You are on page 1of 12

INTERNATIONAL

JOURNAL OF
IMPACT
ENGINEERING
PERGAMON International Journal of Impact Engineering 26 (2001) 53-64
www.elsevier.com/locate/ijimpeng

CONSTITUTIVE MODELING FOR DUCTILE METALS BEHAVIOR


INCORPORATING STRAIN RATE, TEMPERATURE
AND DAMAGE MECHANICS
NICOLA BONORA and PIETRO PAOLO MILELLA
DiMSAT- Dept. of Mechanics, Structures and Environment, University of Cassino
Via G. Di Biasio 43 -03043 Cassino, Italy - e-mail: nbonora@unicas.it

Abstract ---One the most serious limitation to an extensive use of computational techniques in
simulating and predicting structures and components behavior under dynamic loading is given
by the inadequacy of constitutive models to fairly represent failure process. In this paper a new
constitutive model for ductile metals has been developed using the innovative solid state
equation proposed by Milella (1998) and the non-linear damage model proposed by Bonora
(1997). These two models are physically based and require a very limited number of constants
that can be experimentally identified according to the procedures given by the authors. The
implementation of the model in commercial finite element code is simple and cost effective
with respect to similar nucleation and growth (NAG) models with the additional feature that
hydrostatic pressure effect on ductile damage is correctly taken into account. © 2001 Elsevier
Science Ltd. All rights reserved.

Keywords: ductile damage, strain rate, solid state equation, impact, dynamic loading

INTRODUCTION
In the recent years, the use of traditional metals and alloys in structural application has gone in
the direction to exploit, as much as possible, the material capability to plastically deform. For
instance, a wide plastic deformation range is expected for those structural components that have
to absorb energy such as shock absorbers, containment, blast shields, and so on. In the range of
large plastic deformation, the mechanism of failure is intimately connected to the material
constitutive response and no reliable design criteria can be developed without incorporating the
information relative to it. The situation becomes even more complicated when the deformation of
solids occurs under dynamic loading, either low or high strain rate, and in temperature. For
instance, this is the case of blast shields and shaped charges where the liner is deformed with a
strain rate, that varies radially, by the propagation of the pressure loading, generated by the
explosive detonation.
In 1982, Zukas et al. [1] pointed out that the most serious limitation to an extensive use of
computational techniques in simulating and predicting structures and components behavior under
dynamic loading was given by the inadequacy of constitutive models to fairly represent failure
process.
From the computational point of view, it is well know that what affects most the results are
the flow stress and failure criteria, Gupta and Misey [2]. Constitutive models that do not
incorporate damage usually lead to predictions far from the experiment. This behavior can be
observed even when strain rate effects are negligible. For instance, this is the case of the necking
process of a round bar in tension. Here, the global response, applied load versus diameter
reduction at the necking section, cannot be numerically predicted without the use of a damage
model responsible for the localization of plastic deformation and progressive loss of load

0734-743X/01/$ - see front matter © 2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 7 3 4 - 7 4 3 X ( 0 1 ) 0 0 0 6 3 - X
54 N. Bonora, P,P. Milella / International Journal of lmpact Engineering 26 (2001) 53-64

carrying capability, (Bonora, [3]).


Sometimes a posteriori failure criteria can improve the quality of the predictions but too often
they are used as adjustable variable to match existing experimental data.
Thus, in order to predict material behavior under these conditions, damage mechanisms
associated with the increase of plastic deformation and the variation of the constitutive response
true stress-true plastic strain with strain rate and temperature should be given and, if possible,
incorporated in a unified model. Since, many models have been presented in the literature on
both issues mentioned above, a number of additional requirements should be fulfilled when
possible: the model should be valid for a wide class of metals showing a material independence.
The material constants introduced in the model should have a clear physical meaning. A simple
and reliable identification procedure for the material parameters should be possible. Finally, the
numerical implementation of the model in finite element codes should be as much as possible
straightforward and computationally cost effective.
Rencently, Bonora [4] proposed a continuum damage mechanics model for ductile failure,
derived in the framework proposed initially by Lemaitre [5], that with respect to other similar
formulations, resulted to be material independent and capable to predict material behavior under
any general state of stress conditions. On the other side, Milella [6] proposed a new solid state
equation (SSE) for the evolution of plastic flow with strain rate and temperature, showing
dependence of strain rate on temperature.
On these premises, a unified constitutive model for ductile metals that incorporate damage,
strain rate and temperature effects has been developed.

DUCTILE FAILURE MODELING

Plasticity damage in ductile metals is mainly due to the formation and growth of microcavities
that initiate either due to the debonding from the ductile matrix or fracture of brittle inclusions,
such as carbides and sulfides. Microvoids initiation and growth progressively reduces the
material capability to carry loads leading the material to fall. This void growth process controls
ductile failure both under quasi static and dynamic loading conditions. McClintock [7] firstly
recognized the role of microvoids in ductile failure process and tried to correlate the mean radius
of the nucleated cavities to the plastic strain. Rice and Tracy [8] analytically studied the evolution
of spherical voids in a elastic-perfectly plastic matrix. In these pioneering studies the interaction
effects between near microvoids, microvoids coalescence process and hardening effects were
neglected. Failure was postulated to occur when the cavity radius would reach a critical value
(namely, critical cavity growth). Seaman [9] reviewed a number of failure models. He came to
the conclusions that damage should be incorporated into any constitutive model since there is no
instantaneous jump from undamaged to fully separated material and material softening caused by
damage modifies the wave propagation through the material.
In the last decades, two different approaches have been proposed: the Gurson's model based
approach and the continuum damage mechanics (CDM). In the first case, the plasticity damage
due to microvoid formation is taken into account in a modified yield condition as proposed by
Gurson [10] were a porosity term, when activated, softens the material plastic response. Later,
Koplik and Needleman [11] and Needleman and Tvergaard [12] modified the Gurson's model in
order to include the acceleration in the failure process induced by voids coalescence. The
Gurson's model, and the models derived by its modification, are limited as a result of the fact
that a very large number of material constants (up to nine, Brocks and Bemauer, [13]) is required
and none of them is directly measurable on the material. Continuum damage mechanics approach
was initially developed by Lemaltre [5] and Chaboche [14] starting from the initial study of
Kachanov [15]. Here, the complete set of constitutive equations is derived making the strain
equivalence hypothesis (Lemaitre [5]) and assuming the existence of damage dissipation
potential similarly to plasticity. Following this approach and suggesting the use of special forms
N. Bonora, PP Milella / International Journal of lmpact Engineering 26 (2001) 53-64 55

for the damage dissipation potential, many damage models have been proposed (Chandrakanth
and Pandey, [16], Tay and Yang,[17]). Shockey et al. [18], Seaman et al. [19] and Elrich et al.
[20] proposed a CDM model starting from the experimental observations of complex damage
patterns that occur under elevated strain rates deformation process.
Recently Bonora [4] proposed a general non linear CDM model for ductile damage
independent of the material under investigation, Figure 1. In this model, the material parameters
have a clear physical meaning and can be directly measured on the material performing simple
experimental tests without the necessity of any iterative process. In addition the model resulted to
correctly take into account the effects caused by stress triaxiality, Bonora [21].

1,0 .
• A12024exp. data
0,9- 0 Copper 99.9*,4exp. data ~ /
A Steel1091exp.data ~ /
0,8 present model A~m J A

~" 0,4] _

~ 0,3 ,

0,2-

0,1

0,0 I I I " I I ' I I I I


0,0 0,1 0,2 0,3 0,4 0,5 0,6 0,7 0,8 0,9 1,0

(~'%Y(sc~)

Fig. 1. Basic damage evolutions with plastic strain for aluminum copper and steel

NON LINEAR CDM MODEL

Lemaitre [5] firstly defined the CDM framework for plasticity damage. Here, damage is one
of the thermodynamics state variables that takes into account the progressive loss of load
carrying capability of the material as a result of some irreversible modifications, such as
microvoids formation and growth, that initiate in the materials when intense plastic deformation
occurs. Physically speaking, damage can be defined as the reduction of the nominal section area
of a given reference volume element (RVE) as a result of microcavities formation and growth:

where, for a given normal n, ~ ' ) i s the nominal section area of the RVE and at") is the
effective resisting one reduced by the presence of micro-flaws and their mutual interactions. It is
clear that this definition naturally leads to a tensor formulation of the damage variable, as
proposed either by Chaboche [14]. These formulations, even though formally correct, are very
difficult to be use in practical applications as a result of the impossibility to experimentally
identify material constants, damage components evolution and, most of all, their mutual
interactions.
Here, for sake of simplicity and effectiveness, damage is assumed to be isotropic. This
assumption, in many cases, is not too far from reality as a result of the random shapes and
56 N. Bonora, P.P. Milella / International Journal of Impact Engineering 26 (2001) 53-64

distribution of the included particles and precipitates that trigger plasticity damage initiation and
growth. In addition, the assumption that the strain associated with a damage state, under a given
applied stress, is equivalent to the strain associated with its undamaged state under the
corresponding effective stress (also known as the strain equivalence hypothesis, Lemaitre, [5]),
leads to the following definition for the damage variable D, using material stiffness reduction
concept:

D = I ~,H (2)
&
where Eo and E,H are the Young's modulus of the undamaged and damaged material,
respectively. The complete theoretical framework for plasticity damage can be defined by means
of the following additional assumptions:

• A dissipation potential fo , similarly to the one used in plasticity theory, is assumed for
damage;
• Absence of any coupling between damage and plasticity dissipation potentials is assumed. It
follows that the total dissipation potential is given by linear superposition of the dissipation
potential associated to plastic deformationfp (that is the function of the actual stress tensor, ~,
and of the isotropic and kinematic hardening back stress, R and X , respectively) and of the
damage dissipation potential,fo, (function of the damage associated internal variable Y and of
the accumulated effective plastic strain, p):
f = fp(ff,R,X;D)+fo(Y;p,D) (3)

• Damage variable, D, is coupled with plastic strain. The relation between damage and the
irreversible strain in the micro/meso scale is expressed in the kinetic law of damage
evolution where the rate of the plastic multiplier ~. is proportional to the rate of the
effective accumulated plastic strain, /~.
• Plastic damage dissipation is associated to the micro-cavities growth that is a highly non
linear process; thus, the damage dissipation potential has to depends on the effective
accumulated plastic strain that plays the role of an indicator of which growth phase is
dominating microvoids growth process.
• Damage processes are localized in the material micro-scale and their effects remain
confined until the complete failure of several elementary volume elements, with the
consequent appearance of a macroscopic crack, occurs;
• the same set of constitutive equations for the virgin material can be used to describe the
damaged material replacing only the stresses with the effective ones and assigning a state
equation for D.

Bonora [4] proposed the following expression for the damage dissipation potential,

= • . ~ (4)
Io L2 ~ So) p,

Here, De, is the critical value of the damage variable for which ductile failure occurs, S O is a
material constant and n is the material hardening exponent. ~ is the damage exponent that
determines the shape of the damage evolution curve and is related to the nature of the bound
between brittle inclusions and the ductile matrix. The complete set of constitutive equation for a
plastically deformed damaged material can be derived. Here the set is summarized for an
isotropie hardening material (no kinematic hardening):
N. Bonora, P.P Milella /International Journal of lmpact Engineering 26 (2001) 53-64 57

strain decomposition
"~ = Eij"" +e~7
~'ij (5)
elastic strain rate
l + v c ,j v
E 1-D E1----D (6)
plastic strain rate
.e=~o3fp = ~ 3 s;/ 1
(7)
e~ c)oq 2 1- D (~eq
plastic multiplier
b = _)~Offp = ~ = p ( 1 - D ) (8)
3/?
kinetic law of damage evolution

D = - ~ t - - ~ = a . ( D c r - D ° ~ . f ( (r" I . ( D . - D ~ . ~ (9)
ln(e:/e,h) ~(~eq) P
more detailed description on the derivation of these Equations can be found elsewhere, (Bonora,
[4], Bonora and Newaz, [22]). In Equation (9) stress triaxiality effects are taken into account by
the function f(on/O,q) given by,

f ( ~ q ~ ) = 2(1 + v ) + 3" (1 - 2v)" ~ e q )/2


(trH (10)

derived from the assumption that ductile damage mechanism is governed by the total elastic
strain energy, Lemaitre [23]. Here, oH = Okk/3 is the hydrostatic part of the stress tensor and v is
the Poisson's ration. The model needs five material parameters in order to be applied: the strain
threshold (in uniaxial monotonic loading) eth, at which damage processes are activated. The
theoretical failure strain, under complete uniaxial state of stress conditions, ef at which ductile
failure occurs.
The initial amount of damage present in the material, Do. The critical damage, De,, at which
failure occurs and the damage exponent, ct, that control the shape of damage evolution with
plastic strain. Since, for the virgin material Do can be assumed equal to zero, the parameters can
be reduced to four. Bonora [3] and Bonora et al. [24] proposed a simple experimental procedure
where these parameters can be assessed through a uniaxial tensile test using an appropriate
hourglass shaped tensile specimen with rectangular cross section. In order to obtain damage
evolution with plastic strain, Eqn. (9) needs to be numerically integrated, However, for the case
of proportional loading, wherefloH/treq)is constant, Eqn.(9) can be analytically integrated leading
to:

{E I/11
D= Do +(Dc.-Do). 1- 1--1n(p/ P'h) . f
In(e//e,h) k q ]-J J
This expression can be once again specialized for the very special case of pure uniaxial state
(11)

of stress condition wheref(crn/Oeq)-~l:

D~-no+(ncr-Do). l - 1 {I' ln~i/--,h) j J (12)


58 N. Bonora, P.P. Milella /International Journal of lmpact Engineering 26 (2001) 53-64

SOLID S T A T E EQUATION: BACKGROUND

It is now well recognized that metals with a body-centered cubic (BCC) lattice, like a-iron,
ferritic steels, niobium, tantalum etc., have a mechanical behavior strongly dependent on the
temperature T and the strain rate ~. This, in general, is not observed in metals having a face-
centered cubic (FCC) structure, like austenitic steels, aluminum, copper, silver, etc., in particular
as far as temperature is concerned, which does not seem to have the same pronounced effect on
the yield strength, at least below the creep regime. Between the two, there are metals with a
closely-packed hexagonal lattice (HCP), like titanium and zinc, that show a dependence of the
yield strength on temperature and strain rate, but less pronounced than in BCC materials. It is
rather difficult to describe the plastic behavior of all metals through a generalized process that
leads to a unified theory. Many attempts have been made in the past to establish what is known
as the mechanical equation of the solid state that, like the gas state where pressure and volume
are linked with temperature, relates the yield strength Cryto the temperature T and the strain rate

cry = f(e,~,T) (13)

Zener and Hollomon [25], Manjoine [26], Cottrell [27-29, 31], Cottrell and Bilby [30],
Hollomon [32], Yokobori [33], Seeger [34-35], Campbell [36], Conrad [37], Conrad et al.[38-
39], Petch [40], Petch and Helsop [41] and others, have tried to derive that equation through an
empirical or physical approach, based on experiments, or a theoretical one based on either the
activation energy concept or dislocation models. More recently, Johnson and Cook [42], Johnson
et al. [43], Klopp et al.[44], Hoge and Mukherjee [45], Zerilly-Armstrong [46-48], Follansbee
[49], Follansbee and Kocks [50] have addressed the issue and developed empirical or physically-
based constitutive equations. Often, the solid state or constitutive equations proposed were not
consistent at all, but in general, it has been found an exponential dependence of the yield strength
cry on temperature and an equivalence between temperature and strain rate. This latter aspect, in
particular, is well shown by the Zener and Hollomon equation:
cry = f ( ~ . e Q~Rr) (14)

where Q is the so-called activation energy and R the universal gas constant. Milella [6] proposed
a single equation of state able to describe the plastic behavior of materials as a function of strain
rate and temperature. Its thermal component is based on dislocations behavior and, in particular,
on the concept of Cottrell's atmosphere that surrounds each dislocation, supported by the granite
solidity of experimental evidences obtained on a wide variety of steels and metals having
different microstructure and strengths. Its strain rate component, instead, is derived from the
characteristic trend of the material response at constant ~ and variable temperature.

T H E M E C H A N I C A L E Q U A T I O N OF SOLID STATE: T E M P E R A T U R E AND STRAIN


R A T E E F F E C T ON YIELD S T R E N G T H

It must be explicitly said that the temperature field considered in this analysis is that below the
creep range. To deform a metal beyond the elastic limit or yield strength cry, means to activate
and move its dislocations through the crystal. Dislocations encounter obstacles as they move
through the lattice. Some of these, also called long-range obstacles, are due to the structure of the
material and cannot be overcome introducing thermal energy into the crystal. Therefore they
contribute to the yield strength with a component that is non-thermally activated. Others, on the
contrary, can be overcome by thermal energy. These latter are normally called short-range
barriers. We can, therefore, consider the yield strength of a material oy as been made of two
components, one athermal crlr(long-range component) and the other cr,r (short-range component)
N. Bonora, PP. Milella / International Journal of lmpact Engineering 26 (2001) 53-64 59

thermally activated:
try = at,(structure) + asr(T,~) (15)

It is worth noting that the thermal component trsr in Eqn. (15) has been written as a function of
both temperature T and strain rate ~, since, as it will be described later, the strain rate has a
direct effect on the yield strength and it can reduce the actual capability of temperature to provide
thermal energy to overcome short range barriers. The most important short-range obstacle is the
so-called Peierls-Nabarro barrier. Statistical mechanics approach is usually used to quantify the
amplitude of these barriers but it does not provide any physical interpretation.
Milella ascribed the difference in mechanical behavior shown by BCC metals and alloys with
respect to FCC ones to the dislocations size and the corresponding concentration of Cottrell's
atmosphere. During plastic deformation, dislocations are displaced across the lattice and by
doing so also their atmosphere of interstitial atoms shall be moved. This cloud of atoms will not
move freely. As a result, a drag-force is generated that opposes the motion. The higher the
concentration mismatch between the Cottrel's atmosphere and the solute atoms in the lattice, the
greater this force and, therefore, the larger the component of the yield strength produced by this
effect. Solute atoms concentrations, in both the lattice and the atmosphere, depend on
temperature. Using the Wert expression to describe the concentration with temperature of solute
atoms in the lattice and the atmosphere as well, Milella obtained the following expression
relating the yield strength to the temperature:
B

t~y(T ) = C.e r (16)


where C and B are material constants. This linear dependence of the yield stress from l/T, on the
log-log plane, is confirmed for many metals, as given in Figure 2. Here, the slope of the linear
trend is given by the constant B. It has been found that, for a given material, the slope B changes
linearly with strain rate as given in Figure 3 for a A533B steel. It is worthwhile to note that all the
lines merge to the same point on the y-axis that can be considered the athermal component of the
yield strength. This lead to the following general equation of state:

where C, D and E are material constants that can be determined running three tests only
according to the procedure given by Milella [6]. Equation 17 indicates that the general expression
for the yield strength, as a sum of two components, given by Eqn. 15, is only a first
approximation. In fact, the term o~r may not depend directly on temperature, but it does depend
on strain rate and since the latter is influenced, as we have seen, by temperature it turns out that
also ok shall, indirectly, depend on temperature. Therefore, the athermal component of yield
strength is that on which temperature has no effect any more and is given by Eqn. 17 when the
term in brackets becomes equal to one.

UNIFIED C O N S T I T U T I V E M O D E L

The two models summarized in the previous sections take into account of strain rate,
temperature and ductile damage, separately. Since they both resulted successful in predicting the
evolution of the yield strength and damage for a wide class of metals and loading conditions, can
be used in a unified framework. The model proposed by Milella gives the evolution of the yield
strength as a function of the strain rate and temperature, pointing out for the first time a direct
influence of the temperature on the effect due to the strain rate. The modification of the entire
plastic flow curve is not addressed in the model. However, it is possible to assume that the entire
60 N. Bonora, P.P. Milella / International Journal of lmpact Engineering 26 (2001) 53-64

plastic flow curve is scaled with respect to the actual yield strength value similarly to the Johnson
and Cook [51] model.
7,8 18 Ni (1750)
7,6
10NiCrMoCo
7,4- 18 Ni (1260)
7,2'
7,0-
2 6,8:
c: 6,6-
6,4-
6,2-
6,0"
5,8- 1020 A ~uz
I I I I I

0,000 0,004 0,008 0,012 0,016


1/T (°K-l)
Fig. 2. Experimental results of Oy, solid and open points, and their trend, solid lines, in a In vs
1/T diagram for several classes of steels
6,6
• 530 s-1

6,4 1 s-1
0.1 s-1
6,2 10-3 s-1
v
13
t'-

6,0 '

5,8 '

I I I I I

0,000 0,001 0,002 0,003 0,004 0,005


1/T (l/K)
Fig.3. Trend of lnoy versus 1/T in A 533 B steel, for four different strain rates ~.

Thus, assuming a power law for the plastic flow curve, it is possible to write:

o ' ( e p , T , g ) - - o ' r ( T , i ) 14 --e, (18/


ay(T,e)

where n is the hardening exponent and K the hardening slope and try(T,~) is given by Eqn. (17).
If compared to the model proposed by Milella, strain rate effect predicted with the Johnson and
Cook model leads to very close results as shown in Figure 4 where the yield strength predicted by
the two models is given for a wide strain rate range. In Figure 5 the comparison between the
predicted flow curves is given for a strain rate of 1000 see-1 and two temperature, 25°C and
N. Bonora, P.P Milella / International Journal of lmpact Engineering 26 (2001) 53-64 61

200°C, respectively. In Figure 5 the Johnson and Cook parameters have been calibrated matching
the yield strength at low strain rate (i.e, 10E-04 sec -1 and Room Temperature) while the Milella
parameters come from experimental data fit.
1.5
i i
I

Mi,e,,al /
14 I i •
i
1.3. i

1
I
I
.~, 1 . 2
% i
r i

D J
1.1

1 . 0 - -
I
F Y !
f
i

0.9 ' '"""i . ,HH. . ,.HW


' ...... 'l .......
1E-4 1E-3 0.01 0.1 1 10 100 1000 10000 100000

strain rate ~ ~'sec "t]


Fig. 4. Yield strength evolution for a A533B steel with strain rate: comparison between Milella
and Johnson and Cook models.

CDM damage model, and porosity based damage models as well, do not incorporate effect
related to strain rates or temperature. Ping [52] theoretically analyzed the modification resulting
from dynamic loading onto the growth rate of cavities in a ductile matrix including inertial effect,
but without the support of experimental evidences. Experimental observations confirm that under
ductile failure conditions the micro-mechanism of failure is always given by the formation and
growth of microcavities no matter what both the temperature and the strain rate are. This means
that no additional damage dissipation processes are introduced as a result of strain rate or
temperature effect, thus damage formulation should not explicitly depend on these variables. On
the contrary, damage parameters, that represent material specific physical quantities, should show
some variability in association with them. In the model proposed by Bonora, four are the material
damage parameters needed: e.~, ef, Dcr and a. Since the threshold and failure strain are measure of
the capability of the material to plastically deform, a variability with load rate and temperature is
expected:
~,h = E,h (~,T)
ez = e/(g,T) (19)

Johnson and Cook [51] investigated the behavior of several classes of metals under severe
load rates and temperature. They analyzed the dependence of strain to failure with e and T.
While it came out that the role played by temperature was considerable, strain rate effect resulted
to be weak for all materials investigated. For instance, the increase of ductility for OFHC copper,
caused by temperature was of the order of 112% at 50% of homologous temperature while only
an increase of 20%, approximately, for a strain rate of 10000 sec l was found. The threshold
strain is controlled by the nature of the bond between the included particles and the ductile
matrix. Since the strain rates barely affect the strain to failure, it can be assumed that it has no
effect on threshold strain as well. In steels, voids usually start after particle breaking. When
temperature increases, the matrix softens and the stresses at the interface are expected to reduce
resulting in an increase of the threshold strain. The critical damage De, is related to the minimum
62 N. Bonora,P.P. Milella/ InternationalJournalof lmpactEngineering26 (2001)53-64
net resisting section that can still sustain load in the reference volume element, prior failure.
Once again, the experimental observations of Johnson and Cook confirm that the area at failure
in specimen tested with Hopkinson bar, is scarcely sensitive to the strain rate while temperature
has a higher effect. Recent measures on A533B steel performed by Bonora et al. [53] in the low
strain rate range using round notched bar in tension, supported this statement. From these
considerations, it is possible to express the modification induced by strain rate and temperature
onto damage parameters in the following form:

Ieth] Ver°ll A° •l IAt] 1


+ 1+ :
LDOrj[ Leo] Co j, LC,j )
where A0, AI, B0, BI, Co, Cl are material constants, T* is the homologous temperature, e0is a
reference strain rate, while e,h,er and
0 0
DO are the reference damage parameters under quasi-static
loading condition (low strain rate)• However, the constants in each set are expected to have
values very close to each others (i.e, A~Bo=C0) with the first set, that account for the strain rate
effect, globally much less than 1.0 (i.e. A0=B0=Co<0.002). In a first approximation, A0, Al, B0,
Bl, Co, C] can be reduced to two constant only, N (=A0=Bo=Co) and M (AI=B~=C]). If the strain
rate effect is neglected (N<0.002), equation (20) is reduced to:

[ ',,I

Do, J
r:ll(l + M. : )
L:d
(2l)

750
f
I
i
i
•i ~ l j ooq oe,eoeoe" ,;:;...aee

700 - ~ ~=1C~03~c 4 T=2~C

600-
i oeoOl -
"';"""
~, lOE-o<3sec T--200"C

550,,
~ •
500 . . . . ~ .......... ~-.-.- . . . . . . . . . . ~-~-----'-- . . . . .

450 - , " " .... ! A533B steel, T=25=C I


.o i
400 -.' I - - - reference 10E-O4 see "1 K = 1 0 0 n=0.21
j strain rate: 1 0 E + 0 3 sec 4
350 _" -!1 ,,, Milella C = 3 6 5 D=I 19 E=7.2
= • J&C Co=0.0295 A = 4 1 2 . 0 1 5 m=0.9:
I
J i f i i
0.0 0.2 0.4 0.6 0.8 1.0

True plastic strain


Fig. 5. Strain rate and temperature effect on plastic flow curve: comparison between Milella and
Johnson and Cook models.

The damage exponent ct gives the shape of the damage accumulation process as a function of
plastic strain. This value is identified from experimental damage measurements and possible
effect resulting from strain rate and temperature should be experimentally investigated. Bonora et
a/.[53] performed the fist attempt to evaluate possible load rate effect on ductile damage
N. Bonora, PP. Milella / International Journal of lmpact Engineering 26 (2001) 53-64 63

parameters in a A533B steel. The strain rates investigated were in the low range (<1.0 sec -I) but
they were sufficient to measure a clear effect on the flow curve and yield strength showing no
effect on damage parameters. Even though, it cannot be excluded a priori that the same behavior
is maintained at higher strain rates, these preliminary results represent a first encouraging support
to the proposed model. It has to be underlined that, at the moment, the possibility to perform
damage measurements under severe strain rates is made very difficult by the necessity to run
accurate strain controlled tests on which measure the progressive stiffness loss. Finally, the
model can be easily implemented in any commercial finite element code assigning a user defined
plastic flow curve as a function o f the two current variables T and strain rate. In theory, given
(e~, e, T ) for any gauss point of any element in the model, damage calculation requires to solve
the set of equations in (5-9). Neglecting the elastic strain contribution with respect to the plastic
one in both the micro and macro scale, the set of Eqn. (5-9) is reduced to the integration of Eqn.
(9), only.

CONCLUSIONS

The possibility to accurately investigate computationally the response of structural


components under dynamic condition depends on the accuracy of the material constitutive model
to incorporate strain rate and temperature effect. In this paper a new constitutive model with
damage for ductile metals has been developed. Here, strain rate and temperatures have been
taken into account in the global material elastic-plastic response using the innovative solid state
equation as proposed by Milella. The non-linear damage model, as proposed by Bonora, has been
used to describe progressive material softening and failure. The limited experimental data,
relative to ductile damage, available in the literature seems to confirm the model predictions. The
proposed model has the valuable feature to incorporate advanced modeling characteristics, such
as mutual strain rate and temperature effect on yield strength and nucleation and growth (NAG)
modeling, including stress triaxiality effect, without the complexity that usually characterizes
similar formulations.

Acknowledgments This work was funded by the U.S. Air Force contract n ° F61775-99-WE066.

REFERENCES

[1] Zukas J., et al., Impact Dynamics, Wiley, New York, reprinted by Krieger Pub. Co., Malabar, Florida, 1992.
[2] Gupta, A.D., and Misey, J.J, AIAA paper 80-0403. Proc. AIAA 1 ~ h Aerospace Science Meeting, 1980,
Pasadena, USA.
[3] Bonora, N., Ductile damage parameters identification and measurement. Journal of Strain Analysis, 1999, 34
no 6: 463-478.
[4] Bonora, N., A non-linear CDM model for ductile failure. Engineering Fracture Mechanics, 1997, 58:11-28.
[5] Lemaitre, J., A continuous damage mechanics model for ductile fracture. Journal of Engng. Mat. and Tech.,
1985, 107: 83-89.
[6] Milella P.P., Temperature and strain rate dependence of mechanical behavior of body-centered cubic structure
materials. Proc. TMS Fall Meeting '98, Chicago, Illinois, 11-15 Oct., 1998 to be published.
[7] McClintock, F.A., A criterion for ductile fracture by the growth of holes. J. Appl. Mech., 1968, 35:363-371
[8] Rice, J.R. and Tracy, D.M., On ductile enlargement of voids in triaxial stress fields. J. Mech. Phys. Solids,
1969, 17: 210-217.
[9] Seaman L., in W. Pilkey and Merrill, R.B. Shock and Vibration Computer Programs: Review and Sumaries,
1975, Shock & Vibration information Center, Washington, D.C.
[ 10] Gurson, A.L., Continuum theory of ductile rupture by void nucleation and growth: Part I- yield criteria and flow
rules for porous ductile media, J. Engn. Mat. Tech., 1977, 99: 2-15.
[1 l]Koplik, J., and Needleman, A., Void growth and coalescence in porous plastic solids. Int. Journal of Solids
Structures, 1988, 24, n°8: 835-853.
[12] Needleman, A., and Tvergaard, V., An analysis of ductile rupture in notched bars. Journal of Mechanics and
Physics of Solids, 1984, 32" 461.
64 N. Bonora, P P Milella / International Journal of lmpact Engineering 26 (2001) 53-64

[I 3] Brocks, W. and Bernauer, G., Determination of the Gurson parameters by numerical simulations. Proc. of the
2naGriffith Conference, 1995, Sheffield, U.K.
[14]Chaboche J.L., Anisotropic creep damage in the framework of the continuum damage mechanics, Nuclear
Engineering and Design, 1984, 79" 309-319.
[15]Kachanov, L.M., Introduction to continuum damage mechanics, 1986, Martinus,Nijhoff Publisher, Boston-
Dordrecht, 90-247-3319-7
[16]Chandrakanth, S., and Pandey, P.C., A New Ductile Damage Evolution Model. Int. J. of Fracture, 1993, 60:
R73-R76.
[17]Tai, H.W. and Yang, B.X., A new microvoid-damage model for ductile fracture, Engng. Frac. Mech., 1986, 25,
n ° 3:377-384
[ 18] Shockey, D.A., Curran, D.R., and DeCarli, P.S., Journal of Applied Physics, 1975, 46" 3766.
[19] Seaman, L., Curran, D.R., and Shockey, D.A., Journal of Applied Physics, 1976, 47: 4814.
[20] Erlich, D.C. et al., Ballistic Research Laboratory, 1980, ARBRL-CR-00416.
[21] Bonora, N., On the effect of triaxial state of stress on ductility using nonlinear CDM Model. Int. Journal of
Fracture, 1998, 88" 359-371.
[22] Bonora, N. and Newaz, M.G., Low cycle fatigue life estimation for ductile metals using a non-linear continuum
damage mechanics model. Int. J. Solid and Structures, 1997, 35:1881-1894.
[23] Lemaitre J., A course in damage mechanics, 1992, Springer Verlag, New York, pp. 44-46.
[24] Bonora, N., Cavallini, M., Iacoviello, F. and Marcbetti, M., Crack Initiation in AI-Li Alloy Using Continuum
Damage Mechanics in Localized Damage III Computer-Aided Assessment and Control, 1994, Eds.
M.H.Aliabadi, A. Carpinteri, S. Kalisky and D.J. Cartwright, Computational Mechanics Publication,
Southampton Boston, pp. 657-665.
[25] Zener C., Hollomon J.H., Effect of Strain Rate upon Plastic Flow. J. Appl. Ph.,1944, 15: 22-32.
[26] Manjoine M.J., Influence of rate of strain and temperature on yield stress of mild steels. J. Appl. Mech., 1944,
11-4:A211-A218.
[27] Cottrell A.H., Effects of solute atoms on the behavior of dislocations, Ph. Soc., 1948: 30.
[28] Cottrell A.H., Dislocations and Plastic Flow in Cristals, 1953, Oxford
[29] Cottrell A.H., Theory of brittle fracture in steels and similar metals, Trans. AIME, 1959, 212:192-203
[30] Cottrell A.H. and Bilby B.A., Dislocation theory of yielding and strain aging of iron, Ph. Soc., 1949, 62:49
[31]Cottrell A.H., Deformation of solids at high rates of strain, Proc. of Conf. on Prop. of Mat. at High Rates of
Strain, 1957:1-12.
[321Hollomon J. H., The mechanical equation of state, Trans. A1ME, 1947, 171: 535-545.
[33] Yokobori T., The Cottrell-Bilby theory of yielding of iron, Ph. Rev., 1953, 88: 1423.
[34] Seeger A., The temperature dependence of the critical shear stress and of work hardening of metal crystals, Phil.
Magaz., 1954, 45: 771.
[35]Seeger A., The generation of lattice defects by moving dislocations and its applications to temperature
dependence of the flow stress in FCC crystals, Phil Magaz,, 1955, 46:1194
[36] Campbell J.D., Yield delay time of mild steel, Trans. ASM, 1959.51:659
[37] Conrad, H., On the mechanism of yielding and flow in iron. J. Iron Steel Inst., 1961,198: 364.
[38]Conrad, H., Hayes W., Correlation of the thermal component of the yield stress of the BCC metals, 1962,
Aerosp. Corp.
[39] Conrad H., Wiedersich H., (1960)"Activation Energy for Deformation of Metals at Low Temperatures", Acta
Metall., 8, p. 128.
[40] Petch, N.J., The fracture of metals, Progr. in Metal Phys., 1964, 5:1
[41] Petch, N.J., and Helsop, J., The stress to move a free dislocation in alpha iron. Phil Magaz., 1956, 47:886.
[42]Johnson G.R. and Cook W.H., Proc. 7fh Inter. Symp. Ballistic, 1983, Am. Def. Prep. Org. (ADPA),
Netherlands: 1-7.
[43]Johnson G.R., Hoegfeldt J.M., Lindotm U.S., and Nagy A., ASMEJour. Eng. Mater. Tech., 1983, 105:42
[44] Klopp R.W., Clifton R.J., and Shawki T.G., Mech. Mater., 1985, 4" 633
[45] Hoge K.G., and Mukherjee A.K., J. Mater Sci., 1977, 12:1666
[46] Zerilli F.. J., Armstrong R.W., Acta Met. Mat., 1992, 40: 1803.
[47] Zerilli F. J., Armstrong R.W., J. Appl. Phys., 1987, 61: 1816.
[48]Zerilli F. J., Armstrong R.W., J. Appl. Phys., 1990, 68: 1580.
[49] Follansbee P.S., in Metallurgical Applications of Shock-Waves and High Strain Rate Phenomena, 1986, Eds.
L.E. Murr, K.P. Staudhammer, and M.A. Meyers, Dekker, New York, p.451
[50] Follansbee P.S., and Kocks U.F., Acta Metallurgica, 1988, 36: 81.
[51] Johnson, G.R. and Cook, W.H., Fracture characteristics of three metals subjected to various strains, strain rates,
temperatures and pressures. Engineering Fracture Mechanics, 1985, 21, n ° 1:31-48.
[52] Ping, Wang-Ze, Void growth and compaction relations for ductile porous materials under intense dynamic
general loading conditions, International Journal of Solids and Structures, 1994, 31, no. 15:2139-2150.
[53]Bonora, N., Milella, N., Gentile, D., Newaz, G. and Iacoviello, F. (2000), Ductile damage evolution under
different strain rate conditions, to be presented at ASME Winter Meeting, Orlando Nov. 2000.

You might also like