Вы находитесь на странице: 1из 18

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/264427526

The finite element analysis of austenite


decomposition during continuous cooling in
22MnB5 steel

Article in Modelling and Simulation in Materials Science and Engineering July 2014
Impact Factor: 2.17 DOI: 10.1088/0965-0393/22/6/065005

CITATIONS READS

2 287

5 authors, including:

Xiangjun Chen Namin Xiao


Hunan University Chinese Academy of Sciences
8 PUBLICATIONS 4 CITATIONS 26 PUBLICATIONS 337 CITATIONS

SEE PROFILE SEE PROFILE

Guangyong Sun
Hunan University
88 PUBLICATIONS 1,200 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate, Available from: Guangyong Sun
letting you access and read them immediately. Retrieved on: 03 June 2016
Home Search Collections Journals About Contact us My IOPscience

The finite element analysis of austenite decomposition during continuous cooling in 22MnB5

steel

This content has been downloaded from IOPscience. Please scroll down to see the full text.

2014 Modelling Simul. Mater. Sci. Eng. 22 065005

(http://iopscience.iop.org/0965-0393/22/6/065005)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 129.78.56.208
This content was downloaded on 21/08/2014 at 07:40

Please note that terms and conditions apply.


Modelling and Simulation in Materials Science and Engineering

Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 (16pp) doi:10.1088/0965-0393/22/6/065005

The finite element analysis of austenite


decomposition during continuous cooling
in 22MnB5 steel
Xiangjun Chen1,2 , Namin Xiao2 , Dianzhong Li2 ,
Guangyao Li1 and Guangyong Sun1
1
State Key Laboratory of Advanced Design and Manufacturing for Vehicle Body,
Hunan University, Changsha 410082, Peoples Republic of China
2
Shenyang National Laboratory for Materials Science, Institute of Metal Research,
Chinese Academy of Sciences, Shenyang110016, Peoples Republic of China

E-mail: gyli@hnu.edu.cn and nmxiao@imr.ac.cn

Received 26 December 2013, revised 22 May 2014


Accepted for publication 18 June 2014
Published 31 July 2014

Abstract
The hot stamping process has been increasingly used in newly designed vehicles
to improve crash worthiness and fuel efficiency. In this study, a finite element
model based on a subroutine of the commercial software ABAQUS is developed
to predict the interactive influence of temperature field and phase transformation
on high-strength boron steel. JMAK-type equations with the incubation time
and additivity hypothesis are adopted to describe the austenite decomposition
into ferrite, pearlite and bainite, while the Koistinen and Marburger (KM)
model is used to describe the displacive transformation of matensite. The
simulation results show that the introduction of incubation time into a JMAK
equation can provide a more reasonable prediction of the transformation kinetics
than if the equation is unmodified. A comparison between the simulation and the
standard Jominy end-quenching test demonstrates the capability of the present
model for the prediction of transformation kinetics, microstructure distribution
and mechanical properties. Furthermore, the adoption of experimentally
measured microhardness values for the individual phase constituent can produce
improved accuracy of the hardness predictions compared to the empirical
hardness equations.

Keywords: phase transformation, 22MnB5, JMAK model, Jominy end-


quenching test, hot stamping
(Some figures may appear in colour only in the online journal)

0965-0393/14/065005+16$33.00 2014 IOP Publishing Ltd Printed in the UK 1


Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

1. Introduction

Ultra-high-strength steels play a significant role in the reduction of vehicle size and the
improvement in passive safety in the automotive industry. One technology to produce
open-section ultra-high-strength steel components is the hot stamping technique using boron
steels. Different from cold stamping, the blanks are austenitized and subsequently formed
and quenched in the die tools during the hot stamping process. Clearly, hot stamping is a
typical thermomechanical process accompanied by interactions among the temperature, phase
transformation and mechanical response. During the continuous cooling in the forming tools,
the products of austenite decomposition include the ferrite, pearlite, bainite or martensite,
according to its thermal history. The type of microstructure determines the final mechanical
properties, so the high-accuracy prediction of austenite decomposition kinetics is meaningful
in designing the hot stamping process.
In the past, empirical modeling of phase transformations during the cooling of steels
has received considerable attention due to its good usability in industry applications. Due
to differences in the transformation mechanisms, the diffusional transformation and the
displacive transformation are interpreted separately by the different mathematical models.
For the diffusional phase transformation including ferrite, pearlite and bainite, some of the
most commonly used models have been implemented to calculate the phase transformation
kinetics in the arbitrary cooling paths. Woodard et al [1] originally adopted the Scheil
additivity hypothesis to calculate the incubation period of nucleation during continuous cooling
from the isothermal incubation. Since then, the isothermal transformation kinetics with the
additivity rule has been wildly adopted to the nucleation and overall transformation kinetics.
Furthermore, the isothermal kinetics of diffusional transformation in steels can be described
by Avrami [2] type formulas or Hillert [3] type formulas. Accordingly, two kinds of empirical
mathematical models with the additivity hypothesis are proposed, i.e. the JohnsonMehl
AvramiKolmogorov (JMAK) type model [2] and the KirkaldyVenugopalan [4] (KV) type
model. Since these pioneering works, many investigations have been made into modeling the
phase transformation to improve the accuracy of numerical simulations. For example, the basic
JMAK equation is modified to consider the effect of prior austenite grain size and the external
stress. In recent years, the phase transformation calculation has been coupled with the hot
stamping simulation to predict the final mechanical response of high-strength steel. Most of
these studies are based on the KV type model. In the KV model, the austenite decomposition
reaction rate is written in a general form as dXi /dt = f (T )f (Xi )/ [f (G)f (C)] [5]. In
this equation, f indicates a general functional relation that describes the effects of different
factors on the transformation, e.g. the austenite grain size f (G), alloy composition f (C),
temperaturef (T ) and the current fraction formed f (Xi ). The empirical parameters in
these functions are calibrated by the time-temperature-transformation (TTT) or continuous
cooling transformation (CCT) curves. kerstrom [6] developed an austenite decomposition
model of boron steel based on the material subroutine of finite element code LS-DYNA.
Their model modified the original KV kinetics model to incorporate the boron effects in
the ferrite, pearlite and bainite transformations by including a boron term in the functional
expression of the reaction rate. The simulation results show an acceptable agreement with
the experimental temperature histories, hardness profiles and volume fractions of different
transformation products. Bok et al [7] compared the three KV type phase transformation
models using continuous cooling dilatometry and formed experiments of a modified B-
pillar part, which shows that Lis modified model [5] provided the best prediction of the
hardenability and hardness values. The empirical Koistinen and Marburger (KM) model [8]
has been widely adopted to describe the displacive phase transformation kinetics, e.g. austenite

2
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

decomposition into martensite. Carlone [9] also adopted a parabolic law to handle complete
austenite decomposition into martensite. Recently, Lee [10] developed a new martensite
transformation kinetics equation by considering the effects of the austenite grain size and
chemical composition.
In this paper, we focus on simulating the influence of temperature histories on the phase
transformation kinetics of high-strength boron steel, which has been widely adopted as a
basic material of hot stamping products. JMAK-type equations with the incubation time and
additivity hypothesis are adopted to describe the austenite decomposition into ferrite, pearlite
and bainite during continuous cooling, while the KM model is calibrated by the experiments
to the martensite transformation. The model is developed based on the user subroutine of the
commercial FE software ABAQUS. The effect of the incubation time in the JMAK equation
on the transformation kinetics is discussed in detail. The Jominy end-quenching experiment
is carried out to verify the prediction of the temperature histories and phase volume fractions.
The hardness profile is selected as the mechanical property to be evaluated, because it is a
common measurement used to test part quality in actual production.

2. Computational model

2.1. Diffusional phase transformation model

Under the isothermal condition, the JMAK equation has the following general form [2] to
describe the relationship between the transformed time and the phase volume fraction:
F = 1 exp[b t n ]. (1)
Considering the incubation time, the modified form derived from Hawbolt et al [11] is expressed
as follows:
F = 1 exp[b(t ts )n ], (2)
where F indicates the phase volume fraction, t indicates the elapsed time from the beginning
of the transformation, and ts indicates the real starting time of the phase transformation, i.e. the
incubation time for the isothermal transformation. b is a constant depending on the temperature,
composition of the parent phase and the grain size, and n is also a constant dependent on the
mechanism of phase transformation. For the basic form, b and n can be directly obtained from
any two points on the measured isothermal transformed kinetics curve as follows:
ln(1 Fa% )
b= (3)
(ta% ts )n
ln[ln(1 Fa% )/ ln(1 Fb% )]
n= . (4)
ln[(ta% ts )/(tb% ts )]
In most cases of diffusional austenite decomposition, the kinetics of ferrite, pearlite and bainite
transformation is dependent primarily upon some microstructure features, e.g. the local carbon
centration of austenite, the prior austenite grain size and the dissolution of second particles.
In this study, it is assumed that the austenite is completely homogenous while the effects of
second particles also can be neglected for the 22MnB5 steel. As far as the austenite grain size
is concerned, the finer prior austenite grain size will provide a more effective ferrite nucleation
site and then accelerate the transformation kinetics. According to the study of Umemoto
et al [12], a modified term can be introduced into the parameter b when considering the effect
of prior austenite grain size on the transformation kinetics:
   
DT T T m
F = 1 exp b (t ts ) ,
n
(5)
D
3
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

Figure 1. Schematic TTT diagram using the JMAK equation and additive rule.

where DTTT is the austenite grain size used to construct the TTT diagrams, D is the austenite
grain size before the transformation and m is a constant equal to 1 for ferrite and 2 for pearlite
transformations.
Scheils additivity hypothesis is adopted to describe the non-isothermal transformation
behavior [1]. Briefly, in the FE simulation, the transformation time tj is taken as the sum of
the time interval tj and the time obtained based on the volume fraction of each phase at the
previous time step F j 1 , according to the JMAK equation, as follows:
 1
ln(1 F j 1 ) n
tj = tj + . (6)
b
Based on this transformation time, the volume fraction of each phase can be calculated using
equation (2). Both the nucleation and growth stages of non-isothermal transformations are
determined by Scheils additivity rule. The cooling curve is subdivided into a number of small
time steps as shown in figure 1. Then, ti /i is calculated and summed for each time step and
the transformation is assumed to begin when the following equation is fulfilled:

m
ti
= 1, (7)
i
i

where i is the TTT incubation time or the time required to transform the fraction isothermally
at the current isotherm Ti .

2.2. Calculation of incubation time

The incubation time during isothermal soaking can be calculated using any three points on
the TTT curve, e.g. t1% , t50% and t99% at the specific temperatures shown in figure 1. By
substituting t1% , t50% and t99% into equation (2), respectively, ts can be obtained:
 
1
ts = t1% exp (ln t50% k ln t99% ) (8)
1k
4
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

1  
ts t1% exp ln(t50% ts ) k ln(t99% ts ) , (9)
1k

where k = 0.779076. Clearly, the incubation time can be evaluated using the dilation
experiment of isothermal transformation kinetics.

2.3. Martensite transformation model

For the case of diffusionless martensite transformation, the volume fraction of martensite is
calculated using the empirical model established by Koistinen and Marburger [8]. The amount
of martensite can be totally expressed as a function of the temperature as follows:
Fm = Fa {1 exp[0.011 (Ms T )]}, (10)
where Fm is the volume fraction of the martensite and Fa is the volume fraction of austenite
remaining from the previous transformations. Ms and T are the martensite starting temperature
and the current temperature, respectively. It should be noted that the Ms temperature is
dependent on the prior austenite grain size and the previous transformed products. For example,
Guimaraes and Shyne [13] revealed that the Ms temperature increased with increasing austenite
grain size, especially in FeNiC alloys. So the value of Ms is calibrated by the experimentally
dilatational curves with the prior austenite grain size measurement for better accuracy in the
simulation.

2.4. Heat transfer model

The heat conduction in the sample during quenching can be calculated as follows:
T
cp = T + Q, (11)
t
where is the conductivity, is the density, cp is the heat capacity, T is the temperature and Q
is the internal heat generation. Note that Q includes the heat generated by external heat sources
as well as transformed latent heat Qth . The latent heat due to solid phase transformation has
been taken into account as
Fi
Qth = Hi , (12)
t
where Hi is the heat generated due to transformation, Fi is the transformed volume fraction
and t is the duration of the time step.
During the quenching process, the heat exchange occurs between the sample and the
quenching medium, which can be written as



= Hc (Ts Tw ) + Hs Ts4 Tw4
n 
 
= Hc + Hs Ts2 + Tw2 (Ts + Tw ) (Ts Tw )
= H (Ts Tw ) , (13)
where n is the outer normal of the boundary surface,  is the face of the sample, Ts and Tw are
the samples temperature and the quenching medium temperature, respectively, Hc and Hs are
the convection coefficient and the radiation coefficient and H = Hc + Hs (Ts + Tw )(Ts2 + Tw2 )
is the integrated heat transfer coefficient (IHTC).
5
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

Table 1. Chemical composition for 22MnB5 in weight percent.

Component C (%) Si (%) Mn (%) P (%) S (%) Cr (%) Ni (%) B (%)


Measured 0.27 0.29 1.25 0.007 <0.005 0.22 0.013 0.0039

2.5. Hardness calculation

The hardness distribution along the specimen was calculated based on the mixture rule
f+p+b+m
Hv = (Xf + Xp ) Xvf + Xb Hvb + Xm Hvm . Hv is the hardness in Vickers; Xm ,
Xb , Xf and Xp are the volume fraction of martensite, bainite, ferrite and pearlite, respectively;
f+p
and Hvm , Hvb and Hv are the hardness of martensite, bainite and the mixture of ferrite and
pearlite, respectively. Maynier et al [14] proposed empirically based formulas as functions of
the steel composition for calculating the individual phase:
Hvm = 127 + 949C + 27Si + 11Mn + 8Ni + 16Cr + 21 log Vr (14)
Hvb = 323 + 185C + 330Si + 153Mn + 65Ni + 144Cr + 191Mo
+ (89 + 53C 55Si 22Mn 10Ni 20Cr 33Mo) log Vr (15)
Hvf+p = 42 + 223C + 53Si + 30Mn + 12.6Ni + 7Cr + 19Mo
+(10 19Si + 4Ni + 8Cr + 130V) log V r, (16)
where Vr is the cooling rate at 700 C in degrees Celsius per hour. It should be noted that
the effect of the boron content is not considered in the above equations. Bok et al [7] carried
out microhardness experiments and measured the Vickers hardness of the individual phase in
f+p
boron steel, i.e. Hvm = 514.4, Hvb = 283.8 and Hv = 188.5. The experimentally measured
hardness values are also adopted in this paper for comparison with the empirical modeling.

3. Experiments and modeling parameters

The steel used in this study is boron steel 22MnB5, which is widely used in the hot stamping
process. The measured chemical composition is shown in table 1.
The TTT curve and the critical transformation temperatures are measured in the
dilatometry test using the LINSEIS RITA quenching dilatometer L78. During the measurement
of the TTT curve, the specimens are machined as a cylinder (3 mm10 mm) and austenitized
at 900 C for 5 min with a heating rate of 2 C s1 . The prior austenite size (DTTT ) of 75 m
can be obtained after austenitization. Then the specimens are rapidly cooled to the specified
temperatures with a cooling rate of 100 C s1 and held for a specified time. The measured
TTT curves are shown in figure 2. The reference transformation temperatures are required to
start the conversion of isothermal to anisothermal kinetics based on the additivity principle.
These measured values, including Ae3 , A1 , Bs , Ms and Mf , are listed in table 2.
Based on the TTT curve, the incubation time of diffusional phase transformation then can
be obtained according to equation (9). For convenience of the numerical implementation in
the FE simulation, non-linear polynomial fitting of discrete experimental values is adopted to
build the relationship between the temperature and incubation time, as shown in figure 3.
It should be noted that the pearlite transformation occurs when the proeutectoid ferrite
precipitation is still in progress in this case. So it is difficult to distinguish the pearlite starting
point using only the dilation test. In this model, the formation of ferrite and pearlite is
considered to be two stages of one transformation process. Only one JMAK equation is
adopted to calculate the total volume fraction of proeutectoid ferrite and pearlite. Following

6
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

750

700 1%
50%
650 A+F+P 99%

Temperature (oc)
600
A
F+P
550

500
A+B
450 B

400 M
350
10 100 1000
Time (s)

Figure 2. The TTT diagram for the 22MnB5. The letters A, F, P, B and M indicate
austenite, ferrite, pearlite, bainite and martensite, respectively.

Table 2. The material parameters of 22MnB5 used in the simulation.

Reference temperatures ( C) Ae3 = 851 A1 = 718 Bs = 610 Ms = 420 Mf = 210


Latent heat (108 J m3 ) QF = 5.9 QP = 6.0 QB = 6.2 QM = 6.56
Austenite size (m) DTTT = 75 D = 122

750

700
Temperature (C)

650

600
Ferrite
550 Bainite
Virtual point--Bainite
500

450

400
0 20 40 60 80
Time (s)

Figure 3. The calculated incubation time of ferrite and bainite during the isothermal
soaking in 22Mnb5 steel based on the TTT curve.

the model proposed by Pan and Gu [15], the maximum ferrite fraction can be calculated
based on the lever rule of the phase diagram. Furthermore, the quantitative measurement
of the ferrite fraction from the microstructure metallograph of the isothermal transformation
also provides this maximum value. If the total fraction is greater than the maximum ferrite
fraction, the difference between them will be equal to the pearlite fraction. In this model,
the isothermal treatment at 600, 625 C, 650 C, 675 C and 700 C will get a maximum
ferrite fraction of 5%, 9%, 14%, 38% and 90%, respectively, according to the lever rule and
the quantitative metallographic measurement. On the other hand, the mixed microstructure of

7
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

0.008 4.0

0.007 3.5
n--Ferrite
0.006 n--Bainite 3.0

0.005 2.5

Parameter --b

Parameter --n
0.004 2.0

0.003 1.5

0.002 1.0
b--Ferrite
0.001 b--Bainite 0.5

0.000 0.0
400 450 500 550 600 650 700 750 800

Temperature (C)

Figure 4. Parameters b and n in the JMAK equation as function of temperature.

1000

950 Specific Heat 50

Thermal Conductivity (Wm-1K-1)


Specific Heat (10-6m2s-1)

900 Thermal Conductivity


40
Martensite
850
+Bainite Ferrite+Pearlite
800 30

750

700 Austenite 20

650
Austenite 10
600 Martensite+Bainite
550 0
Ferrite+Pearlite
500
0 200 400 600 800 1000

Temperture (C)

Figure 5. The specific heat and thermal conductivity for 22MnB5.

ferrite, pearlite and bainite can form at the intermediate temperature region (about 575600 C)
during isothermal soaking. For competitive pearlite and bainite transformation, identification
of the starting time of bainite is still a challenge for both theoretical predictions and experimental
observations [16]. So the incubation time of bainite in this special temperature region will be
set up artificially, and then calibrated using the inverse calculation in the subsequent prediction
and comparison with the experimental CCT curves (the marked virtual point in figure 3). The
kinetic parameters in the JMAK equation also can be obtained from the experimental TTT
curve. Figure 4 shows the kinetic parameters b and nof ferrite and bainite as a function of
temperature.
In this paper, both thermal conductivity and specific heat measurements are conducted
based on the laser pulse method and the reference benchmark specimen using the FlashlineTM-
5000 Thermal Properties Analyzer of ANTER Inc. The measured thermal conductivity
and specific heat values with individual phase microstructure, e.g. the austenite in the high
temperature range and the martensite and bainite in the low temperature range, are adopted.
Then these measured values are interpolated and extended to those temperature points
undergoing phase transformation, as shown in figure 5. During the actual calculation, the
simple linear mixture rule is applied for the multi-phase case. The heat transfer coefficient of
air cooling and water cooling are obtained from the experiments, as shown in figure 6. Latent

8
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

Heat transfer coefficient of water cooling(W/m2K)


Heat transfer coefficient of air cooling(W/m2K)
120 25000

100 air cooling


20000
water cooling
80
15000

60

10000
40

5000
20

0 0
0 200 400 600 800 1000
O
Temperature ( C)

Figure 6. The heat transfer coefficient of air and water cooling.

Figure 7. Schematic illustration of the Jominy end-quenching test.

heat for ferrite, pearlite, bainite and martensite formation is introduced using the following
values [17]: QF = 5.9E8, Qp = 6.0E8, QB = 6.2E8 and QM = 6.5E8 J m3 .
The Jominy end-quenching test is widely used to characterize the hardenability of steels.
In this work, the standard Jominy test is carried out for the benchmark verification of
computational modeling. A schematic illustration of the Jominy end-quenching test is shown
in figure 7. The boron steel cylinder (diameter: 30 mm; length: 110 mm) is heated to 900 C
austenization temperature with a pre-set holding time. The prior austenite size (D ) of 122 m
can be obtained after austenization. Then, the sample is quenched from its end surface with a
controlled and standardized jet of water. Four points (points A, B, G and H in figure 7) with a
K-type thermocouple are set to record the temperature change during quenching. The IHTC

9
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

1.0

0.8
Experiment

Fraction 0.6 Calculation (1%-50%)


Calculation (1%-99%)
Calculation(50%-99%)
0.4

0.2 P

FF
0.0

0 50 100 150 200


Time (S)

Figure 8. The phase transformation kinetics without considering incubation time in


the JMAK equation and the corresponding microstructure metallograph (F: ferrite, P:
pearlite).

of the air cooling and water quenching is calculated based on the experiment measurement and
reverse calculation, as shown in figure 6. After quenching, the steel cylinder is cut into several
parts for the metallographical microstructure observation and hardness measurement.

4. Simulation results

4.1. The influence of incubation time on the transformation kinetics

As described in section 3, the material parameters n and b in the JMAK model represent the
transformation kinetics. If the incubation time ts is not considered in equations (3) and (4),
n and b can be evaluated by any two points in the transformation kinetics curve, e.g. the
starting (F = 1%), the middle (F = 50%) and the finishing (F = 99%) points. We try to
calculate n and b using the different point pairs. Then, the isothermal austenite decomposition
kinetics into ferrite and pearlite at a soaking temperature of 650 C is calculated based on
the fitting parameters, as shown in figure 8. It should be noted that the calculated fraction
is the total amount of ferrite and pearlite. The calculated results show that the predicted
kinetics deviates from the experiment result. For example, when the point pair of F = 1% and
F = 50% is adopted, the predicted transformation kinetics at late stage is obviously faster than
the experiment. However, if the incubation time is considered, a good agreement is obtained
between the simulation and experiment for any point pair, as shown in figure 9. This implies
that the introduction of incubation time into the JMAK equation can provide a more reasonable
prediction for the transformation kinetics than the unmodified one.
Then, according to the isothermal kinetics and additivity rule, computational simulations
of continuous cooling are performed to validate the austenite decomposition model at a constant
cooling rate. In figure 10, computational results are shown and compared with the experimental
continuous cooling transformation (CCT) diagram. The result shows that the simulated starting
time and ending time of austenite decomposition into ferrite, pearlite and bainite agree well
with the CCT diagram. The predicted critical cooling rate for full martensite transformation
(20 s1 ) is also in agreement with the experimental observation. This implies that this model
shows a good capability to simulate austenite decomposition into different products.

10
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

1.0

0.8
Experiment

0.6
Calculation (1%-50%)
Fraction Calculation (1%-99%)
Calculation(50%-99%)
0.4

0.2 P

F
0.0

0 50 100 150 200


Time (S)

Figure 9. The phase transformation kinetics considering incubation time in the JMAK
equation and the corresponding microstructure metallograph (F: ferrite, P: pearlite).

750

700 A

650 o
F+P
20 /s
600
Temperature (oC)

550

500
B
450

400
Ms=420o Experiment (1%)
350 Experiment (99%)
M B+M
Calculation (1%)
300
Calculation (99%)
250 o
Mf=220
200
10 100 1000 10000

Time (s)

Figure 10. The predicted starting and finishing time of the CCT curve compared with
the experiment.

4.2. Simulation of the Jominy end-quenching test

The simulation is performed using a subroutine (USDFLD and HETVAL module) in the FE
simulation program ABAQUS combined with the phase transformation model described in
section 2. The finite element sample is discretized onto a mesh of 515 diffusive heat transfer
elements of DCAX4 type. The end surface of the cylinder is set as the water cooling area and
the rest as the air cooling area. The medium temperature of air and water is given as a constant
of 20 C.
The variations in the simulated temperature of the pre-set surface point with cooling time
are shown in figure 11. Clearly, the points near the quenched end have a higher cooling rate.
The recalescence effect of latent heat liberation on the temperature evolution is especially
obvious at low cooling rates. The temperature range of recalescence implies that austenite
decomposition into ferrite occurs.

11
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

1000

Calculation
800
Experiment
point-F

Temperature (oC)
point-E
600 point-D

400
point-C
point-B
point-A

200

0
0 200 400 600 800 1000
Time (s)

Figure 11. Comparison of temperature histories obtained from this FE simulation and
experiments.

(a) 1000 1.0


Temperature(10mm)

800 Martensite(10mm) 0.8


Bainite(10mm)
Temperature (oC)

Volume fraction
0.6
600

0.4
400

0.2
200

0.0

0
0 100 200 300 400 500 600 700

Time (s)

(b) 1000 1.0


Temperature(50mm)
Bainite(50mm)
0.8
800 Pearlite(50mm)
Ferrite(50mm)
Temperature (oC)

Volume fraction

0.6
600

0.4
400

0.2
200

0.0

0
0 50 100 150 200 250 300 350 400

Time (s)

Figure 12. The predicted transformation kinetics (A-10 mm, B-50 mm).

Figure 12 shows the predicted transformation kinetics of points located at the quenched
ends 10 mm and 50 mm, respectively. Clearly, martensite, bainite, pearlite and ferrite occur
in turn with decreased cooling rates. Figures 13 and 14 predict the final microstructure
components and the corresponding fraction along the Jominy bar. The results of the

12
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

Figure 13. The predicted final microstructure distribution (F: ferrite, P: pearlite, B:
bainite, M: martensite).

1.2

1.0

0 10 20 30 40 50 60 70 80 90 100 110 120 130 140


Volume fraction

0.8 Distance (mm)

0.6 Ferrite
Pearlite
Bainite
0.4 Martensite

0.2

0.0

0 20 40 60 80 100
Distance (mm)

Figure 14. The simulated final microstructure components and the corresponding
fraction along the Jominy bar.

metallographical observation of different locations away from the quenched end are shown in
figure 15. Although the microstructure components are difficult to characterize quantitatively,
the comparison implies a qualitative agreement between the modeling and experiments.
Finally, the hardness distribution can be predicted based on the mixture rule of the phase
component. Figure 16 shows the predicted and measured Vickers hardness profile as a function
of the distance from the quenched end. Both empirical formulas and the experimentally
measured microhardness are adopted for the prediction. The comparison shows that both
Maynier empirical formulas and the microhardness values provide a good agreement in the
intermediate to high cooling rate range, i.e. the region with the martensite or martensite/bainite

13
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

M
M

B B

F
P
P

F F
B

P
B

Figure 15. The corresponding metallographical observation of locations x = 0 mm (a),


12 mm (b), 22 mm (c), 32 mm (d), 42 mm (e) and 52 mm (f ). (F: ferrite, P: pearlite, B:
bainite, M:martensite).

mixture. However, the prediction based on the experimental microhardness shows a better
agreement than the Maynier empirical formulas for the low cooling rate range. This derivation
seems to result from the absence of boron content in the empirical formulas. So experimentally
measured hardness values for the individual phase constituent have been shown to produce
improved accuracy in the predictions compared to empirical hardness equations.

5. Conclusions

A computational model that describes the interactive influence between temperature field and
phase transformation is developed for simulating microstructure evolution in high-strength

14
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

550

500
Simulation with Microhardness test
450
Simulation with Maynier equation
Experiment

Hardness HV
400

350

300

250

200

150
0 20 40 60 80 100

Distance (mm)

Figure 16. Measured and calculated hardness as a function of distance from the
quenched end.

boron steel. The JMAK-type equations with the incubation time and additivity hypothesis
are adopted to describe the austenite decomposition into ferrite, pearlite and bainite during
continuous cooling, while the KM model is calibrated by the experiments to the martensite
transformation. The simulation results show that the introduction of incubation time into
the JMAK equation can provide a more reasonable prediction for the transformation kinetics
than the unmodified one. The comparison between the simulation and the standard Jominy
end-quenching test implies the superiority of this model for the prediction of transformation
kinetics, microstructure distribution and mechanical properties. Furthermore, the adoption
of experimentally measured microhardness values for the individual phase constituent can
produce a higher accuracy in hardness predictions compared to empirical hardness equations.
In the hot stamping process of the steel sheet, the interaction between the stress and
phase transformation has a great influence on the mechanical response of the materials, such
as residual stress distribution and dimensional precision of parts. Generally speaking, the
prediction accuracy of most thermo-microstructural-mechanical models is strongly dependent
on the constitutive law of the individual phases and their relationship with the global mechanical
responses. Unfortunately, the experimental constitutive law of individual phases as function
of strain rate and temperature, especially the phases in the intercritical region such as ferrite,
pearlite and bainite, are somewhat incomplete. Furthermore, the mixed global mechanical rule
deduced from the individual phases and its application range should be assessed carefully for
better accuracy in the model. Future perspectives in research may mainly focus on the influence
of phase transformation strain and transformation plasticity on the mechanical responses,
especially in the case of multiple phase transformation and complex thermomechanical
circumstances.

Acknowledgments

This work is supported by The National Natural Science Foundation of China (61232014,
11202072) and The Open Foundation of State Key Laboratory of Advanced Design and
Manufacturing for Vehicle Body (Hunan University, China).

15
Modelling Simul. Mater. Sci. Eng. 22 (2014) 065005 X Chen et al

References

[1] Woodard P, Chandrasekar S and Yang H 1999 Analysis of temperature and microstructure in the
quenching of steel cylinders Metall. Mater. Trans. B 30 81522
[2] Avrami M 1939 Kinetics of phase change: I. General theory J. Chem. Phys. 7 1103
[3] Hillert M 1969 The mechanism of phase transformations in crystalline solids J. Inst. Met. 97 23147
[4] Kirkaldy J S and Venugopalan D 1983 Prediction of microstructure and harden-ability in low alloy
steels Int. Conf. on Phase Transformations in Ferrous Alloys, Philadelphia, PA, 1983) ed A R
Marder and J I Goldstein, pp 12548
[5] Li M V, Niebuhr D V, Meekisho L L and Atteridge D G 1998 A computational model for the
prediction of steel hardenability Metall. Mater. Trans. B 29 66172
[6] kerstrom P 2006 Modelling and simulation of hot stamping, Lule tekniska universitet/Tillampad
fysik, maskin-och materialteknik/Hllfasthetslara
[7] Bok H-H, Lee M-G, Pavlina E J, Barlat F and Kim H-D 2011 Comparative study of the prediction
of microstructure and mechanical properties for a hot-stamped B-pillar reinforcing part Int. J.
Mech. Sci. 53 74452
[8] Koistinen D P and Marburger R E 1959 A general equation prescribing the extent of the austenite-
martensite transformation in pure iron-carbon alloys and plain carbon steels Acta Metall. 7 5960
[9] Carlone P, Palazzo G S and Pasquino R 2010 Finite element analysis of the steel quenching process:
temperature field and solidsolid phase change Comput. Math. Appl. 59 58594
[10] Lee S-J and Lee Y-K 2008 Finite element simulation of quench distortion in a low-alloy steel
incorporating transformation kinetics Acta Mater. 56 148290
[11] Hawbolt E, Chau B and Brimacombe J 1983 Kinetics of austenitepearlite transformation in
eutectoid carbon steel Metall. Trans. A 14 180315
[12] Umemoto M, Hiramatsu A, Moriya A, Watanabe T, Nanba S, Nakajima N, Anan G and Higo Y
1992 Computer Modelling of Phase Transformation from Work-hardened Austenite ISIJ Int.
32 30615
[13] Guimaraes J and Shyne J 1970 On the mechanical sensitization f austenite in FeNiC alloys Scr.
Metall. 4 101921
[14] Maynier P, Dollet J and Bastien P 1978 Hardenability Concepts with Applications to Steels ed
D V Doane and J S Kirkaldy (New York: AIME) pp 51844
[15] Pan J S and Gu J F 2009 Mathematical fundamentals of thermal process modeling of steels Handbook
of Thermal Process Modelling of Steels (Boca Raton, FL: CRC) pp 2630
[16] Reti T, Fried Z and Felde I 2001 Computer simulation of steel quenching process using a multi-phase
transformation model Comput. Mater. Sci. 22 26178
[17] Buchmayr B and Kirkaldy J 1990 Modeling of the temperature field, transformation behavior,
hardness and mechanical response of low alloy steels during cooling from the austenite region
J. Heat Treat. 8 12736

16

Вам также может понравиться