You are on page 1of 7

Mechanics of Materials 41 (2009) 982–988

Contents lists available at ScienceDirect

Mechanics of Materials
journal homepage: www.elsevier.com/locate/mechmat

A constitutive model for high strain rate deformation in FCC metals


based on irreversible thermodynamics
Mingxin Huang a,*, Pedro E.J. Rivera-Díaz-del-Castillo a, Olivier Bouaziz b,
Sybrand van der Zwaag a
a
Faculty of Aerospace Engineering, Delft University of Technology, Kluyverweg 1, 2629 HS, Delft, The Netherlands
b
ArcelorMittal Research, Voie Romaine-BP30320, 57283 Maizières-lès-Metz Cedex, France

a r t i c l e i n f o a b s t r a c t

Article history: The present work extends a recent model for plastic deformation of polycrystalline metals
Received 27 April 2008 based on irreversible thermodynamics. A general dislocation evolution equation is derived
Received in revised form 27 May 2009 for a wide range of strain rates. It is found that there is a transitional strain rate (103 s1)
over which the phonon drag effects play a dominant role in dislocation generation result-
ing in a significant raise in the dislocation density and flow stress. The model reduces to the
classical Kocks–Mecking model at low strain rates.
Ó 2009 Elsevier Ltd. All rights reserved.

1. Introduction at such high strain rate regime. However, Armstrong and


co-workers (2007) proposed that such flow stress increase
There are many engineering applications of metals is caused by the dislocation generation at the shock front,
related to high strain rate deformation such as high speed not by a retarding effect of dislocation drag.
machinery, armour systems, high speed transportation In order to understand the physical mechanism for such
vehicles and spacecraft. In order to optimize the perfor- high strain rate deformation, a large number of physically
mance of metals undergoing plastic deformation at such based constitutive models involving dislocations have
dynamic loading conditions, it is necessary to understand been proposed such as the Zerilli–Armstrong model (Zerilli
its physical mechanism, which is different with the coun- and Armstrong, 1987), the mechanical threshold stress
terpart at low strain rates. For instance, TEM micrographs model (Follansbee and Kocks, 1988), the model incorporat-
show that the dislocation structures are substantially dif- ing phonon drag effects (Nemat-Nasser et al., 2001;
ferent between low strain rates and high strain rates Nemat-Nasser and Yulong, 1998) and the model proposed
(Hirsch and Plesek, 2006). Moreover, when the strain rate by Meyers and co-workers (2002, 2003).
exceeds 103 s1, many investigators found experimen- The present work is an attempt to develop a unified
tally that the flow stress increases dramatically in many physically based constitutive model for both low and high
metals and alloys such as copper (Follansbee and Kocks, strain rate deformation in FCC metals. This formulation is
1988), tantalum (Kapoor and Nemat-Nasser, 2000), alu- derived by incorporating the viscous drag effects and by
minium (Lee and Chen, 2006; Lee et al., 2000) and stainless relating the entropy generation to the glide and accumula-
steels (Nemat-Nasser et al., 2001). Viscous drag effects on tion of dislocations. Such relationship is based on the the-
dislocation movement are thought by some researchers ory of irreversible thermodynamics which is widely
(Follansbee and Weertman, 1982; Nemat-Nasser and Yu- employed in chemical and mechanical engineering but less
long, 1998) as the main origin of the increase of flow stress commonly applied to physical metallurgy (Poliak and
Jonas, 1996). Poliak and Jonas (1996) employed irrevers-
ible thermodynamics to predict the initiation of dynamic
* Corresponding author. Current address: ArcelorMittal Research, Voie recrystallization for nickel and steels. Such process occurs
Romaine-BP30320, 57283 Maizières-lès-Metz Cedex, France. Tel.: +33 (0)
387704253; fax: +33 (0) 387704712.
when the entropy production rate reaches a minimum
E-mail address: mingxin.huang@arcelormittal.com (M. Huang). value. Ghoniem et al. (2000) proposed a model based on

0167-6636/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mechmat.2009.05.007
M. Huang et al. / Mechanics of Materials 41 (2009) 982–988 983

irreversible thermodynamics to illustrate the motion of where qm is the average mobile dislocation density per
three dimensional interacting dislocation loops. The unit volume and v is the average velocity of mobile dislo-
authors have successfully employed the theory of irrevers- cations. The relation between strain rate and mobile dislo-
ible thermodynamics to model plastic deformation in met- cation density is expressed by the well known Orowan
als. For instance, the steady state deformation of FCC relationship (Haasen, 1996) as
metals over a wide range of temperatures and strain rates
c_ ¼ qm vb ð6Þ
has been quantitatively described (Huang et al., 2007). The
stress–strain behaviour of single crystals and coarse where c_ is the shear strain rate and b is the magnitude of
grained polycrystals of FCC metals at low strain rates has the Burgers vector. X3 is the general force for dislocation
also been modelled (Huang et al., 2008a,b). Moreover, the glide expressed as
effect of grain size has been incorporated to predict the
ðs  sf Þb
characteristic strength and ductility of ultrafine grained al- X3 ¼ ð7Þ
T
loys (Huang et al., 2008a,b). Those results have been ob-
tained with a very limited, but yet physically rooted, where s is the shear flow stress and sf is the friction stress
number of parameters, adding confidence to the possibility for dislocation glide. (s  sf) represents the net stress
of describing plastic deformation employing irreversible acting on mobile dislocations. Therefore, inserting Eqs.
thermodynamics theory. (2)–(7) into Eq. (1) the entropy generation rate can be
rewritten as
 
2. Theory di S 1 dqþ dq
¼ ðs  sf Þc_ þ E þE : ð8Þ
dt T dt dt
2.1. Entropic analysis
An increase of the average dislocation density dq during
During plastic deformation in metals, three irreversible the time interval dt is the sum of the average dislocation
processes are generally considered to take place (Hirth and generation and annihilation:
Lothe, 1982; Huang et al., 2008a,b, 2007): (1) dislocation dq dqþ dq
generation, (2) dislocation annihilation (dynamic recovery) ¼  : ð9Þ
dt dt dt
and (3) dislocation glide. According to the theory of irre-
Inserting Eq. (9) into Eq. (8), it can be rewritten as
versible thermodynamics, the entropy generation rate
 
can be expressed as the sum of the products of generalised di S 1 dq dq
forces and their corresponding fluxes in the irreversible
¼ ðs  sf Þc_ þ E þ 2E : ð10Þ
dt T dt dt
processes (Prigogine, 1961). Therefore, the average entropy
generation rate per unit volume due to the three irrevers- The entropy generation rate is proportional to the en-
ible processes is expressed as ergy dissipation rate and can therefore be expressed in an-
other form (Huang et al., 2008a,b, 2007; Prigogine, 1961):
di S
¼ J1 X 1 þ J2 X 2 þ J3 X 3 : di S 1 dW diss
dt |ffl{zffl} |ffl{zffl} |ffl{zffl} ¼ ð11Þ
dislocation generation dislocation annihilation dislocation glide dt T dt
ð1Þ where Wdiss is the dissipative energy per unit volume
transformed into heat during plastic deformation. Invoking
The general flux J1 represents the average dislocation
the energy conservation law, the dissipative energy can be
generation rate per unit volume
written as
dqþ
J1 ¼ ð2Þ dW diss dW in dU s
dt ¼  ð12Þ
dt dt dt
where dq+ is the length of dislocations per unit volume
where dWin is the input energy by loading during the time
generated during a time interval dt. The general force for
interval dt and is expressed as
dislocation generation or annihilation can be expressed as
E dW in ¼ s dc ð13Þ
Xi ¼ ð3Þ
T where c is the shear strain. dUs is the increase in stored en-
where i = 1 for generation and i = 2 for annihilation. E is the ergy due to the dislocation density increase (the sum of
potential energy of the dislocation per unit length and T is generation and annihilation) during dt and is expressed as
the absolute temperature. J2 corresponds to the average dU s ¼ E dq: ð14Þ
dislocation annihilation rate per unit volume:
Therefore, inserting Eqs. (12)–(14) into Eq. (11) yields
dq  
J2 ¼ ð4Þ di S 1 dq
dt ¼ sc_  E : ð15Þ
dt T dt
where dq is the length of dislocations per unit volume
annihilated during a time interval dt. J3 is the general flux Eqs. (10) and (15) are two different expressions to de-
corresponding to dislocation glide expressed as scribe the entropy generation. They should be equal, offer-
ing an expression for the evolution of the average
J 3 ¼ qm v ð5Þ
dislocation density:
984 M. Huang et al. / Mechanics of Materials 41 (2009) 982–988

dq sf c_ dq The cross-slip length L and its attempt frequency vcs


¼  : ð16Þ
dt 2E dt can, respectively, be rewritten as:
16pvV
2.2. Dislocation annihilation
L¼ ð20Þ
lb3
At high temperature, the dislocation annihilation mech- and
anism is mainly dislocation climb caused by the diffusion
lb 4
of vacancies along the dislocation line (Argon and Moffatt, vcs ¼ vD : ð21Þ
16pvV
1981). However, at low temperature, the dislocation anni-
hilation mechanism is mainly due to the cross-slip of dislo- Pcs is the probability that the cross-slip attempt is suc-
cations (Puschl, 2002). The current work deals with the cessful and may be expressed as (Nix et al., 1985)
deformation of FCC metals at room temperature such that  
DG
the annihilation mechanism considered here is the cross- Pcs ¼ exp  ð22Þ
slip of dislocations. In FCC metals, the most common kT
cross-slip mechanism is compact cross-slip (i.e. Friedel–Es- where k is the Boltzmann constant and DG is the activation
caig mechanism (Bonneville and Escaig, 1979)) shown energy for the cross-slip. Therefore, the dislocation annihi-
schematically in Fig. 1: the cross-slip segment L immedi- lation rate (Eq. (17)) can be rewritten as

ately re-dissociates in the cross-slip plane (e.g. ð1 1 1Þ)  
which is also a compact plane. More details about the mod- dq lb4 DG
¼ vD exp  q: ð23Þ
elling of the cross-slip mechanism may be found in the re- dt 8pvV kT
view paper by Puschl (2002). The dislocation annihilation
The activation energy for cross-slip effectively increases
process due to the dislocation cross-slip can be expressed
with increasing temperature but decreases with strain rate
as (Nix et al., 1985)
(Kocks and Mecking, 2003); it may be expressed as
dq
¼ N2Lvcs Pcs ð17Þ vcs kT
dt DG ¼ AkT ln  sV ð24Þ
c_ lb3
where N is the number of sites per unit volume where such
cross-slip events can occur and is approximately N = q/L. where A is a constant in the order of 1 for a given temper-
The factor 2 accounts for two screw dislocation segments ature. The term (kT/lb3)sV accounts for the reduction in
with opposite sign of Burgers vector which annihilate activation energy due to the applied stress; this is consis-
simultaneously at one single annihilation event. vcs is the tent with the well known Friedel–Escaig model (Bonneville
attempt frequency for such cross-slip event and is taken and Escaig, 1979; Puschl, 2002). Inserting Eq. (21) into Eq.
to be vcs = (b/L)vD (Nix et al., 1985) where vD(1013 s1) is (24) offers
the Debye frequency. The cross-slip length L is related to !
lb4 vD kT
its activation volume V (Fig. 1): DG ¼ AkT ln  3 sV: ð25Þ
16pvV c_ lb
V
Lffi ð18Þ Inserting Eq. (25) into Eq. (23), it can be rewritten as
db
" ! #
where d is the stacking fault width of screw dislocations dq lb4 lb4 vD sV
which is determined by the stacking fault energy v (Puschl, ¼ vD exp A ln þ 3 q: ð26Þ
dt 8pvV 16pvV c_ lb
2002):
Referring to Eq. (16), the friction stress sf for pure FCC
lb 2 metals may be expressed as
d¼ : ð19Þ
16pv pffiffiffiffi
sf ¼ blb q þ sd ð27Þ
where the first term in the right hand side represents the
athermal friction stress due to other existing dislocations
(Haasen, 1996). b is a constant accounting for the interac-
tions between dislocations and l is the shear modulus. sd
is the friction stress due to viscous drag effects, which may
be caused by phonon and electron scattering (Hirth and
Lothe, 1982). At moderate temperatures, phonon effects
dominate over electron effects (Hirth and Lothe, 1982). The
contribution of electron effects can only be seen at very
low temperatures (about 10% of Debye temperature) (Hirth
and Lothe, 1982). In the current work, such low temperature
deformation is not considered so that electron effects are ne-
glected. Therefore, sd is taken as (Nemat-Nasser et al., 2001)
Fig. 1. Schematic illustration of the cross-slip mechanism (Friedel–Escaig
 L is
model): a segment L re-dissociates in the cross-slip plane (e.g. ð1 1 1Þ).
  
c_
the cross-slip length and d is the stacking fault width (Bonneville and sd ¼ s0 1  exp  _ ð28Þ
Escaig, 1979; Puschl, 2002).
c0
M. Huang et al. / Mechanics of Materials 41 (2009) 982–988 985

where s0 is a material related constant which can be mea-


sured at a very high strain rate. For instance, it is found that
s0 = 140 MPa for stainless steels (Nemat-Nasser et al.,
2001). c_ 0 represents an effective damping coefficient
2
affecting the dislocation glide (c_ 0 ¼ qm b sy =B where sy is
the yield stress at high temperature and B is the phonon
drag coefficient). In the present work s0 and c_ 0 are fitting
parameters to be chosen according to the experimental
data. The physical explanation of sd is presented in the lit-
erature (Nemat-Nasser et al., 2001). Therefore, inserting
Eqs. (23), (27), and (28) into Eq. (16), substituting
c_ ¼ dc=dt, and approximating E = lb2/2 (Humphreys and
Hatherly, 2004), the dislocation evolution Eq. (16) can be
rewritten as
  
dq s0 c_ b pffiffiffiffi lb4 vD
¼ 2 1  exp  þ q 
dc lb c_ 0 b 8pvV c_ Fig. 2. Model predictions (solid lines) and experimental data (symbols)
" ! # for pure copper deformed at 295 K and at three strain rates 1.5  102,
lb4 vD sV 8.5  103 and 6.4  105 s1 (two experimental data sets for the highest
 exp A ln þ 3 q: ð29Þ strain rate).
16pvV c_ lb
The first term in the right hand side of the evolution Eq. dW diss sdc  Edq
(29) stands for the dislocation generation caused by pho- dT ¼ ¼ ð31Þ
mq C V mq C V
non drag effects; the second term accounts for dislocation
generation due to existing dislocations, which serve as where mq is the mass density and CV is the heat capacity.
obstacles for dislocation glide; the last term incorporates Here it is assumed that the heat flow in the bulk is fast en-
the effects of dynamic recovery due to the cross-slip of ough such that no temperature gradients develop in the
screw dislocations. bulk. The temperature increase will decrease the shear
modulus and increase the cross-slip rate.
2.3. Stress–strain behaviour
3. Application
By solving Eq. (29) with an initial dislocation density q0
before deformation, the evolution of the average disloca- In Fig. 2, the model predictions are compared to exper-
tion density can be obtained. Then, the stress–strain imental data in oxygen free electronic (OFE) copper
behaviour can be predicted by utilizing the relationship (0.9999 Cu) with a well annealed equiaxed grain structure
between the shear flow stress and the average dislocation and an average grain size of 40 lm deformed at 295 K at
density (Bailey and Hirsch, 1960; Kocks and Mecking, different strain rates. The parameters used in the simula-
2003; Kuhlmann-Wilsdorf, 1999): tion are shown in Table 1. Fig. 2 shows that the model pre-
pffiffiffiffi dictions are in good agreement with experimental data by
s ¼ alb q ð30Þ
Follansbee and Kocks (1988) considering the results cover
where a is a constant describing the interactions between a wide range of strain rates from 102 to 105 s1. It is worth
dislocations. to note that the stress and strain in Fig. 2 are true normal
The quasi-static deformation is assumed to be isother- stress r (r = Ms) and normal strain e (e = c/M) where M is
mal, while for high strain rate deformation (say strain rates the Taylor factor.
larger than 102 s1) it is assumed to be adiabatic. The tem- Fig. 3 shows the flow stress and average dislocation
perature increase dT of the bulk during the strain interval density at e = 0.15 for wide ranges of strain rates. The
dc for adiabatic deformation is related to the dissipative experimental data are obtained from Follansbee and Kocks
energy dWdiss and is approximated as (1988) and the strain rate is the normal strain rate e_ ¼ c_ =M.

Table 1
Values of parameters used in the model.
mq (kg/m3) CV (J/(kg K)) l (GPa) b (nm) V (b3) a
4
8940 386 47.4  exp(3.97  10 T) 0.256 300 0.4
Molinari and Molinari and Mecking et al. (1986) Nes (1997) Puschl (2002) Huang et al.
Ravichandran (2005) Ravichandran (2005) (2007)

M q0 (m2) b v (J m2) s0 (MPa) c_ 0 ðs1 Þ A


3.06 5  1012 0.024 47  103 10 5  105 0.98
Humphreys and Humphreys and Kocks and Nes (1997)
Hatherly (2004) Hatherly (2004) Mecking (2003)
986 M. Huang et al. / Mechanics of Materials 41 (2009) 982–988

The initial dislocation density q0 = 5  1012 m2 is a rea-


sonable value for the annealed samples (Huang et al.,
2008a,b; Humphreys and Hatherly, 2004). The activation
volume V = 300b3 is the experimentally measured value
for pure Cu (Puschl, 2002). One may argue that the activa-
tion volume may slightly change with the progress of
deformation. However, as a first order approximation, the
present work assumes that the activation volume is con-
stant during deformation. s0 = 10 MPa is one order magni-
tude lower than the one employed for stainless steels
(Nemat-Nasser et al., 2001). Referring to Eq. (29), the value
of b determines the rate of dislocation generation due to
existing dislocations. In the present work, b is set as
0.024 which is close to the value used in the Kocks–Mec-
king model (Kocks and Mecking, 2003). s0 and c_ 0 are the
Fig. 3. Flow stress and dislocation density as a function of strain rate at
parameters describing the dislocation movement at high
the strain e = 0.15 for OFE copper deformed at 295 K.
strain rates and related to phonon drag effects. Both are fit-
ted to the experimental data in the current work. Another
fitting parameter related to the activation energy is A
It shows that there is a transitional strain rate (103 s1) which determines the slope of the curve at low strain rates
over which the phonon drag effects play a dominant role in Fig. 3. c_ 0 affects the transitional strain rate over which
in dislocation accumulation (Eq. (29)) resulting in an the dislocation density increases dramatically (Fig. 3). s0
important raise in the average dislocation density and flow determines the magnitude of the dislocation density at
stress with strain rate. the high strain rate regime (Fig. 3). In conclusion, all
parameter values are taken from literature except for s0,
4. Discussion c_ 0 and A which are obtained by fitting.
For low stacking fault energy metals such as Cu, the
The phonon drag becomes important when the disloca- twinning may also contribute to the plastic deformation
tion velocity goes beyond the range 103 or 102  Ct (Hir- at high strain rate (Meyers et al., 1995, 2002; Petit and
th and Lothe, 1982), where Ct is the shear wave velocity. For Dequiedt, 2006). However, as a first order approximation,
pure copper C Cu 3
t ¼ 2:9  10 m s
1
(De Hosson et al., 2001). the present work assumes that the dislocation glide is
The velocity at which phonon drag becomes prominent de- the main mechanism for plastic deformation and the con-
pends on the metallic system. In the present work, a critical tribution of twinning is neglected.
dislocation speed of 102  C Cut is assumed. The correspond- The current work predicts that the dislocation genera-
ing strain rate for such dislocation velocity can be calculated tion and flow stress saturate when the strain rate is larger
from e_ ¼ c_ =M ¼ qm vb=M ¼ qm ð102 C Cu t Þb=M. Assuming a than 106 s1 (Fig. 3). This prediction does not agree with
mobile dislocation density qm = 5  1011 m2 (Follansbee some recently published works. For instance, Meyers and
and Weertman, 1982), e_ is found to be 103 s1. Beside the co-workers (2003) and Armstrong and co-workers (2007)
experimental evidence, this result could be seen as another found that the flow stress of pure Cu still increases with
proof to support what the current model predicts: the tran- strain rate when the strain rate is higher than 107 s1.
sitional strain rate is 103 s1 beyond which the phonon Discrete dislocation dynamics simulations (Shehadeh
drag effect on plasticity appears (Fig. 3). Experimental et al., 2006) and experiments (Meyers et al., 2003) also
observations confirmed that the dislocation generation at show that the dislocation density can still increase even
very high strain rates increases significantly in shock-de- when it is higher than 1017 m2. However, in our model,
formed copper (Meyers et al., 2003; Murr, 1981). This tran- the dislocation density is approaches an asymptotic value
sitional strain rate is also found experimentally in of about 1.5  1015 m2. This discrepancy is caused by
tantalum (Kapoor and Nemat-Nasser, 2000), aluminium the use of the relationship between strain rate and the fric-
alloys (Lee and Chen, 2006; Lee et al., 2000) and Al-6XN tion stress induced by the phonon drag effect (Eq. (28)).
stainless steel (Nemat-Nasser et al., 2001). For aluminium This relation is obtained from the literature (Nemat-Nasser
alloys deformed at strain rates ranging from 103 to et al., 2001) and predicts that the friction stress caused by
6  103 s1, the average dislocation density measured by the phonon drag is saturated when the strain rate is higher
TEM is also found to increase rapidly with strain rate (Lee than a critical value, which in turn results the saturation of
and Chen, 2006). Such increase is comparable to the one pre- dislocation density and the flow stress at a very high strain
dicted by the current model for pure copper (Fig. 3). Three rate. This discrepancy could be circumvented by choosing a
dimensional dislocation dynamics (DD) simulations also re- model incorporating the relation between the strain rate
ported that the dislocation density increases dramatically at and friction stress induced by the phonon drag (De Hosson
high strain rates for shock compression in copper single et al., 2001). Such relationship is expressed as sd = Bv/b,
crystals (Shehadeh et al., 2005). where the phonon drag coefficient B is introduced by the
A detailed analysis of the parameters employed in the so called phonon wind and fluttering mechanisms, i.e.
model (Table 1) is required. The physical parameters mq, B ¼ fB0wind þ B0flutter =½1  ðv=C t Þ2 gT=hD where hD is the
CV, l, b, a, M and v are taken directly from the literature. Debye temperature, and B0wind and B0flutter account for the
M. Huang et al. / Mechanics of Materials 41 (2009) 982–988 987

phonon wind and fluttering mechanism at the Debby tem- the dislocation density and flow stress. The model
perature, respectively, when dislocation velocity ap- reduces to the classical Kocks–Mecking model at low
proaches 0. Thus, incorporating the relation between the strain rates.
dislocation velocity and the strain rate (Eq. (6), Orowan
relationship), the friction stress sd can be expressed as Acknowledgements
" #
0 B0flutter T c_
sd ¼ Bwind þ 2 h 2
: ð32Þ The authors employed at the TU Delft are grateful to
1  ðc=bqm C t Þ
_ D b q
m ArcelorMittal Research for financial support.
In this expression, sd always increases with strain rate.
Thus, the discrepancy (saturation of dislocation density References
and flow stress) in the current model could be solved. This
Argon, A.S., Moffatt, W.C., 1981. Climb of extended edge dislocations. Acta
solution will be explored elsewhere. Metall. 29, 293–299.
The current model predicts that the cause of the flow Armstrong, R.W., Arnold, W., Zerilli, F.J., 2007. Dislocation mechanics of
stress upturn at high strain rate is due to the increase of shock-induced plasticity. Metall. Mater. Trans. A 38, 2605–2610.
Bailey, J.E., Hirsch, P.B., 1960. The dislocation distribution, flow stress, and
dislocation density. Moreover, such increase of dislocation stored energy in cold-worked polycrystalline silver. Philos. Mag. 5,
density is induced by a new dislocation generation mech- 485–497.
anism: new dislocations are generated due to the interac- Bonneville, J., Escaig, B., 1979. Cross-slipping process and the stress-
orientation dependence in pure copper. Acta Metall. 27, 1477–1486.
tions between fast moving dislocations and phonons De Hosson, J.T.M., Roos, A., Metselaar, E.D., 2001. Temperature rise due to
(Eq. (29)). Such mechanism may be related to the energy fast-moving dislocations. Philos. Mag. A 81, 1099–1120.
dissipated by the high speed movement of dislocations. Follansbee, P.S., Kocks, U.F., 1988. A constitutive description of the
deformation of copper based on the use of the mechanical threshold
This amount of dissipative energy may be large enough stress as an internal state variable. Acta Metall. 36, 81–93.
to create new dislocations in the neighbourhood of the Follansbee, P.S., Weertman, J., 1982. On the question of flow stress at high
moving dislocations. For instance, molecular dynamics strain rates controlled by dislocation viscous flow. Mech. Mater. 1,
345–350.
simulations have confirmed that new dislocations can be
Ghoniem, N.M., Tong, S.H., Sun, L.Z., 2000. Parametric dislocation
generated around a moving dislocation when its velocity dynamics: a thermodynamics-based approach to investigations of
approaches 70% of the velocity of sound (Koizumi et al., mesoscopic plastic deformation. Phys. Rev. B 61, 913–927.
Haasen, P., 1996. Physical Metallurgy, third ed. Cambridge University
2002).
Press, Cambridge.
Different to the current model, Armstrong and co-work- Hirsch, E., Plesek, J., 2006. A theoretical analysis of experimental results of
ers (2007) proposed that the flow stress upturn at high shock wave loading of OFE copper relating the observed internal
strain rate is due to the generation of dislocations at the structure to the deformation mechanism. Int. J. Impact Eng. 32, 1339–
1356.
shock front. Such generation of dislocations is thermally Hirth, J.P., Lothe, J., 1982. Theory of Dislocations. John Wiley & Sons, New
activated. Phonon effects do not play any role in the flow York.
stress upturn at high strain rate. Despite the difference be- Huang, M., Rivera-Díaz-del-Castillo, P.E.J., Bouaziz, O., van der Zwaag, S., 2008a.
Predicting the strength and ductility of ultrafine grained interstitial free
tween the mechanisms, both Armstrong’s model and the steels using irreversible thermodynamics. In: Proceeding of New
current model can reproduce the experimental results. Developments on Metallurgy and Applications of High Strength Steels,
Thus, the real mechanism for the origin of the flow stress Buenos Aires, Argentina, 2008, pp. 805–812 (CD-ROM).
Huang, M., Rivera-Díaz-del-Castillo, P.E.J., Bouaziz, O., van der Zwaag, S.,
upturn at high strain rate is still an open question. The real 2008b. Irreversible thermodynamics modelling of plastic deformation
picture may be more complex than the models of metals. Mater. Sci. Technol. 24, 495–500.
descriptions. Huang, M., Rivera-Díaz-del-Castillo, P.E.J., van der Zwaag, S., 2007.
Modelling steady state deformation of fcc metals by non-
At low strain rates, the first term in the right hand side
equilibrium thermodynamics. Mater. Sci. Technol. 23, 1105–1108.
of Eq. (29) vanishes. Therefore, the current model reduces Humphreys, F.J., Hatherly, M., 2004. Recrystallization and Related
to the classical Kocks–Mecking model (Kocks and Mecking, Annealing Phenomena, second ed. Elsevier, Oxford.
Kapoor, R., Nemat-Nasser, S., 2000. Comparison between high and low
2003) at low strain rates. In other words, the current irre-
strain-rate deformation of tantalum. Metall. Mater. Trans. A 31, 815–
versible thermodynamics model may provide a thermody- 823.
namics basis for the Kocks–Mecking model. At high strain Kocks, U.F., Mecking, H., 2003. Physics and phenomenology of strain
rates, it naturally incorporates the influence of phonon hardening: the FCC case. Prog. Mater. Sci. 48, 171–273.
Koizumi, H., Kirchner, H.O.K., Suzuki, T., 2002. Lattice wave emission from
drag effects on the accumulation of dislocations. The low a moving dislocation. Phys. Rev. B 65, 214104.
and high strain rate regimes are physically integrated into Kuhlmann-Wilsdorf, D., 1999. The theory of dislocation-based crystal
a single formulation. plasticity. Philos. Mag. A 79, 955–1008.
Lee, W.-S., Chen, T.-H., 2006. Rate-dependent deformation and dislocation
substructure of Al–Sc alloy. Scr. Mater. 54, 1463–1468.
5. Summary Lee, W.-S., Shyu, J.-C., Chiou, S.-T., 2000. Effect of strain rate on impact
response and dislocation substructure of 6061-T6 aluminum alloy.
Scr. Mater. 42, 51–56.
Based on the theory of irreversible thermodynamics, a Mecking, H., Nicklas, B., Zarubova, N., Kocks, U.F., 1986. A ‘‘universal”
physically based constitutive model with a very limited, temperature scale for plastic flow. Acta Metall. 34, 527–535.
but yet physically rooted, number of parameters, has been Meyers, M.A., Andrade, U.R., Chokshi, A.H., 1995. The effect of grain size
on the high-strain, high-strain-rate behavior of copper. Metall. Mater.
proposed to describe the evolution of the average disloca- Trans. A 26, 2881–2893.
tion density in copper over a wide range strain rates. It is Meyers, M.A., Benson, D.J., Vohringer, O., Kad, B.K., Xue, Q., Fu, H.H., 2002.
found that there is a transitional strain rate (103 s1) Constitutive description of dynamic deformation: physically-based
mechanisms. Mater. Sci. Eng. A 322, 194–216.
over which the phonon drag effects play a dominant role Meyers, M.A., Gregori, F., Kad, B.K., Schneider, M.S., Kalantar, D.H.,
in dislocation generation resulting in a significant raise in Remington, B.A., Ravichandran, G., Boehly, T., Wark, J.S., 2003. Laser-
988 M. Huang et al. / Mechanics of Materials 41 (2009) 982–988

induced shock compression of monocrystalline copper: Poliak, E.I., Jonas, J.J., 1996. A one-parameter approach to determining the
characterization and analysis. Acta Mater. 51, 1211–1228. critical conditions for the initiation of dynamic recrystallization. Acta
Molinari, A., Ravichandran, G., 2005. Constitutive modeling of high-strain- Mater. 44, 127–136.
rate deformation in metals based on the evolution of an effective Prigogine, I., 1961. Introduction to Thermodynamics of Irreversible
microstructural length. Mech. Mater. 37, 737–752. Process. John Wiley & Sons, New York.
Murr, L.E., 1981. In: Meyers, M.A., Murr, L.E. (Eds.), Shock-Wave and High- Puschl, W., 2002. Models for dislocation cross-slip in close-packed
Strain-Rate Phenomena in Metals. Plenum Press, New York, p. 607. crystal structures: a critical review. Prog. Mater. Sci. 47,
Nemat-Nasser, S., Guo, W.-G., Kihl, D.P., 2001. Thermomechanical 415–461.
response of AL-6XN stainless steel over a wide range of strain rates Shehadeh, M.A., Bringa, E.M., Zbib, H.M., McNaney, J.M., Remington, B.A.,
and temperatures. J. Mech. Phys. Solids 49, 1823–1846. 2006. Simulation of shock-induced plasticity including homogeneous
Nemat-Nasser, S., Yulong, L., 1998. Flow stress of F.C.C. polycrystals with and heterogeneous dislocation nucleations. Appl. Phys. Lett. 89,
application to OFHC Cu. Acta Mater. 46, 565–577. 171913–171918.
Nes, E., 1997. Modelling of work hardening and stress saturation in FCC Shehadeh, M.A., Zbib, H.M., Diaz de la Rubia, T., 2005. Multiscale
metals. Prog. Mater. Sci. 41, 129–193. dislocation dynamics simulations of shock compression in copper
Nix, W.D., Gibeling, J.C., Hughes, D.A., 1985. Time-dependent deformation single crystal. Int. J. Plast. 21, 2369–2390.
of metals. Metall. Trans. A 16, 2215–2226. Zerilli, F.J., Armstrong, R.W., 1987. Dislocation-mechanics-based
Petit, J., Dequiedt, J.L., 2006. Constitutive relations for copper under shock constitutive relations for material dynamics calculations. J. Appl.
wave loading: twinning activation. Mech. Mater. 38, 173–185. Phys. 61, 1816–1825.

You might also like