

3.3 Analyses for one-dimensional and three-dimensional diffusion

Equation (20) for n = 2 was numerically calculated under the initial and boundary conditions of eqs. (3a)(3d) and (4a)(4d), respectively, by the technique reported in a previous study.31) One of the results is shown in Fig. 10. In this figure, like Fig. 6, the dependencies of rθγ, rγα and l on t for r0 = 0.5 µm at T = 1073 K are indicated as solid, dashed and dotted curves, respectively. As the annealing time increases from t = 0 s to t = 0.17 s, rθγ decreases from 0.5 to 0 µm, but rγα increases from 0.5 to 1.2 µm. Thus, td = 0.17 s in this case. This value is almost one order of magnitude smaller than td = 1.9 s in Fig. 6. According to eq. (24), l = rγα for rθγ = 0 µm. Hence, at td = 0.17 s, a spherical γ particle with radius of l = 1.2 µm is produced in the α matrix owing to the complete dissolution of the θ particle with r0 = 0.5 µm. The values of td with n = 2 were obtained for r0 = 0.15, 0.25 and 0.5 µm at T = 1000–1173 K in a similar manner to Fig. 10.

Fig. 10

The distances rθγ and rγα and the thickness l of the γ phase for n = 2 versus the annealing time t shown as solid, dashed and dotted curves, respectively.

On the other hand, the dissolution time td for n = 0 corresponding to eq. (19) for n = 2 is described as follows16) using the parameter Ω defined by eq. (17).   

\begin{equation} t_{\text{d}} = \frac{(r_{0})^{2}(1 + \varOmega)}{2D^{\gamma}\omega^{\gamma \alpha}\varOmega^{2}} \end{equation} (30)
For n = 0, the dependence of td on T was calculated from eq. (30) for r0 = 0.15, 0.25 and 0.5 µm at T = 1000–1173 K. The result is shown in Fig. 11(a). In contrast, Fig. 11(b) indicates the corresponding result calculated from eq. (12) using the values of λγα and λθγ in Fig. 5. Furthermore, the dependence of td on T with n = 2 was calculated from eq. (19). The result is represented in Fig. 12(a). Figure 12(b) shows the corresponding result for the numerical calculation mentioned above. In Figs. 11 and 12, the ordinate and the abscissa indicate T and td, respectively, and dotted, dashed and solid curves represent the results of r0 = 0.15, 0.25 and 0.5 µm, respectively. Hereafter, Figs. 11 and 12 are called time-temperature-dissolution (TTD) diagram. As can be seen in Fig. 11(b), td monotonically decreases with increasing temperature T. At each temperature, the smaller r0 is, the smaller td becomes. In Fig. 11(a), however, td is not a monotone decreasing function of T. Thus, eq. (30) is not reliable approximation of eq. (12). In contrast, Fig. 12(a) provides the result similar to Fig. 12(b). For each value of r0 and T, td is almost one order of magnitude smaller in Fig. 12 than in Fig. 11(b). Hereafter, td is designated td-P, td-A and td-S for Figs. 11(b), 12(a) and 12(b), respectively. Here, td-P is the td value for n = 0 calculated from eq. (12) of the analytical model, td-A is that for n = 2 estimated from eq. (19) of the approximate model proposed by Akbay et al.,16) and td-S is that for n = 2 evaluated from eq. (20) of the numerical model.

Fig. 11

The time-temperature-dissolution (TTD) diagrams for n = 0: (a) eq. (30), and (b) eq. (12).

Fig. 12

The time-temperature-dissolution (TTD) diagrams for n = 2: (a) eq. (19), and (b) eq. (20).

The ratio Ψ = td-A/td-S was calculated for each curve in Fig. 12(a) and 12(b). The results for r0 = 0.15, 0.25 and 0.5 µm are shown as dotted, dashed and solid curves in Fig. 13. In this figure, the ordinate and the abscissa indicate Ψ and T, respectively. As can be seen, all the curves coincide with one another. This means that Ψ does not depend on r0. At T = 1000 K, Ψ = 0.99. As the temperature increases, Ψ deceases and then takes the minimum value of 0.91 at T = 1100 K. On the other hand, at T > 1100 K, Ψ increases with increasing temperature and then reaches to 0.98 at T = 1173 K. Thus, unlike eq. (30) for n = 0, eq. (19) for n = 2 is rather acceptable approximation for evaluation of td. According to the result in Fig. 13, the evaluation error of eq. (19) is less than 10%.

Fig. 13

The ratio td-A/td-S (Ψ) versus the temperature T shown as dotted, dashed and solid curves for r0 = 0.15, 0.25 and 0.50 µm, respectively.

In contrast, the ratio Φ = td-S/td-P was calculated for each curve in Fig. 11(b) and 12(b). The results for r0 = 0.15, 0.25 and 0.5 µm are indicated as dotted, dashed and solid curves in Fig. 14. In this figure, the ordinate and the abscissa indicate Φ and T, respectively. Like Fig. 13, all the curves coincide with one another also in Fig. 14. Hence, Φ is insensitive to r0. Although Φ gradually increases with increasing temperature, it takes values close to 0.1 at T = 1000–1173 K. The dependence of Φ on T is expressed by the following equation of the same formula as eq. (22).   

\begin{equation} \varPhi = b_{0} + b_{1}T + b_{2}T^{2} + b_{3}T^{3} \end{equation} (31)
The values of the coefficient bi (i = 0, 1, 2, 3) estimated from the result in Fig. 14 are listed in Table 2. From eq. (12) and the definition of Φ, we obtain the following equation.   
\begin{equation} t_{\text{d-S}} = \varPhi t_{\text{d-P}} = \frac{\varPhi}{4D^{\gamma}}\left(\frac{r_{0}}{\lambda^{\theta \gamma}} \right)^{2} \end{equation} (32)
As previously mentioned, td-P and td-S are the values of td for n = 0 and 2, respectively. Combination of eq. (32) with eq. (31) provides an analytical technique for evaluation of td-S for n = 2 at T = 1000–1173 K in the binary Fe–C system.

Fig. 14

The ratio td-S/td-P (Φ) versus the temperature T shown as dotted, dashed and solid curves for r0 = 0.15, 0.25 and 0.50 µm, respectively.

Table 2 The values of the coefficient bi (i = 0, 1, 2, 3) in eq. (31) estimated from the result in Fig. 14.

3.4 Comparison with observation

The values of td in the binary Fe–C system were experimentally determined by various researchers.3,9,33) Hereafter, the experimental value of td is denoted by td-E. For instance, Li et al.33) reported the td-E value for n = 0 at T = 1073 K with r0 = 0.01 µm, Miyamoto et al.9) published that for n = 2 at T = 1073 K with r0 = 0.5 µm, and Bain3) gave out that for n = 2 at T = 1118 K with r0 = 4.5 µm. Table 3 summarizes their results. The corresponding values of td-P and td-S are also shown in Table 3. For n = 0 at T = 1073 K with r0 = 0.01 µm, td-E = 3.5 s33) and td-P = 0.76 ms. Thus, td-E is more than three orders of magnitude greater than td-P. According to the observation by Li et al.,33) the γ phase grows owing to the migration of the interface between the γ phase and the lamellar-pearlite region. Hence, under their experimental conditions,33) the growth of the γ phase occurs due to the reverse pearlite reaction but not the reactive diffusion between the α and θ phases. In such a case, the interface has to migrate over long distances to complete the reverse pearlite reaction, and thus td-E becomes much greater than td-P. In contrast, for n = 2 at T = 1073 K with r0 = 0.5 µm, td-E = 6.7 s9) and td-S = 0.17 s. Hence, td-E is more than one order of magnitude greater than td-S. For n = 2 at T = 1118 K with r0 = 4.5 µm, td-E = 8–15 s3) and td-S = 5.3 s. In this case, td-E is several times greater than td-S. For spherical particles of the θ phase dispersed in the matrix of the α phase, there exists a certain distribution of the particle size. In such a case, td-E is predominantly determined by the maximum particle size but not by the mean particle size. In Table 3, r0 = 0.5 and 4.5 µm are obtained from the mean particle size. If information on the maximum particle size is available, we may estimate a value of td-S closer to td-E.

Table 3 Experimental td-E value for n = 0 reported by Li et al.33) and those for n = 2 published by Bain3) and Miyamoto et al.9) Corresponding values of td-P and td-S are also shown.

4. Conclusions

The influence of the morphology of the cementite (θ) phase on the growth behavior of the austenite (γ) phase during austenitization of the binary Fe–C alloy with the ferrite (α) + θ two-phase microstructure was quantitatively analyzed using the different kinetic models. The various assumptions were adopted to simplify the quantitative analysis. For the flat plate of the γ phase produced between the α and θ lamellae, the one-dimensional diffusion of C takes place along the direction normal to the α/γ and γ/θ interfaces. In such a case, the analytical solution of the diffusion equation is available under the present assumptions. On the other hand, for the spherical particle of the θ phase distributed in the matrix of the α phase, the θ phase particle is covered with a spherical shell of the γ phase. As a result, the three-dimensional diffusion of C in the spherical coordinate system occurs along the radial direction. In this case, no analytical solution is known for the diffusion equation, and hence the numerical and approximate techniques were adopted for the quantitative analysis. The analysis provides the time-temperature-dissolution (TTD) diagrams for the flat sheet and the spherical particle of the θ phase. According to these TTD diagrams, the dissolution time td of the θ phase into the γ phase is much smaller for the spherical particle than for the flat sheet. Combining the analytical equation for the flat sheet with the result of the numerical calculation for the spherical particle, we can evaluate the TTD diagram for the spherical particle in an analytical manner.

REFERENCES
 
© 2020 The Japan Institute of Metals and Materials
feedback
Top