Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

MODELING OF HOT DEFORMATION OF CARBON STEELS IN THE

INTERCRITICAL ZONE
J. Cancio Jiménez L, J. Federico Chávez1, José-María Cabrera2, J. M. Hallen L1.
1*
Department of Metallurgy and Materials Engineering, Instituto Politécnico Nacional -
ESIQIE, UPALM, 07738 México City, México
2* Departamento de Ciencia de Materiales e Ingeniería Metalúrgica, EEBE,

Universidad Politécnica de Cataluña, España

Keywords: hot deformation, hot deformation modeling, Constitutive equations


Intercritical region

ABSTRACT
During the last years, significant advances have been made in the prediction of the hot
deformation behavior of metallic materials, steels in particular. Although many models
and types of constitutive equations are reported in literature, the most accepted model
is the physical basis model based on the single-variable approximation, which calculates
the evolution of the density of dislocations, taking into account the coexistence of
hardening deformation and softening by dynamic restoration. Finally, when applying the
relationship between stress and density of dislocations, a phenomenological equation is
obtained which relates the stress and the deformation as a function of temperature and
strain rate. The effect of the chemical composition on the hot deformation of austenite
and ferrite has recently been included in the equations.
In this work the parameters of Fe-C-Mn-Si alloys for the ferritic, austenitic and intercritic
models are determined and applied in a generic constitutive model that predicts the
behavior of hot deformation in the biphasic zone (α+‫)ﻻ‬. To do this, the model is modified
by the mixing law, as it was initially developed for single-phase steels. The percentages
of each of the present phases are determined as a function of temperature and chemical
composition. The developed constitutive model is validated by comparing the theoretical
stress-strain curves with experimental curves carried out by isothermal uniaxial
compression tests of 1008 and 1035 carbon steels at different temperatures and strain
rates.
The compression tests were carried out in a dilatometer able to apply the compression
load and with temperature control of ± 0.5 °C, at speeds of 10-3, 10-2 and 10-1 s-1. By
combining the behavioral dependence models during hot deformation in both austenite
and ferrite as a function of the chemical composition, the behavior of the steels in the
intercritical interval is predicted, which compared with the experimental data, presents a
minimum dispersion, which is considered very acceptable and that allows to analyze the
different stages of deformation and hardening, by this reason a corrected model is
proposed.
In addition, a detailed microstructural characterization of the samples used for the
experimentation is presented, both at the initial condition, and after quenched at the end
of each experiment, all in order to relate the results of the curves with the microstructure
of the steels and the percentages of phases.
I INTRODUCTION

Steel is one of the most researched materials throughout history, even today certain
aspects of its behavior under certain conditions continue to be ambiguous. For instance,
during the shaping at high temperatures (T> 0.5 Tm, where T and Tm are the
temperature and the absolute melting temperature expressed in 0K, respectively [1, 2,
17] .There is a great agreement regarding the role that the C, Si and Mn present on the
mechanical properties of carbon steels, especially when their behavior is analyzed when
they are subjected to cold work, however, the same case does not occur with regard to
the behavior of the this elements during the operations of deformation at high
temperatures [1, 10, 17], given the multiplicity of variables involved in this type of
operations (speed and mode of deformation, temperature, active diffusion mechanisms,
chemical composition, initial austenitic grain size, dislocation structure, energies of
activation and stacking failure, etc.) it is expected that the analysis becomes complex
and greater efforts are required to discriminate or attribute the effect of the composition
chemistry. (2).

The phenomena involved in the processes of deformation at high temperatures are


controlled by phenomena of diffusion, movement of dislocations and atomic interaction
with the environment. Examples of this are strain hardening and dynamic restoration
phenomena; Recovery and dynamic recrystallization. As a consequence of the atomic
nature of the species that diffuses through the crystal lattice, these phenomena are
affected in one way or another. [1-3]

Thus, for example, C affects the hot behavior of steels depending on their possibilities of
diffusion through the network and their interaction with other atoms and / or dislocations.
However, his behavior is not entirely clear (1). It is believed that the addition of C causes
distortions in the crystal lattice which results in an increase in the kinetics of the self-
diffusion mechanisms of the matrix atoms. In turn, there is not much agreement in the
literature about its effect on maximum yield stresses; there are authors who report a
decrease with the addition of C, an increase and even that there is no effect of that
element (2). It should be remembered that it is not enough to limit the study of the effect
of a specific element to the changes in the values of the maximum stresses, since its
effect can also influence the mechanisms of hardening and softening that take place
during shaping, as well as the diffusion processes. [1-3]

The relative size of the solute atoms relative regard of the matrix determines whether a
diffusion process operates interstitially or substitutionally. In the case of Si and Mn, it has
been observed that increasing its contents leads to hardening of the steel as in cold
working. However, some results question whether both elements play a similar role in
terms of softening mechanisms. Some authors (1) relate this behavior to the modification
of both the stacking fault energy and the activation energy necessary for the process.
The addition of solute clearly limits mobility
and the migration of grain boundaries of austenite, directly affecting recovery kinetics
and dynamic recrystallization (16). With all this, it should be noted that the possible
combined effect between the different alloying elements is even less analyzed.

The model taken as the basis for this work (1) has been tested in the brass alloy
CuZn40Pb2, in which 2 phases are presented. The correlation of the curves obtained
with it, are considered acceptable with respect to the experimental ones. Which were
determined as a function of temperature, deformation speed and% of phases. To get to
this, it is necessary to know the deformation and the softening mechanisms that act in
each of the phases. Varela (2) studies the hot creep behavior of so-called construction
steels and their dependence on chemical composition, specifically C, Si and Mn. This
behavior was analyzed by isothermal uniaxial compression in the temperature range
between 900 and 1100 ℃ at true deformation rates in the range of 10-3 s-1, 10-2s-1 and
10-1s-1 and its implementation in physical-based constitutive models. P. Rodríguez-
Calvillo et al. (3) show that the flow curves of all ferritic materials (5) show the
characteristics of materials that only have dynamic recovery, therefore the
experimentally obtained values of the maximum stress are analyzed using the model of
Garafalo and Zener-Hollomon (11)

For the above reasons, it was considered pertinent to apply this model in the present
study, to determine the hot flow curves that indicate the behavior of the Fe-C-Mn-Si
alloys in the intercritical zone corresponding to the Iron-Carbon diagram.

The present study is focused on obtaining a generic constitutive model that predicts the
hot deformation behavior of carbon steels in the intercritical zone (α + ‫ )ﻻ‬according to
their contents of the basic elements Fe-C-Mn-Si This establishes a scenario that
promotes greater rigor in the control of deformation processes so that their behavior can
be defined a priori under microstructural criteria. Thus, the establishment of a constitutive
equation that predicts such behavior in function of the basic parameters that govern the
softening and hardening phenomena [6, 8, 9, 14] is essential to optimize the hot forming
processes, mainly those that are develop in the intercritical zone. The modeling of the
maximum values obtained from the stress σss, from the flow curves, in the steady state,
as well as from the whole flow curve using the model proposed by Estrin and Mecking
and Bergstrom, [12,15] presents good correlation with the experimental data.

At the same time, the hot-flow behavior is studied experimentally by isothermal uniaxial
compression [4, 7], over a wide range of temperature and deformation velocity through
the analysis of the curves obtained with steels 1008 and 1035, thus obtaining parameters
in the intercritical interval, applicable to a generic constituent model of hot deformation.
These results are also supported by microstructural studies (13) of the percentages of
phases present in the deformed and tempered samples by optical microscopy and SEM.
Finally we propose a corrected generic constitutive model that describes the hot creep
behavior of Fe-C-Si-Mn type alloys in the intercritical zone (α + ‫)ﻻ‬, based on the law of
mixtures of austenite and ferrite deformation, where all the characteristic parameters of
the same are a function of the chemical composition of the alloy.

2 Experimental procedure

Two commercial steels 1008 and 1035 with diameter of ¼ inch were selected, with which
there is a wide intercritical interval, table 1 shows the chemical composition of them.
From the real experimental tests in the Dilatometer, corresponding to the temperature
intervals between the Ac1 and Ac3 temperatures of the coexistence of the phases (α + ‫)ﻻ‬
of the Fe-C diagram, the temperatures for the experimental tests were defined

Table 1 CHEMICAL COMPOSITION STEELS 1008 AND 1035

Steel %C %Si %Mn %P %S


SAE 1008 0.075 0.145 0.351 0.035 0.034
SAE 1035 0.308 0.251 0.787 0.042 0.035

The transformation curves were determined with DILATOMETER brand BAHR DIL 805
A/D whose results are shown in figure 1.
Time

Temperature

Fig. 1 Cooling curve obtained with Dilatometer


Time

Temperature

Fig. 2 heating curve obtained with Dilatometer.

The experimental test plan was carried out taking into account the actual heating and
transformation temperatures determined in the dilatometer, to say Ac1 and Ac3 remaining
as shown in Table 2.

Table 2 Experimental plan of tests.


STEEL έ TEMPERATURE (°C)
1·10-3 815 800 785 770
1008 1·10-2 815 800 785 770
1·10-1 815 800 785 770
1·10-3 785 770 755 740
1035 1·10-2 785 770 755 740
1·10-1 785 770 755 740

In general, it can be said that the thermal cycle established for hot compression tests
can be divided into six different stages:

1. Heating to total solubilization temperature of austenite.


2. Maintenance at that temperature
3. Cooling to test temperature
4. Maintenance at said temperature for homogenization of the microstructure
5. Compression test itself
6. Temple of the test tube

Cylindrical samples of 10 X 5 mm were used.


Fig. 3 Stages of testing the sample both low and high carbon.

2.1 Proposed generic model

It should be noted that in principle the ferritic model is considered and the corresponding
spreadsheet is developed to obtain the yield curves of the steels mentioned according
to the reference [3].

From (2), the spreadsheet for the austenitic model is developed with which the yield
curves of the steels in question are obtained. It should be emphasized that these models
are developed to be used only in single-phase models.

The present investigation is based mainly on the aforementioned articles (1-3), from
which a generic model for Fe-C-Mn-Si alloys is obtained to implement its use in the
biphasic intercritical zone (α+‫)ﻻ‬. The model is evaluated for the verification of the hot
deformation behavior of the Fe-C-Mn-Si alloy in the intercritical zone (α+‫ )ﻻ‬Thus, for a
single phase and using model Cabrera:

{ έ /D(T)}=B (sin h β σ / E (T))n Ec. 1

Where:

έ is the deformation velocity, D (T) is the diffusion coefficient affected by temperature,


Sin h = Hyperbolic sine function Β = Structural Parameter β is the parameter that is equal
to the inverse of the tension and represents the dependency change between the speed
of deformation and the tension. It also represents the change of potential behavior to
exponential and is independent of temperature. σ is True tension E (T) is Young's
modulus affected by temperature. n is the pre-exponential factor. In two-phase zones,
the equation becomes:

Ec. 2

This equation considers the percentages of each of the phases present as a function of
the temperature and chemical composition of the alloy, in the same way its effect in those
conditions of the diffusion coefficient, Young's modulus and the structural parameter and
behavior change from potential to exponential. The% of the present phases is
determined by the lever rule and therefore the composition of the ferrite and the
austenite. The above depends on the temperature and the chemical composition. The
intercritical generic model is obtained considering the spreadsheet in the ferritic zone
and the austenitic zone, in each of them only the composition of each one of the phases,
the deformation conditions and chemical composition is indicated. In the spreadsheet of
STEEL the intercritical model only the percentage of each one of the present phases is
indicated.

3 RESULTS AND DISCUSSION


Los rangos para las temperaturas de ensayos en base a pruebas del dilatómetro fueron
para el acero BC de 770 a 815 0C a y para el AC de 740 a 785 0C respectivamente. El
tamaño de grano antes de ser ensayadas es muy similar en ambos aceros y en una
sección transversal el valor fue de T.G. de 8.5. En la tabla 3 se muestra los valores
medios del TG para las condiciones de prueba.

STEEL 1008

TEMPERATURE 0C

STEEL 1035
TEMPERATURE 0C

Table 3. Grain size of the samples after tempering

It can then be established that the average size for all alloys is around 38.125 μm with
an average standard deviation of ± 10 μm, and it is assumed here that the influence of
this parameter on creep behavior is negligible.

3.1 Generic model for Fe-C-Mn-Si alloys in the intercritical zone (α+‫)ﻻ‬

Once the Ferritic and austenitic models were validated, the intercritical generic model
was obtained, using the Cabrera model, modified according to the law of mixtures.

The calculation of the Young's modulus in the intercritical zone is determined by the
following equation:
E(T)I = E(T)α %Xα + E(T)γ %Xγ
Ec. 3

To calculate the diffusion coefficient in the intercritical zone, the following expression was
used:

D(T)I = D(T)α %Xα + D(T)γ %Xγ


Ec. 4
The structural parameter and the equivalent to the reciprocal tension are
calculated as follows:

BI = Bα %Xα + Bγ %Xγ Ec. 5

ΒI = βα %Xα + βγ %Xγ Ec. 6

The values of these 2 parameters are defined according to the chemical composition (2)
and (3) corresponding the following values for the steady state and peak, see table 4

Table 4, Structural parameter values and equivalent.

The parameters of hardening and softening were determined by the following


relationships:
Ec. 7
ΩI = Ωα %α + Ωγ %γ

UI = Uα %α + Uγ %γ Ec. 8
The yield stress in the intercritical zone is defined by the following
expression:
σI = σα %Xα + σγ %Xγ so it follows that: Ec. 9

Ec 10

The percentages of each of the phases are calculated by the lever rule at the test
temperatures.
3.2 Creep curves

The tests were carried out under the protective atmosphere of Ar and, in order to
minimize friction between the tools of the machine and the specimen, wolfram sheets
were used. Each test was carried out until a true deformation, maximum of 0.8 with the
aim of minimizing the possible heterogeneities in deformation. The cited curves are
represented as the true stress versus the true deformation as a function of temperature
and for different deformation rates: (a) 10-1 s-1, (b) 10-2 s-1 (c) 10-3 s-1. Comparison
between the experimental flow curves (E) and those predicted by the proposed model (I)
at different temperatures and deformation speed.
200
120
180 I
E
TRUE STRESS (MPa) 100 160

TRUE STRESS (MPa)


140
80 I
E 120

60 100

80
40 AC-740-E-0.01
60
AC-740-E-0.001
40
20
20
0
0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN () TRUE STRAIN ()

250
Fig. 4 Experimental versus
200
theoretical curve Steel AC-740
at different speeds.
TRUE STRESS (Mpa)

I
150 E

100

AC-740-E-0.1
50

0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8

TRUE STRAIN (

200

120 I 180
E I
160 E
100
TRUE STRESS (MPa)
TRUE STRESS (MPa)

140

80 120

100
60
80

40 AC-755-E-0.001 60
AC-755-E-0.01
40
20
20

0 0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8

TRUE STRAIN () TRUE STRAIN ()

250

200
Fig. 5 Experimental versus
TRUE STRESS (MPa)

I
150 E
theoretical curve Steel AC-755
at different speeds.
100
AC-755-E-0.1

50

0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN ()
200

120
I 180 I
E E
160
100

TRUE STRESS (Mpa)


TRUE STRESS (MPa) 140

80 120

100
60
80
AC-770-E-0.001
40 60
AC-770-E-0.01
40
20
20

0 0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN () TRUE STRAIN ()

250

I Fig. 6 Experimental versus


E
200 theoretical curve Steel AC-770
TRUE STRESS (MPa)

at different speeds.
150

100

AC-770-E-0.1

50

0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN ()

200
120
I 180
E I
100 160 E
TRUE STRESS (MPa)

TRUE STRESS (MPa)

140
80
120

60 100

80
40 AC-785-E-0.01
AC-785-E-0.001 60

40
20
20
0
0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN ()
TRUE STRAIN ()

250

I
E Fig. 7 Experimental versus
200

theoretical curve Steel AC-785


TRUE STRESS (MPa)

150 at different speeds.

100

AC-785-E-0.1

50

0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8

TRUE STRAIN ()


50
130
45 I
120
E I
40 110 E
100
TRUE STRESS (MPa)
35

TRUE STRESS (MPa)


90
30
80
25 70

20 60

BC-770-E-0.001 50
15
40
10 30 BC-770-E-0.01
5 20
10
0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0
TRUE STRAIN () 0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8

TRUE STRAIN ()

130
120 I
110 E Fig. 8 Experimental versus
100 theoretical curve Steel BC-770
TRUE STRESS (MPa)

90
80
at different speeds.
70
60
50
40 BC-770-E-0.1
30
20
10
0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN ()

50 80

45 I
E 70 I
40 E
60
TRUE STRESS (MPa)

TRUE STRESS (MPa)

35

30 50

25 40

20
30 BC-785-E-0.01
BC-785-E-0.001
15
20
10

5 10

0 0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN ()
TRUE STRAIN ()

130
Fig. 9 Experimental versus
120 I
110 E theoretical curve Steel BC-785
100 at different speeds.
TRUE STRESS (MPa)

90
80
70
60
50
40 BC-785-E-0.1
30
20
10
0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN ()
50
80
45
I I
70
40 E E
TRUE STRESS (MPa)

35 60

TRUE STRESS (MPa)


30 50

25
40
20
30
15 BC-800-E-0.001 BC-800-E-0.01
10 20

5 10

0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN ()
TRUE STRAIN ()

130 Fig. 10 Experimental versus


120 I theoretical curve Steel BC-800
110 E
100 at different speeds.
TRUE STRESS (MPa)

90
80
70
60
50
40
BC-800-E-0.1
30
20
10
0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8

TRUE STRAIN ()

50
60
I
45 I
E
E
40 50
TRUE STRESS (MPa)

35
TRUE STRESS (MPa)

40
30

25
30
20

15 BC-815-E-0.001 20 BC-815-E-0.01

10
10
5

0
0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN ()
TRUE STRAIN (

130
120 I
110 E
100
TRUE STRESS (MPa)

90
Fig. 11 Experimental versus
80
70
theoretical curve Steel BC-815
60 at different speeds.
50
40
BC-815-E-0.1
30
20
10
0
0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8
TRUE STRAIN ()
STRAIN SPEED (ε)
TEMPERATU
RE 0C

AC
ER
O

Table 5, Peak tension and steady state values.

Table 5 shows the comparison of the previous curves, in which the peak peak tension is
indicated for all the curves (experimental and theoretical), in order to more easily quantify
the discrepancies between the two.

Fig. 6 shows an acceptable correlation between experimental and


theoretical stress values.

In fig. 6 The graphical comparison of the experimental and theoretical saturation tension
using the modified Cabrera model.
From the previous yield curves it is observed that those obtained by the intercritical
generic model, both the BC and the AC, have an acceptable tendency with respect to
the experimental curves, which demonstrates the reliability of the proposed model.

3.3 Microstructures before and after the hot deformation of the samples taken by optical
microscopy

The nomenclature used is E (Tested), T (Quenching), 101, 102 and 103 equivalent to
0.1, 0.01 and 0.001 s-1 which correspond to the speed of deformation. From figure 7
show micrographs of quenching carbon steel (T), at the same magnifications and treated
at different deformation rates. It is worth noting the effect of the deformation speed, it is
observed that at a higher deformation speed there is a grain refinement. (Ferrite white
grains, black martensite). In all the figures the same phenomenon of grain size is
presented as in the low carbon. During the deformation, the grain size increases, from
the center towards the periphery. This phenomenon of grain size is similar to that of high
carbon steel. Figure 7, 8 and 9, shows the variation of the grain size from the center of
the same towards the periphery,

Fig. 7 BC steel quenching to 800 0C deformed to 0.1 s-1 (101), 0.01 s-2 (102) and 0.001
s-3 (103) to 200X.
Fig. 8 BC steel quenching at 800 0C deformed at various deformation speeds.

Fig. 9 AC steel quenching at 740 0C deformed at various deformation speeds.


Additionally, a characterization by scanning electron microscopy was performed on the
deformed and quenching samples at the end of the compression test, in order to verify
the microstructure at the end of these tests.
The figures from 10 to 15 were taken with the high resolution Jeol 6701 A microscope,
in order to confirm the characteristics of the structure.
In Fig. 10 the sample of BC steel deformed at 800 ° C and a velocity of 10-1 s-1, in which
a microstructure with a ferritic matrix with black polymorphic grains, the amount of
martensite that is appreciated in white. In figures 11 and 12 the microstructure is
presented as in fig. 30, the only variation is the deformation velocity 10-2 and 10-3 s-1,
respectively. Slightly more deformed and larger grains are observed as the rate of
deformation decreases. A growth of the austenite grains transformed to martensite is
also observed. The effect can be considered in inverse form, generating a refining of
grain as the speed of deformation increases. From Figure 13 to 15, a greater presence
of martensite needles is observed in a polymorphic ferritic matrix. With the decrease of
the deformation speed, grain growth is observed.

Fig. 10, Steel BC-800-E-101 showing the presence of martensite in a ferritic matrix.
Taken at 1000X and 4000X increases (MEB) respectively.

Fig. 11 Steel BC-800-E-102 showing the presence of martensite in a ferritic matrix.


Taken at 1000X and 4000X increases respectively (SEM)
Fig. 12 Steel BC-800-E-103 showing the presence of martensite in a ferritic matrix.
Taken at 1000X and 4000X increases respectively (SEM)

Fig. 13 Steel AC-785-E-101 showing the presence of martensite in a ferritic matrix. Taken at
1000X and 4000X increases respectively (SEM).

Fig. 14 Steel AC-785-E-102 showing the presence of martensite in a ferritic matrix.


Taken at 1000X and 4000X increases respectively (SEM)

Fig. 15 Steel AC-785-E-103 showing the presence of


martensite in a ferritic matrix. Taken at 1000X and 4000X increases respectively
(SEM)
4 Conclusions

Based on the results obtained with the Generic Model developed in the present work for
the prediction of behavior under hot deformation in the intercritical zone of carbon steels
and its validation with tests performed on a dilatometer, the following conclusions can be
drawn:
I. In the true stress-strain curves σ-ε, result of the isothermal uniaxial compression
tests, the variation of the creep stress was determined and quantified by effect of
modifying the temperature and the true deformation speed. The trends are
expected, that is, an increase in temperature or a decrease in the speed of
deformation leads to a decrease in the yield stress and vice versa.
II. For the 2 steels tested and under all deformation conditions, the values of true
stress and true peak deformation σp and εp increase with the increase of the
deformation speed and decrease when the temperature increases and vice
versa. As a general trend, it is observed that the higher yield stresses are
recorded for the AC steel at higher deformation rates, while the lower ones are
for the BC steel mainly at low deformation rates.
III. Microstructurally, the amount of phases present depends on the chemical
composition of the steel and test temperature. The final mechanical properties of
the treated steels will depend on these conditions.
IV. The divergence between the experimental results and those predicted by the
model is acceptable and the phenomena that occur during a hot compression
test, basically dependent on the hardening and softening, can be identified with
certainty through the corresponding parameters in the temperature ranges and
deformation speeds applied, reflecting the goodness of the same in the
intercritical zone, which to date had not been contemplated with a generic
constitutive model.
V. Based on the foregoing, it can be affirmed that the proposed generic constitutive
model is susceptible to be applied in the constitutive modeling of
thermomechanical shaping operations of Fe-C-Mn-Si metal alloys in the inter-
critical temperature range
5 References
(1) L. Suárez, P. Rodríguez-Calvillo, J.M. Cabrera, A. Martínez-Romay, D. Majuelos
Mallorquín, A. Coma, “Hot working analysis of a CuZn40Pb2 brass on the monophasic
(β) and intercritical (α+β) regions”. Materials Science & Engineering, A 627(2015), pp.
42–50.
(2) Gonzalo Varela Castro. “Efecto de los contenidos de C, Si y Mn en el comportamiento
a fluencia en caliente de aceros de construcción al carbono. Aplicación a la obtención
de productos largos laminados” (Abril 2012). Tesis doctoral, U. Politécnica de Cataluña.
(3) P. Rodriguez-Calvillo, A. Boulaajaj, M. Pérez-Sine, J. Schneider, J.M. Cabrera. “On
the hot working of FeSi ferritic steels”, Materials Science &Engineering A606 (2014)127–
138 Marzo 28, 2014.
(4) B. Mintz, J.R. Banerjee and K.M. Banks: “Regression equation for Ar3 temperature
for coarse grained as cast steels”, Iron making and Steelmaking 201138(3), 197.
(5) B. Mintz and A. Cowley: “Deformation induced ferrite and its influence on the elevated
temperature tensile flow stress-elongation curves of plain C-Mn and Nb containing
steels”, Materials Science and Technology 2006 22(3), 279.
(6) G.E. Dieter, Mechanical Metallurgy, McGraw-Hill Book Company, London, 1988. H.
Mirzadeh, J.M. Cabrera, J.M. Prado, A. Najafizadeh, Mater. Sci. Eng. A 528, (2011)
3876–3882
(7) G. Mohapatra, F. Sommer and E.J. Mittemeijer: Calibration of a quenching and
deformation differential dilatometer upon heating and cooling: thermal expansion of Fe
and Fe-Ni alloys, Thermochimica Acta 2007 453(1), 31
(8) S. Akta, G.J. Richardson, C.M. Sellars, ISIJ Int. 45 (2005) 1686–1695.Y.V.R.K.
Prasad, S. Sasidhara, Hot Working Guide – A Compendium of Processing Maps, ASM
L. Suárez, P.
(9) C.M. Sellars and J.A. Whiteman: Recrystallization and grain growth in hot rolling,
Metal Science 1979 13, 187.
(10) Cowley, R. Abushosha and B. Mintz: Influence of Ar3 and Ae3 temperatures on hot
ductility of steels, Materials Science and Technology 1998 14(11), 1145
(11) C. Zener, J.H. Hollomon, C.W. MacGregor and J.C. Fisher: A velocity-modified
temperature for the plastic flow of metals, Journal of Applied Mechanics-Transactions of
the ASME 1947
(12) Y. Bergstrom and B. Aronsson: The application of a dislocation model to the strain
and temperature dependence of the strain hardening exponent n in the Ludwik-Hollomon
relation between stress and strain in mild steels, Metallurgical, Transactions 1972 3(7),
1951.
(13) Y. Maehara, K. Yasumoto, Y. Sugitani and K. Gunji: Effect of carbon on hot ductility
of as-cast low alloy steels, Transactions of the ISIJ 1985 25(10), 1045.
(14) T. Sakai and J.J. Jonas: Dynamic recrystallization: Mechanical and microstructural
considerations, Acta Metallurgica 1984 32(2), 189.
(15) Y. Estrin and H. Mecking: A unified phenomenological description of work
hardening and creep based on one-parameter model, Acta Metallurgica 1984, 32(1), 57.
(16) T. Sakai and J.J. Jonas: Dynamic recrystallization: Mechanical and microstructural
considerations, Acta Metallurgica 1984 32(2), 189.
(17) H.J. McQueen, S. Yue, N.D. Ryan and E. Fry: Hot working characteristics of steels
in austenitic state, Journal of Materials Processing Technology 1995 53(1-2), 293.
THEORETICAL PEAK TENSION (MPA)

You might also like