Anda di halaman 1dari 11

Yu Cao1

e-mail: yu.cao@chalmers.se

Johan Ahlstrm
Birger Karlsson
Department of Materials and Manufacturing
Technology,
Chalmers University of Technology,
Gothenburg SE-41296, Sweden

Mechanical Behavior of a
Rephosphorized Steel for Car
Body Applications: Effects of
Temperature, Strain Rate, and
Pretreatment
Temperature and strain rate effects on the mechanical behavior of commercial rephosphorized, interstitial free steel have been investigated by uniaxial tensile testing, covering
temperatures ranging from 60 C to 100 C and strain rates from 1 104 s1 to
1 102 s1 encompassing most conditions experienced in automotive crash situations.
The effect of prestraining to 3.5% with or without successive annealing at 180 C for 30
min has also been evaluated. These treatments were used to simulate pressing of the
plates and the paint-bake cycle in the production of car bodies. Yield and ultimate tensile
strengths, ductility including uniform and total elongation and area reduction, thermal
softening effect at high strain rate, and strain rate sensitivity of stress were determined
and discussed in all cases. It was found that the Voce equation s s
0exp / 0 can be fitted to the experimental true stress-true plastic strain data with
good precision. The parameter values in this equation were evaluated and discussed.
Furthermore, temperature and strain rate effects were examined in terms of thermal and
athermal components of the flow stresses. Finally, a thermal activation analysis was
performed. DOI: 10.1115/1.4003491
Keywords: rephosphorized interstitial free steel, tensile properties, temperature effects,
strain rate sensitivity, strain hardening, Voce equation, thermal and athermal components

Introduction

Interstitial free IF steels 17 containing microalloying elements Ti, Nb, and V and very low C content have very good
formability. It is known that Ti, Nb, and V have strong affinity to
C and N and can combine with them to form compounds such as
TiC, TiN, and NbCN. If the content of these microalloying elements in relation to C and N exceeds the stoichiometric ratio of
the compounds mentioned, the dissolved C and N might be removed from their interstitial positions. The steels can then be
considered to be essentially interstitial free. The strain aging that
arises from the segregation of interstitial C or N to the dislocations produced by the deformation is therefore inhibited 79.
However, if the ferrite matrix still contains a certain amount of
solute C or N for some reason, both static and dynamic strain
aging may appear 10,11. The effect of prestrain level on the
yield stress increment in steels with small amounts of interstitial
elements has been found to be weak. Elements, such as P, Si, and
Mn, are normally added to provide solid solution strengthening
12,13, among which P gives the greatest increase in strength per
unit addition. Rephosphorized IF steel, having excellent formability, is a type of steel suitable for deep drawing and is widely used
in car bodies in the automobile industry.
The structural integrity of car bodies in crash situations is
mainly a matter of energy absorption of specific components exposed to loading under demanding conditions such as low temperature and high deformation rates. Furthermore, the manufactur1
Corresponding author.
Contributed by the Materials Division of ASME for publication in the JOURNAL OF
ENGINEERING MATERIALS AND TECHNOLOGY. Manuscript received September 21, 2010;
final manuscript received January 20, 2011; published online March 22, 2011. Assoc.
Editor: Yoshihiro Tomita.

ing process of car bodies including pressing of the plates and the
paint-bake cycle put certain demands on formability and stability.
Temperature and strain rate effects on the mechanical behavior of
the steel in virgin and pretreated states are therefore important to
be known in the above mentioned industrial processes. In addition, structural analyses and safety assessments require constitutive relations between strain and flow stress, which, again, contain
temperature and strain rate as important parameters. Apart from
experimental finding of these relations, an important purpose of
the present study was to fit the experimental data to the phenomenological Voce model and to analyze the obtained model parameters. The results can be implemented in software for stress analysis to accurately model the material behavior in crash situations.

Experimental

2.1 Material. Commercial rephosphorized steel used for deep


drawing taken from industrial regular production was investigated
in this paper. The material was rolled to a thickness of 1.5 mm.
The chemical composition and mechanical properties at room
temperature, as well as grain size, of the alloy are listed in Tables
1 and 2. Data in Table 1 gives the molar ratio between Ti+ Nb
+ V and C + N is 1.38, which is above the critical value required
for the steel to be nominally interstitial free. The material studied
had a polygonal ferritic microstructure with a mean intercept grain
size of 10 15 m with slightly elongated grains along the rolling
direction Fig. 1. Tensile tests showed no noticeable mechanical
anisotropy in the plane of the sheets.
2.2 Testing and Evaluation. In this study, prestraining to
3.5% plastic strain in the rolling direction was done to simulate
the pressing of the plates. This was done at room temperature with
a strain rate of 1 104 s1. Approximately half of the strain

Journal of Engineering Materials and Technology


Copyright 2011 by ASME

APRIL 2011, Vol. 133 / 021019-1

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Table 1 Chemical compositions of the steel studied wt %


C
0.005
V
0.0018

Si

Mn

Cr

Ni

Mo

Cu

0.08

0.39

0.0535

0.0052

0.03

0.02

0.004

0.010

Al

Sn

Ti

As

Nb

Co

0.0296

0.0061

0.0416

0.0017

0.0005

0.0276

0.004

0.0064

hardening capability compared with the strain at geometrical instability was consumed at this strain level. Subsequent heat treatment at 180 C, 30 min in air was performed to simulate the
paint-bake cycle involved in the manufacturing process of automobile body structures. Symbols AR, PS, and PSA will be used to
represent as-received, prestrained, and prestrained and annealed,
respectively. The storage time before tensile testing was approximately 1 yr for AR and 410 months after prestraining for PS and
PSA, respectively. The storage time used will give mechanical
properties representative for the final product.
Tensile specimens were machined with the tensile direction parallel to the rolling direction. The strain rate was varied by six
orders of magnitude from 1 104 s1 to 1 102 s1. In most
cases, duplicate specimens were tested.
A servohydraulic Instron 8032 load frame equipped with a temperature chamber was used for low strain rate testing up to 1
101 s1. The testing temperature ranged from 60 C to
100 C and each specimen was held at the testing temperature for
30 min prior to the tension test. The tests were run in strain rate
control and a clip-on extensometer was used. The high strain rate
tensile testing up to 1 102 s1 was performed at room temperature by using a servohydraulic Instron VHS 8800 load frame,
in which the actuator was accelerated to a predetermined velocity
before gripping of the test bar. All the high strain rate tensile tests
were run at a constant imposed crosshead speed. As normally
experienced in this type of tests, the load signal was somewhat
disturbed due to the elastic waves transmitted through the load

Table 2 Basic properties at room temperature of the steel


studied
Grain size
m

R p0.2
MPa

Rm
MPa

Hardness HV10

Total elongation A10

10 15

230

364

119 1.6

0.41

Fig. 1 Microstructure of the studied steel. Longitudinal section perpendicular to the rolling plane. Rolling direction
vertical.

021019-2 / Vol. 133, APRIL 2011

cell and the so called ringing phenomenon was observed at the


highest strain rate of 1 102 s1. Adjacent averaging was then
used to smoothen the load signal. The strains in the high strain
rate tests were calculated from the displacement of the actuator.
The uniform elongation was obtained from the tensile stressstrain curve at the maximum load. The total elongation was measured between small indentations over a length of 40 mm that had
been inscribed on the specimen surface prior to testing. The area
reduction RA at fracture of the materials at different strain rates
and temperatures was evaluated using the equation
RA =

A0 A f
A0

where A0 is the original cross section area and A f is the area after
fracture. Digital image processing software ZEISS AXIOVISION was
used to measure the A f value by outlining the fracture area under
a Zeiss stereo microscope. An example is shown in Fig. 2.

Mechanical Properties and Comments

3.1 Engineering Stress-Strain Characteristics. The influence of strain rate and temperature on the flow stress as a function
of strain is shown in Fig. 3. There is a significant increase in the
flow stress when the strain rate is increased or when the temperature is decreased, indicating a rather high sensitivity to changes in
strain rate and temperature. An obvious upper yield point is only
observed at low temperature 60 C. It is also interesting to note
the overlapping of two curves with different deformation conditions AR at 20 C, 1 104 s1 and 100 C, 1 101 s1, implying that the effect of increasing temperature is equivalent to the
decrease of strain rate in thermally activated processes.
Compared with the AR material, PS and PSA treatments increase both yield R p0.2 and ultimate tensile strength Rm in the
temperature range from 60 C to 100 C. For example, at a strain
rate 1 104 s1 at room temperature, the rise in R p0.2 for a PS
specimen, which has been stored in air for several months, is
about 85 MPa. Additionally, developing yield phenomena can be

Fig. 2 An example of the fracture area measurement

Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 4 Yield strength, ultimate tensile strength and strain hardening ratio as a function of strain rate and temperature

Fig. 3 Engineering stress-strain curves at different temperatures and strain rates for the a AR, b PS, and c PSA material conditions

observed also at room temperature. Figure 4 gives the strength


levels for the different conditions at different strain rates and temperatures. It can be seen that straining and aging lead to larger
increase in R p0.2 than in Rm, especially at low strain rates. Compared with the PS treatment, further heat treatment PSA only
insignificantly influences R p0.2 and Rm.
The numbers in Fig. 4 represent the strain hardening ratio
Rm / R p0.2, which is often used to describe the strain hardening
ability. This ratio depends on both strain rate and temperature. The
present steel keeps considerable strain hardening ability even after
PS and PSA treatments. Generally, this ability decreases with increasing strain rate for the AR condition while this is less evident
for the PS and PSA conditions. The strain hardening ability is very
strain rate sensitive at low temperatures 60 C due to the difficulty to trigger the plastic deformation at high strain rates 1
101 s1.
In addition to strain hardening, the increased strength after PS
treatment is caused by the strain aging effect at ambient temperature. Although the studied material belongs to IF steel type, during
the production process, a certain amount of interstitial solute atoms still remains in the ferrite. It has been claimed that the segregation of 510 ppm solute is sufficient to promote dislocation
locking in low carbon steels 10,14. The bake-hardening phenomenon has been observed in a Ti-alloyed IF steel with only 4
Journal of Engineering Materials and Technology

ppm solute C 11.


However, no further hardening is observed after subsequent annealing at 180 C for 30 min. The possible reason may be as
follows. Owing to the limited amount of interstitial solute atoms
in the lattice, further segregation of the interstitial solute atoms to
the dislocations is not likely. Although additional fine carbides
may be formed during annealing by withdrawing C from the Cottrell locking atmosphere, the net result is an almost unchanged
tensile behavior.
3.2 Ductility. The uniform elongation of the studied steel at
maximum load is shown in the lower part of Fig. 5. Strain rate
hardening the additional hardening caused by increased strain
rate lowers the value of uniform elongation in all cases. The same
observation for IF steels at room temperature was also reported by
other researchers 15. It is noted, however, that the high strain
rate of 1 102 s1 decreases this deformation ability more severely. There are two reasons that may contribute to this phenomenon. The variation of strain hardening rate = d / d p with true
plastic strain p is shown in Fig. 6. The derivative d / d p is taken
from the true stress-strain curves, which is fitted by four-order
polynomials. The reduced strain hardening rate at strain rate of
1 102 s1 beyond a few percent of strain results in decreased
uniform deformation ability. Furthermore, at high strain rate, there
is not enough time for the heat produced during the plastic deformation to dissipate. Consequently, a portion of this heat remains
within the specimen, causing an increase in temperature and thus
softening. The effect of deformation induced heating is more pronounced at higher strain rates. This may also contribute to the
early start of the necking.
Prestraining reduces the amount of uniform deformation Fig.
APRIL 2011, Vol. 133 / 021019-3

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 5 Uniform and total elongation as a function of strain rate


and temperature

5 as expected due to the decreased strain hardening ability Fig.


4. Essentially similar values are obtained for the specimens with
PS and PSA treatments. Considering the effect of the temperature,
there is no clear difference in the uniform elongation in the temperature range from 60 C to 100 C at the strain rate of 1

104 s1. However, considerably decreased strain hardening


ability at 60 C under strain rate of 1 101 s1 Fig. 4 lowers
the uniform elongation in both virgin and pretreated conditions.
Instead of a monotonous decrease with strain rate, however, the
total elongation shows a minimum at an intermediate strain rate of
1 101 s1 at room temperature, as indicated in the upper part of
Fig. 5. It is known that total elongation is the sum of uniform
elongation and post-uniform elongation in the neck region. While
increased strain rate reduces the uniform elongation, as discussed
before, it has opposite effect on the post-uniform elongation due
to the enhanced strain rate sensitivity of flow stress, which will be
further discussed in Sec. 3.4. Uniform elongation contributes
more to the total elongation at lower strain rate while the postuniform deformation is more important when the strain rate is
beyond the critical value, which is around 1 101 s1 at room
temperature. The influence of the temperature on total elongation
is more complex. At the strain rate of 1 104 s1, the scatter in
the present results is too large to allow a definitive conclusion.
However, similar total elongation is obtained at different temperatures at the strain rate of 1 101 s1.
The total area reduction is also the sum of uniform area reduction and the nonuniform one in the neck region. However, the
latter is generally dominant. Figure 7 gives the total RA as a
function of the temperature and strain rate. It is around 90% in all
conditions and virtually insensitive to variations in strain rate and
temperature. The studied steel is basically a single phase material,
where the RA value depends only on the limit plasticity of ferrite,
which keeps approximately unchanged in the investigated temperature and strain rate range.
3.3 Thermal Softening Effect at High Strain Rate. Under
the strain rate of 1 102 s1, a semi-infinite plate condition for
unsteady state thermal conduction is met approximately as the
Fourier number F0 calculated by using Eq. 2 is 0.05, which is
smaller than the critical value 0.1 16. In this case, the characteristic thermal diffusion distance 2t is 0.66 mm. It is thus reasonable to assume that the deformation at this strain rate is mainly
an adiabatic process. The estimated temperature increase T up to
maximum uniform strain by using Eq. 3 is 17 C for AR material.
F0 = t/L2 =
T =

Fig. 6 Strain hardening rate for AR steel. Isothermal true


stress-strain curve cf. Fig. 8 was used for calculation at 1
102 s1.

c p

k
t

c p L2

where k is the thermal conductivity 73 W m1 K1 17, is the


density of the material 7870 kg m3, c p is the thermal capaci-

Fig. 7 RA at fracture as a function of a temperature and b strain rate

021019-4 / Vol. 133, APRIL 2011

Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 8 a Determination of the thermal softening function, Eq. 4 and b effect of thermal softening on stressstrain curve for a strain rate of 1 102 s1

tivity 452 J kg1 K1 17, is the thermal diffusivity


= k / c p = 2.06 105 m2 s1, t is the duration of the tensile
test, L is the thickness of the specimen 1.5 mm, and is the
fraction of plastic work converted into heat 0.9 was used in this
calculation.
A phenomenological thermal softening function 18, as given
in Eq. 4, is used to estimate the thermal softening effect at high
strain rate, by which the adiabatic stress-strain curve can be transformed to an isothermal one.
FT = 1

T Tr
Tm Tr

where Tr is a reference temperature 213 K in this case, Tm is the


melting temperature of the steel 1808 K, and n is a material
constant. n was determined by regression analysis to be 0.71 and
0.69 for the strain rates 1 104 s1 and 1 101 s1, respectively. Figure 8a shows the evaluation procedure for n while Fig.
8b demonstrates the deviation between isothermal and adiabatic
conditions for the largest strain rate 1 102 s1 considering a
temperature increase according to Eq. 3 and an estimated mean
n-value of 0.70.
Further understanding of the deformation situation in the neck
region is offered by Fig. 9, representing a room temperature test at
the highest strain rate 1 102 s1. Figure 9a gives the variation
of Vickers hardness as a function of the distance from the fracture

surface after the tensile test. It is clear that the material is roughly
30% harder at the fracture point in the neck area. Because the
hardness correlates well with the tensile strength, the hardening
after maximum uniform strain up to fracture can be approximated.
On the other hand, the true strain f at the fracture point can be
estimated by using equation f = ln A0 / A f , where A0 is the original cross section area and A f is the area after fracture. The very
small time of deformation in the neck region in this case of the
order of 0.005 s tells that the deformation is truly adiabatic with
almost no heat loss to neighboring parts of the tensile specimen.
The calculated temperature increase T at the fracture point by
using Eq. 3 is as high as 332 C for AR material, assuming linear
development of the stress after maximum uniform strain, as given
in Fig. 9b. Compared with the region with uniform deformation
Fig. 8b, the situation in the neck region is drastically different
with an adiabatic heating being 1020 times higher. As indicated
in Fig. 9a, there is a marked fluctuation of hardness in the neck
region, indicating softened shear bands. The fracture process itself
in the necked region therefore occurs at higher temperatures and
additionally at higher local strain rates than indicated by the nominal test data.
3.4 Strain Rate Sensitivity. The strain rate sensitivity for
AR material calculated by using = / ln is shown in Fig. 10.
It should be noticed that only average value in the strain rate range

Fig. 9 a Variation of Vickers microhardness 300 g as a function of the distance from the fracture surface. AR condition
at largest strain rate investigated, 1 102 s1 at 20 C. b Schematic development of the stress after necking.

Journal of Engineering Materials and Technology

APRIL 2011, Vol. 133 / 021019-5

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 10 Strain rate sensitivity for AR material. In the strain


rate range of 1 104 1 101 s1 at 60 C, the average value
is given as only two data points were available. For the largest
strain rate 1 102 s1, the transformed isothermal data were
used.

of 1 104 1 101 s1 at 60 C is given. Positive strain rate


sensitivity is observed under all the conditions studied with virtually identical -values for PS and PSA conditions as well. The
-values depend on both strain rate and temperature. As a general
trend, increases with decreasing temperature and increasing
strain rate. A larger strain rate sensitivity on the flow stress of IF
steels at high strain rates at room temperature was also observed
by other researchers 15. The actual strain level has some influence on , depending on the deformation condition. At low temperature 60 C, increases with decreasing strain due to the
obvious yield phenomenon observed at 1 101 s1. However,
the influence of strain is limited at higher temperatures.
In tensile tests, the necking after load instability may be characterized by a diffuse necking superimposed by localized neck
preceding the final fracture 19. The diffuse neck starts at the
instability point, answering to the uniform elongation. As the
strain rate sensitivity rises with the increased strain rate in the
necking zone with strain concentration, it counteracts the tendency to further localized straining and leads to lengthening of the
necking zone. At the same time, the RA corresponding to the final
fracture is fairly independent of temperature and strain rate for the
present material Fig. 7. A necessary consequence is an increase
in the total elongation to fracture at increasing strain rates, as
confirmed by the experimental data in Fig. 5. It is evident that
strain rates exceeding about 1 101 s1 with marked increase of
the parameter Fig. 10 lead to clear increase of the total elongation to fracture Fig. 5. An independent manifestation of this
phenomenon is the increased displacement between the points of
instability and fracture in the recorded load-displacement curves.
It should finally be remarked that increased -values have similar effects on both diffuse and localized necking. In the present
context, however, the effect on the diffuse necking gives large
contribution and results in wedge-shaped tensile bars after tests at
high strain rates cf. Fig. 9.

Fig. 11 Examples of the Voce fitting for AR material. The fitting


was performed for the nominally recorded stress-strain data,
except for the largest strain rate 1 102 s1, where the transformed isothermal data were used.

tion generation and dynamic recovery. The dislocation density and


arrangement are constant in this case. 0 is the extrapolated initial
stress and 0 is a characteristic strain relaxation strain, determined by the dynamic recovery rate, which is related to annihilation of dislocations 21 by the cross-slip of dislocations. Higher
recovery rate corresponds to lower 0 value.
The true stress and plastic strain data of the investigated material conditions were fitted to the Voce equation using the leastsquares method with s, 0, and 0 as free parameters. The curves
were cut after the maximum uniform strain was reached. Some
fitting examples are given in Fig. 11. The Voce equation fits well
in all cases and good correlation between the experimental data
and the model is obtained. The Voce equation can thus be used to
describe the whole tensile flow curves from yielding to uniform
strain at maximum engineering stress. The parameters obtained in
the curve fitting process, all depending on T, and material history, are presented in Fig. 12. High speed testing at 1 102 s1
gives the highest statistic fitting error, which is 3.75 MPa for s
and 0 and 0.26 for 0 PS.
As shown in Fig. 12a, 0 decreases monotonously with temperature but levels out at higher temperature. 0 increases with
increasing strain rates and more so for lower temperatures Fig.
12b. The variation of s Figs. 12c and 12d is generally in
accordance with the case for 0. It is also clear that changing
strain rate has less effect on both s and 0 at 100 C compared
with lower temperatures. The material history has significant influence and the PS and PSA pretreatments raises 0 more than s.
The parameter 0 exhibits a more complex variation with temperature and strain rate, as indicated in Figs. 12e and 12f. 0
generally decreases with increasing temperature. At the highest
strain rates, 0 drops to small values. However, competing effects
on strain hardening and recovery give some irregularities, which
will be discussed later.

5
4

Constitutive Model for the Material

The Voce equation 20 can be used to express tensile flow


stress and work hardening in the linear steady state. It can be
written in the form

= s s 0exp p/0

where p is the true plastic strain. s is the so called saturation


stress at which full plasticity is reached and the strain hardening
rate becomes zero, corresponding to equilibrium between disloca021019-6 / Vol. 133, APRIL 2011

Discussion

The present investigation concerns the influence of temperature,


strain rate, and material history mechanical and thermal pretreatments on the deformation behavior of IF low alloyed steel commonly used in production of car bodies. The influence of these
parameters is even more important for the mechanical behavior of
plates in cars suffering crash situations in service. Here, the car
body is often exposed to extreme loading conditions. The purpose
was therefore to investigate the mechanical behavior of this material in virtually all conditions that can appear in service. An
Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 12 Voce parameters for the different material conditions

important goal of this study was also to describe the mechanical


behavior in parametric form suitable for use in commercial software simulating crash situations.
The discussion will deal with the effect of temperature, strain
rate, and pretreatment on the quantities constituting the successful
Voce equation. Later, the influence of these parameters will be
discussed in physics based terms such as athermal and thermal
components ath and th of the flow stress. Finally, a thermal
activation analysis is performed to further illustrate the role of
temperature and strain rate on the stress-strain behavior of the
Journal of Engineering Materials and Technology

different material conditions. A simple approach is that the true


stress can be separated into the two stress components given by
the summation

true = ath + th

5.1 Effect of Temperature, Strain Rate, and Material History on Voce Parameters. The Voce equation Eq. 5 contains
three parameters that determine the flow stress as a function of
plastic strain from initial yielding and onward through the strain
APRIL 2011, Vol. 133 / 021019-7

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 13 Comparison of the Voce parameter 0 with yield stress


Rp0.2

hardening process. 0 represents the extrapolated initial stress associated with the beginning of plastic deformation. Figure 13
compares Voce parameter 0 with R p0.2 for the present steel in
different conditions and test states. It is concluded from the small
deviation of the yielding stress R p0.2 from the 0 = R p0.2 line that
the Voce equation can adequately describe the onset of plastic
deformation for this material. The parameters 0 and s, on the
other hand, describe the strain hardening.
The Voce equation is consistent with the Kock model that takes
the generation and dynamic annihilation of dislocations into account 21. The characteristic strain 0 is determined by the dynamic recovery, which, in nature, is a thermally activated process
involving movement diffusion of atoms/vacancies to the dislocations. The diffusion rate and characteristic diffusion distances,
being functions of strain rate, temperature, and material history,
are thus important factors. Increase in temperature makes diffusion more rapid, promoting the approach to steady state. The influence of strain rate is more complicated with the following two
opposing effects via its influence on dislocation velocity and diffusion distance of atom/vacancies.
a

At lower strain rate with small dislocation velocity, there


is time for the atoms/vacancies to diffuse, promoting recovery. Equilibrium between dislocation generation and
recovery is thus reached at lower dislocation density or
stress. The converse is true for deformation at higher
strain rate.
On the other hand, the strain rate also influences the diffusion distance, which is related to the dislocation density. It is known that the higher the density, the shorter the
average diffusion distance and thus annihilation events
take place more quickly. It has been found that strain rate
is a key factor for the final configuration of dislocations
22,23. Their density increases significantly at high
strain rates for IF steel, especially at low strains 22. In
addition, temperature increase due to high speed deformation promotes the process of recovery as well. In this
sense, high speed deformation is beneficial to the recovery. For extremely high strain rates 2.8 104 s1,
Lins et al. 24 reported the development of an ultrafine
equiaxed grain structure in adiabatic shear bands due to
the progressive subgrain misorientation prism recrystallization by the formation and mechanical rotation of subgrains during deformation. Such shear bands seem to also
appear in the neck region at the highest strain rates in the
present investigation Fig. 9a.

021019-8 / Vol. 133, APRIL 2011

As a consequence, a complex variation of 0 is obtained in this


study, as shown in Figs. 12e and 12f. There is a combined
influence of temperature, strain rate, and material history.
Let us consider the temperature effect on 0 first. It is seen from
Fig. 12e that at strain rates of 1 101 s1 or lower, increasing
temperature lowers 0 because of increased recovery. This circumstance is most pronounced in the low temperature range. The less
variation of 0 with temperature at 1 104 s1 is caused by the
relatively longer time available for diffusion due to the lower
dislocation velocity here. Some structure change, which is presumed to postpone recovery, might occur at this lower strain rate
when temperature is above a certain value. Copreaux et al. 25
studied the effect of precipitation on the development of dislocation substructure in Ti-alloyed interstitial free steels with varying
carbon contents in the range 0.003 0.017 wt %, comparable to
the case in the present investigation. The volume fraction and the
size of TiC precipitates, which could pin the dislocations in their
study, was of the order of 104 with sizes about 10 nm. It was
found that as the volume fraction of TiC precipitates increases
i.e., reduced distance between particles, the average distance between secondary dislocation cell walls cell size decreases 25.
In the case of our study, C and/or N dissolved in the ferrite
proved by the strain age hardening effect can combine with Ti or
Nb to form carbide at the lower strain rate at 100 C, leading to
the less strain rate sensitivity and higher 0 value due to their
pinning of the dislocations. Prestraining provides more dislocations as preferential nucleation sites, thus facilitating the formation of these carbides, as indicated by the increased gap in 0
between different strain rates in Fig. 12e. At the same time, the
crossing point below which lower strain rate is favorable for the
recovery shifts toward the lower temperature.
Based on the relation between recovery and strain rate discussed above, it is suggested that the variation of 0 with strain
rate consists of three stages IIII at room temperature Fig.
12f. The initial reduction of 0 with strain rate in stage I is
caused by the structure change suggested above, which is time
dependent and thus encouraged by lower strain rate. Predeformation facilitates this process, leading to a steeper variation than that
of the AR material. The reason for the subsequent slight increase
of 0 stage II is mainly the limited diffusion time for the atom/
vacancies with increasing strain rate in this region. The following
decrease of 0 at higher strain rates stage III is resulted from the
shortened diffusion distance and increased temperature and thus
promoted recovery there. Owing to the equivalent effect of increasing temperature with lowering strain rate in the thermally
activated processes, it is conjectured that the curve might shift
toward right higher strain rate side at higher temperatures. The
converse might be the case for lower temperatures while 0 is
more sensitive to the temperature and increased sharply with
strain rate. Further study is needed here to facilitate further interpretation.
5.2 Athermal and Thermal Stress Components. The athermal stress component depends on large obstacles creating longrange stress fields while the thermal component originates from
local obstacles acting only over short distances a few atomic
distances. The former one is weakly dependent on temperature
through the elastic shear modulus and has a strong dependence
on microstructure. The thermal component is a function of plastic
strain rate and temperature and can be overcome with the aid of
thermal activation. The proportionality between the shear stress s
and the main principal stress allows tensile stresses to be separated into athermal and thermal parts, as indicated by Eq. 6, a
procedure followed here.
Illustrated by the overlapping of two engineering stress-strain
curves for AR steel under changed deformation conditions 20 C,
1 104 s1 and 100 C, 1 101 s1 and other stress-strain data
in this study e.g. Fig. 3, it is evident that increasing temperature
is equivalent to the decrease of strain rate in the thermally actiTransactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 14 a Relationship between stress and strain rate compensated temperature Tsrc at different strain levels. For the
largest strain rate 1 102 s1, however, transformed isothermal stress data were used. b Variation of athermal stress
component ath with plastic strain. Material in AR and PS conditions, respectively.

vated processes. Inspired by the constitutive model of Johnson


and Cook 18 and Zerilli and Armstrong 26, the concept strain
rate compensated temperature Tsrc can be used to convert the
strain rate effect to a temperature effect by using Eq. 7 over a
certain strain range.

Tsrc = T 1 + c ln

where T is the test temperature, Tsrc is the strain rate compensated


equivalent temperature, is the actual strain rate, r is the
chosen reference strain rate, and c is an adjustable constant. A
statistical fitting with a reference strain rate of 1 104 s1 gave a
c-value of 0.036 for the present material up to maximum uniform
strain. A similar approach was done in Ref. 27 for ferritic/
martensitic steel.
To identify the thermal stress component, the recorded stresses
are to be normalized by the shear modulus , which is a function
of the temperature 28. The variation of within the actual temperature range from 60 C to +100 C, is, however, limited to
approximately 5%. Calculating the normalized Voce parameter
0 / as a function of the strain rate compensated temperature Tsrc
gives an exponential decay, as shown in the lower part of Fig.
14a. At high Tsrc, 0 tends asymptotically to the values corresponding to the athermal mechanisms. Curve fitting by using exponential decay gives the athermal part 0 / as 0.0028 0.0001.
Figure 14a also presents the variation of true stress with Tsrc
at other strain levels. Exponential decay is followed as well and
the obtained athermal component ath / is shown in Fig. 14b. It
is seen that this component increases with strain demonstrating
how it contributes to the strain hardening. However, ath / gradually saturates at higher strains. Approximately parallel true /
Tsrc curves in Fig. 14a indicate that the contribution of athermal component is dominant for the strain hardening. Figure 14b
also indicates that increased initial dislocation density PS condition gives an upward shift of the athermal stress but with virtually kept strain hardening.
It should be remarked that deviations in grain size would only
influence the athermal part of the flow stress and thus mainly the
parameter 0 in the Voce representation. All thermal parts, influenced by temperature and strain rate variations, should then be
essentially unaffected by changes in grain size. This has important
implication in engineering practice: For instance, IF steel used in
car bodies are virtually always welded with associated grain
coarsening in the heat affected zones. Such grain size variations
would thus not affect the strain rate or temperature dependence,
locally or globally.
Journal of Engineering Materials and Technology

5.3 Thermal Activation Analysis. The thermal activation


volume V is defined as the dislocation-affected region in plastic
deformation, as shown in Eq. 8a, where G is the Gibbs free
energy required by a dislocation to overcome a localized obstacle
and is the thermal stress. Simply replacing the shear stress by
the tensile flow stress provides a macroscopic phenomenological thermal activation volume V, which is supposed to inherit
some of the physical implication of its microscopic counterpart
V . V can be calculated by using Eq. 8b. With approximation
1 / 2, we can obtain V 2V.
V =

G
T=const

V = G/ = kT

ln

8a

= kT/

8b

T=const

where k is the Boltzmann constant 1.38 1023 J K1 and is


the strain rate sensitivity.
The flow stress at 5% tensile plastic strain is used in the present
thermal activation analysis. It has been suggested that the activation volume depends on temperature but is not affected by strain
rate 26. However, the results obtained in this study show that V
is dependent on both the strain rate and the temperature.
Figure 15 shows our results for thermal activation volume
V / b3 as a function of strain rate compensated temperature, using
Eq. 8 and -values from Fig. 10 with b = 3 / 2a = 0.2483 nm.
V / b3 ranges from 10 to 160 under the studied conditions and it
increases exponentially with Tsrc. The activation volume for IF
steel obtained by other researchers 12,29 is in the same range. It
can be concluded that a thermal rate controlling PeierlsNabarro
mechanism with an activation volume of 1020 b3 controls the
deformation of the present material at low Tsrc. V / b3 increases
strongly at higher Tsrc. In the latter region, thermal activation is
less influential and other rate controlling mechanisms, competing
with the Peierls one, are likely to become operative.
It is evident from Fig. 15 that the activation volume is the same
for both AR and PS conditions, meaning, that the increased dislocation density caused by prestraining does not affect the activation volume. This is plausible since the mean distance between
dislocations in both conditions is much larger than the size of
the thermal activation volume V. The consistency of the activation volume at high strain rates and at the low temperatures, as
indicated in Fig. 15, supports the assumption that the rate controlling mechanism at high strain rate is the same as that at low
APRIL 2011, Vol. 133 / 021019-9

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Fig. 15 Variation of thermal activation volume with Tsrc. Only


average values are given for 60 C AR and PS and 100 C
PS in the strain rate range of 1 104 1 101 s1.

7
temperature. It can be concluded that the PeierlsNabarro mechanism controls the deformation rate of the IF steel also at high
strain rates.
For a given temperature the strain rate sensitivity is inversely
proportional to the activation volume V, as indicated by Eq. 8b.
It can thus be deduced that the strain rate sensitivity decreases
exponentially with increasing Tsrc. This does not conflict with the
results in Fig. 10 since strain rate compensated temperature is
considered here.

Conclusions

A rephosphorized, interstitial free steel with 0.005 C, 0.08 Si,


0.030 Al, 0.042 Ti, and 0.028 Nb wt % was studied, regarding
tensile properties that cover a large range of temperatures from
60 C to +100 C and strain rates 1 104 1 102 s1. The
material was characterized in its virgin state AR and after prior
deformation, without PS, and with aging PSA. The Voce equation was successfully fitted to the experimental data. The parameters in this equation were thoroughly evaluated for the different
states, and the thermal and athermal components of flow stress
were analyzed likewise. Based on these results, the following conclusions can be drawn.
1 An increase in yield R p0.2 and ultimate tensile strength
Rm is generally observed for increased strain rates and
decreased temperatures.
2 Prestraining and superimposed annealing lead to increased
strength, decreased strain hardening, and smaller uniform
elongation at all temperatures and strain rates. The rise of
R p0.2 is larger than that of Rm. However, compared with the
PS treatment, further heat treatment PSA only has ignorable influence on the mechanical behavior. The AR material
should be even less sensitive to annealing. During the process of production, the components of a car body may experience deformation ranging from 0 to a certain strain
level in different locations. The negligible sensitivity to annealing makes the results obtained in this study applicable
to the service condition.
3 Increasing strain rate lowers the uniform elongation monotonously, especially at high strain rates 1 102 s1 and
low temperatures 60 C. The total elongation shows a
minimum value at intermediate strain rates. The reduction
in area is around 90% for all conditions and is relatively
independent of temperature and strain rate.
4 The adiabatic deformation condition present at the highest
strain rates could be transformed to an equivalent isother021019-10 / Vol. 133, APRIL 2011

mal one by considering the thermal softening effect. Although this temperature increase never exceeds about 17 C
in the stage of uniform deformation, the corresponding flow
stress in the necking zone needs to be corrected to represent
isothermal conditions.
The strain rate sensitivity = / ln is positive for all
conditions studied. As a general trend, the higher the strain
rate, the higher the strain rate sensitivity. increases with
decreasing temperature and has its largest value at 60 C
and small strains for the present material.
The Voce equation = s s 0exp p / 0 can be
used to describe the whole tensile flow curves from yielding to uniform strain at maximum engineering stress. The
parameters obtained in the curve fitting process depend on
temperature, strain rate, and material history. 0, which describes the beginning of plastic deformation, always increases monotonously with decreasing temperature or increasing strain rate. s and 0 representing the strain
hardening depend on both temperature and strain rate, as
reflected by a balance between dislocation generation and
dynamic recovery.
The athermal stress component, which increases with
strain, essentially determines the strain hardening while the
thermal part of the flow stress is influenced by temperature
and strain rate factors. Analysis of recorded data gave a
quantitative relation between the roles of these two factors,
indicating equivalence between high strain rates and low
temperatures.
The thermal activation volume V depends on both the
strain rate and the temperature. It is likely that at high strain
rates or low temperatures, PeierlsNabarro mechanism controls the deformation.

Acknowledgment
The authors would like to thank Mr. Magnus Johansson, a
former research associate at our department, for providing important parts of the experimental data in this project. Financial support was given by Swedish Agency for Innovation Systems VINNOVA.

References
1 Samet-Meziou, A., Etter, A. L., Baudin, T., and Penelle, R., 2008, Relation
Between the Deformation Sub-Structure After Rolling or Tension and the Recrystallization Mechanisms of an IF Steel, Mater. Sci. Eng., A, 473, pp.
342354.
2 Shen, Y. F., Xue, W. Y., Wang, Y. D., Liua, Y. D., and Zuo, L., 2008, Tensile
Behaviors of IF Steel With Different Cold-Rolling Reductions, Mater. Sci.
Eng., A, 496, pp. 383388.
3 Saimoto, S., and Diak, B. J., 2001, Strain Rate Sensitivity of Ultra-Low
Carbon Steels, Mater. Sci. Eng., A, 319321, pp. 294298.
4 Saha, R., and Ray, R. K., 2007, Formation of Nano- to Ultrafine Grains in a
Severely Cold Rolled Interstitial Free Steel, Mater. Sci. Eng., A, 459, pp.
223226.
5 Majta, J., and Muszka, K., 2007, Mechanical Properties of Ultra Fine-Grained
HSLA and Ti-IF Steels, Mater. Sci. Eng., A, 464, pp. 186191.
6 Li, B. L., Cao, W. Q., Liu, Q., and Liu, W., 2003, Flow Stress and Microstructure of the Cold-Rolled IF-Steel, Mater. Sci. Eng., A, 356, pp. 3742.
7 Tomota, Y., Lukas, P., Harjo, S., Park, J.-H., Tsuchida, N., and Neov, D., 2003,
In Situ Neutron Diffraction Study of IF and Ultra Low Carbon Steels Upon
Tensile Deformation, Acta Mater., 51, pp. 819830.
8 Sachdev, A. K., 1982, Dynamic Strain Aging of Various Steels, Metall.
Trans. A, 13A, pp. 17931797.
9 Tsuchida, N., Baba, E., Nagai, K., and Tomota, Y., 2005, Effects of Interstitial
Solute Atoms on the Very Low Strain-Rate Deformations for an IF Steel and
an Ultra-Low Carbon Steel, Acta Mater., 53, pp. 265270.
10 De, A. K., Vandeputte, S., and Cooman, B. C. D., 1999, Static Strain Aging
Behavior of Ultra Low Carbon Bake Hardening Steel, Scr. Mater., 418, pp.
831837.
11 Dehghani, K. D., and Jonas, J. J., 2000, Dynamic Bake Hardening of
Interstitial-Free Steels, Metall. Mater. Trans. A, 31A, pp. 13751384.
12 Uenishi, A., and Teodosiu, C., 2003, Solid Solution Softening at High Strain
Rates in Si- and/or Mn-Added Interstitial Free Steels, Acta Mater., 51, pp.
44374446.
13 Cahn, R. W., Haasen, P., and Kramer, E. J., 1992, Materials Science and
Technology: A Comprehensive Treatment, Constitution and Properties of

Transactions of the ASME

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Steels, F. B. Pickering, ed., VCH, New York, p. 824.


14 Wilson, D. V., and Russell, B., 1960, The Contribution of Atmosphere Locking to the Strain-Aging of Low Carbon Steels, Acta Metall., 8, pp. 3645.
15 Rana, R., Singh, S. B., Bleck, W., and Mohanty, O. N., 2009, Effect of
Temperature and Dynamic Loading on the Mechanical Properties of CopperAlloyed High-Strength Interstitial-Free Steel, Metall. Mater. Trans. A, 40, pp.
856866.
16 Serth, R., 2007, Process Heat Transfer: Principles and Applications, Elsevier
Academic, Amsterdam, p. 755.
17 Batra, R. C., and Kim, C. H., 1992, Analysis of Shear Banding in Twelve
Materials, Int. J. Plast., 8, pp. 425452.
18 Johnson, G. R., and Cook, W. H., 1983, A Constitutive Model and Data for
Metals Subjected to Large Strains, High Strain Rates and High Temperatures,
Proceedings of the Seventh International Symposium on Ballistics, Belgium,
pp. 541547.
19 Zhang, Z. L., Hauge, M., degrd, J., and Thaulow, C., 1999, Determining
Material True StressStrain Curve From Tensile Specimens With Rectangular
Cross-Section, Int. J. Solids Struct., 36, pp. 34973516.
20 Kocks, U. F., 1976, Laws for Work-Hardening and Low Temperature Creep,
ASME J. Eng. Mater. Technol., 98, pp. 7685.
21 Estrin, Y., and Mecking, H., 1984, A Unified Phenomenological Description
of Work Hardening and Creep Based on One-Parameter Models, Acta Metall., 32, pp. 5770.

Journal of Engineering Materials and Technology

22 Uenishi, A., Teodosiu, C., and Nesterova, E. V., 2005, Microstructural Evolution at High Strain Rates in Solution-Hardened Interstitial Free Steels,
Mater. Sci. Eng., A, 400401, pp. 499503.
23 Lee, W.-S., and Lam, H.-F., 1996, The Deformation Behaviour and Microstructure Evolution of High-Strength Alloy Steel at High Rate of Strain, J.
Mater. Process. Technol., 57, pp. 233240.
24 Lins, J. F. C., Sandim, H. R. Z., Kestenbach, H.-J., Raabe, D., and Vecchio, K.
S., 2007, A Microstructural Investigation of Adiabatic Shear Bands in an
Interstitial Free Steel, Mater. Sci. Eng., A, 457, pp. 205218.
25 Copreaux, J., Lanteri, S., and Schmitt, J.-H., 1993, Effect of Precipitation on
the Development of Dislocation Substructure in Low Carbon Steels During
Cold Deformation, Mater. Sci. Eng., A, 164, pp. 201205.
26 Zerilli, F. J., and Armstrong, R. W., 1987, Dislocation-Mechanics-Based Constitutive Relations for Material Dynamics Calculations, J. Appl. Phys., 615,
pp. 18161825.
27 Sptig, P., Baluc, N., and Victoria, M., 2001, On the Constitutive Behavior of
the F82H Ferritic/Martensitic Steel, A309-310, pp. 425429.
28 Sptig, P., Odette, G. R., Donahue, E., and Lucas, G. E., 2000, Constitutive
Behavior and Fracture Toughness Parameters of the F82H Ferritic/Martensitic
Steel, J. Nucl. Mater., 283287, pp. 721726.
29 Uenishi, A., and Teodosiu, C., 2004, Constitutive Modelling of the High
Strain Rate Behaviour of Interstitial-Free Steel, Int. J. Plast., 20, pp. 915
936.

APRIL 2011, Vol. 133 / 021019-11

Downloaded From: http://materialstechnology.asmedigitalcollection.asme.org/ on 12/17/2015 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Anda mungkin juga menyukai