Sunteți pe pagina 1din 10

PAPER 6 Yngve Bergström February 2011

___________________________________________________________________________

A DISLOCATION BASED MODEL FOR THE WORK HARDENING


BEHAVIOUR OF DUAL PHASE STEELS

INTRODUCTION
The properties of DP steels are characterized by a low yield strength due to the absence of
Lüders bands and a high rate of work hardening, which results in a high tensile strength and
good formability (1). DP steels also show a high energy absorbing ability which implies a
good crashworthiness (2–5). The explanation to the materials behaviour is to be found in the
microstructure, which in DP steels mainly consists of two phases: ferrite and martensite.

Over the years many attempts have been made to describe the stress-strain behaviour of DP
steels (6-10). In most cases empirical relationships have been applied but also physically
based models have been developed. One frequently used concept is the Ashby-model (11)
which is based on the assumption that the dislocations may be divided into two types:
statistically stored (SSD) and geometrically necessary dislocations (GND) with pile ups of
dislocations in arrays. The latter type of dislocations are supposed to eliminate stress
concentrations and strain gradients originating from differences in hardness between the
ferrite and martensite phases. Unfortunately, these arrays have never been observed in high
stacking fault materials like ferrite suggesting that this type of explanation is not applicable to
steel (12). A large number of strain gradient plasticity (SGP) theories have also been proposed
with moderate success. It is quite obvious, therefore, that a new simple physically based
dislocation theory for the stress-strain behaviour of DP steels is needed. It is also obvious that
such a theory must be based on the in-homogeneous behaviour of the plastic deformation
process involved in this type of materials. Such a model was recently presented by Bergström
et al (13) and we will below give a short presentation of it and apply it to a DP800 steel tested
at room temperature and a strain rate of 10-2 s-1.

THEORY (13)
Background
Quite often uniaxial stress-strain curves are used to investigate the plastic deformation
behaviour in metals and it is customary to assume that the whole testing volume of the tensile
test specimen takes part in the plastic deformation process. This is a reasonable assumption
for single phase materials but it is definitely a wrong one for DP steel consisting of a soft
ferritic phase and a hard martensitic phase. In fact, experiments clearly indicate that the
martensite phase in a DP steel normally does not undergo plastic deformation for strains
below necking. Hence the plastic deformation process is localised to the ferrite implying that
the plastic deformation is indeed inhomogeneous and gives rise to large local strains in the
specimen.
Strain
Töjningin the
i αferrite Macroscopic strain
Makroskopisk töjn

a)
∆l α ∆l
l0 l0
l0 ∆l
∆l 2∆l ∆l
b) = M α
l0 / 2 l0 l0
l0/2 ∆l
∆l 2∆l α ∆l
c) = M

l0 / 2 l0 l0
∆l

Fig.1: a) an elongation ∆l of a specimen of length, l0, containing 100% ferrite


b-c) an elongation ∆l of a specimen containing 50% ferrite and 50% martensite
distributed in two different ways.

1
Yngve Bergström homepage: www.plastic-deformation.com/
PAPER 6 Yngve Bergström February 2011
___________________________________________________________________________
This is illustrated in Fig.1 where the macroscopic, or global strain and the corresponding local
strain in the ferrite is compared for three different cases . Specimen a/ in Fig. 1 contains 100%
ferrite and, since the plastic deformation is homogeneous i.e. containing 100% ferrite, we can
∆l
see that the macroscopic strain, , after an elongation ∆l equals the local strain in the
l0
ferrite. In specimens b/ and c/ consisting of 50% ferrite and 50% martensite distributed in two
different ways we notice that the local strain in the ferrite, for the same amount of elongation,
is two times larger than the macroscopic strain. The explanation is of course that the
martensite does not take part in the plastic deformation and that all plastic deformation is
localised to the ferrite. This simple example therefore explains why the rate of strain
hardening is higher in a DP steel than in a single-phase steel.

It has also been experimentally verified that the plastic deformation in the ferrite of a DP steel
is in-homogeneous. This can be seen from the EBSD-pictures in Fig. 2 (13). The blue colour
indicates small miss-orientations (small plastic deformation) and red colour large miss-
orientations(large plastic deformation). The colours green and yellow lie in between.

The results in Fig. 2b emanate from an un-deformed specimen of a DP 800 steel. The green
colour close to the martensite particles suggests that the stresses caused by the transformation
of austenite to martensite gives rise to a local plastic deformation in the ferrite close to the
martensite surfaces. Inside the ferritic grains the colour is blue indicating small or negligible
plastic deformation.

Fig. 2b. shows result from the same DP steel being strained 8%. In this case we can see that
the colours in the ferrite, close to the martensite, have changed from green to yellow and even
red. Inside the ferrite grain the colour have changed from dark blue to red, yellow, green and
light blue. The results also indicate the possibility that small volume fractions ferrite, inside
the ferrite grains, have not undergone plastic deformation at all – almost dark blue spots.
Hence, the results suggest that there are pronounced strain gradients also inside the ferrite
grains with large miss-orientations close to the martensite particles and smaller ones in the
centre of the grains.

It is therefore reasonable to assume that local volume fractions of the ferrite controls the
plastic deformation process in DP steels. It is also obvious that these volume fractions are
strain dependent.

Fig. 2a; EBSD-analysis of undeformed DP800. The colours indicate degree of plastic deformation (13). Blue is
small deformation and red is large deformation. Green and yellow represent deformations in between. See text.

2
Yngve Bergström homepage: www.plastic-deformation.com/
PAPER 6 Yngve Bergström February 2011
___________________________________________________________________________

Figur 2b: EBSD-analys av 800DP strained to 8%. The colours indicate degree of plastic deformation (13). Blue
is small deformation and red is large deformation. Green and yellow represent deformations in between. See
text.

We will now take these two types of plastic in-homogeneity into account and derive a simple
physically based dislocation theory for the stress-strain behaviour of DP steels (13). The
theory is discussed in the light of results from experimental uniaxial testing, light optical and
SEM/EBSD-studies of the microstructure. Since the new theory is developed from the
original Bergström dislocation theory, it is appropriate to start with a short summary.

The original Bergström theory: a summary (14)


It is well established that the true flow stress, σ(ε), in metals and alloys is related to the total
dislocation density ρ(ε) as

σ (ε ) = σ i 0 + σ def (ε ) = σ i 0 + α ⋅ G ⋅ b ⋅ ρ (ε ) (1)

also known as the Taylor equation(15), where σ i 0 , is the friction stress, σ def (ε ) , is the work
hardening component of the flow stress, α is a dislocation strengthening constant, G is the
shear modulus, b is the nominal value of the Burgers vector, and ε is the true strain. If the
validity of eqn(1) is accepted the main problem in formulating a dislocation theory for the true
stress-true strain behaviour of metals and alloys is to derive a relationship for ρ(ε).

For single-phase metals and alloys Bergström (14) derived the following expression for the
strain dependence of ρ(ε):

dρ m
= − Ω ⋅ ρ (ε )
d ε b ⋅ s(ε ) (2)

where ρ(ε) is the total dislocation density, s(ε) is the mean free path of dislocation motion, b is
the nominal value of the dislocation Burger´s vector, Ω is a strain independent material
constant representing the remobilisation (dynamic recovery) of immobile dislocations and m
is the Taylor constant.

The following relationship for s(ε) has proved to give an excellent description of the strain
dependence of the dislocation mean free path, s, in single phase metals (16)

s (ε ) = s0 + ( s1 − s0 ) ⋅ e− k ⋅ε (3)

3
Yngve Bergström homepage: www.plastic-deformation.com/
PAPER 6 Yngve Bergström February 2011
___________________________________________________________________________
where s1 and s0 are the initial and final values of s. k is the rate constant determining the rate
at which s(ε) goes from s1 to s0.

A dislocation theory for the uniaxial stress-strain behaviour of DP-steel.(13)


As mentioned in the introduction there are strong experimental evidences indicating that the
plastic deformation process in DP steel is in-homogeneous. In deriving a dislocation theory
for this type of metal we will therefore make the following assumptions:

1. The theory is valid for uniaxial strain and for tensions up to necking
2. The work hardening process starts in the ferrite close to the phase interface of ferrite
and martensite, partly as a result of the presence of high residual stresses caused by the
transformation from austenite to martensite, partly due to the incompatibility
conditions which give rise to stress concentrations in these areas. The work hardening
process is thus initially located in the ferrite close to the martensite surfaces. As a
result of increasing work hardening in these areas, the deformation process is forced to
proceed inwards the ferrite grains due to a lower hardening resistance there.
3. The investigated DP steels consist of the following volume fractions:
(i) martensite, fm, which deforms elastically but not plastically
(ii) active ferrite, f, which deforms elastically and plastically
(iii) non-active ferrite, fundef, which deforms elastically and successively
deforms plastically
4. A non-homogeneity factor f is introduced which is defined as de active part of the total
volume fraction of ferrite, f0, participating in the plastic deformation process. At small
strains, that is, in the beginning of the deformation process, f << f0. With increasing
strain, f is assumed to grow towards f0. As a result of the continuously increasing
volume fraction of active ferrite, the rate of dislocation creation and work hardening
decreases with increasing strain.
5. The un-deformed fraction of ferrite, that is, the continuously decreasing amount f0 - f,
may be considered as hard, not participating in the deformation process and can in
that respect be compared with the martensite fraction. Therefore, stress concentrations
may arise in between the deformed and undeformed ferrite as well. This reminds of
the Lüder-phenomenon where a front separates deformed ferrite from undeformed.
6. The plastic deformation process does not necessarily start in all ferrite grains at the
same time but may randomly and continuously be initiated at different locations
during the plastic deformation process.
7. Although the dislocation density will show local variations due to the inhomogeneous
plastic deformation, eqn(1) is assumed to hold for the average dislocation density in
the active ferrite volume.
8. The mean free path of mobile dislocations is assumed to decrease exponentially with
strain in accordance with eqn(3).

The plastic strain dependence of the dislocation density, ρ, in the active ferrite may according
to the chain-rule be written

d ρ d ρ dε
= ⋅ (4)
dε f dε dε f

where εf is the local strain in the active ferrite and ε is the global strain. Since the global
strain, ε, is related to the local strain, εf, as

4
Yngve Bergström homepage: www.plastic-deformation.com/
PAPER 6 Yngve Bergström February 2011
___________________________________________________________________________

dε ≅ f ⋅ dε f (5)

the global strain dependence of the total dislocation density for a DP steel may be expressed
as (compare with eqn(2)

d ρ (ε ) 1  m 
= ⋅ − Ω ⋅ ρ (ε )  (6)
dε f (ε )  b ⋅ s(ε ) 

We will now assume that the strain dependence of f(ε) may be written, see (16),

f (ε ) = f 0 + ( f1 − f 0 ) ⋅ e− r⋅ε (7)

where r is a material parameter that controls the rate of this progress, f1 is the initial active
volume fraction of the ferrite taking part in the deformation process and f0 is the total amount
of ferrite. f(ε) thus starts from the value of f1 and hereafter grows with increasing strain
towards f0.

Using eqn(3) and (7) in eqn(6) we finally achieve an expression for the inhomogeneous
deformation behaviour of ferrite in DP steel (13)

d ρ (ε ) 1  m 
= ⋅ − Ω ⋅ ρ (ε )  (8)
dε f0 + ( f1 − f 0 ) ⋅ e− r⋅ε  b ⋅ ( s0 + ( s1 − s0 ) ⋅ e
− k ⋅ε
) 

Eqn(8) gives the dislocation strain dependence for a DP steel expressed in the ferrite phase
parameters, that is, a description of the continuous localisation process in the DP steel. A
comparison between eqn(8) and eqn(2), which is the original Bergström expression for the
“homogeneous” ρ-ε behaviour in pure ferrite, shows that eqn(8) transcends into eqn(2) if
f0=f1=1, that is, if we assume that fm=0 and that the deformation is “homogeneous”.

EXPERIMENTS, RESULTS AND DISCUSSION

Experiments
In order to demonstrate the possibilities of the dislocation model presented above we shall
analyse a true stress-strain curve from a DP800 steel recorded at room temperature and at a
strain rate of 10-2 s-1. The martensite volume fraction in the investigated steel varies in the
range 25-30%.

Fitting procedure
A special Matlab subroutine, based on the Matlab Curve Fitting Toolbox, was designed for
the purpose of this type of study. In the fitting procedure the following parameters were kept
constant: α = 0.5, G = 80000 MPa, b = 2.5·10-10 m, m = 2. The following parameters were
allowed to vary freely within certain physical limits: Ω, σi0, ρ0, f1, f0, r, s1, s0 and k. The
start values were chosen on the basis of the experience obtained from previous analyses of
this type of DP steel.

5
Yngve Bergström homepage: www.plastic-deformation.com/
PAPER 6 Yngve Bergström February 2011
___________________________________________________________________________

Results
The results from this fitting procedure are presented in Fig. 3, 4 and 5. The left graph in Fig. 3
shows the experimental and theoretical stress-strain curves. The theoretical red curve is
covering the experimental data. The red cross represents the theoretically calculated true
strain to necking and the corresponding true flow stress. The error of fit is shown in the right
graph where medium and running medium values are given. We can see that the fit is god
over the whole strain interval. At smaller strains, however, the accuracy of fit is slightly
worse. This is presumably due to the presence of residual stresses caused by the austenite –
martensite transformation. It seems that these stresses have been released after ~ 2% strain
which seems reasonable.

Fig 3. Experimental and theoretical σ – ε curves for the investigated DP 800 steel – left, and
corresponding accuracy of fit – right. For details, see the text.

Fig 4. A summary of the parameter values obtained in the fit – left, and the theoretical ρ-ε
curve

The parameter values recorded in the fit are presented to the left in Fig. 4 and to the right the
corresponding theoretically calculated ρ-ε curve is shown.

The variations of f and s with strain are depicted in Fig. 5. The left graph indicates that the
effective volume fraction of ferrite increases from an initial value f1=0.15 towards a

6
Yngve Bergström homepage: www.plastic-deformation.com/
PAPER 6 Yngve Bergström February 2011
___________________________________________________________________________
maximum value f0 = 0.73. The rate constant r is equal to 21.6. In the right graph we can see
that the mean free path s of dislocation motion decreases from the initial value s1 = 0.86 µm to
a final value s0 = 5.7 µm. The rate constant k is equal to 297.8 implying that s reaches s0 after
around 1% of strain.

In Fig. 5 The corresponding theoretically calculated f(ε)-ε curve is shown to the left while the
corresponding s(ε)-ε curve is depicted to the right.

Discussion
Let us now discuss the reasonableness of the parameter values obtained in the fitting
procedure presented above, see Fig. 4.

The dislocation remobilisation parameter, Ω, takes a value of 5 in the fit. This is a typical
room temperature Ω - value for ferrite (14) and supports the assumption that the plastic
deformation process is localised to the ferritic phase in the investigated DP steel.

The high value of the “grown-in” dislocation density ρ0 = 1.12·1014 m-2 is not surprising since
the austenite-martensite transformation gives rise to large stresses in the ferrite. The above
figure is an average value and it is reasonable to believe that the “grown-in dislocation density
is even higher close to the martensite surfaces.

f1 is the initial volume fraction of active ferrite and the low value of 0.15 tells us that initially
only 15% of the ferrite takes part in the plastic deformation process. At necking f has reached
a value of approximately 0.6, see Fig 5. If we neglect the small effect of a strain dependent
mean free path s this means that the rate of dislocation generation has decreased by a factor of
~4 and the rate of work hardening by a factor of ~2 after 7% global strain. The high initial
rate of work hardening is thus caused by a low active volume fraction ferrite. It is also
obvious from Fig. 5 that f does not reach the value of f0 at necking and that therefore there
may exist ~ 10% un-deformed ferrite in the testing specimen. This may then explain the blue
dots observed in the centre of the ferrite grains in Fig 2b/.

The mean free path, s, of dislocation motion decreases very rapidly (due to a high k value)
from s1 = 0.86 µm to s0 = 0.57 µm within 1% of deformation. This means that the dislocation
cell diameter attains a value of approximately 0.6 µm after this strain. These s-values are in
fair agreement with the dislocation cell diameters measured by Korzekwa et al (17) in a DP

7
Yngve Bergström homepage: www.plastic-deformation.com/
PAPER 6 Yngve Bergström February 2011
___________________________________________________________________________
steel by using transmission electron microscopy. They are also in agreement with the distance
measured between the precipitates observed in the ferrite matrix in the investigated DP steel.

A simple and useful way of checking the reasonableness of the friction stress, σi0, and, ρ0, is
to proceed from eqn(1) and (8) and assume that the global strain is small. The flow stress may
then approximately be written

m ⋅ε
σ (ε ) ≈ σ i 0 + α ⋅ G ⋅ b ⋅ ρ0 + (9)
b ⋅ s1 ⋅ f1

where the involved parameters are defined above. At a global strain of 0,002 and the
parameter values presented in Fig. 4. the corresponding true flow stress is calculated to be
~675 MPa which is in good agreement with the corresponding experimentally observed flow
stress value, see Fig. 4. It is also interesting to note that the actual local strain in the ferrite at
this stage is approximately equal to 0.002/0.15=0.013 which also explains the high initial rate
of work hardening in DP steels.

In a recently published paper (13) the following four types of DP steels were investigated with
the aid of the present dislocation theory: DP500, DP600, DP800 and DP 980. The objective of
that study was to investigate the impact of martensite content and ferrite grain size on the
work hardening behaviour. The theory proved to be very useful for this purpose.

CONCLUSIONS
The above discussed dislocation theory accurately describes the plastic deformation behaviour
of DP steel. Results from microstructure investigations as well as results from analysis of the
stress-strain behaviour in terms of the present theory, tell us that the plastic deformation
process in DP steel is inhomogeneous By introducing the concept of a non-homogeneity
parameter f (ε), that specifies the volume fraction of ferrite that takes active part in the plastic
deformation process, it is possible to give a precise physical description of the deformation
behaviour until necking.

REFERENCES
(1) C. Federici, S.Maggi, and S. Rigoni, The Use of Advanced High
Strength Steel Sheets in the Automotive Industry, Fiat Auto—
Engineering & Design, Turin, Italy, 2005.

(2) J. R. Fekete, J. N. Hall, D. J. Meuleman, and M. Rupp,


“Progress in implementation of advanced high-strength steels
into vehicle structures,” Iron & Steel Technology, vol. 5, no. 10,
pp. 55–64, 2008.

(3) R. Koehr, ULSAB-AVC Advenced Vehicle Concepts Overview


Report, in Safe, Affordable, Fuel Efficient Vehicle Concepts for the
21st Century Designed in Steel, Porsche Engineering Services,
Detroit, Mich, USA, 2002.

(4) P. Tsipouridis, E. Werner, C. Krempaszky, and E. Tragl,


“Formability of high strength dual-phase steels,” Steel Research
International, vol. 77, no. 9-10, pp. 654–667, 2006.

8
Yngve Bergström homepage: www.plastic-deformation.com/
PAPER 6 Yngve Bergström February 2011
___________________________________________________________________________
(5) M. Takahashi, “Development of high strength steels for
automobiles,” Nippon Steel Technical Report, no. 88, pp. 2–7,
2003.

(6) J. Bouquerel, K. Verbeken, and B. C. De Cooman,


“Microstructure-based model for the static mechanical
behaviour of multiphase steels,” Acta Materialia, vol. 54, no.
6, pp. 1443–1456, 2006.

(7) V. Colla, M. De Sanctis,A. Dimatteo,G. Lovicu, A. Solina, and


R. Valentini, “Strain hardening behavior of dual-phase steels,”
Metallurgical and Materials Transactions A, vol. 40, no. 11, pp.
2557–2567, 2009.

[8] Z. H. Cong, N. Jia, X. Sun, Y. Ren, J. Almer, and Y. D. Wang,


“Stress and strain partitioning of ferrite andmartensite during
deformation,” Metallurgical and Materials Transactions A, vol.
40, no. 6, pp. 1383–1387, 2009.

(9) Z. Jiang, Z. Guan, and J. Lian, “Effects of microstructural


variables on the deformation behaviour of dual-phase steel,”
Materials Science and Engineering A, vol. 190, no. 1-2, pp. 55–
64, 1995.

(10) C. Kim, “Modeling tensile deformation of dual-phase steel,”


Metallurgical Transactions A, vol. 19, no. 5, pp. 1263–1268,
1988.

(11) M. F. Ashby, “The deformation of plastically nonhomogeneous


alloys,” in Strengthening Method in Crystals, A.
Kelly and B. Nicholson, Eds., Applied Science Publishers LTD,
London, UK, 1971.

(12) H. Mughrabi, “Dual role of deformation-induced geometrically


necessary dislocations with respect to lattice plane
misorientations and/or long-range internal stresses,” Acta
Materialia, vol. 54, no. 13, pp. 3417–3427, 2006.

(13). Y. Bergström, Y. Granbom and D. Sterkenburg, ”A dislocation-


based theory for the deformation hardening behaviour of DP steels:
Impact on martensite content and ferrite grain size,”
Journal of Metallurgy, vol 2010, Article ID 647198, 16 pages, 2010

(14) Y. Bergström, “The plastic deformation of metals—a dislocation


model and its applicability,” Reviews on PowderMetalurgy
and Physical Ceramics, vol. 2, no. 2-3, pp. 79–265, 1983.

(15) G. I. Taylor, “The mechanism of plastic deformation of


crystals. Part I. Theoretical,” Proceedings of the Royal Society
A, vol. 145, pp. 362–387, 1934.

9
Yngve Bergström homepage: www.plastic-deformation.com/
PAPER 6 Yngve Bergström February 2011
___________________________________________________________________________
(16) Y. Bergström, “The plastic deformation of metals—a dislocation
model and its applicability,” Reviews on PowderMetalurgy
and Physical Ceramics, vol. 2, no. 2-3, pp. 79–265, 1983.

(17) D. A. Korzekwa, D. K. Matlock, and G. Krauss, “Dislocation


substructure as a function of strain in a dual-phase steel,”
Metallurgical Transactions A, vol. 15, no. 6, pp. 1221–1228,
1984.

10
Yngve Bergström homepage: www.plastic-deformation.com/

S-ar putea să vă placă și