Вы находитесь на странице: 1из 25

Thermomechanical response of DH-36 structural steel over

a wide range of strain rates and temperatures


Sia Nemat-Nasser
*
, Wei-Guo Guo
Center of Excellence for Advanced Materials, Department of Mechanical and Aerospace Engineering, University of California,
San Diego, 9500 Gilman Drive, La Jolla, CA 92093-0416, USA
Received 1 August 2002; received in revised form 22 August 2002
Abstract
To understand and model the thermomechanical response of DH-36 Naval structural steel, uniaxial compression
tests are performed on cylindrical samples, using an Instron servohydraulic testing machine and UCSDs enhanced
Hopkinson technique. True strains exceeding 60% are achieved in these tests, over the range of strain rates from 0.001/s
to about 8000/s, and at initial temperatures from 77 to 1000 K. The microstructure of the undeformed and deformed
samples is examined through optical microscopy. The experimental results show: (1) DH-36 steel displays good ductility
and plasticity (strain >60%) at low temperatures (even at 77 K) and high strain rates; (2) at relatively high temperatures
and low strain rates (especially below about 0.1/s), its strength is not temperature sensitive, indicating that the material
has good weldability; (3) dynamic strain aging (DSA) occurs at temperatures between 500 and 1000 K and in the range
of strain rates from 0.001/s to 3000/s, the peak value of the stress shifting to higher temperatures with increasing strain
rates (it is about 600 K at 0.001/s, about 650 K at 0.1/s, and about 800 K at 3000/s); (4) adiabatic shearbands develop
when the strain exceeds about 30% at 77 K, and at higher strains for higher temperatures; and (5) the microstructural
evolution of the material is not very sensitive to changes in strain rates and temperatures. Finally, based on the
mechanism of dislocation motion, and using our experimental data, the parameters of a physically-based model de-
veloped earlier for AL-6XN stainless steel [J. Mech. Phys. Solids 49 (2001) 1823] are estimated and the model pre-
dictions are compared with various experimental results, excluding the dynamic strain aging eects. Good agreement
between the theoretical predictions and experimental results is obtained. In order to further verify the model inde-
pendently of the experiments used in the modeling, additional compression tests at a strain rate of 8000/s and various
initial temperatures are performed, and the results are compared with the model predictions. Good correlation is
observed. As an alternative to this model, the experimental data are also used to estimate the parameters in the
JohnsonCook model [G.R. Johnson, W.H. Cook, 1983, in: Proceedings of the Seventh International Symposium on
Ballistic, The Hague, The Netherlands, p. 541] and the resulting model predictions are compared with the experimental
data, again excluding the dynamic strain-aging eects. These and related results suggest that the physically based model
has a better prediction capability over a broader strain rate and temperature range.
2003 Elsevier Ltd. All rights reserved.
Keywords: Structural steel; Strain rate; Aging; Modeling
*
Corresponding author. Fax: +1-858-534-2727.
E-mail address: sia@ucsd.edu (S. Nemat-Nasser).
0167-6636/$ - see front matter 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/S0167-6636(02)00323-X
Mechanics of Materials 35 (2003) 10231047
www.elsevier.com/locate/mechmat
1. Introduction
DH-36 is a high strength structural steel used in
naval and other structural applications. As ship
hull steel, especially in high-speed sealift vessels
(HSS), it may be subjected to high-rate loading
due to collision, impact, or explosion, which can
be exacerbated by cold temperatures. Therefore, it
is required that, this steel has high toughness and
high strength under variable conditions including
various temperatures and loading rates. Addi-
tionally, the material must have good weldability.
In this paper, we examine the response of DH-36
steel over a wide range of strain rates and tem-
peratures, and using the experimental results, we
develop constitutive relations for this material.
Iron, the major component of steel, exists below
its melting point in two crystal forms, body-cen-
tered cubic (bcc), which is stable from below room
temperature to 912 C (known as a-ferrite), and
from 1394 C to the melting point of 1530 C
(known as d-ferrite); and the face-centered cubic
(fcc), which is stable between 912 and 1394 C
(know as austenite or c-iron) (Krauss, 1990). In
most common usage, the operating temperature is
below 912 C, with the steel having a bcc structure.
As a traditional metal, steel has been the subject
of extensive studies in the past few decades, both
experimentally and theoretically. However, to our
knowledge, DH-36 steel has not been studied sys-
tematically over a broad range of strain rates and
temperatures. The present paper reports the results
of systematic experiments over strain rates ranging
from 0.001/s to about 8000/s, and initial tempera-
tures ranging from 77 to 1000 K. Using the results
of these experiments, a physically-based model is
developed and its predictions are compared with
the experimental results. To further assess and
compare the model results with other models, we
have examined the JohnsonCook model, and
have compared its predictions with our experi-
mental results.
2. Experimental procedure and results
2.1. Material and samples
All tests are carried out on a DS-36 baseplate.
The chemical composition of this structural steel
baseplate is given in Table 1.
All samples have a 5 mm nominal diameter and
5 mm height. To reduce the end friction on the
samples during the low- and high-strain deforma-
tion, the sample ends are rst polished using
waterproof silicon carbide paper, 1200 and 4000
grit, and then they are greased for low- and room-
temperature tests. A molybdenum-powder lubri-
cant is used for the high-temperature experiments.
An argon atmosphere is used in the heating fur-
nace in order to prevent oxidization. To examine
the microstructure of the undeformed and de-
formed samples, the samples are sectioned along
the loading and transverse directions, and then
polished and etched, as required by standard me-
tallography. The etching reagent is the Nital: 98 ml
alcohol and 2 ml HNO
3
. Figs. 1 and 2 show the
microstructure of an undeformed sample. Fig. 1
displays the microstructure perpendicular to the
rolling direction of the DH-36 baseplate, and Fig.
2 that along the rolling direction. In these gures,
white regions are ferrites and black regions are
pearlites. It is known that ferrites have lower
strength and hardness, but higher plasticity and
toughness, whereas pearlites have a reversed
property. Figs. 1 and 2 show that the microstruc-
ture of this DH-36 plate is not the same in the two
considered directions. Pearlites in Fig. 1 form
black strips, while pearlites in Fig. 2 have a non-
uniform distribution. Hence, the microstructure of
the plate is not uniform. Since material properties
usually depend on the microstructure, we prepared
10 samples along various directions of the base-
plate, and tested them under dierent loading
conditions. Fig. 3 shows some of the results for
indicated strain rates. The samples for this gure
Table 1
Major alloy content of DS-36 (%)
C Mn Cu Si Cr Mo V Ti Al Nb P S
0.14 1.37 0.14 0.22 0.08 0.03 0.001 0.003 0.017 0.03 0.007 0.001
1024 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
are identied in the following manner according to
their orientation and location: RDrolling direc-
tion of the plate; PDperpendicular to the rolling
direction; TOPtop part of the plate thickness;
MIDmiddle part of the plate thickness; and
BOTbottle part of the plate thickness. As is seen
there are no remarkable dierences among each
set. These experimental results show that the me-
chanical properties of this DH-36 plate do not
depend on either the orientation or the through-
the-thickness location of the samples. This killed
structural steel has an average grain size of about
9 lm.
2.2. Low and high strain-rate experiments
Compression tests at strain rates of 0.001/s and
0.1/s are performed using an Instron hydraulic
testing machine, over the temperature range from
77 to 800 K, with true strains exceeding about
60%. Elevated temperatures are attained with a
high-intensity quartz lamp, in a radiant-heating
furnace of an argon environment. The temperature
is measured using a thermocouple arrangement,
and is maintained constant to within 2 C. The
deformation of the specimen is measured by
LVDT, mounted in the testing machine, and is
calibrated and compared with the results of a
standard extensometer before the test. The low
temperature of 77 K is obtained by immersing the
specimen and the testing xture (AL
2
O
3
ceramic
bars) in a bath of liquid nitrogen. Typical true
stress-true strain curves of DH-36 at strain rates of
0.001/s and 0.1/s are displayed in Figs. 4 and 5,
respectively.
Dynamic tests at a strain rate of 3000/s are
performed using UCSDs recovery Hopkinson
technique (Nemat-Nasser et al., 1991; Nemat-
Nasser and Isaacs, 1997) at temperatures of 77
1000 K, and strains exceeding 60%. For the high
strain-rate tests at elevated temperatures, it is
necessary to heat the sample to the required
temperature while keeping the incident and
transmission bars of the Hopkinson device at a
suitably low temperature. To do this, Nemat-
Nasser and Isaacs (1997) have developed a novel
enhancement of the compression recovery Hop-
kinson technique (Nemat-Nasser et al., 1991) for
high-temperature tests, where a furnace is em-
ployed to preheat the specimen, while keeping the
transmission and incident bars outside the fur-
nace. These bars are then automatically brought
into gentle contact with the specimen, just before
the stress pulse reaches the specimen-end of the
incident bar. The temperature is measured by a
thermocouple that also holds the specimen inside
the furnace.
The true stresstrue strain curves at a strain rate
of 3000/s are shown in Fig. 6. UCSDs recovery
Hopkinson technique makes it possible to obtain
an isothermal ow stress at high strain rates and
various temperatures. The isothermal ow stress
Fig. 1. DH-36 plate; microstructure transverse to the rolling
direction of the plate.
Fig. 2. DH-36 plate; microstructure in the rolling direction of
the plate.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1025
of DH-36 at a strain rate of 3000/s and tempera-
tures of 77500 K, is given in Fig. 7, which also
includes the corresponding adiabatic curves (heavy
solid curves).
2.3. Experimental results and discussion
DH-36 is a high-strength low-alloy structural
steel, with a carbon content of about 0.14%. Based
0
200
400
600
800
1000
1200
0.00 0.10 0.20 0.30 0.40 0.50 0.60
True Strain
T
r
u
e
S
t
r
e
s
s
(
M
P
a
)
DH-36, 0.001/s, 296K
DH-36, 3,000/s, T
0
= 296K
Sample name: RD, PD, TOP, MID, BOT
Fig. 3. Stresssstrain curves of samples cut at indicated directions and locations from a DH-36 baseplate.
0
400
800
1200
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e
S
t
r
e
s
s
(
M
P
a
)
77K
296K
400K
500K
600K

800K
DH-36, 0.001/s
Fig. 4. True stresstrue strain curves at indicated initial temperatures and a strain rate of 0.001 1/s.
1026 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
on the iron-carbon phase diagram (Krauss, 1990),
DH-36 steel has the bcc structure characteristics.
The performance of steels mainly depends on the
properties associated with their microstructure. It is
well known that bcc metals have high temperature
and strain-rate sensitivity, and their mechanical
properties are strongly aected by impurities, gen-
erally attributed to the rate-controlling mechanism
0
200
400
600
800
1000
1200
1400
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e
S
t
r
e
s
s
(
M
P
a
)
77K
296K
400K
500K
600K
700K
800K
DH-36, 0.1/ s
Fig. 5. True stresstrue strain curves at indicated initial temperatures and a strain rate of 0.1 1/s.
0
200
400
600
800
1000
1200
1400
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e
S
t
r
e
s
s
(
M
P
a
)
296K
400K
500K
600K
700K
800K
1000K
900K
DH-36, 3,000/s
T
0
= 77K
Fig. 6. Adiabatic stressstrain curves at indicated initial temperatures and a strain rate of 3000/s.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1027
of the thermal component of the ow stress. Figs.
46 show the strong dependence of the ow stress
on temperature. Figs. 810 present the ow stress as
a function of temperature for indicated strain rates
and strains. When the temperature increases from
77 K (liquid nitrogen) to about 800 K, the ow
stress rst decreases quickly, but at about 400 K
this decrease slows down and is followed by an
Fig. 7. Comparison between adiabatic and isothermal ow stress at 3000/s.
0
400
800
1200
1600
0 100 200 300 400 500 600 700 800 900 1000
Temperature, (K)
F
l
o
w
S
t
r
e
s
s
(
M
P
a
)
Strain = 0.05
Strain = 0.1
Strain = 0.2
Strain = 0.3
Strain = 0.4
Strain = 0.5
Strain = 0.55
DH-36, 0.001/ s
Fig. 8. Flow stress as a function of temperature for indicated strains and 0.001/s strain rate.
1028 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
increase with increasing temperature. Beyond 400
K, the ow stress is not sensitive to temperature
changes (except for dynamic strain aging), showing
that this material has good weldability. As is seen in
0
400
800
1200
1600
0 100 200 300 400 500 600 700 800 900 1000
Temperature , (K)
F
l
o
w
s
t
r
e
s
s
(
M
P
a
)
Strain = 0.05
Strain = 0.1
Strain = 0.2
Strain = 0.3
Strain = 0.4
Strain = 0.5
Strain = 0.55
DH-36, 0.1/ s
Fig. 9. Flow stress as a function of temperature for indicated strains and 0.1/s strain rate.
300
500
700
900
1100
1300
0 200 400 600 800 1000 1200
Temperature, (K)
F
l
o
w
S
t
r
e
s
s
(
M
P
a
)
Strain = 0.05
Strain = 0.1
Strain = 0.2
Strain = 0.3
Strain = 0.4
Strain = 0.5
DH-36, 3,000/s
Fig. 10. Flow stress as a function of temperature for indicated strains and 3000/s strain rate.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1029
Figs. 810, beyond 400 K, a peak for the ow
stress appears within the temperature range from
600 to 900 K, depending on the strain rate. When
the strain rate is 0.001/s, the peak stress occurs at
about 600 K; for 0.1/s, it occurs at about 650 K;
and for 3000/s, it is at about 900 K. It can be ex-
trapolated that a further increase in the strain rate
may move this peak stress to even higher tempera-
tures.
2.3.1. Dynamic strain aging
The phenomenon of increasing ow (yield)
stress with increasing temperature is due to dy-
namic strain aging (DSA). The term static strain
aging (SSA) generally refers to the transient stress
peaks observed in dilute alloys when a prestrained
specimen is totally or partially unloaded and aged
for a prescribed time and then reloaded at the
same prestraining strain rate. It is commonly ac-
cepted that this eect is related to the pinning of
dislocations by diusing solute atoms during the
aging period. In general, dynamic strain aging is
dened as recurrent pinning (serrated ow) of dis-
locations while arrested at obstacles during their
motion that results in plastic straining (Kubin et
al., 1992). Dynamic strain aging is generally at-
tributed to the additional resistance to dislocation
motion produced by the mobility of solute atoms
that can diuse to dislocations above a certain
temperature (Beukel and Kocks, 1982) while the
dislocations are waiting at their short-range
barriers. During this waiting period, a Cottrell
atmosphere or a core atmosphere can be formed at
dislocations depending on the temperature and
strain rate (Nakada and Keh, 1970). In steel, the
occurrence of static and dynamic strain aging re-
sults from the diusion of C and N in the tem-
perature range of 150300 C. Cho et al. (2000)
report that the substitutional elements Cr and Ni
in 304 stainless steel produce static strain aging at
temperatures of 9001100 C. To explain the strain
dependence of the peak value of the stress, two
models have been examined, that due to Cottrell
and that due to Kocks et al. (1975). The Cottrell
model assumes that diusion of solute atoms is
assisted by excess vacancies generated during
plastic deformation. In interpreting their results,
Kocks et al. abandon the vacancy hypothesis
and propose a new model according to which the
solute transport and arrest at dislocations is con-
trolled by pipe diusion via forest dislocations that
intersect the slip plane. In this model, the solute
hardening process is localized at obstacles associ-
ated with forest junctions, and the resulting
hardening eect is inuenced by the evolution of
the forest dislocation density, i.e., it is related to
the strain hardening. A key feature of thermally
activated dislocation motion is that the disloca-
tions spend most of their time to interact with their
local obstacles, such as forests of dislocations,
vacancies, and solute atoms (Cheng et al., 2001).
The dislocation moves in a jerky way. The segre-
gation of solute atoms to the dislocation core oc-
curs during the time when the dislocations are
waiting at the local obstacles. Because a strong
interaction force exists between dislocations and
solute atoms in the dislocation core area, the dif-
fusion of the core atmosphere becomes signicant
at lower temperatures than is possible for the sol-
ute atoms situated outside the core area, provided
that the shear stress is high.
With increasing strain rates, higher tempera-
tures are required to drive the solutes to disloca-
tions at sucient speeds. As the solutes catch up
with moving dislocations and pinned them down,
the ow stress increases. In Fig. 6, when the true
strain is less than about 6%, the ow stress at a
temperature of 800 K is lower than that at 600 K.
But when the true strain is greater than 6%, the
ow stress at 800 K is higher than that at 600 K.
This shows that a certain amount of plastic
straining must occur before dynamic strain aging
is observed.
In order to investigate the microstructural evo-
lution with temperature, two interrupted tests are
performed; see Figs. 11 and 12. In Fig. 11, samples
1 and 2 are loaded to a true strain of 60% at 77 and
296 K, respectively. Then sample 3 is loaded to a
true strain of about 28% at a temperature of 296 K
and is unloaded to zero. It is then cooled to 77 K,
and reloaded at the same strain rate. As is seen, the
composite ow stress curve of sample 3 then fol-
lows basically rst the ow stress curve of sample
2, and then that of sample 1. In Fig. 12, sample 4 is
loaded to a true strain of about 18% at a strain
rate of 0.001/s and at 296 K (dashed curve which
1030 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
actually is an isothermal curve). Then it is reloaded
at a strain rate of 3000/s and initial temperature of
500 K, resulting in an adiabatic ow stress (again
shown by a dashed curve). Sample 5 is loaded to a
true strain of about 16% at a strain rate of 3000/s
and initial temperature of 500 K, then it is un-
loaded, brought back to 500 K, and reloaded at the
same strain rate (again the solid curve). The dashed
0
400
800
1200
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e
S
t
r
e
s
s
(
M
P
a
)
DH-36, 0.001/s
296K
77K
Sampl e 1
Sample 3
Sample 2
Fig. 11. Eect of temperature jump from 296 to 77 K on ow stress.
0
200
400
600
800
1000
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strai n
T
r
u
e
S
t
r
e
s
s
(
M
P
a
)
3,000/s, 500K
0.001/s, 296K
DH-36
Sample 4
Sample 5
Fig. 12. Eect of temperature jump from 296 to 500 K on ow stress at indicated strain rates.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1031
and solid curves show that the prior dierent de-
formation histories of these two samples do not
substantially aect their subsequent response. Re-
sults in Figs. 11 and 12 show that, the micro-
structure of DH-36 does not evolve with its
preloading in a manner which would aect its
further mechanical properties.
2.3.2. Strain-rate eect on ow stress
It is known that the ow stress of most mate-
rials increases with increasing strain rates. In Fig.
13, the stresstemperature curves at various indi-
cated strain rates are compared for a xed strain of
10%. As is seen, the ow stress strongly depends
on the strain rate, especially when the temperature
exceeds 400 K and the dynamic strain-aging peak
shifts to higher temperatures with an increasing
strain rate. In Figs. 14 and 15, the strain-rate eect
on the ow stress is further examined by changing
the strain rate from 0.001/s to 0.1/s and 3000/s,
respectively. In these two gures, the stressstrain
curves at 0.001/s, 0.1/s, and 3000/s are rst ob-
tained at the common initial temperature of 296 K,
and then two samples (designated as samples 6 and
7, respectively) are loaded at same strain rate of
0.001/s to a true strain of about 20% at 296 K.
Then sample 6 is reloaded at a strain rate of 0.1/s
and 296 K. This interrupted reloading stressstrain
curve follows closely the stressstrain curve that is
obtained by continuous loading at 0.1/s. Sample 8
is loaded at a strain rate of 3000/s and initial
temperature of 296 K to about 16% strain, un-
loaded, brought back to 296 K, and then reloaded
at the same strain rate. Sample 7 is reloaded at a
strain rate of 3000/s and initial temperature 296 K.
It is seen that there are no essential dierences
between the responses of samples 7 and 8 in re-
loading. These interrupted test results verify that,
pre-straining at dierent strain rates does not seem
to result in microstructural changes of sucient
degree to change the materials subsequent re-
sponse.
2.3.3. Microstructural observations
In order to check more closely the microstruc-
tural changes in DH-36 steel that may occur with
changes in the temperature, strain rate, and plastic
deformation, the microstructure of several de-
formed samples is examined by optical micro-
scope. Typical results are displayed in Figs. 1618.
400
600
800
1000
1200
1400
0 200 400 600 800 1000 1200
Temperature, (K)
F
l
o
w
S
t
r
e
s
s
(
M
P
a
)
Strain-Rate = 0.001/s
Strain-Rate = 0.1/s
Strain-Rate = 3,000/s
Strain-Rate = 8,000/s
DH-36,

= 0.10
Fig. 13. Eect of strain rates on ow stress at a true strain of 10%.
1032 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
0
200
400
600
800
1000
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e
S
t
r
e
s
s
(
M
P
a
)
0.1/s
0.001/s
DH-36, 296K
Sample 6
Fig. 14. Eect of strain rate jump from 0.001/s to 0.1/s on ow stress.
0
200
400
600
800
1000
1200
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e
S
t
r
e
s
s
(
M
P
a
)
3,000/s
0.001/s
DH-36, 296K
Sample7
Sample8
Fig. 15. Eect of strain rate jump from 0.001/s to 3000/s on ow stress.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1033
In Fig. 16, a sample is tested at a strain rate of
0.001/s and temperature of 77 K, to a true strain
exceeding 70% (isothermal). The ferrite and pear-
lite strips that were originally directed in the ver-
tical loading direction are now highly deformed
(the loading direction in all subsequent micro-
graphs is the same as that in Fig. 16). When two
samples are tested at the common temperature of
77 K and at strain rates of 3000/s and 0.001/s,
respectively, to a true strain of about 60% (adia-
batic, for the high strain rate), shearbands are seen
to have formed at an angle of 45 to the loading
direction (Figs. 17 and 18), and within the shear-
bands, the grains are seen to have been extended in
the shearband direction. This suggests that the
microstructure of DH-36 steel does not show re-
markable dierences when the material is tested at
dierent strain rates and dierent temperatures.
2.3.4. Flow stress at 77 K
Fig. 5 shows that when the true strain reaches
about 30%, the ow stress of DH-36 steel at 77 K
begins to drop with increasing strain, suggesting
some kind of microstructural damage at this
strain. To check this, the sample was examined by
optical microscope at 1000 amplications, and no
microcracks or other damages were noticed. In
Fig. 19, the ow stress curves of samples 9 and 10
are seen to be essentially the same, conrming that
the drop in the stress is repeatable. To further in-
vestigate this unusual result, sample 11 is loaded to
a true strain of about 30%, unloaded, and then
reloaded again at 0.1/s and 77 K. The resulting
ow stress is higher than that of samples 9 and 10,
as well as the initial part of sample 11. Sample 12 is
rst loaded to a true strain of about 20% and
unloaded. Then it is reloaded. The reloading curve
now follows closely the true stressstrain curve of
sample 9 or 10. Fig. 20 is the microstructure of
sample 12. In this gure, no shearbands are seen at
45 to the loading direction, but, instead, grains
Fig. 16. Microstructure of a sample strained to c 74% at 800
K and 0.001/s.
Fig. 17. Microstructure of a sample strained to c 59% at 77
K and 3000/s.
Fig. 18. Microstructure of a sample strained to c 58% at 77
K and 0.001/s.
1034 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
are seen to have extended perpendicular to the
loading direction. No microcracks or other dam-
ages are detected at this strain rate and tempera-
ture. This unusual phenomenon has not been
observed at other strain rates at 77 K. The drop in
the ow stress at a strain rate of 0.1/s and a tem-
perature of 77 K could be due to dynamic strain
aging.
3. Physically based constitutive model
3.1. Evaluation of plastic workheat conversion
factor
Plastic deformation generates heat, which is
either dissipated to the surroundings or is used to
increase the temperature of the material. When the
rate of heat generation is greater than the rate of
heat loss, the temperature of the material in-
creases. This generally happens at high strain
rates. For materials whose ow stress is tempera-
ture dependent, a continuous rise in temperature
during deformation results in simultaneous low-
ering of the ow stress. The temperature rise can
be calculated from
DT
_
c
0
b
q
0
C
V
sdc; 3:1
where q
0
is the mass density (7.8 g/cc), C
V
is the
temperature-dependent heat capacity (taken as 0.5
J/g K at room temperature), c is the plastic strain, s
is the ow stress in MPa, and b is the fraction of
the plastic work which is converted into heat. The
value of b is determined experimentally. Data re-
ported by Kapoor and Nemat-Nasser (1998) for
0
400
800
1200
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
DH-36, 0.1/s, 77K
Sample 9
Sample 11
Sample 12
Sample 10
Fig. 19. Flow stress at 0.1/s strain rate and 77 K temperature.
Fig. 20. Microstructure of a sample strained to c 20% un-
loaded, then reloaded to c 48% at 77 K and 0.001/s.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1035
several metals suggest that, for large strains (e.g.,
c P20%), b is essentially 1. This has also been
veried to be the case for several other polycrys-
talline metals, see Nemat-Nasser et al. (1999),
Nemat-Nasser and Isaacs (1997), and Nemat-
Nasser and Guo (1999). In the present case, we
have also found that b % 1:0.
To examine whether or not b % 1:0 for DH-36
steel, an indirect experiment is performed. The
area under the true stressstrain curve gives the
plastic work per unit volume in uniaxial defor-
mation. Three samples (designated as a, b, and c,
respectively) are loaded at the same strain rate of
3000/s. Sample a is loaded to a true strain of about
60% at an initial temperature of 23.6 C (room
temperature). The corresponding true stress-true
strain curve is displayed by a thin dashed curve in
Fig. 21. This is essentially an adiabatic true stress
true strain relation for DH-36. The temperature
rise in this adiabatic test is calculated using Eq.
(3.1), with b % 1:0.
Samples b and c are rst loaded to a true strain
of 24%, starting at room temperature (23.6 C).
Their true stressstrain relations are shown by
thick solid curves in Fig. 21. These curves fall on
the curve corresponding to sample a, showing the
reproducibility of the test results. The temperature
rise at a true strain of 24% is 50.6 C, calculated by
Eq. (3.1) with b % 1:0. Then sample b is heated to
74.2 C (50.6 +23.6) that corresponds to the initial
temperature of 23.6 C, and is reloaded at the
same strain rate, producing the second thin dashed
curve, shown in Fig. 21. This curve follows closely
the adiabatic curve of sample a. As a check, sam-
ple c is reloaded at its initial room temperature
(23.6 C), and the corresponding true stressstrain
curve is displayed by the thick solid curve marked
sample c. The stress dierence between the adia-
batic curve and this isothermal curve is measured
to be about 45 MPa, for a strain increment of 24%.
It is clear that this stress dierence (45 MPa) is due
to thermal softening of the material. Two impor-
tant conclusions are drawn from these results: (1)
if there was any recovery between unloading and
reloading, it did not aect the ow stress notice-
ably, as the interrupted curve of sample b follows
the uninterrupted curve of sample a; and (2) es-
sentially the entire plastic work is converted to
heat with a negligibly small amount being stored in
the sample as the elastic energy of the dislocations
and other defects, or lost through sample bound-
aries.
Fig. 21. Verication of heat conversion.
1036 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
3.2. Physically based constitutive model
The experimental results described above reveal
the following characteristics for DH-36 steel: (1)
The plastic ow stress of this DH-36 strongly de-
pends on the temperature and strain rate. Its
temperature sensitivity is greater for temperatures
below about 400 K, and minimal when the tem-
perature exceeds 400 K; (2) Dynamic strain aging
occurs at almost all strain-rates over a temperature
range from 500 to 1000 K, with the peak stress in
dynamic strain aging shifting to higher tempera-
tures with increasing strain rates; and (3) The mi-
crostructure of the material does not evolve in a
manner to aect its ow stress, as the temperature
and strain rate are changed.
A suitable constitutive model for this material
should therefore include all the above eects.
Based on the concept of dislocation kinetics, and
guided by experimental results, a physically based
model is developed by Nemat-Nasser and Isaacs
(1997), Nemat-Nasser et al. (1999) and Nemat-
Nasser and Guo (1999) and applied to several
polycrystalline metals. A similar model which in-
cludes all the characteristics observed in DH-36
structural steel does not exist. In the present work
we seek to incorporate the experimental under-
standing presented above for DH-36 steel, into the
constitutive model suggested by Nemat-Nasser
and co-workers. However, we will not include the
dynamic strain aging eects in the model.
Consider the plastic ow in the range of tem-
peratures and strain rates where diusion and creep
are not dominant, and the deformation occurs ba-
sically by the motion of dislocations. Here, for DH-
36, we assume that the ow stress, s, consists of two
parts: one part essentially due to the short-range
thermally activated eect which may include the
Peierls stress, point defects such as vacancies and
self-interstitials, other dislocations which intersect
the slip plane, alloying elements, and solute atoms
(interstitial and substitutional). We denote this by
s

. The second part is the athermal component, s


a
,
mainly due to the long-range eects such as the
elastic stress eld of dislocation forests and grain
boundaries. Thus, the ow stress is written as
s s
a
s

: 3:2
In this formulation, the total ow stress of a ma-
terial, s, is a function of the strain rate, _ cc, tem-
perature, T, and some internal microstructural
parameters. The microstructure here refers to the
grain sizes, the distribution of second-phase par-
ticles or precipitates, and the distribution and
density of dislocations. In general, the most com-
monly used microstructural parameter is the av-
erage dislocation density, q. The microstructure
can evolve dierently for dierent loading condi-
tions, that is, for dierent values of _ cc and T.
3.3. Athermal stress component, s
a
The athermal part, s
a
, of the ow stress, s, is
independent of the strain rate, _ cc. The temperature
eect on s
a
is only through the temperature de-
pendence of the elastic modulus, especially the
shear modulus, lT (Conrad, 1970). s
a
mainly
depends on the microstructure of the material, e.g.,
the dislocation density, grain sizes, point defects,
and various solute atoms such as those listed in
Table 1. Based on linear elasticity, s
a
would be
proportional to lT. Hence, we set
s
a
f q; d
G
; . . .lT=l
0
; 3:3
where q is the average dislocation density, d
G
is the
average grain size, the dots stand for parameters
associated with other impurities, and l
0
is a ref-
erence value of the shear modulus. In a general
loading, the strain c represents the eective plastic
strain (see Section 5) which is a monotonically
increasing quantity in plastic deformation. In the
present case, c denes the loading path and is also
a monotonically increasing quantity, since _ cc > 0.
Therefore, it can be used as a load parameter to
dene the variation of the dislocation density, the
average grain size, and other parameters which
aect s
a
, i.e., we may set
s
a
f qc; d
G
c; . . .lT=l
0

^
ff clT=l: 3:4
Further, as a rst approximation, we may use a
simple power-law representation of
^
ff c, and
choose an average value for l
0
so that lT=l
0
% 1.
Then, s
a
may be written as
s
a
% a
0
a
1
c
n
; 3:5
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1037
where a
0
, a
1
and n are free parameters which must be
xed experimentally. We emphasize that the eective
plastic strain, c, or any plastic-strain components
cannot in general represent the microstructure, and
here c is used strictly as a load parameter.
To identify the constitutive parameters for the
athermal stress in Eq. (3.4), we examine the vari-
ation of the ow stress with temperature, as shown
in Figs. 810, and 13. These results suggest that the
ow stress is essentially independent of the tem-
perature, beyond a certain critical temperature,
e.g., 400 K at strain rates of 0.001/s and 0.1/s, and
700 K at a strain rate of 3000/s. Because dynamic
strain aging occurs at all strain rates and at greater
temperatures, we evaluate the parameters of the
athermal stress component using the results of Fig.
10, and arrive at the following nal expression
s
a
750c
1=4
: 3:6
From Eq. (3.4), we observe that the function f
depends at least on two length scales, one associ-
ated with the average dislocation density, and the
other with the average grain size. Within the range
of our experimental results, both the dislocation
density and the grain sizes change with deforma-
tion. Since the elastic interaction forces between
two isolated dislocations is inversely proportional
to their spacing, it is often assumed, after Taylor
(1934, 1938), that the ow stress (here only the
athermal part) should also display a similar rela-
tion and hence be proportional to the square root
of the average dislocation density. Be that as it
may, there is no reason to expect that the average
dislocation density, qc, should have any pre-
dened dependence on the load parameter, i.e., the
eective plastic strain c. Indeed, our experimental
results give the relation (3.6), with a fourth root,
for the present material, and for other materials,
dierent exponents have been obtained experi-
mentally; see, e.g., (Nemat-Nasser et al., 1991,
1999, 2001; Nemat-Nasser and Isaacs, 1997).
3.4. Thermally activated component of the ow
stress, s

As mentioned before, the thermally activated


part of the ow stress, s

, represents the resistance


to the motion of dislocations by the short-range
barriers, such as the Peierls stress, point defects,
and other dislocations which intersect the slip
planes. The quantity s

, in general, is a function of
temperature, T, strain rate, _ cc, and internal vari-
ables characterizing the microstructure of the
material. As is discussed in connection with the
interrupted test results given in Figs. 11, 12, 14,
and 15, the microstructure of this material does
not seem to be very sensitive to temperature and
the strain-rate histories. Therefore, s

in the pre-
sent case is expected to depend on _ cc and T, and the
structure of the short-range barriers.
To obtain a relation between _ cc, T, and s

, let DG
be the energy that a dislocation must use through
its thermal activation in order to overcome its
short-range barrier. Kocks et al. (1975) suggest the
following relation between DG and s

, representing
a typical barrier encountered by a dislocation:
DG G
0
1
_

s

^ ss
_ _
p
_
q
; G
0
^ ssbk ^ ssV

;
3:7
where 0 < p 61 and 1 6q 62 dene the prole of
the short-range barrier, ^ ss is the shear stress above
which the barrier is crossed by a dislocation with-
out any assistance from thermal activation, and G
0
is the free energy required for a dislocation to
overcome the barrier solely by its thermal activa-
tion; b is the magnitude of the Burgers vector; k and
are the average eective barrier width and spac-
ing, respectively; and V

is the activation volume.
We note in passing that

V

3
p
provides a nat-
ural length scale in this physics-based model.
We dene the plastic strain rate by _ cc bq
m
tt
bq
m
x
0
expDG=kT, and set
_ cc _ cc
r
exp
_

DG
kT
_
; 3:8
where _ cc
r
q
m
b tt; q
m
is the average density of the
mobile dislocations and tt x
0
expDG=kT is
their average velocity, with x
0
being the attempt
frequency; and k is the Boltzmann constant. From
Eq. (3.7) and (3.8), obtain
s

^ ss 1
_
_

_

kT
G
0
ln
_ cc
_ cc
r
_
1=q
_
_
1=p
: 3:9
1038 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
In Eq. (3.9), the parameters p and q dene the
prole of the short-range energy barrier to the
motion of dislocations. Ono (1968) and Kocks et
al. (1975) suggest that p 2=3 and q 2 are
suitable values for these parameters for many
metals. Nemat-Nasser and co-workers (1997
1999,2001) have veried this for several metals.
Here, for DH-36, we also use the same values for p
and q in (3.9). The parameters k=G
0
and _ cc
r
char-
acterize the temperature and strain-rate sensitivity
of the material. Greater temperature sensitivity is
associated with the larger k=G
0
, whereas larger _ cc
r
corresponds to smaller strain-rate sensitivity. The
product k=G
0
= ln _ cc
r
can be estimated directly
from the experimental data of Fig. 22. The steps
are as follows. The experimental data in Fig. 22 are
obtained by subtracting s
a
, given by Eq. (3.6),
from the data in Fig. 10. The results represent the
variation of s

with temperature, for indicated


strains and a strain rate of 3000/s. It is seen that
dynamic strain aging occurs beyond a temperature
of 800 K at this strain rate. We therefore exclude
the experimental data for temperatures exceeding
800 K, as we do not wish to include the dynamic
strain aging in our model; see Cheng and Nemat-
Nasser (2000) for modeling that includes DSA
eects. Therefore, for temperatures less than 800
K, the results in Fig. 22 can be represented by a
single curve (solid curve in Fig. 22), given by
s

15001 0:00105T
1=2

3=2
: 3:10
Comparing Eqs. (3.9) and (3.10), we conclude that
^ ss 1500 MPa, and that

k
G
0
ln
_ cc
_ cc
r
0:00105: 3:11
To obtain the values of G
0
and _ cc
r
from Eq. (3.11),
it is necessary to estimate _ cc
r
or have a measured
value for G
0
. Nemat-Nasser et al. (2001) have
found that k=G
0
% 6:6 10
5
K
1
and _ cc
r
% 2
10
10
=s are suitable values for AL-6XN stainless
steel. Here we have found that the same values
may be used for DH-36 steel. The corresponding
parameters, used in Eq. (3.11), then are: G
0
% 1:3
eV per atom, and b O10
10
m, x
0

O10
12
=s, q
m
O10
13
m
2
, and
0
O10
3

lattice spacing.
Now, the nal constitutive relation (denoted as
the PB-model, physics-based) for this material
becomes, for T 6T
c
,
-200
0
200
400
600
800
1000
1200
0 200 400 600 800 1000 1200 1400
Temperature, T (K)
F
l
o
w

S
t
r
e
s
s
,

*

(
M
P
a
)
Strain = 0.05
Strain = 0.2
Strain = 0.3
Strain = 0.4
Strain = 0.5
DH-36, 3,000/s
1/2
1500 [1-(0.00105T) ]

3/2
Fig. 22. Thermal part of stress: experimental results and its mathematical t at indicated strains.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1039
s 750c
1=4
1500 1
_
_
_

_
6:6 10
5
T ln
_ cc
2 10
10
_
1=2
_
_
_
3=2
;
where
T T
0
0:267
_
c
0
sdc 3:12
and for T > T
c
, we have
s 750c
1=4
; 3:13
where
T
c
6:6 10
5
ln
_ cc
2 10
10
_ _
1
;
and b=q
0
C
V
0:267 K/MPa.
Figs. 2326 compare the experimental results
with the PB-model predictions at strain rates of
0.001/s to 3000/s, for indicated initial tempera-
tures. To further verify the predictability of this
model, independent tests at an 8000/s strain rate
and various initial temperatures are performed,
and the results are displayed in Fig. 27, together
with the corresponding model predictions. As is
seen, good correlation between these data and the
model predictions is obtained.
As pointed out before, the PB-model does not
include the dynamic strain aging eects, which oc-
cur in the temperature range of 300 to 800 K, at the
low strain rates of 0.001/s and 0.1/s. In Figs. 23 and
24 we have shown the experimental results for these
low strain rates. Aside from the eect of dynamic
strain aging, the model predictions are in reason-
able agreement with the experimental results.
Here we note that Eq. (3.10) and the experi-
mental data of Fig. 22 suggest that the activation
volume V

and hence the corresponding length
scale

are essentially constant in the present case.


This is not, however, generally the case for other
materials, as has been shown for titanium and
copper in (Nemat-Nasser and Isaacs, 1997; Ne-
mat-Nasser and Li, 1998).
4. Assessment by JohnsonCook model
In order to assess the PB-model, we now con-
sider the JohnsonCook (JC) (Johnson and Cook,
1983) model that has enjoyed much success in
various applications (Liang and Khan, 1999), and
0
200
400
600
800
1000
1200
1400
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
77K
296K
400K
DH-36, 0.001/s
Point Curves: Experiments
Solid Curves: PB Model Predictions
Fig. 23. Comparison of PB-model predictions with experimental results at a strain rate of 0.001/s and indicated initial temperatures.
1040 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
apply it to model our experimental results for DH-
36. As suggested by Johnson and Holmquist
(1988), care must be exercised when using extra-
polation to estimate the parameters of this model.
In the JC-model, the von Mises eective stress, s, is
expressed as,
s A Bc
n
1 C ln _ ee

1 T
m
; 4:1
0
200
400
600
800
1000
1200
1400
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
) T
0
= 77K
T
0
= 600K
T
0
= 500K
T
0
= 400K
T
0
= 296K
DH-36, 3,000/s
Point Curves: Experiments
Solid Curves: PB Model Predictions
Fig. 25. Comparison of PB-model predictions with experimental results at a strain rate of 3000/s and indicated initial temperatures.
0
200
400
600
800
1000
1200
1400
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
77K
296K
400K
DH-36, 0.1/s
Point Curves: Experiments
Solid Curves: PB Model Predictions
Fig. 24. Comparison of PB-model predictions with experimental results at a strain rate of 0.1/s and indicated initial temperatures.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1041
where c as before is the eective plastic strain,
_ ee

_ cc=_ cc
0
is the dimensionless strain rate, ( _ cc
0
is
normally taken to be 1.0/s), and
T


T T
r
T
m
T
r
; 4:2
0
400
800
1200
1600
2000
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
T
0
= 296K
T
0
= 500K
T
0
= 77K
Point Curves: Experiments
Solid Curves: PB Model Predictions
DH-36, 3,000/s
Fig. 26. Comparison of PB-model predictions with experimental results at a strain rate of 3000/s and indicated (isothermal) tem-
peratures.
0
200
400
600
800
1000
1200
1400
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
T
0
= 77K
T
0
= 296K
T
0
= 500K
T
0
= 800K
DH-36, 8,000/s
Point Curves: Experiments
Solid Curves: PB Model Predictions
Fig. 27. Comparison of PB-model predictions with experimental results at a strain rate of 8000/s and indicated initial temperatures.
1042 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
where T
r
is a reference temperature and T
m
( 1773
K) is the melting temperature of the material. T
r
must be chosen as the lowest temperature of in-
terest or the lowest temperature of experiments
because the parameter m is normally less than one,
and T should be greater than or equal to T
r
for Eq.
(4.1) to be valid; in the present case, we take
T
r
50 K. The ve material constants, A, B, n, C
and m, are obtained such as to achieve good cor-
relation with the experimental results. Table 2
gives the nal values of these parameters.
The nal expression for the JohnsonCook
model for DH-36 now is
s 10201 1:5c
0:4
1 0:015 ln _ ee

1 T
0:32
;
T


T 50
1723
: 4:3
In this equation, s, c, and _ ee are axial true stress,
true strain, and the numerical value of the strain
rate (since it is normalized by the 1/s strain rate),
respectively. Figs. 2832 compare the experimental
results with the model predictions at strain rates of
0.001/s to 8000/s, for indicated initial tempera-
tures.
Both models show good agreement with the
Hopkinson bar data. Because of the eect of dy-
namic strain aging, which is not incorporated into
either model, both models tend to under-predict
the ow stress for an initial temperature of 800 K,
as is seen in Figs. 27 and 32.
Compared with the PB-model, it is seen that, at
low strain rates of 0.001/s and 0.1/s, the JC-model
has less accurate predictions for temperatures
between room and 400 K. The reason for this is
that the strain-rate constant C has been chosen
such that good correlation with the high strain-
rate data is achieved, as the primary application
of the model is at higher strain rates. This then
Table 2
Values of the parameters in the JC-model for DS-36
A B n C m T
r
(K) T
m
(K)
1020 1530 0.4 0.015 0.32 50 1773
0
200
400
600
800
1000
1200
1400
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
77K
296K
400K
DH-36, 0.001/s
Point Curves: Experiments
Solid Curves: JC Model Predictions
Fig. 28. Comparison of JC-model predictions with experimental results at a strain rate of 0.001/s and indicated initial temperatures.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1043
leads to a poor prediction of the ow stress at low
strain rates. Since the PB-model is based on var-
ious aspects of the kinetics and kinematics of
dislocation motion, it seems to have a better
predictive capability over a broader range of
strain rates.
0
200
400
600
800
1000
1200
1400
1600
1800
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
77K
296K
400K
DH-36, 0.1/s
Point Curves: Experiments
Solid Curves: JC Model Predictions
Fig. 29. Comparison of JC-model predictions with experimental results at a strain rate of 0.1/s and indicated initial temperatures.
0
200
400
600
800
1000
1200
1400
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
T
0
= 77K
T
0
= 600K
T
0
= 500K
T
0
= 400K
T
0
= 296K
DH-36, 3,000/s
Point Curves: Experiments
Solid Curves: JC Model Predictions
Fig. 30. Comparison of JC-model predictions with experimental results at a strain rate of 3000/s and indicated initial temperatures.
1044 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
5. Application of model to three-dimensional defor-
mation
To apply the model for the three-dimen-
sional calculations, we view s and _ cc as the ef-
fective von Mises stress and strain rate, dened
by
s
3
2
r
0
ij
r
0
ij
_ _
1=2
; _ cc
2
3
D
p
ij
D
p
ij
_ _
1=2
; 5:1
0
400
800
1200
1600
2000
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
T
0
= 296K
T
0
= 500K
T
0
= 77K
Point Curves: Experiments
Solid Curves: JC Model Predictions
DH-36, 3,000/s
Fig. 31. Comparisonof JC-model predictions with experimental results at a strain rate of 3000/s and indicated(isothermal) temperatures.
0
200
400
600
800
1000
1200
1400
1600
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
True Strain
T
r
u
e

S
t
r
e
s
s

(
M
P
a
)
T
0
= 77K
T
0
= 296K
T
0
= 500K
T
0
= 800K
DH-36, 8,000/s
Point Curves: Experiments
Solid Curves: JC Model Predictions
Fig. 32. Comparison of JC-model predictions with experimental results at a strain rate of 8000/s and indicated initial temperatures.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1045
where r
0
ij
, i; j 1; 2; 3, are the rectangular Carte-
sian components of the deviatoric part of the true
stress tensor, and D
p
ij
are the components of the
deviatoric part of the plastic deformation rate
tensor. In (5.1), repeated indices are summed over
1, 2, and 3. For uniaxial tests, s is the axial stress
and _ cc is the axial strain rate.
To illustrate the use of model results in three-
dimensional settings, consider a plasticity model in
which the Jaumann rate of true stress, r
D
ij
, is re-
lated to the elastic deformation rate tensor by
r
D
ij
C
ijkl
D
kl
D
p
kl
:
r
D
ij
_ rr
ij
W
ik
r
kj
r
ik
W
kj
;
5:2
where C
ijkl
is the instantaneous elasticity tensor, D
ij
and W
ij
are the deformation rate and spin tensors,
given in terms of the velocity gradient L
ij

ot
i
ox
j
, by
D
ij

1
2
L
ij
L
ji
, and W
ij

1
2
L
ij
L
ji
, respec-
tively. We now consider the simplest model for the
deviatoric plastic deformation rate, D
p
ij
, as follows:
D
p
ij
_ cc
l
ij

2=3
_ ; l
ij

r
0
ij

2=3
_
s
; 5:3
with _ cc given by (3.8). For DH-36 and using the PB-
model parameters, we have
_ cc 2 10
10
exp
_
_
_

10
5
6:6T
1
_

s 750c
1=4
1500
_ _
2=3
_
2
_
_
_
;
c
_
t
0
2
3
D
p
ij
D
p
ij
_ _
1=2
dt; 5:4
where s is measured in MPa, and t measures the
actual time.
6. Conclusions
To understand and model the thermomechani-
cal response of DH-36 structural steel, uniaxial
compression tests are performed on cylindrical
samples. True strains exceeding 60% are achieved
in these tests, over the range of strain rates from
0.001/s to about 8000/s, and at initial temperatures
from 77 to 1000 K. In an eort to understand the
underlying deformation mechanisms, some inter-
rupted tests with temperature and strain rate
jumps are also performed. The microstructure of
undeformed and deformed samples is examined.
Several noteworthy conclusions are as follows.
1. The experimental results show that the mechan-
ical behavior of the DH-36 baseplate does not
depend on the sample orientation (or location
through the thickness).
2. This steel displays good ductility and plasticity
(strain>60%) at low temperatures (even at 77
K) and high strain rates, without displaying any
noticeable damage or microcracks, but shear-
bands are detected at true strains exceeding 60%.
3. When the temperature exceeds the room temper-
ature, the ow stress of this material decreases at
a lower rate (especial at low strain rates, i.e., be-
low about 0.1/s) with increasing temperature.
Thus, the strength of this material is not temper-
ature-sensitive at high temperatures, showing
that the material has good weldability.
4. Dynamic strain aging occurs at temperatures
between 500 and 1000 K and the range of
strain-rates from 0.001/s to 3000/s, with the
peak value of the stress shifting to higher tem-
peratures with increasing strain rates. For ex-
ample, to about 600 K for a 0.001/s strain
rate, to about 650 K for 0.1/s strain rate, and
to about 800 K for a 3000/s strain rate.
5. The microstructure of this material is not very
sensitive to the changes of strain rate and tem-
perature.
6. Based on the experimental results, a physically
based model is developed. In the absence of dy-
namic strain aging, the model predictions are in
goodagreement withthe experimental results over
a wide range of temperatures and strain rates.
7. As an alternative to this model, the Johnson
Cook model is considered and its free parame-
ters are estimated using our data. Both models
show good agreement with the Hopkinson bar
data, with the physics-based model having bet-
ter correlation with the experimental results
over a broader range of strain rates.
Acknowledgements
The authors would like to thank Mr. Philip
Dudt for supplying the material and data in
1046 S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047
Table 1, and Mr. Jon Isaacs for his assistance in
sample preparation. This work was supported by
Naval Surface Warfare Center, Carderock Divi-
sion (NSWCD) under a program directed by Mr.
Philip Dudt, through contract N00167-02-M-0346
to the University of California, San Diego.
References
Beukel, A.V.D., Kocks, U.F., 1982. The strain dependence of
static and dynamic strain-aging. Acta Metall. 30, 10271034.
Conrad, H., 1970. The athermal component of the ow stress in
crystalline solids. Mater. Sci. Eng. 6, 260264.
Cho, Sang-Hyun, Yoo, Yeon-Chul, Jonas, J.J., 2000. Static and
dynamic strain aging in304austenitic stainless steel at
elevated temperature. J. Mater. Sci. Lett. 19, 20192022.
Cheng, J.Y., Nemat-Nasser, S., 2000. A model for experimen-
tally observed high-strain-rate dynamic strain aging in
titanium. Acta Mater. 48, 31313144.
Cheng, J.Y., Nemat-Nasser, S., Guo, W.G., 2001. A unied
constitutive model for strain-rate and temperature depen-
dent behavior of molybdenum. Mech. Mater. 33, 603616.
Johnson, G.R., Cook, W.H., 1983. A constitutive model and
data for metals subjected to large strains, high strain rates
and high temperatures. In: Proceedings of the Seventh
International Symposium on Ballistic, The Hague, The
Netherlands, pp. 541547.
Johnson, G.R., Holmquist, T.J., 1988. Evaluation of cylinder-
impact test data for constitutive model constants. J. Appl.
Phys. 64 (8), 39013910.
Krauss, G., 1990. Microstructures, processing, and properties
of steel. ASM Handbook 1, 126139.
Kocks, U.F., Argon, A.S., Ashby, M.F., 1975. Thermodynam-
ics and kinetics of slip. Progress Mater. Sci. 19, 1271.
Kubin, L.P., Estrin, Y., Perrier, C., 1992. On static strain aging.
Acta Metall. 40, 10371044.
Kapoor, R., Nemat-Nasser, S., 1998. Determination of tem-
perature rise during high strain rate deformation. Mech.
Mater. 27, 112.
Liang, R.Q., Khan, A.S., 1999. A critical review of experimen-
tal results and constitutive models for BCC and FCC metals
over a wide range of strain rates and temperatures. Int. J.
Plast. 15, 963980.
Nakada, Y., Keh, A.S., 1970. Serrated ow in NiC alloys.
Acta Metall. 18, 437443.
Nemat-Nasser, S., Guo, W.G., 1999. Flow stress of commer-
cially pure niobium over a broad range of temperatures and
strain rates. Mater. Sci. Eng. A 284, 202210.
Nemat-Nasser, S., Isaacs, J.B., 1997. Direct measurement of
isothermal ow stress of metals at elevated temperatures
and high strain rates with application to Ta and TaW
alloys. Acta Mater. 45, 907919.
Nemat-Nasser, S., Li, Y.L., 1998. Flow stress of fcc polycrys-
tals with application to OFHC Cu. Acta Mater. 46, 565
577.
Nemat-Nasser, S., Isaacs, J.B., Starrett, J.E., 1991. Hopkinson
techniques for dynamic recovery experiments. Proc.R. Soc.
Lond 435 (A), 371391.
Nemat-Nasser, S., Guo, W.G., Liu, M.Q., 1999. Experimen-
tally based micromechanical modeling of dynamic response
of molybdenum. Scr. Mat. 40, 859872.
Nemat-Nasser, S., Guo, W.G., Kihl, D.P., 2001. Therome-
chanical response of AL-6XNstainless steel over a wide
range of strain rates and temperatures. J. Mech. Phys. Solids
49, 18231846.
Ono, K., 1968. Temperature dependence of dispersed barrier
hardening. J. Appl. Phys. 39, 18031806.
Taylor, G.I., 1934. The mechanism of plastic deformation of
crystalsI, II. Proc. R. Soc. London A 145, 362404.
Taylor, G.I., 1938. Plastic strain in metals. J. Inst. Metals 62,
307324.
S. Nemat-Nasser, W.-G. Guo / Mechanics of Materials 35 (2003) 10231047 1047

Вам также может понравиться