You are on page 1of 31

Accepted Manuscript

Micromechanical analysis of two heat-treated dual phase steels:


DP800 and DP980

A. Ch.Darabi , H.R. Chamani , J. Kadkhodapour , A.P. Anaraki ,


A. Alaie , M.R. Ayatollahi

PII: S0167-6636(16)30463-X
DOI: 10.1016/j.mechmat.2017.04.009
Reference: MECMAT 2733

To appear in: Mechanics of Materials

Received date: 6 November 2016


Revised date: 6 April 2017
Accepted date: 26 April 2017

Please cite this article as: A. Ch.Darabi , H.R. Chamani , J. Kadkhodapour , A.P. Anaraki , A. Alaie ,
M.R. Ayatollahi , Micromechanical analysis of two heat-treated dual phase steels: DP800 and DP980,
Mechanics of Materials (2017), doi: 10.1016/j.mechmat.2017.04.009

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service
to our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and
all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
Highlights

 Comparison of combined in-situ test with DIC technique results, with FEM

results.

 Just 3D micromechanical model is able to estimate macro experimental result

as well

 Three types of strain localization occur at low and high applied strains.

T
IP
 The micromechanical models have negligible dependency on the boundary

CR
conditions.

US
AN
M
ED
PT
CE
AC

1
ACCEPTED MANUSCRIPT

Micromechanical analysis of two heat-treated dual phase steels:

DP800 and DP980


A. Ch.Darabia, H.R. Chamanib, J. Kadkhodapoura,*, A. P.Anaraki a, A. Alaiec,

M.R. Ayatollahib
a
Department of Mechanical Engineering, Shahid Rajaee Teacher Training

T
University, Tehran, Iran

IP
b
Fatigue and Fracture Lab., Center of Excellence in Experimental Solid Mechanic
and Dynamics, School of Mechanical Engineering, Iran University of Science and

CR
Technology, Narmak, 16846, Tehran, Iran.
c
Department of Mechanical Engineering, Isfahan University of Technology,
Isfahan, Iran

US
AN
Abstract

Dual phase (DP) steels, which have an intricate microstructure, consist of ferrite and
M

martensite phases and because of interaction between the ferrite and martensite phases, those
have complex behaviors. In this study, the flow curves of two heat-treated DP steels, DP800
ED

and DP980, were compared with those of the 2D and 3D micromechanical models based on
actual microstructure under symmetric and periodic boundary conditions. The effects of
martensite phase distributions on the flow curves of 2D and 3D micromechanical models
PT

were investigated. Moreover, two different microstructures of DP980 with two different
distributions of martensite phases were compared together to investigate the effect of
CE

martensite phase distribution in 2D and 3D micromechanical models on the stress-strain


curve of DP980. Also, an in-situ tensile setup test was employed in order to investigate the
AC

initiation of localized strain patterns in the ferrite phase. The stress-strain curve, localization
patterns and initiation of slip bands obtained from 2D and 3D micromechanical models were
compared with the experimental results. The numerical results revealed that the stress-strain
curves obtained from 3D models in comparison with 2D models are in closer agreement with
experimental data.

* Corresponding author at: Department of Mechanical Engineering, Shahid Rajaee Teacher Training University,
Tehran, Iran. Tel.: +98 21 22970052.
javad.kad@gmail.com
2
ACCEPTED MANUSCRIPT

Keywords: Dual phase steel, microstructure, micromechanical modeling, mechanical


behavior.

1. Introduction

Nowadays, several modern high-strength steels have been produced to reduce the weight of
automobiles, improve passive safety features, and consider energy saving. These types of
steels like dual phase (DP) and transformation induced plasticity (TRIP)include different

T
phases in their microstructure[1].Having high strength besides good formability and high

IP
energy absorption in case of accidents cause these groups of steels to be widely employed in

CR
automotive industry. These mechanical properties originate from the different phases in
microstructures with considerable differences from aspect of ductility and strength. As a
result, favorable properties of each phase of multiphase steels can be combined into one

US
material with a pure ferrite matrix and then about5–35% martensite islands can be dispersed
through the second phase. This behavior is like that of composite materials where the ferrite
AN
matrix supports the formability of the material while mostly the martensite islands are the
strengthening elements[2].
M

Strain localization is the first stage of failure process in multi-phase steels. After initiation of
strain localization, fracture quickly initiates in the localized region, voids grow and, therefore,
ED

failure would happen in ferrite–martensite interfaces [3].Several experimental investigations


have been done on damage mechanisms in DP steels. Kadkhodapour et al.[4]examined the
damage mechanism of DP steel by scanning electron microscopy (SEM) analysis and
PT

reported the incident of void nucleation on both ferrite-ferrite grains and ferrite-martensite
grains interfaces. They showed that the morphology and distribution of martensite phase play
CE

significant role in the failure mechanism of DP steels. Also, the localized pattern have an
influence on the failure mechanism. Ahmad et al. [5]reported three types of void nucleation:
AC

brittle fracture in martensite phase, void formation at the ferrite–martensite interfaces, and
void formation at the ferrite–ferrite interfaces. They also found that from low to intermediate
martensite volume fraction (Vm), the voids format at the ferrite–martensite interfaces whereas
the other failure mechanisms occur at martensite volume fraction above 32%. Tasan et al.
[6]studied the fracture surface of the tensile test specimens in order to investigate hardness.
They performed some indentation tests and reported that the failure was affected by shear
banding, strain localization, and damage.

3
ACCEPTED MANUSCRIPT
The purpose of all modifications during steelmaking is to adjust microstructural
configuration, chemical composition, thermo-mechanical processing routes, or heat
treatments. Different types of failure could be observed in DP steels, ductile failure in ferrite
phase, brittle fracture of martensite phase, and interface debonding, which happens between
martensite and ferrite phases[7, 8]. In the last decade, mechanical properties of DP steels
have been developed and several types of steel in terms of higher ultimate strength and work
hardening rate with a considerable increase in formability have been introduced[9, 10].
Macroscopic properties of DP steels depend on several parameters, including mechanical
properties of ferrite and martensite phases, volume percentage of phases, and grain size of

T
ferrite phase [8, 11]. Researchers have found different mechanical properties for ferrite and

IP
martensite phases of DP steels [12-14]. Since the macroscopic mechanical behaviors of DP

CR
steels depend on their microstructures, modeling of DP steel must be done at the
microstructure level. Therefore, in order to find out the local deformation mechanisms, strain
partitioning, and strain localization of DP steels, micromechanical models have been used
[15-19]. US
AN
One of the most important reasons to study the micromechanical modeling of DP steels is to
obtain a DP-type microstructure with optimum mechanical features. Therefore,
understanding local deformation mechanisms is necessary for design of an optimized
M

microstructure for DP steels. Shen et al.[20] investigated the strain distribution between the
ferrite and martensite grains. Kang et al.[21] and Alaie et al.[22,23]coupled a SEM with
ED

digital image correlation (DIC) technique and investigated the microscopic strain distribution
in both phases of the dual phase steel. Papaefthymiou et al. [24] applied a FE approach in
PT

order to consider the influence of DP microstructure. They used a representative volume


element (RVE) was within the framework of continuum mechanics. The RVE model
CE

illustrated a unit cell, representing a cut-out of the whole microstructure. This model made it
possible to recognize the influences of each modeled phase on the overall strength of
materials [16]. Sun et al. [17, 18]predicted the slip bands and ultimate ductility of DP steels
AC

by a micromechanical modeling procedure under different loading conditions using the


plastic strain localization theory. As a result, the plastic strain localization theory predicted
ductile failure due to the incompatible deformation between the hard martensite phase and
the soft ferrite one. Choi et al. [19] conducted a similar micromechanical model analysis.
Many authors have used 2D RVE modeling using real microstructure of material. For
example, Sun et al. [17]investigated the failure modes and ductility of DP steel in a 2D RVE.
Plastic work hardening properties were determined by the in-situ synchrotron-based high-

4
ACCEPTED MANUSCRIPT
energy X-ray diffraction technique. Ramazani et al. [25]studied the effect of microstructure
banding on the mechanical behavior of DP steel using a 2D RVE and compared the
mechanical behavior between the equiaxial and banded microstructures. Uthaisangsuk et
al.[16] attempted to describe the effect of multiphase microstructure on the failure
mechanism using 3D RVE within the framework of continuum mechanics. Ayatollahi et al.
[26] described the ductile damage in the ferritic matrix by Gurson-Tvergaard-Needleman
damage model. Different models under tensile testing were simulated and good agreement
between the 3D model and experimental results were achieved.

T
In this study, representative volume elements, which were generated based on the 2D and

IP
randomly extruded 3D models, were constructed for DP800 and DP 980 steels. These RVE
models were used to(i) investigate the effect of boundary conditions for 2D and 3D models,

CR
(ii) examine the effect of loading direction for 2D and 3D models, and (iii) describe the
localization patterns and initiation of slip bands for 2D and 3D models. In order to validate

US
the numerical results, both models were compared with the experimental results. Using an in-
situ test setup combined with DIC technique, the localization pattern and initiation of slip
AN
bands were compared with the results of microstructure-based finite element analysis.
Finally, the development of strain localization in 3D micromechanical model was compared
with that of the 2D micromechanical model under the periodic boundary condition.
M

2. Experiment
ED

2.1 Materials
A low carbon steel sheet with 3 mm thickness and chemical compositions given in Table 1
PT

was used to produce DP800 and DP980 steels. Intercritical annealing process at and
was accomplished and the steel samples with DP microstructures including a specific
CE

martensite phase fraction were prepared. The annealing temperatures ranged from Ac1 to Ac 3
temperatures of the investigated steel. The holding times after approaching the intercritical
AC

temperatures were 10 and 15min, respectively. An overview of the accomplished intercritical


annealing processes is displayed in Figure 1. Finally, the samples were quickly quenched in
water to room temperature.

After heat treatment, the samples were prepared for the metallographic analysis. The samples
were grounded by silicon carbide sheets with the grit sizes of 100, 240, 400, 600, 800, 1000,
1500, and 2000, respectively. Then, they were polished using the alumina solution of 0.5µm.

5
ACCEPTED MANUSCRIPT
After polishing, the specimens were etched by 2% nital for 3–5 s. The microstructures of
DP800 and DP980 resulted in optical micrographs shown in Figure 2.

In these micrographs, the light gray zones are ferrite phases and the dark regions are
martensitic phases. The volume fractions of martensite phase, Vm, were also determined
according to ASTM E562 [27]. Using mean linear intercept method, the volume martensite
phase fractions (MPF), Vm, were found to be approximately 25% and 33% for DP800 and
DP900, respectively.

T
IP
2.2 Mechanical testing
After the metallographic analysis, tensile specimens were prepared from the heat-treated

CR
sheets of steel for tensile testing. The tensile test was performed by means of universal testing
machine with the strain rate of .Six specimens from DP980 and DP800were

US
tested. During the tensile tests, displacement was recorded using an extensometer. The
average value of the true stress-strain curves of DP800 and DP980 from the tensile tests are
AN
shown in Figure 3.

In order to experimentally investigate the deformation of different phases of DP steels, the


M

tensile in-situ test setup was created in two separate components. The first component was a
screw-driven fixture in order to draw and hold the sample during the tension test. This
ED

component was installed inside the chamber of the SEM. The second component made
enough force for deformation of tensile sample in the next step. During the tensile test, after
PT

taking the images for each deformation level, the first component was removed from the
chamber and, then, using the second component situated outside the chamber the sample was
CE

stretched to the next step. Using this method, the test requires more time than the
conventional in-situ setups; however, this technique allows the users to perform the test using
AC

small- or medium-sized SEM chambers.

The tensile in-situ test was performed using the designed test setup inside a Philips XI30
scanning electron microscope. The samples were clamped at both ends and stretched
gradually under displacement control (Figure 4a and b). The change in the macroscopic
dimensions was measured in each step using TESA TS300 profile projector measurement
device, and the test was continued until the specimen became very close to failure. At the
beginning of each step, an algorithm was used to find the location of previous points and
capture the images in the new deformational state. More details of this test method could be
6
ACCEPTED MANUSCRIPT
found in Refs.[22, 23].The test could be continued until very close to the final failure without
the risk of specimen rupture under the microscope lens. Therefore, the deformational
behavior of microstructure could be observed at the last moments before the final failure of
the specimen (Figure 4c and d). It should be noted that, the amount of local plastic
deformation was analyzed using image processing techniques.

3. Micromechanical modeling
3.1 Representative volume element (RVE)

An RVE was generated based on the actual microstructure to describe the inhomogeneous

T
structure of investigated DP steels. An RVE should contain both phase combination and

IP
microstructural configuration. It should also be large enough to represent the most important
microstructure properties of the studied DP steel and parameters such as shape, morphology,

CR
size, and randomness of constituent phases should represent the investigated microstructure.
This method provides a realistic description of the material deformation in the micro scale.

US
Over 20 microstructure images were analyzed and volume fraction of martensite phase (Vm)
was computed. Then the microstructure image in which the area fraction of martensite phase
AN
was closest to Vm, was chosen. 3D RVE was generated such that its volume fraction and the
area fraction at each layer (cross section) were equal to Vm. Also, the microstructure image
M

was chosen to ensure that RVE’s mechanical properties converged towards homogenized
mechanical properties.
ED

3.2 2D RVE
PT

A two-dimensional (2D) RVE was constructed by an FE code in ANSYS based on the actual
microstructures obtained by the image processing. The 2D RVE of DP980 steel
CE

(microstructure-B) with 33% martensite phase and 90,000 plain strain elements is illustrated
in Figure 5, where the dark elements refer to the martensite phase and the white regions show
AC

the ferrite phase.

3.3 3D RVE
In this model, a 3D RVE was statistically constructed in the out-of-plane direction by a
random redistribution of the phases in the 2D model in a way that the volume fraction of
martensite phase remained constant in each layer through the thickness of 3D RVE. The
random distributions of martensite phase were generated by image processing. The 3D RVE

7
ACCEPTED MANUSCRIPT
was cubic with 152,000 SOLID185 elements in ANSYS software. Figure 6 shows the 3D
random RVE generated for the microstructure-A DP980 steel with 33% martensite phase.
The symmetric and periodic boundary conditions were considered for both 2D and 3D RVE
and the sizes of 2D and 3D RVE were 100×100 µm and 100 ×100×100 µm, respectively.

3.4 Material behavior

In this paper, a dislocation theory based on strain hardening method was applied to predict
the stress-strain curves for each phase, which were empirically computed according to the

T
local chemical composition of DP800 and DP980. This method was used to calculate the

IP
stress-strain correlations of single phases. Therefore, the Peierl's stress was calculated

CR
according to this method [28, 29].The approach can be expressed as follows:

1  exp(M .k . ) (1)
   0     .M .. b .
k .L

US
where the third term in this equation describes the effects of the cumulated dislocation
AN
strength together with the softening behavior due to recovery. is constant, M is a Taylor
factor, µ is the shear modulus, b is the Burger's vector, L is the dislocation mean free path,
M

and k is the recovery rate. For ferrite and martensite phases, the two latter parameters and
are different in each phase of steel[28, 29].The first term in Eq. (1) is the Peierl's stress (σ0)
ED

that depends on the chemical constituents of each phase in the material and it is presented in
Eq. (2).

 0  77  80  (% Mn)  750  (% p)  80  (%Cu) 60  (%Si) 


PT

(2)
45  (% Ni)  11 (% Mo)  60  (%Cr)  5000  (N ss )
CE

In Eq. (1), represents the material strengthening by carbon content which could be
estimated for ferrite phase by the following equation:
AC

 (max)  5000  (%Cssf ) (3)

and for martensite phase, it could be estimated as follows:


 (max)  3065  (%Cmss )  161 (4)

where and indicateweight percent of carbon content in ferrite and martensite


phases, respectively. In Table 2, all the parameters for both of DP steels and individual
phases are listed. Some of the parameters were extracted from[30].Figures7 and 8 show the
computed flow curves for ferrite and martensite phases of DP800 and DP980.
8
ACCEPTED MANUSCRIPT

3.5 Boundary conditions of RVE


In the micromechanical model, a suitable boundary condition (BC) should be defined for the
model. In this model, two types of boundary conditions, symmetric and periodic, are
employed for describing the micromechanical deformation.

3.5.1 Periodic boundary conditions

The general boundary condition for unit cells in continuum micromechanics is periodicity

T
which could explain any geometrical deformation of unit cells; therefore, the heterogeneous

IP
material could be modeled. Since in periodic boundary condition, the multipoint constraints

CR
need to be applied (Eq. (5)), especially in three-dimensional problems, this boundary
condition needs high CPU time and memory. Eq. (5) represents periodicity boundary
conditions for the mechanical material in the small strain range as follows[31]:
u DC (X )  u AB (X ) , US
u BC (Y )  u AD (Y )
(5)
AN
u DC (Y )  u AB (Y )  u D (Y ), u BC (X )  u AD (X )  u B (X )

where u is considered as the displacement in different directions (X and Y), point A is


M

assumed to be fixed and (X, Y) are local coordinates used to denote corresponding points on
the pairs of face. AB, BC, CD and AD are the outer edges of a 2D RVE. Eq. (5) directly
ED

implies that:
uC ( X )  u B (X ) , uC (Y )  u D (Y ) (6)
PT

This boundary condition is displayed in Figure 9 where the tensile loading is applied at point
C.
CE

Periodic boundary conditions were implemented inside the FE software using constraint
AC

equations. For this purpose, a code was written in a way that selects two opposing nodes
locating on two opposing faces (e.g. the left and right faces) of the 3D RVE model.
Afterwards, the nodal displacements of these nodes were constrained using constraint
equations.

3.5.2 Symmetry boundary conditions

9
ACCEPTED MANUSCRIPT
For some unit cells, such as rectangular and hexahedral ones, where the shape of the unit cell
is compatible with the symmetry planes, this type of boundary condition is suitable.
Symmetry boundary condition is a simple model of periodic boundary conditions[31], as
shown in Figure 10.Formulation of this boundary condition is given as follows:

u BC (X )  u B (X ) uCD (Y )  u D (Y ) u AD (X )  0 u AB (Y )  0 (7)

where u is considered as the displacement component in X and Y directions.

T
IP
4. Influence of mesh size
Mesh sensitivity was studied for both the 2D and 3D RVE models. Two element sizes were

CR
employed in the 3D random RVE to investigate the sensitivity of the element size. In other
words, two models with a fine mesh (512000 elements) and a coarse mesh (64000) with the

US
same volume fraction of martensite were used (Figure 11).As shown in Figure 12, the
variation in the stress-strain curve was negligible due to sufficient element refinement. Thus,
AN
the 3D RVE model was not sensitive to mesh refinement when the volume and area fraction
of martensite phase were constant.
M

A similar study was conducted for the 2D RVE model and five different element sizes were
investigated. As shown in Figure13, the microstructures were meshed with 400×400,
ED

300×300, 200×200, 100×100, and 50×50 elements, respectively. The increased number of
elements can help the simulations to have smoother grain boundaries between ferrite and
martensite phases. Figure14 shows the stress-strain curves obtained for different numbers of
PT

elements.
CE

5. Result and discussion


AC

5.1. Loading direction and boundary condition analysis

In this study, mechanical behaviors of DP steels with ferrite-martensite composition were


considered. Therefore, factors such as the stress-strain curve, localization pattern by in-situ
tests, and micromechanical modeling of the complex 2D and 3D models behaviors were
investigated. Figure 15 illustrates the stress-strain curves of 2D micromechanical models
under tension in X and Y directions and under both the periodic and symmetry boundary

10
ACCEPTED MANUSCRIPT
conditions for microstructures A, B, and C. These results indicated that the 2D models have
anisotropic behavior.
Figure 16 shows the effect of loading direction in the 3D micromechanical model on true
stress-strain curve for three microstructures (A, B, and C). These investigations were done
under both the periodic and symmetry boundary conditions and in X, Y, and Z directions. For
DP980 steel, the 3D model with loading in Z direction predicted a slightly lower value. In the
3D model presented in this work, martensite phase was distributed along z axis; as a result,
each cross-section along Z direction has the constant area fraction of martensite, as shown in
Figure 17. The distribution of martensite phases in X and Y directions in every cross-section

T
has different area fractions. However, the volume fractions of martensite phase in all three

IP
directions were the same. The results indicated that the increased martensite phase area

CR
fraction in some cross-sections, with the total of constant martensite phase volume fraction,
leads to an increase in its strength and the strength of its surrounding cross-sections.

US
In order to investigate the effect of boundary condition, both the symmetric and periodic
boundary conditions were employed for the 2D and 3D RVE models and all of the models
AN
were analyzed in Y direction. Figure 18 illustrates the effect of boundary conditions in 2D
RVE models for the three microstructures. The results demonstrated that, for DP980,the
M

symmetric boundary condition predicts the true stress-strain curve of DP steel slightly more
than the periodic boundary condition, whereas for DP800, both boundary conditions predict
ED

the same values. Figure 19 shows the effects of boundary conditions for 3D models, which
predicted the same flow curves for DP980 and DP800 under both boundary conditions. The
symmetric boundary condition needs the fewer constraint equations than the periodic one
PT

and, consequently, the solution of this model was less time consuming. Thus, in other
investigations for the 3D model, the symmetric boundary condition can be used.
CE

5.2. Validating simulation results by experimental data


AC

Since the 2D micromechanical models predicted roughly similar outcomes in X and Y


directions under both boundary conditions, the experimental results were compared with
those of 2D micromechanical models in Y direction and with periodic boundary condition
(Figure 20). It is obvious that 2D models are not able to predict the experimental stress-strain
curves of dual phase steels. In Figure 21, the results of 3D micromechanical models obtained
under periodic boundary conditions under loading in Y direction are compared with the
experimental ones. It could be observed that the 3D micromechanical models have the
acceptable results.

11
ACCEPTED MANUSCRIPT

5.3. Initiation of strain localization pattern


In this section, the localized strain regions in ferrite phase are investigated both numerically
and experimentally in detail. The localized deformations in ferrite phase could be divided into
three types: (1) in the small ferrite grains located between martensite islands; (2) near the
martensite island due to the crack in martensite grains; and (3) in the middle of large ferrite
grains.
Figure 22 shows three kinds of localizations that could happen in ferrite phase. In Figure 22a,

T
localization occurred in a ferrite grain located between two closely spaced martensite grains.

IP
The existence of some harder phase grains close to each other resulted in stress/strain
localization. In addition, at high strain, the inhomogeneity between two phases causes some

CR
stress concentration which accelerates the localization in these regions. The second type of
localized deformation was observed in ferrite grains within the neighborhood of martensite

US
islands, in which the martensite grains have much less deformation than the neighboring
ferrite phase and the mismatch between the values of deformation in two phases causes
deformed bands on the boundary. This phenomenon that happened in martensite grains A and
AN
B is shown in Figure 22b. Generally, dislocation density near the boundary and center of a
grain is more than that of other regions. At low strains, dislocation density prevents a large
M

deformation in the boundary, whereas in the center of a large ferrite grain, this phenomenon
leads to plastic strain (Figure 22c).Also, the martensite grains located in the lower layer and
ED

middle of large ferrite grain can affect this phenomenon.


In Figure 23, the numerical results of 2D and 3D micromechanical models were compared
with the in-situ test results. In this comparison, six regions were examined in both numerical
PT

and experimental studies in detail. There were some accepted numerical results; however,
some regions (like regions A) were not identified in the experimental results. These
CE

discrepancies occurred because of changing in the material's behavior or the low accuracy in
the image processing procedure. Figure23d shows that the 2D micromechanical model
AC

predicted a critical slip band in the region B, whereas it could not be seen in the 3D
micromechanical model (Figure 23e) and in-situ test result (Figure 23c). Although 2D and 3D
micromechanical models predicted the same regions for slip bands, the value of strain in the
2D micromechanical model was more than that of the 3D micromechanical model and
experimental result. As can be seen in regions D and E, when ferrite grains are located
between martensite islands, the value of strain is more than that of the other regions. In
regions C and F, micromechanical models and experiments experience local plastic strain
localization in the boundary between ferrite and martensite grains.
12
ACCEPTED MANUSCRIPT
Figure 24 shows the development of localization pattern in the ferrite phase. The equivalent
plastic strains are shown in the 2D and 3D micromechanical models for different
unidirectional strains from 0.04 to 0.13 in four different steps. In Figure 24a, the initiation of
the slip bands in the microstructure is illustrated. Figure 24b shows that some of the slip
bands grew to join each other, while some other remained unchanged. Most of the bands
growth occur through the ferrite phase and in the direction of the neighbor of martensite
phase. In Figure 24c, the propagation of bands is presented. Finally, in Figure 24d, the final
slip bands can be observed. As reported previously [32], the final slip is caused by some
shear type coalescence due to the stress field in the microstructure. In the 2D models, the slip

T
bands in the out-of-plane direction (Z direction) are essentially similar, which can lead to an

IP
earlier failure than that of experimental results (Figure 20). Also, the 3D model shown in

CR
Figure 24has fewer regions containing low plastic strain than the 2D model, which can be due
to the fact that the microstructures of different layers and their effects on the generation of
strain in them vary. In other words, grains of martensite in sub-layers of ferrite grains

US
decrease the produced strain in the ferrite phase. Therefore, the slip bands produced in the Z
direction can result in a failure delay compared to the 2D model. Also, in Figure 24, different
AN
kinds of localization pattern (as specified and numbered in Figure 22) are shown for 3D
models.
M

Different type of boundary conditions imposes various constraints on the displacements of


nodes locating on boundaries of RVEs. Therefore, the type of boundary conditions affects the
ED

development of shear strain localization inside the RVEs. For example, Figure 25 illustrates
the distribution of plastic strain in 2D micromechanical model due to εyy=0.13 and under
symmetry and periodic boundary conditions. As can be seen, although the overall shear band
PT

pattern for both symmetry and periodic boundary conditions are similar, however, the
localized shear bands are different.
CE
AC

6. Conclusion
In this paper, the micromechanical behavior of DP steels was investigated using experimental
and numerical methods. Two different micro structural methods were used as numerical
approaches: (i) 2D micromechanical models (plane strain) generated based on the real 2D
microstructural patterns of DP800 and DP980 steels to provide acceptable predictions of the
material behavior which was experimentally obtained under tensile loading and (ii) 3D model
in which martensite phase was distributed randomly in ferrite phase for each layer in the out-
of-plane direction. As shown in Section 5, the randomly extruded 3D model was able to

13
ACCEPTED MANUSCRIPT
provide close agreement (for both DP800 and DP980 steels) with the experimental results.
The FE results indicate that both micromechanical models have negligible dependency on the
boundary conditions, and both models have an isotropic behavior.
The results also revealed that, for DP800 and DP980 steels, the stress-strain curves predicted
by the 3D random RVE model are approximately 20% and 24% higher than those of the 2D
plane strain RVE model, respectively. In 3D models, the grains of martensite in sub-layers of
ferrite grains decrease the produced strain in the ferrite phase. This could eventually result in
the reinforcement of ferrite phase by martensite phase in the adjacent layer. When the slip
bands form in the ferrite phase of the RVE layers due to being located next to the martensite

T
phase of the adjacent layer, the slip band growth will be prevented and are not formed

IP
consistently. This inconsistency of the slip bands can cause a delay in the rupture because the

CR
slip bands move towards each other. Consequently, the 3D model can provide higher strength
than the 2D model until the slip bands approach each other and eventually the model
ruptures.

US
Finally, the numerical results showed that the 3D model was able to provide good estimates
AN
for the microstructural deformation behavior obtained from the experimental result (in-situ
test). Also investigated in this paper were three types of initiation of strain localization that
occurred at low and high applied strains. At low strains, the strain localization takes place
M

because of softening localization initiated in the middle of the large ferrite grains. However,
at high strains, two effective localizations are observed during the in-situ tensile test: (i) at the
ED

interface of martensite and ferrite phases and (ii) at the small ferrite grains trapped between
the martensite grains for which the amount of strain localization is much higher than other
PT

localization patterns.
CE

Acknowledgements
This work was supported by Shahid Rajaee Teacher Training University under contract
AC

number 27773.

References
[1] W. Bleck, S. Papaefthymiou, A. Frehn, Microstructure and tensile properties in dual phase
and trip steels, Steel research international, Vol. 75, No. 11, pp. 704-710, 2004.
[2] T. Furukawa, M. Tanino, H. Morikawa, M. ENDO, Effects of composition and processing
factors on the mechanical properties of as-hot-rolled dual-phase steels, Transactions of the
Iron and Steel Institute of Japan, Vol. 24, No. 2, pp. 113-121, 1984.

14
ACCEPTED MANUSCRIPT
[3] S. K. Paul, Real microstructure based micromechanical model to simulate microstructural
level deformation behavior and failure initiation in DP 590 steel, Materials & Design, Vol.
44, pp. 397-406, 2013.
[4] J. Kadkhodapour, A. Butz, S. Ziaei Rad, Mechanisms of void formation during tensile testing
in a commercial, dual-phase steel, Acta Materialia, Vol. 59, No. 7, pp. 2575-2588, 2011.
[5] E. Ahmad, T. Manzoor, K. L. Ali, J. Akhter, Effect of microvoid formation on the tensile
properties of dual-phase steel, Journal of materials engineering and performance, Vol. 9, No.
3, pp. 306-310, 2000.
[6] C. Tasan, J. Hoefnagels, M. Geers, A critical assessment of indentation-based ductile damage
quantification, Acta materialia, Vol. 57, No. 17, pp. 4957-4966, 2009.
[7] M. Sarwar, R. Priestner, Influence of ferrite-martensite microstructural morphology on tensile
properties of dual-phase steel, Journal of Materials Science, Vol. 31, No. 8, pp. 2091-2095,
1996.
[8] M. Erdogan, The effect of new ferrite content on the tensile fracture behaviour of dual phase

T
steels, Journal of Materials Science, Vol. 37, No. 17, pp. 3623-3630, 2002.

IP
[9] M. Rashid, GM 980 X- A Unique High Strength Sheet Steel With Superior Formability, Fuel,
Vol. 2014, pp. 10-21, 1976.
[10] M. Rashid, E. Cprek, Relationship between microstructure and formability in two high-

CR
strength, low-alloy steels, Formability Topics–Metallic Materials, pp. 174-190, 1978.
[11] M. Erdogan, R. Priestner, Effect of martensite content, its dispersion, and epitaxial ferrite
content on Bauschinger behaviour of dual phase steel, Materials science and technology, Vol.
18, No. 4, pp. 369-376, 2002.
[12]

[13]
US
F. Zhang, A. Ruimi, P. C. Wo, D. P. Field, Morphology and distribution of martensite in dual
phase (DP980) steel and its relation to the multiscale mechanical behavior, Materials Science
and Engineering: A, Vol. 659, pp. 93-103, 2016.
M. Radu, J. Valy, A.-F. Gourgues, F. L. Strat, A. Pineau, Continuous magnetic method for
AN
quantitative monitoring of martensitic transformation in steels containing metastable
austenite, Scripta materialia, Vol. 52, No. 6, pp. 525-530, 2005.
[14] E. Fereiduni, S. G. Banadkouki, Improvement of mechanical properties in a dual-phase
ferrite–martensite AISI4140 steel under tough-strong ferrite formation, Materials & Design,
M

Vol. 56, pp. 232-240, 2014.


[15] V. Uthaisangsuk, U. Prahl, W. Bleck, Stretch-flangeability characterisation of multiphase
steel using a microstructure based failure modelling, Computational Materials Science, Vol.
ED

45, No. 3, pp. 617-623, 2009.


[16] V. Uthaisangsuk, U. Prahl, W. Bleck, Micromechanical modelling of damage behaviour of
multiphase steels, Computational Materials Science, Vol. 43, No. 1, pp. 27-35, 2008.
[17] X. Sun, K. S. Choi, W. N. Liu, M. A. Khaleel, Predicting failure modes and ductility of dual
PT

phase steels using plastic strain localization, International Journal of Plasticity, Vol. 25, No.
10, pp. 1888-1909, 2009.
[18] X. Sun, K. S. Choi, A. Soulami, W. N. Liu, M. A. Khaleel, On key factors influencing ductile
CE

fractures of dual phase (DP) steels, Materials Science and Engineering: A, Vol. 526, No. 1,
pp. 140-149, 2009.
[19] K. S. Choi, W. N. Liu, X. Sun, M. A. Khaleel, Microstructure-based constitutive modeling of
TRIP steel: Prediction of ductility and failure modes under different loading conditions, Acta
AC

Materialia, Vol. 57, No. 8, pp. 2592-2604, 2009.


[20] H. Shen, T. Lei, J. Liu, Microscopic deformation behaviour of martensitic–ferritic dual-phase
steels, Materials science and technology, Vol. 2, No. 1, pp. 28-33, 1986.
[21] J. Kang, Y. Ososkov, J. D. Embury, D. S. Wilkinson, Digital image correlation studies for
microscopic strain distribution and damage in dual phase steels, Scripta Materialia, Vol. 56,
No. 11, pp. 999-1002, 2007.
[22] A. Alaie, J. Kadkhodapour, S. Z. Rad, M. A. Asadabad, S. Schmauder, Formation and
coalescence of strain localized regions in ferrite phase of DP600 steels under uniaxial tensile
deformation, Materials Science and Engineering: A, Vol. 623, pp. 133-144, 2015.
[23] A. Alaie, S. Z. Rad, J. Kadkhodapour, M. Jafari, M. A. Asadabad, S. Schmauder, Effect of
microstructure pattern on the strain localization in DP600 steels analyzed using combined in-
situ experimental test and numerical simulation, Materials Science and Engineering: A, Vol.
638, pp. 251-261, 2015.
15
ACCEPTED MANUSCRIPT
[24] S. Papaefthymiou, W. Bleck, U. Prahl, C. Acht, J. Sietsma, S. van der Zwaag,
Micromechanical damage simulations of TRIP steels, in Proceeding of, Trans Tech Publ, pp.
1355-1360.
[25] A. Ramazani, K. Mukherjee, U. Prahl, W. Bleck, Modelling the effect of microstructural
banding on the flow curve behaviour of dual-phase (DP) steels, Computational Materials
Science, Vol. 52, No. 1, pp. 46-54, 2012.
[26] M. Ayatollahi, A. C. Darabi, H. Chamani, J. Kadkhodapour, 3D Micromechanical Modeling
of Failure and Damage Evolution in Dual Phase Steel Based on a Real 2D Microstructure,
Acta Mechanica Solida Sinica, Vol. 29, No. 1, pp. 95-110, 2016.
[27] A. E562-08, “Standard Test Method for Determining Volume Fraction by Systematic Manual
Point Count, 2008.
[28] C. Thomser, U. Prahl, H. Vegter, W. Bleck, Modelling the mechanical properties of
multiphase steels, Computer Methods in Materials Science, Vol. 7, pp. 42-46, 2007.
[29] R.-M. Rodriguez, I. Gutiérrez, Unified formulation to predict the tensile curves of steels with

T
different microstructures, in Proceeding of, Trans Tech Publ, pp. 4525-4530.

IP
[30] J. Kadkhodapour, A. Butz, S. Ziaei-Rad, S. Schmauder, A micro mechanical study on failure
initiation of dual phase steels under tension using single crystal plasticity model, International
Journal of Plasticity, Vol. 27, No. 7, pp. 1103-1125, 2011.

CR
[31] H. J. Böhm, A short introduction to basic aspects of continuum micromechanics, Cdl-fmd
Report, Vol. 3, 1998.
[32] A. B. J. Kadkhodapour, S. Ziaei Rad, Mechanisms of void formation during tensile testing in
a commercial, dual-phase steel, Acta Materialia, Vol. 59, No. 7, pp. 2575-2588, 2011.

Figures:
US
AN
M
ED
PT

Figure 1. Intercritical annealing process.


CE
AC

16
ACCEPTED MANUSCRIPT

20 μm 20 μm

T
(a) (b)

IP
CR
US
AN

20 μm
M

(c)
ED

Figure 2. Optical micrograph of dual phase steels, (a) Microstructure-A of DP 980 with
, (b) Microstructure-B of DP980 with , and (c) Microstructure-C of
PT

DP800 with

1200
CE

1000
True Stress (MPa)

800
AC

600

400
DP 980
200 DP 800

0
0 0.02 0.04 0.06 0.08 0.1 0.12
True Strain (-)

Figure 3. Experimental true stress-strain curves of DP 800 and DP980 steels

17
ACCEPTED MANUSCRIPT

a b

T
IP
c d

CR
US
AN

Figure 4. (a) Portable fixture that is located inside the scanning electron microscope, (b)
M

Measurement of specimen dimensions for different stages using video projector and profile
projector devices, (c) and (d) Deformational behavior of microstructure close to failure.
ED

1
ELEMENTS
MAT NUM
PT
CE
AC

Y
Z X

Figure 5. 2D RVE model of DP980 microstructure-B with 90,000 elements (elements of


ferrite phase are removed for better illustration).

18
ACCEPTED MANUSCRIPT

Z X

T
IP
Figure 6. 3D RVE model with 512,000 elements

CR
700

600
US
AN
500
True Stress (MPa)

400

300
M

200 25%-MPF

100 33%-MPF
ED

0
0 0.1 0.2 0.3 0.4 0.5
True Strain (-)
PT

Figure 7. Flow curves for ferrite phase of DP980 and DP800


CE

3000
AC

2500
True Stress (MPa)

2000

1500

1000
25%-MPF

500 33%-MPF

0
0 0.01 0.02 0.03 0.04 0.05 0.06
True Strain (-)

19
ACCEPTED MANUSCRIPT
Figure 8. Flow curves for martensite phase of DP980 and DP800.

T
IP
Figure 9. Deformation of periodic boundary condition under uniaxial loading

CR
US
AN

Y
M

X
ED

Figure 10: Deformation of symmetric boundary condition under uniaxial loading


PT
CE
AC

Z X

(a) (b)

Figure 11. FE mesh of 3D RVE model of DP980 steel, (a) coarse mesh (64000 elements),
and (b) fine mesh (512000 elements)

20
ACCEPTED MANUSCRIPT

T
IP
Figure 12. Comparison of stress-strain curves between the periodic 3D RVE models

CR
with 512000 elements (fine mesh) and 64000 elements (coarse mesh)

US
AN
M
ED

(a) (b) (c)


PT
CE
AC

(d) (e)

Figure 13. 2D RVE of DP980 steel by periodic boundary condition with (a) 400×400
elements, (b) 300×300 elements, (c) 200×200 elements, (d) 100×100 elements, and (e) 50×50
elements.

21
ACCEPTED MANUSCRIPT

T
IP
CR
Figure 14. Comparing stress-strain curves between the periodic 2D-RVE models with
different mesh sizes.

US
AN
M
ED
PT

(a) (b)
CE
AC

(c) (d)

22
ACCEPTED MANUSCRIPT

T
(e) (f)

IP
Figure 15. The true stress-strain curves obtained from 2D micromechanical models in
loading directions X and Y with periodic and symmetric boundary conditions, (a) DP980
microstructure A-under periodic B.C, (b) DP980 microstructure A-under symmetric B.C, (c)

CR
DP980 microstructure B-under periodic B.C, (d) DP980 microstructure B-under symmetric
B.C, (e) DP800 microstructure C-under periodic B.C and (f) DP800 microstructure C- under
symmetric B.C.

US
AN
M
ED
PT
CE

(a) (b)
AC

23
ACCEPTED MANUSCRIPT
(c) (d)

T
IP
(e) (f)

CR
Figure 16.The true stress-strain curves obtained from 3D micromechanical models in loading
directions X and Y with periodic and symmetric boundary conditions, (a) DP980

US
microstructure A-under periodic B.C, (b) DP980 microstructure A-under symmetric B.C, (c)
DP980 microstructure B-under periodic B.C, (d) DP980 microstructure B-under symmetric
B.C, (e) DP800 microstructure C-under periodic B.C and (f) DP800 microstructure C- under
AN
symmetric B.C.
M
ED
PT
CE

Figure17. Distribution of martensite area fraction in directions X, Y, and Z, in DP980


AC

microstructure-A

24
ACCEPTED MANUSCRIPT
1000 1000

800 800

True Stress (MPa)


True Stress (MPa)

600 600

400 400
Periodic B.C Perioidc B.C
200 Symmetric B.C 200
Symmetric B.C

0 0
0 0.02 0.04 0.06 0.08 0.1 0 0.02 0.04 0.06 0.08 0.1
True Strain (-) True Strain (-)

T
(a) (b)

IP
1000

CR
800
True Stress (MPa)

600

US
400
Periodic B.c
200
Symmetric B.C
AN
0
0 0.02 0.04 0.06 0.08 0.1 0.12
True Strain (-)

(c)
M

Figure 18.The true stress-strain curves obtained from 2D micromechanical models in loading
direction Y under the periodic and symmetric boundary conditions, (a) DP980
ED

microstructure-A, (b) DP980 microstructure-B, and (c) DP800 microstructure-C,


PT

1200 1200
CE

1000 1000
True Stress (MPa)
True Stress (MPa)

800 800

600 600
AC

400 400
Periodic B.C Perioidc B.C

200 Symmetric B.C 200 Symmetric B.C

0 0
0 0.02 0.04 0.06 0.08 0.1 0 0.02 0.04 0.06 0.08 0.1
True Strain (-) True Strain (-)

(a) (b)

25
ACCEPTED MANUSCRIPT
1200

1000

True Stress (MPa)


800

600

400
Periodic B.C

200 Symmetric B.C

0
0 0.02 0.04 0.06 0.08 0.1 0.12
True Strain (-)

T
(c)

IP
Figure 19.The results of 3D micromechanical models in loading direction Y under periodic
and symmetric boundary condition, (a) DP980 microstructure A, (b) DP980 microstructure B,

CR
and (c) DP800

1200

1000 US
True Stress (MPa)

800
AN
600

400 Simulation microstructure A


M

Simulation microstructure B
200 Experiment

0
ED

0 0.02 0.04 0.06 0.08 0.1


` True Strain (-)

(a) (b)
PT

Figure 20. Comparing experimental and 2D simulated models' flow curve in (a) DP980
microstructure A and microstructure B, and (b) DP800 micristructure C.
CE
AC

26
ACCEPTED MANUSCRIPT

T
IP
(a) (b)

CR
Figure 21. Comparing the experimental and predicted stress-strain curves based on 3D RVE
models of (a) DP980 microstructures A and B, (b) DP800microstructure-C

a b
US c
AN
M
ED

Figure 22. Different kinds of localized deformation in DP steels: (1) in a small ferrite grains
between martensite phases, (2) near martensite phase due to crack in martensite grains, and
PT

(3) in the middle of large ferrite grains


CE
AC

27
ACCEPTED MANUSCRIPT

T
IP
.108

d .11
e
.096

.0977

CR
.084

.0855
B
.072

.0733 B
A
A
.06

.0611
MN

US
MX

.048

C E .0488 C E
X

.036
Y

.0366
D D
.024

.0244
AN
.012

.0122
F
F
0

0
M

Figure 23. Deformed shape of microstructure (applied strain is 13%), (a) initial
microstructure, (b) deformed shape of microstructure, (c) strain distribution (εyy) in the in-situ
test results, (d) strain distribution εyy in 2D micromechanical model, and (e) strain distribution
ED

(εyy) in 3D micromechanical model.


AUG 2 2015
AUG 2 2015 AUG 2 2015 13:13:34
.108

13:11:31 AUG 2 2015 .11


13:12:13
PT

13:12:52
.096

.0977
2
.084

3 .0855
2
.072

MN
MN MN MN .0733
MX
MX
.06

MX MX .0611
1
MN

3
MX
CE

.048

.0488
X

.036
Y

.0366
1
.024

.0244
2
.012

.0122
1
3D
0

0
AC

0 .026667 .053333 .08 .106667


.013333 .04 .066667 .093333 .12
0 .026667 .053333 .08 .106667
.013333 .04 .066667 0 .093333 .026667.12 .053333 0.08 .026667
.106667 .053333 .08 .106667
.013333 .04 .066667 .013333
.093333 .04.12 .066667 .093333 .12
.108

.11
.096

.0977
.084

.0855
.072

.0733
.06

.0611
MN

MX

.048

.0488
X

.036
Y

.0366
.024

.0244
.012

.0122

2D
0

28
ACCEPTED MANUSCRIPT

Figure 24. Development of strain localization in 2D and 3D micromechanical models under periodic
boundary conditions: (a) , (b) , (c) , and (d) .

.108
.11

.096
.0977

.084
.0855

.072
.0733

.06
.0611

MN

MX

.048
.0488

.036
Y

Z
.0366

T .024
.0244

IP .012
.0122

0
0

CR
(a)
US (b)
AN
Figure 25: Development of strain localization in 2D micromechanical models under (a) symmetry
boundary conditions,(b) periodic boundary conditions. .
M
ED
PT
CE
AC

29
ACCEPTED MANUSCRIPT

Tables

T
IP
Table 1. Chemical compositions of low carbon steel, heat-treated in this study.

CR
Element C Mn Si P S Cr Ni Al Mo V Cu Co

0.01
Mass contents% (-). 0.2 1.1 0.22 0.004 0.02 0.157 0.08 0.04 0.008 0.121 0.019
1

US
Table 2. Summary of the model constants for the flow curve prediction of individual phases
AN
α[30] M [30] b[30] L [30] K [30]
0 µ[30]
Steel Phase Fraction (%)
(MPa) (MPa) (MPa)
(-) (-) (m) (m) (-)
M

0 6
Ferrite 67 155.6 175 0.33 3 80,000 1
DP980
0 6
Martensite 33 155.6 1423 0.33 3 80,000 38 10 41
ED

0 6
Ferrite 75 155.6 155 0.33 3 80,000 0.67
DP800
0 6
Martensite 25 155.6 2124 0.33 3 80,000 38 10 41
PT
CE
AC

30

You might also like