Effect of Prior Austenite Grain Size On Pearlite Transformation in A Hypoeuctectoid Fe-C-Mn Steel

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/260702710

Effect of Prior Austenite Grain Size on Pearlite Transformation in a


Hypoeuctectoid Fe-C-Mn Steel

Article  in  Metallurgical and Materials Transactions A · September 2013


DOI: 10.1007/s11661-013-1996-0

CITATIONS READS
23 1,889

5 authors, including:

María M. Aranda Bij-Na Kim


Spanish National Research Council Delft University of Technology
15 PUBLICATIONS   149 CITATIONS    6 PUBLICATIONS   106 CITATIONS   

SEE PROFILE SEE PROFILE

Carlos Capdevila
Centro Nacional de Investigaciones Metalúrgicas (CENIM)
242 PUBLICATIONS   3,536 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

In-line generation of ultrafine spheroidised pearlitic microstructures in medium-high carbon steel by means of intercritical rolling View project

Divergent Pearlite in a Fe-C-Mn-Al Quaternary System View project

All content following this page was uploaded by Carlos Capdevila on 17 March 2014.

The user has requested enhancement of the downloaded file.


Effect of prior austenite grain size on pearlite transformation in a hypo-euctectoid

Fe-C-Mn steel

M. M. Aranda1, B. Kim2, R. Rementeria1, C. Capdevila1, C. García de Andrés1


1
Materalia Group, Centro Nacional de Investigaciones Metalúrgicas (CENIM-CSIC), Avda. Gregorio del

Amo, 8; 28040 Madrid, Spain.


2
Department of Materials Science and Metallurgy, University of Cambridge, Pembroke Street,

Cambridge CB2 3QZ, UK.

Abstract

The aim of this work is to evaluate the influence of the prior Austenite Grain Size (AGS) on the

austenite-to-pearlite isothermal decomposition in a Fe-C-Mn hypo-eutectoid steel. Due to the strong

influence grain boundaries have on pearlite transformation kinetics, morphological aspects of pearlite

from two conditions with very different AGS were studied and characterized. Results allow us to

conclude that the formation of pearlite and ferrite are favoured for small AGS values, whereas a

larger AGS led to an increase in the total amount of pearlite volume fraction. Furthermore, the

average size of pearlitic colonies increased with increasing AGS, and it appears that the interlamellar

spacing of the pearlite does not depend on AGS, but instead, is controlled by the isothermal

decomposition temperature. Finally, it was observed that the ratio between lamellar thickness of

ferrite and cementite depended on AGS.

Keywords: Austenite Grain Size, pearlite, hypo-eutectoid steel, transformation kinetics.


1
Introduction

Pearlite is a microconstituent formed during the eutectoid decomposition of austenite by the

cooperative growth of ferrite (α) and cementite (θ) lamellae. It was one of the first structures in

metals to be described in considerable detail, and over the years it has been the subject of many

investigations. Seminal works of Zener [1], Hillert [2-3], Mehl [4] and Avrami [5] established the basis

of thermodynamics and rate-controlling mechanism for the growth of pearlite. Besides all the work

reported, the pearlite reaction is still attracting the interest of many researchers, and nowadays,

topics such as the non-steady growth (divergent pearlite) [6-7-8] are discussed in literature.

Moreover, there is a wide variety of research on the effects of microstructural parameters on the

mechanical properties of pearlitic steels [9-10-11].

A well-known fact is that the active nucleus for pearlite formation can be either ferrite or cementite,

depending on the temperature and composition. The nucleation sites can be grain boundaries or

inclusions [12], and once either one of the constituent phases is nucleated, the conditions

surrounding the new nucleus are ripe for the nucleation of the other, and pearlite grows in a co-

operative manner.

Although the close relationship between grain-boundary area and potency of nucleation sites is well-

known, the effect of austenite grain size (AGS) on pearlite transformation has not been described in

detail yet. Therefore, the goal of this work is to study the effect of AGS on pearlite volume fraction

and morphology parameters such as interlamellar spacing and pearlite colony size.

Materials and Experimental techniques


2
The studied alloy is a hypo-eutectoid steel and the chemical composition is shown in Table 1.

The value of AGS was determined on cylindrical dilatometric test pieces of 2 mm diameter and 12 mm

length that were heat treated with an Adamel Lhomargy DT1000 high-resolution dilatometer at 5°K/s

up to 1473K (1200˚C) and 1173K (900˚C) for 180 s, followed by subsequent quenching to room

temperature. The method of thermal etching was used for revealing the prior-austenite grain [13] and

the Feret diameter was measured in order to estimate the AGS by an image analyzer program. Values

of 120 and 5 µm were obtained after austenitizing at 1473K (1200˚C) and 1173K (900˚C), respectively.

The pearlite transformation heat treatments were designed based on an isopleth diagram which was

calculated by means of ThermoCalc using the TCFE7 thermodynamic database. After austenitizing to

reach the aimed AGS, samples were quenched to different temperatures and were subsequently held

isothermally during various times, to monitor the evolution of pearlitic transformation. The samples

were mounted and polished using standard metallographic techniques and were etched with picric

acid [14]. Pearlite morphology was characterized by optical microscopy (OM) and Scanning electronic

microscopy (SEM HITACHI S-4800). The following properties were measured from scanning electron

micrographs [15]: interlamellar spacing ( ), area per unit volume of the pearlite colonies, ( ), and

the edge length of the pearlite colonies (aP).

The kinetics of pearlite transformation was studied by means of dilatometry. The pearlite volume

fraction was estimated from the change in length recorded by the dilatometric curves, which is

related to the relative volume change due to the allotropic decomposition of austenite during

isothermal holding, as indicated by the following relationships:

(1)

3
(2)

Following the method established by Onik et al. [16] and Lusk et al. [17], a forward fitting method for

fitting the kinetic parameters of internal state variable models for austenite decomposition is

proposed. Rather than converting dilatometry data to phase fraction for the purpose of fitting, the

decomposition of austenite into pearlite (for Ta<A1) can be described by the following series of

reactions: γ→α+ γ′ (reaction 1) and γ′ →α + θ (reaction 2). The relative volume change for the

austenite-to-pearlite transformation is presented as follows:

(3)

The relative volume changes are given in terms of the specific volume per Fe-atom in each phase,

which is defined as the unit-cell volume, Vi, of phase i (for i=ferrite (α), cementite (θ) and austenite

(γ)) divided by the number of Fe atoms in the unit cell. The parameter ni indicates the degree

transformation for reaction i. The unit cell volumes are calculated from the tabulated lattice

parameters in Table2, assuming that for ferrite, for austenite, and

for cementite.

With the aim of the predictions obtained for the evolution of the volume fraction of phases during

isothermal holding, a comparison has been carried out between experimental values following

standard quantitative metallographic techniques and the values derived from dilatometric curves
4
following the proposed model. The results are presented in Figure 2. As illustrated in Figure 2a, there

is a good agreement between the predicted and theoretical calculations at several temperatures,

bearing in mind the fast kinetics attained during isothermal austenite-to-pearlite decomposition at

low temperatures. A more in-detail comparison for a full transformation at 873K (600˚C) and the

theoretical values for this condition are shown in Figure 2b. Figure 2c shows the volume fraction

evolution for both ferrite and pearlite formation during austenite-to-pearlite isothermal

decomposition. Experimental metallographic data points match well with the theoretical points;

hence the factorizing method considered allows studying the evolution of the amount of ferrite and

pearlite with isothermal time with minimum metallographic observation tests.

Finally, the thickness for both ferrite and cementite were measured from scanning electron (SEM)

micrographs that were analyzed with MATLAB software (see Figure 3). SEM images of 8-bit greyscale

were obtained, where each pixel represents a value of intensity (28 values) ranging from 0 (black) to

255 (white). In order to avoid the brightness /contrast effect from image-to-image and thus apply the

same threshold level for all images, these were normalized. Each pixel was expressed within the range

from 0 to 1, where the former fits in with the minimum intensity value of the image, and the latter

with the maximum intensity, with any fractional values in between. A threshold level was determined

to binarize the images, i.e., according to their values, grey pixels were rounded off to either 1 (white,

cementite) or 0 (black, ferrite). As the etchant employed to reveal de microstructure is more sensitive

to the carbide, the cementite layers come off thinner. To balance this effect, the threshold value

selected was 0.35 so that the remaining shade surrounding the cementite is considered to belong to

the carbide. Once the images were binarized, cementite and ferrite were distinguished as different

objects and their geometrical properties were calculated.


5
Each object (cementite or ferrite) was approximated to an ellipse with minor semiaxis length a1 and

major semiaxis b1, which have the same normalized second central moments as the object. Assuming

that the objects can be approximated to rectangles with a base width of a2 and height b2, the

relationship between the rectangle and the ellipse parameters can be determined as follows. The

normalized second central moments of the ellipse and the rectangle with respect to a horizontal axis

through the centroid of the given shape are, respectively,

(4)

(5)

Therefore, . Consequently, knowing the ellipse minor semiaxis length, it is possible to

determine the width of the equivalent rectangle, hence, the average thickness of the cementite or

ferrite layers.

Results and Discussion

In hypoeutectoid steels, pearlite grows at the later stages of transformation from the already formed

proeutectoid ferrite, so the final microstructure will be a mixture of proeutectoid ferrite and pearlite.

Figure 4 illustrates the microstructure achieved after isothermal holding at different temperatures for

both AGS tested. It is clear from this figure that, both isothermal temperature and AGS have a strong
6
influence on the pearlite volume fraction formed. More precisely, Figure 5 illustrates the evolution of

proeutectoid ferrite and pearlite for AGS values of 120 and 5 µm during isothermal austenite-to-

pearlite transformation at 933K (660oC). It was noticed that there is a significant difference in the

amount of proeutectoid ferrite and pearlite at the end of the isothermal transformation. The higher

the AGS, the higher the total amount of pearlite volume fraction formed in the final microstructure.

This result is consistent with the decrease in austenite grain boundary area, and hence on nucleation

sites, as austenite grain size increases. Therefore, the nucleation sites for proeutectoid ferrite are

substantially decreased for coarse AGS, decreasing its final volume fraction. On the other hand, small

AGS promotes faster kinetics for both pearlite and ferrite formation. So, faster kinetics in small AGS is

related to more abundant nucleation sites.

Moreover, the differences in proeutectoid ferrite volume fraction will promote significant differences

in composition in the untransformed austenite before pearlite formation, as it indicated in Figure 6(a).

In this figure, it is clearly observed that while the composition of untransformed austenite for fine

AGS falls inside the Hultgren’s extrapolation, it is not the case for coarse AGS.

In this work, the austenite-to-pearlite transformation is studied during two distinct stages: nucleation

of pearlite and continuous growth of pearlite. Regarding nucleation of pearlite, since proeutectoid

ferrite is usually the first phase to develop on isothermal heat treatment in hypoeutectoid steels,

pearlite nodules nucleate on the proeutectoid ferrite-austenite (α/γ) interface. It has been observed

that the formation of pearlite requires the establishment of cooperative growth of ferrite and

cementite [12, 18]. The previous formation of proeutectoid ferrite enriches the surrounding austenite

in carbon, promoting the formation of cementite nucleus at the γ/α interface. The local reduction of

carbon content in the austenite that surrounds the cementite nucleus leads to the ferrite formation of

7
pearlite aggregate. The simultaneous ferrite and cementite formation process yields to the

characteristic lamellar structure of pearlite. Aaronson et al [19] analyzed the conditions under which

nucleation is feasible at moving disordered interface boundary: the migration rate of the γ/α

boundary at which nucleation may take place must not exceed that which displaces this boundary a

distance equal to the austenite lattice parameter in the time required for an embryo to develop to the

critical nucleus size. The critical value for a moving α/γ interface, , at which a crystal of cementite

can nucleate is:

(6)

where aγ is the austenite lattice parameter; is the carbon concentration in austenite at α/γ

interface; kB is the Boltzmann constant; T is the isothermal temperature; DCγ is the carbon diffusion

coefficient in austenite at the isothermal temperature; K is the ratio between the volume of the

double spherical cap critical nucleus and that of a sphere of the same radius; ∆GV is the volume free

energy change; cosψ is the ratio between the interfacial energies of disordered α/γ (σαγ) and α/θ

(σαθ) boundaries, and is defined as cosψ =σαγ / 2σαθ.

Assuming cementite nucleation will only take place at moving α/γ interfaces at a velocity below ,

a very simple one-dimensional model for ferrite growth might help understand the effect of AGS on

the differences attained on proeutectoid ferrite and pearlite volume fractions [20-21-22].

8
The one-dimensional growth of planar grain boundary allotriomorphs from opposite sides of an

austenite grain is illustrated schematically in Figure 6(b). This process may be considered in two

stages. The first stage involves a growth from both sides of the grain, following the assumption that

austenite has a semi-infinite extent with constant boundary conditions. In this stage, the carbon

concentration in austenite far from the α/γ interface remains the same as the overall carbon content

of the steel. During the second stage, soft-impingement occurs, where the growth rate considerably

decreases and the austenite is considered to have a finite extent beyond the α/γ interface. Therefore,

the carbon concentration in the center of the austenite grain is given by balancing the amount of

carbon enrichment in the austenite against the corresponding depletion in ferrite. It is assumed that

α/γ interface moves in z-direction normal to the interface plane, and austenite is considered to have a

finite size L in that direction. Therefore, the austenite grain diameter (dγ) could be assumed to be 2L.

The position of the interface at any time t is defined by z = Z, being Z = 0 at t = 0. In this initial state,

the carbon concentration in the austenite is uniform and corresponds to the overall carbon

composition ( ). The position of the interface at the onset of soft impingement is defined by z = Z1

and t = t1. At that moment, carbon concentration rises at every point in the austenite located ahead

of the interface. At the subsequent stage (z=Z2 and t=t2), the concentration of carbon in the center of

the austenite grain increases from to xL. Finally, when z = Z3 and t = t3, the carbon content in

austenite is uniform and equal to that of austenite at the α/γ interface (xγα).

The instantaneous interfacial carbon mass balance is described as follows:

(7)

9
Assuming that for a coarse AGS, soft-impingement regime is not reached because the nucleated

allotriomorphs are far enough to avoid the overlapping of carbon diffusion fields ahead of the moving

α/γ interface, and applying the mass conservation at both sides of the interface, the interface velocity

derived from equation (7) at the onset of the soft impingement is expressed by:

(8)

On the other hand, the interface velocity once the soft-impingement conditions are fulfilled (fine AGS)

is given by:

(9)

Therefore, in both cases, the interface velocity expressed in equations (8) and (9) are varying with the

inverse of austenite grain diameter. Thus, the interface velocity for a coarse AGS is significantly higher

than that for a fine AGS. Since cementite is not formed until the moving α/γ interface is lower

than , the thickness of the allotriomorphs before cementite nucleation is higher for ferrite formed

in fine AGS than that for coarse AGS, which is consistent with the experimental differences in the

amount of ferrite obtained.


10
On the other hand, the fact that untransformed austenite before pearlite reaction at 933K (660oC)

(Figure 6a) for a coarse AGS presents a composition outside the Hultgren’s extrapolation leads to the

interesting question about the value of carbon diffusion field ahead of the moving interface: is it

achieving the extrapolated γ/γ+θ phase boundary? According to thermodynamics, if carbon

concentration at the austenite-cementite interface is lower than the extrapolated γ/γ+θ phase

boundary, carbon should diffuse away from cementite, and cementite is expected to stop growing,

leading to the formation of so-called degenerated pearlite. However, metallographic observation of

pearlite in Figure 7 clearly shows a lamellar microstructure that might be formed with a composition

lower than that for continuous growth of pearlite indicated by γ/γ+θ phase boundary. This would only

be explained if the net flux of carbon between ferrite, cementite and austenite in the forming pearlite

is positive. This assumption needs to be validated through experimental measurements of carbon

diffusion fields ahead of the moving α/γ interface, which will be addressed in a future work.

Besides the effect of AGS on pearlite volume fraction, the fact that grain boundary nucleation is

almost suppressed for coarse AGS, has allowed us to study the effect of isothermal temperature on

pearlite volume fraction without the interference of grain-boundary nucleated phases, such as

proeutectoid ferrite and bainite. As expected, the total amount of pearlite increases with decreasing

isothermal temperature due to the increasing free energy change accompanying the transformation.

Furthermore, a fully pearlitic microstructure was achieved (Figure 8) in samples austenitizated at

1473K (1200˚C) after isothermal holding at 833 K (560oC). Calculations in Figure 9 show that the alloy

was simultaneously super saturated regarding the ferrite and cementite (black solid lines in the

figure). Therefore, and according to Hultgren’s extrapolation concept, pearlite is the only

11
decomposition product of austenite at this temperature. The result indicated in Figure 8 is a nice

experimental validation of this concept.

The isothermal transformation temperature or, more specifically, the undercooling, ΔT, has a strong

influence on the pearlite interlamellar spacing; the larger the ΔT, the smaller the interlamellar

spacing. When the interlamellar spacing is large, diffusion distances for the transport of solute are

larger, and therefore, the growth of pearlite is slowed down. The growth rate increases at higher

undercoolings since the free energy change accompanying the transformation increases. However

since the reaction is diffusion-controlled, the diffusion distances must be reduced to compensate for

the decrease in diffusivity. Consequently, the interlamellar spacing is reduced as the transformation

temperature decreases. The experimental measurements of σ0 and the edge length of the pearlite

colonies, aP, are presented as a function of isothermal decomposition temperature in Figure 10a.

In order to set up the relationship between PAGS-isothermal temperature and interlamellar spacing-

pearlitic colony size, different samples were characterized. It was found that the higher the isothermal

temperature, the higher the pearlitic colony size. However, it appears that the interlamellar spacing of

the pearlite does not depend on Tγ , but instead, is controlled by the isothermal decomposition

temperature of austenite, T. The measurements are recorded in Figure 10b. It can be observed that

the two different AGS, at the same isothermal temperature, yield almost identical values of σ0. On the

other hand, the austenite-to-pearlite transformation temperature has a significant influence on σ0, as

shown in Figure 10a, which is fully consistent with previous reported works in literature [9-10].

Another important issue on the effect of AGS on carbon enrichment of untransformed austenite is the

effect on the as-transformed pearlite morphology. The different composition of untransformed

austenite for fine and coarse AGS as indicated in Figure 6(a), might lead to microstructures with a

12
different ratio between lamellar thickness of cementite and ferrite. Table 3 lists the corresponding

values of lamellae thickness for both austenitization conditions and isothermal temperature

decomposition of 933K (660˚C). It was found that the thickness of cementite lamellar is bigger for fine

AGS, and the ratio between both ferrite and cementite is lower than that for coarse AGS (see Figure

11), which is consistent with the lower carbon content of pearlite for coarse AGS, as calculations

illustrate in Figure 6.

Conclusions

The effect of PAGS on pearlite volume fraction and on pearlitic morphology in a hypoeuctectoid Fe-C-

Mn steel during austenite-to-pearlite isothermal decomposition was studied. The main conclusions

achieved are listed as follows:

1. The PAGS affects the total volume fraction of pearlite formed. The higher the PAGS, the higher

the total amount of pearlite volume fraction formed in the final microstructure. On the other

hand, small PAGS leads to faster kinetics of both pearlite and ferrite formation.

2. Fully pearlitic microstructures were obtained in samples austenitized at 1473K (1200˚C) at

isothermal temperatures below 873K (600˚C).

3. It was found that the lamellar thickness of cementite is bigger when the austenitization

temperature is lower, and the ratio of lamellar thickness of both phases is higher as

austenitization temperature increases.

13
References

1. C. Zener: Trans AIME, 1945, vol. 167, pp. 550–595.

2. M. Hillert: Met. Trans, November 1972, vol.3, pp.2729–2741.

3. M. Hillert: Solid-Solid Phase Transformation, TSM-AIME, Warrendale, PA, USA, 1982, pp. 789-

806.

4. J. Johnson, R. Mehl: Trans AIME, 1939, vol.135, pp. 416–442.

5. M. Avrami: J.Chem Phys, 1939, vol. 7, pp. 1103-1113.

6. John W. Cahn and W.C. Hagel: Acta Metall, 1963, vol. 11, pp. 561-574.

7. C.R. Hutchinson, R.E. Hackenberg, G.J. Shiflet: Acta Mater, 2004, vol. 52, pp.3565-3585.

8. A.S. Pandit and H. K.D.H. Bhadeshia: Proc. R. Soc. A, 2011, vol. 467, pp. 2948-2961.

9. Maria das Gragas Mendes da Fonseca Games, Luiz Henrique de Almeida, Luiz Claudio F. C.

Games, and Iain Le May: Materials Characterization, 1997, vol. 39, pp. 1-14.

10. Xiaodan Zhanga, Andy Godfreya, Xiaoxu Huangb, Niels Hansenb, Qing Liuc: Acta. Mater, May

2011, vol. 59, pp. 3422–3430.

11. Akhmad A. Kordaa, Y. Mutoha, Y. Miyashitaa, T. Sadasueb: Mater. Sci. Eng. A, July 2006, vol.

428, pp. 262–269.

12. M. Hillert: Decomposition of Austenite by Diffusional Processes, pp. 197-249, Interscience,

New York, 1962.

13. C. García De Andrés, M.J. Bartolomé, C. Capdevila, D. San Martín, F.G. Caballero, V. López:

Mater Char, 2001, vol.46, pp. 389-398.

14. F.G. Caballero, C. García De Andrés, C. Capdevila: Mater Char, 2000, vol. 45, pp. 111-116.

14
15. F.G. Caballero, C. Capdevila and C. García De Andrés: ISIJ International, 2003, vol. 43, no5, pp.

726-735.

16. M. Onink, C.M. Brakrnan, F.D. Tichelaar, E.J. Mittemeijer, S. Van der Zwaag: Scripta. Metall.

Mater, 1993, vol. 29, pp. 1011-1016.

17. Mark Lusk and Herng-Jeng Jou: Metall. Mater. Trans A, 1997, vol. 287A, pp. 287-291.

18. C. Capdevila, F.G. Caballero, C.G. De Andrés: Acta Mater, 2002, vol. 50, pp. 4629- 4641.

19. H.I. Aaronson, M.R. Plichta, G.W. Franti, K.C. Russell: Metall. Trans., 1978, vol. 9A, pp. 363-371.

20. J. Gilmour, G. Purdy, J. Kirkaldy: Metal. Mater. Trans. B, 1972, vol. 3, pp. 1455-1464.

21. C. Capdevila, F.G. Caballero, C.G. De Andrés: Metal. Mater. Trans A, 2001, vol. 32, pp. 661-669.

22. F.G. Caballero, C. Capdevila, C.G. De Andrés: Script. Mater, 2000, vol. 42, pp. 1159-1165.

15
Figures

Figure 1. Isopleth diagram for the Fe-C-Mn alloy system, showing the range of isothermal decomposition

temperatures.

16
a)
1 833K570 ºC
933K660 ºC
0.8
963K690 ºC
Metallography

Vol. Fraction
0.6

0.4

0.2

0
1 10 100 1000 10000

Time / s
b) 3.0E-03

2.5E-03

2.0E-03
∆l / lo

1.5E-03

1.0E-03

5.0E-04 Experimental
Calculated
0.0E+00
0 50 100 150 200
Time / s

c) 1
0.9
0.8
0.7
Volume Fraction

0.6
ferrite
0.5
0.4 pearlite
0.3
Experimental
0.2
0.1
0
0 50 100 150 200
Time /s

Figure 2. a) Values of pearlite volume fraction at different isothermal temperatures by Lusk’s and Onik’s models

and by metallography techniques. b) Experimental and theoretical dilatometric curves at 933K (660˚C) and c)

Comparison between calculated (dashed and solid lines) and measured (red points) volume fraction of ferrite

and pearlite values.

17
a)

b)

c)

Figure 3. a) SEM micrograph of a colony of pearlite b)image of pearlite after binarization c) approximation of
the cementite to an ellipse with MATLAB software.

18
(a) (b)

(c) (d)

Figure 4. Optical micrographs showing the effect of the isothermal decomposition temperature and

PAGS on the pearlitic microstructure: a) Tγ=1473K(Tγ=1200˚C) and T=963K(T=690˚C) after holding for t=36 s, b)

Tγ=1473K (Tγ=1200˚C) and T=933K (T=660˚C) after holding for t=150 s, c) Tγ=1173K (Tγ=900˚C) and T=988K

(T=715˚C) after holding for t=1800 s and d) Tγ=1173K (Tγ=900˚C) and T=933K (T=660˚C) after holding for

t=1800 s.

19
50µm 25µm

(a) (b)

c)

Figure 5. Microstructure of the samples obtained at 933K (660˚C) and austenitized at: a) 1473K (1200˚C) and b)

1173K (900˚C). c) Volume fraction of pearlite and ferrite for two different austenitizing temperatures as a

function of decomposition time.

20
(a)

γα t=0 t1 t2 t3
x
CARBON

xL

α γ
αγ
x
Z 1 Z2 Z3 L
Distance z

(b)

Figure 6. (a) Isotherm for 933K (660˚C) indicating the untransformed austenite composition prior to pearlite

transformation for the case of the two AGS studied, and (b) schematic representation of the growth of grain

boundary allotriomorphs illustrating the carbon concentration profile in austenite and ferrite.

21
a)

b)

Figure 7. SEM micrographs illustrating the effect of AGS on pearlitic colony size for sample austenitizing at: a)

1473K (1200˚C) and b) 1173K (900˚C). In both cases, the isothermal temperature decomposition was 933K

(660˚C).

22
1.2

0.8
Vol. Fraction

0.6

0.4

0.2

0
793
520 570
843 620
843 670
943 720
993

Isothermal
Isothermal Temeprature
temperature/ K/ oC

Figure 8. Evolution of the amount of pearlite as a function of isothermal temperature (AGS= 120 µm).

23
Figure 9. Fe-C-Mn isotherm at 833K (560˚C) indicating that the alloy is simultaneously supersaturated regarding

ferrite and cementite (Hultgren’s extrapolation).

24
σ0 (µm)

a (µm)
p
773 823 873 923 973 1023 1073
Isothermal Temperature /K
Isothermal Temperature (˚C)

(a)

12
0.2
10

8 0.15
σo/ µm
ap /µm

6
0.1

4
0.05
2

0 0
1
1473K
1
1473K 1173K 1173K
Austenitization Temperature /K Austenitization Temperature /K

(b)

Figure 10. a) Variation of aP and σ0 with isothermal temperature. All the samples were austenitized at 1473K

(1200˚C), and b) variation of aP and σ0 as a function austenitizing temperature for isothermal temperature of

933K (660˚C).

25
Lamellar thickness /μm 0.200

0.150

0.100 cementita
Cementite
Ferrite
ferrita

0.050

0.000
1173K 1473K

Figure 11. Values of lamellae thickness for both austenitization conditions and isothermal temperature

decomposition of 933K (660˚C).

26
Tables

Table 1. Composition of the studied steel (at. %)

C Mn Si Fe

2 1 0.35 Balance

27
Table 2. Lattice parameters of ferrite, austenite and cementite as a function of temperature T (in K) and carbon

concentration (in atomic fraction).

Phase Lattice parameter (nm)

α a = 0.28863 x ( 1+ 17.5 x x [T-800])

γ a =( 0.36306 + 0.078χ) x {1+ (24.9-50χ) x x [T-1000]}

a = 0.45234x {1+(5.311x - 1.942x T + 9.655x )x [T-293]}

b = 0.50883x {1+(5.311x - 1.942x T + 9.655x ) x [T-293]}


θ

c = 0.67426x {1+(5.311x - 1.942x T + 9.655x ) x [T-293]}

28
Table 3. Measurements of lamellar thickness of ferrite and cementite for both austenitization temperatures.

AGS=120 µm AGS=5 µm

ferrite cementite ferrite cementite

Thickness
0.160 0.057 0.139 0.078
(µm)

Std 0.032 0.012 0.035 0.021

Ratio α/θ
2.800 1.785
thickness

29

View publication stats

You might also like