2004-A Conventional-Strain Gradient Plasticity Theory
2004-A Conventional-Strain Gradient Plasticity Theory
2004-A Conventional-Strain Gradient Plasticity Theory
www.elsevier.com/locate/ijplas
Abstract
There exist two frameworks of strain gradient plasticity theories to model size effects
observed at the micron and sub-micron scales in experiments. The first framework involves
the higher-order stress and therefore requires extra boundary conditions, such as the theory of
mechanism-based strain gradient (MSG) plasticity [J Mech Phys Solids 47 (1999) 1239;
J Mech Phys Solids 48 (2000) 99; J Mater Res 15 (2000) 1786] established from the Taylor
dislocation model. The other framework does not involve the higher-order stress, and the
strain gradient effect come into play via the incremental plastic moduli. A conventional theory
of mechanism-based strain gradient plasticity is established in this paper. It is also based on
the Taylor dislocation model, but it does not involve the higher-order stress and therefore falls
into the second strain gradient plasticity framework that preserves the structure of conven-
tional plasticity theories. The plastic strain gradient appears only in the constitutive model,
and the equilibrium equations and boundary conditions are the same as the conventional
continuum theories. It is shown that the difference between this theory and the higher-order
MSG plasticity theory based on the same dislocation model is only significant within a thin
boundary layer of the solid.
# 2003 Elsevier Ltd. All rights reserved.
Keywords: Strain gradient plasticity; Dislocation model; Conventional theory
0749-6419/$ - see front matter # 2003 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijplas.2003.08.002
754 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
1. Introduction
Recent experiments have repeatedly shown that metallic materials display strong
size effect at the micron and sub-micron scales. For example, the micro-indentation
hardness of metallic materials (e.g., aluminum, copper, silver, tungsten) increases by
a factor of 2–3 as the indentation depth decreases to microns and sub-microns (e.g.,
Nix, 1989, 1997; De Guzman et al., 1993; Stelmashenko et al., 1993; Atkinson, 1995;
Ma and Clarke, 1995; Poole et al., 1996; McElhaney et al., 1998; Suresh et al., 1999;
Saha et al., 2001; Tymiak et al., 2001; Swadener et al., 2002). Plastic work hardening
increases significantly in micro-torsion tests of thin copper wires as the wire diameter
decreases from 170 to 12 mm (Fleck et al., 1994). Similar increase has been observed in
micro-bend tests of micron-thick nickel foils (Stolken and Evans, 1998) and sub-micron
thick aluminum beams (Haque and Saif, 2003), in aluminum matrix reinforced by
micron-size silicon carbide particles (Lloyd, 1994), and in a micro-electro-mechanical
system (MEMS) called digital micromirror device (Douglass, 1998). These size
effects have been attributed to geometrically necessary dislocations associated with
nonuniform plastic deformation. Discrete dislocation simulations have confirmed
the micron-scale size effect in metallic materials (e.g., Cleveringa et al., 1997, 1998,
1999a,b).
The constitutive models of conventional plasticity theories possess no intrinsic
material lengths, therefore cannot explain the above size-dependent material beha-
vior at the micron and sub-micron scales. There are however many dislocations at
the micron scale (around 30 in each direction for a dislocation density of 1015 m2)
such that their collective behavior on plastic work hardening of materials should be
characterized by a continuum (but not conventional) plasticity theory. For this
reason, strain gradient plasticity theories have been developed from the notion of
geometrically necessary dislocations in the dislocation theory (Nye, 1953; Cottrell,
1964; Ashby, 1970; Arsenlis and Parks, 1999; Gurtin, 2000), though there are early
phenomenological theories of strain gradient plasticity (e.g., Aifantis, 1984; Lasry
and Belytschko, 1988; Zbib and Aifantis, 1988; Mühlhaus and Aifantis, 1991; de
Borst and Mühlhaus, 1992; Aifantis, 1992; Sluys et al., 1993; Zhu and Zbib, 1995,
1997). The strain gradient plasticity theories are intended for applications to mate-
rials and structures whose dimension controlling plastic deformation falls roughly
within the range from 1=10 to 10 mm, such as in micro-components and MEMS,
micro-electronic packages, micro-machining, thin films, and composite materials.
There are two frameworks of strain gradient plasticity theories. The first one is
Mindlin’s (1964, 1965) framework of higher-order continuum theories, and it
involves the higher-order stress as the work conjugate of strain gradient. The order
of equilibrium equations are higher than that in the conventional continuum
theories, therefore additional boundary conditions are required. Examples in this
class include Fleck and Hutchinson (1993, 1997, 2001), Fleck et al. (1994), Gao et al.
(1999a,b), Chen and Wang (2000, 2002), Huang et al. (2000a,b), Qiu et al. (2001,
2003), Gurtin (2000, 2002), Hwang et al. (2002, 2003a), Mariano et al. (2002), Wang
et al. (2003). The other framework of strain gradient plasticity theories does not
involve the higher-order stress, and requires no additional boundary conditions. The
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 755
plastic strain gradient (or inverse elastic strain gradient) comes into play through the
incremental plastic modulus. Examples in this class include Acharya and Bassani
(2000), Acharya and Beaudoin (2000), Bassani (2001), Beaudoin and Acharya
(2001), Evers et al. (2002), and Dai and Parks (2003).
From dimensional considerations, intrinsic material lengths have been introduced
to scale with strain gradients in both classes of strain gradient plasticity theories.
Based on the Taylor dislocation model (Taylor, 1934, 1938), Nix and Gao (1998)
and Gao et al. (1999b) identified the intrinsic material length in strain gradient
plasticity as ð=Y Þ2 b, where is the shear modulus, Y the yield stress, and b the
Burgers vector. For typical metallic materials is two orders of magnitude larger
than Y, and b is on the order of 0.1 nm. Therefore, this intrinsic material length is
on the order of microns, which is indeed the scale where size effects are observed in
2
aforementioned experiments. It is noted that this intrinsic material length 2 ð=Y Þb
2
is similar to material lengths ð=II Þ b identified by Seeger (1958) and =ðY II Þ b
d
by Lücke and Mecking (1973) and Kocks (1985), respectively, where II ¼ d" p is the
stage-II plastic work hardening modulus. Nix and Gao (1998) further showed that
the micro-indentation hardness estimated from the Taylor dislocation model agreed
very well with the experimental data of copper (McElhaney et al., 1998) and silver
(Ma and Clarke, 1995). Gao et al. (1999b) and Huang et al. (2000a) adopted a
multiscale, hierarchical framework to develop a macroscale theory of mechanism-
based strain gradient (MSG) plasticity from the microscale Taylor dislocation
model. The MSG plasticity theory agrees well with micro-bend experiments of nickel
foils (Stolken and Evans, 1998) and aluminum films (Haque and Saif, 2003), and
with micro-torsion of copper wires (Fleck et al., 1994) (see Gao et al., 1999a). The
MSG plasticity theory also agrees well with McElhaney et al.’s (1998) micro-inden-
tation hardness data for bulk copper (Huang et al., 2000b), with Stelmashenko et
al.’s (1993) hardness data for bulk tungsten (Qiu et al., 2001), with Saha et al.’s
(2001) hardness data for an aluminum thin film on a glass substrate, and with
Lloyd’s (1994) experiments of an aluminum matrix reinforced by silicon carbide
particles (Xue et al., 2002a). The MSG plasticity theory was used by Xue et al.
(2002b) to study Douglass’ (1998) digital micromirror device, and by Jiang et al.
(2001) to explain the mechanism of brittle fracture in ductile materials observed by
Elssner et al. (1994).
The MSG plasticity theory is a higher-order continuum theory that involves
higher order governing equations and requires additional boundary conditions. It is
therefore more complex than conventional plasticity theories, and the finite element
method (FEM) for the higher-order continuum theory is also more complex than
the standard FEM (Shu and Fleck, 1999; Huang et al., 2000b). Shi et al. (2001)
recently used the singular perturbation method to investigate a solid subject to a
constant body force, and showed that the effect of higher-order stress is significant
only within a thin layer near the boundary of the material. The thickness of
boundary layer is on the order of 10 nm, which is much smaller than the character-
istic length scale (e.g., microns) the strain gradient plasticity theories are intended
for. Therefore, for material properties that represent an average over the micron
scale and above, such as the micro-indentation hardness, the higher-order stress has
756 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
little or essentially no effect (Huang et al., 2000b; Saha et al., 2001). However, it is
important to separate the effect of higher-order stress from the strain gradient effect;
the former is within a thin boundary layer (thickness on the order of 10 nm) and the
latter comes from the Taylor dislocation model and is important at the micron scale.
Since the effect of higher-order stress is negligible away from the thin boundary
layer, is it possible to develop a strain gradient plasticity theory based on the
dislocation model to incorporate the strain gradient effect without the higher-order
stress? The governing equations in such a theory would be essentially the same as the
conventional plasticity theories, and therefore require no additional boundary
conditions. Its difference with the corresponding higher-order MSG plasticity theory
would be significant only within the thin boundary layer. The purpose of this paper
is to establish such a theory from the dislocation model.
Before such a theory is established, we present an approach in Section 2 to
approximately represent the rate-independent stress–strain curve in uniaxial tension
by a viscoplastic-like relation. Such an approach enables us to develop a rate-inde-
pendent conventional theory of mechanism-based strain gradient plasticity (CMSG)
without the higher-order stress from the dislocation model of Taylor (1934, 1938)
and Bailey and Hirsch (1960). The dislocation model is summarized in Section 3,
and the new strain gradient plasticity theory is developed in Section 4. The results of
several examples are compared with the higher-order MSG plasticity theory (Gao et
al., 1999b; Huang et al., 2000a,b) in Section 5.
qffiffiffiffiffiffiffiffiffiffiffiffi qffiffiffiffiffiffiffiffiffiffiffiffi
3 0 0
:p 2 :p :p
where e ¼ 2 ij ij is the von Mises effective stress, and " ¼ 3 "ij " ij is the equiva-
lent plastic strain rate. In conventional, rate-independent plasticity theories,
: : 3 0 : : :
" p is related to the effective stress rate e ¼ 2ije 0ij by " p ¼ e =hp during plastic loading
:
e > 0, where hp ¼ dp ¼ Y f 0 ð"p Þ is the incremental plastic modulus obtained
d"
from the stress–plastic strain relation ¼ Y fð"p Þ in uniaxial tension. Here Y is
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 757
the initial yield stress, f is a non-dimensional function of plastic strain "p which takes
the form
E"p N
fð"p Þ ¼ 1 þ ð2:3Þ
Y
for a power-law hardening solid, E is the Young’s modulus, and N is the plastic
work hardening exponent (04N < 1).
As discussed in the next section, for non-uniform plastic deformation at the
micron and sub-micron scale, the dislocation model of Taylor (1934, 1938) and
Bailey and Hirsch (1960) gives the flow stress dependent on both plastic strain "p
: @ : p @ : p
and plastic strain gradient p, i.e., =("p, p). Its rate form, ¼ @" p " þ @ p ,
:p :
gives the plastic strain rate " that depends on both stress rate and rate of plastic
:
strain gradient p . When substituted into (2.2), it does not yield a self-contained
:
constitutive model because of the term p . In fact, this is the reason the MSG
plasticity theory (Gao et al., 1999b; Huang et al., 2000a,b) based on the Taylor
dislocation model involves the higher-order stress.
:
One alternative way to determine the plastic strain rate, " p , is to use a viscoplastic
:p :
formulation which gives " in terms of the effective stress e rather than its rate e ,
such as the power-law viscoplastic model (Hutchinson, 1976; Pan and Rice, 1983;
Peirce, et al., 1983; Canova and Kocks, 1984; Asaro and Needleman, 1985)
m
: : e
"p ¼ "0 ; ð2:4Þ
Y fð"p Þ
:
where "0 is a reference strain rate, Yf("p) represents the stress–plastic strain curve in
uniaxial tension, and m is the rate-sensitivity exponent which usually takes a large
value (520). In the limit m=1, (2.4) is equivalent to = Yf("p) in uniaxial tension.
Such a formulation, however, displays a strain rate sensitivity and dependence on
:
the normalized time "0 t, even though this dependence is rather weak for a large rate
sensitivity exponent m (520).
In order to eliminate the strain rate and time dependence, Kok et al. (2002a,b,c)
developed a viscoplastic-limit by replacing the reference strain rate with the effective
strain rate. Such a replacement is merely for the mathematical convenience to
eliminate the rate dependence. The difference with and without this replacement is
negligible for a large exponent m (e.g., 20).
:
We follow the same approach to replace the reference strain rate "0 in (2.4) by the
:
effective strain rate ". As to be shown in Figs. 1 and 2, the effect of such replacement
is negligible. Eq. (2.4) becomes
m
: : e
"p ¼ " ; ð2:5Þ
Y fð"p Þ
qffiffiffiffiffiffiffiffiffiffiffiffi
: 2 :0 :0 : : 1 :
where " ¼ 3 "ij "ij , and "0ij ¼ " ij 3 "kk ij is the deviatoric strain rate. Eqs. (2.1), (2.2)
and (2.5) give a rate-independent elastic–plastic constitutive relation. As to be shown
758 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
Fig. 1. Uniaxial stress–strain relation for rate sensitivity exponent m=5, 20 and 1; Y is the initial yield
stress; power-law hardening exponent N=0.2, the ratio of yield stress to Young’s modulus Y =E ¼ 0:2%,
Poisson’s ratio =0.3. The limit m=1 corresponds to the conventional power-law hardening relation
(2.3).
in Section 4, (2.5) can be modified to account for the strain gradient effect via the
dislocation model of Taylor (1934, 1938) and Bailey and Hirsch (1960). Eqs. (2.1),
(2.2) and the modification of (2.5) then lead to a strain gradient plasticity theory
without the higher-order stress in Section 4.
We examine (2.1), (2.2) and (2.5) in uniaxial tension in order to determine whether
they can accurately represent the uniaxial stress–strain relation in (2.3). For 11=
and other stresses =0, (2.1) and (2.2) give the volumetric and deviatoric strain rates
as
:
: :
" 11 þ 2"22 ¼ ; ð2:6Þ
3K
:
2 : : :
ð" 11 "22 Þ ¼ þ "p ; ð2:7Þ
3 3
and (2.5) gives the plastic strain rate
m
: 2 : :
" p ¼ ð"11 "22 Þ ; ð2:8Þ
3 Y fð"p Þ
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 759
Fig. 2. Loading and unloading stress-strain curve in uniaxial tension; Y is the initial yield stress. The
stress increases monotonically to 1.5 Y, at which it unloads completely and reloads to a higher stress
2 Y. The material parameters are the same as those in Fig. 1. The solid curve (m=20) corresponds to the
present plasticity theory. The dashed curve (m=1) corresponds to the conventional power-law hardening
with a linear unloading relation (a straight line with the slope of Young’s modulus E).
: : :
where we have used " ¼ 23 ð"11 "22 Þ in uniaxial tension. Eqs. (2.6), (2.7), and (2.8)
are ordinary differential equations for "11, "22 and "p, and they can be solved by the
standard numerical method (e.g., Press et al., 1986). The resulting strains "11, "22
and "p depend only on the stress , not time t, since (2.1), (2.2) and (2.5) give a
rate-independent plasticity theory.
Fig. 1 shows the uniaxial stress–strain relation ( vs. "11) for rate sensitivity
exponent m=5, 20 and 1, where is normalized by the initial yield stress Y,
power-law hardening exponent N=0.2, the ratio of yield stress to Young’s modulus
Y =E ¼ 0:2%, Poisson’s ratio =0.3, and m=1 corresponds to the conventional
power-law hardening relation (2.3). It is observed that all curves are very close, and
there is almost no difference between the curves for m=20 and m=1. Therefore,
for m520, (2.6), (2.7) and (2.8) can represent the power-law hardening in uniaxial
tension very well. Even though such an approach looks complex, it paves the way
for the latter establishment of a strain gradient plasticity theory without the higher-
order stress, as shown in Section 4.
Similar to viscoplastic models, the constitutive relations (2.6), (2.7) and (2.8) do
not separate elastic and plastic deformations, and therefore can be applied to both
loading and unloading. Fig. 2 shows the loading and unloading stress–strain curve
in uniaxial tension. The stress is normalized by the initial yield stress Y, and the
760 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
material parameters are the same as those in Fig. 1; N=0.2, Y =E ¼ 0:2%, =0.3;
and m=20 and m=1. Here m=1 corresponds to the conventional power-law
hardening relation = Yf("p) and it becomes a straight line with the slope E in Fig. 2
during unloading. The stress increases monotonically to =1.5 Y, at which it
unloads completely to =0 and reloads to a higher stress =2 Y. It is observed that
the curve for m=20 is very close to that for m=1 during both loading and
unloading. Therefore, the constitutive relations (2.6), (2.7) and (2.8) can represent
well the loading and unloading behavior of rate-independent materials in uniaxial
tension.
The dislocation model of Taylor (1934, 1938) and Bailey and Hirsch (1960) gives
the shear flow stress in terms of the dislocation density as
pffiffiffi
¼ b ; ð3:1Þ
where is the shear modulus; b is the magnitude of the Burgers vector; and is an
empirical coefficient around 0.3. Nabarro et al. (1964) pointed out that (3.1) can be
derived by multiple methods.
The dislocation density is composed of the density S for statistically stored
dislocations, which accumulate by trapping each other in a random way (Ashby,
1970), and density G for geometrically necessary dislocations, which are required
for compatible deformation of various parts of the material (Nye, 1953; Cottrell,
1964; Ashby, 1970), i.e.
¼ S þ G : ð3:2Þ
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p
flow ¼ Mb S þ r : ð3:5Þ
b
Y fð"p Þ 2
S ¼ : ð3:6Þ
Mb
Its substitution into (3.5) gives the flow stress accounting for the nonuniform
plastic deformation associated with geometrically necessary dislocations as
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
flow ¼ ½Y fð"p Þ 2 þM 2 r2 2 b p ¼ Y f 2 ð"p Þ þ l p ; ð3:7Þ
where
2 2
l ¼ M 2 r2 b ¼ 182 b ð3:8Þ
Y Y
is the intrinsic material length in strain gradient plasticity, M=3.06 and r ¼ 1:90. It
is noted that the intrinsic material length l in (3.8) represents a natural combination
of elasticity (), plasticity ( Y), and atomic nature of solids (b). For typical metallic
materials, the Burgers vector b is on the order of one tenth of a nanometer, =Y is
on the order of 102, and is around 0.3. These give l on the order of microns, which
is indeed the length scale at which the size effects are observed in aforementioned
experiments.
We point out several observations on the flow stress in (3.7).
(i) If the characteristic length of plastic deformation is much larger than the
intrinsic material length l, as in macroscopic deformation, the strain gradient term
l p in (3.7) becomes negligible such that the flow stress degenerates to Yf("p) in
conventional plasticity.
(ii) The flow stress in (3.7) is based on the Taylor dislocation model, which repre-
sents an average of dislocation activities and is therefore only applicable at a scale
much larger than the average dislocation spacing Ld (e.g., 3 4Ld). For a typical
dislocation density of 1015 =m2 , the average dislocation spacing is around 30 nm such
that the flow stress in (3.7) holds at a scale above 100 nm.
(iii) Even though the intrinsic material length l in (3.8) depends on the choice of
initial yield stress Y, the flow stress in (3.7) is, in fact, independent of Y. This is
because both terms inside the square root in (3.7), namely the uniaxial stress–plastic
strain relation Yf("p) and Y2 l p ¼ 182 2 b p , are independent of Y.
762 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
:
: 0ij 3":p 0
" 0ij ¼ þ : ð4:2Þ
2 2e ij
:
The equivalent plastic strain rate " p in (2.5), which involves the flow stress Yf("p)
in uniaxial tension, has not accounted for the strain gradient effect. A natural way to
introduce the strain gradient effect is to replace the uniaxial flow stress Yf("p) in
(2.5) by the flow stress flow in (3.7) based on the Taylor dislocation model, i.e.,
m " #m
:p : e : e
" ¼" ¼" pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : ð4:3Þ
flow Y f 2 ð"p Þ þ l p
The substitution of (4.3) into (4.1) and (4.2) gives the strain rate
: : :
: kk 0ij 3" e m 0
" ij ¼ ij þ þ ij
9K 2 2e flow
: : : !m
kk 0ij 3" e
¼ ij þ þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ij0 : ð4:4Þ
9K 2 2e Y f 2 ð"p Þ þ l p
This constitutive relation accounts for the strain gradient effect but does not
involve the higher-order stress. It can be inverted to give the stress-rate in terms of
the strain-rate as
:
: : :0 3" e m 0
ij ¼ K"kk ij þ 2 "ij ij
2e flow
( : " #m )
: :0 3" e
¼ K" kk ij þ 2 " ij pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ij0 : ð4:5Þ
2e Y f 2 ð"p Þ þ l p
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 763
It is observed that the effective plastic strain gradient p reduces the plastic strain
rate, which is consistent with the aforementioned experiments. Similar to Section 2,
(4.4) or (4.5) gives a unified constitutive relation for both loading and unloading.
It is interesting to point out the relation between the above constitutive model
(4.5) and two frameworks of strain gradient plasticity theories. Even though (4.5)
does not involve the higher-order stress, as to be shown in Section 5, it gives the
same results with the higher-order MSG plasticity theory (Gao et al., 1999b; Huang
et al., 2000a,b) except within a thin boundary layer of the solid. It belongs to the
aforementioned second framework of strain gradient plasticity theories (e.g.,
Acharya and Bassani, 2000; Acharya and Beaudoin, 2000) since the effective plastic
strain gradient reduces the plastic incremental moduli. Moreover, it bears similarity
with Evers et al.’s (2002) and Dai and Parks’ (2003) viscoplastic theory of strain
: :
gradient plasticity because, if the reference strain rate " 0 were not replaced by "
in (2.4) and (2.5), the present constitutive relation (4.5) would have also been
viscoplastic.
p
4.2. Effective plastic strain gradient
The flow stress flow in (4.4) and (4.5) is a function of the equivalent plastic strain
"p and effective plastic strain gradient p as given in (3.7), whereÐ "p is a non-
:
decreasing function of deformation obtained from (4.3) by "p ¼ "p dt. In the
following, we discuss two definitions of the effective plastic strain gradients
established from the models of geometrically necessary dislocations by Gao et al.
(1999b) and by Hwang et al. (2003b).
Following Fleck and Hutchinson (1997), Gao et al. (1999b) used three quadratic
invariants of the plastic strain gradient tensor to represent p. The coefficients of the
quadratic invariants were determined by three models of geometrically necessary
dislocations for bending, torsion, and void growth. The resulting effective plastic
strain gradient is given by
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p 1 p p
¼ ; ð4:6Þ
4 ijk ijk
where
p
ijk ¼ "pik;j þ "pjk;i "pij;k ; ð4:7Þ
Ð :
and "pij ¼ "pij dt.
Nye’s (1953) dislocation tensor (e.g., Kondo, 1952, 1955; Bilby et al., 1955;
Eshelby, 1956; Kröner, 1960; Fox, 1966) leads to another expression of the effective
plastic strain gradient p. Even though the total deformation gradient F=Fe.Fp
(Lee, 1969) corresponds to a compatible deformation field, its plastic part Fp (or
equivalently, the inverse elastic part, Fe1) is generally incompatible, and therefore
requires geometrically necessary dislocations to accommodate the incompatible
plastic deformation (Steinmann, 1996; Arsenlis and Parks, 1999; Shizawa and Zbib,
764 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
1999; Acharya and Bassani, 2000; Acharya and Beaudoin, 2000; Beaudoin and
Acharya, 2001; Cermelli and Gurtin, 2001). Steinmann (1996) and Cermelli and
Gurtin (2001) introduced a tensor of geometrically necessary dislocations as
1 p
detFp F ðr Fp Þ, where det stands for the determinant, and ! is the gradient
operator. For infinitesimal deformation which is consistent with the simple sum of
dislocation densities in (3.2), F=I+"p+"e+’, where I is the second-order identity
tensor, "p and "e are the tensors of plastic and elastic strains, respectively, and ’ is
the rotation tensor. The above tensor of geometrically necessary dislocations
1 p
detFp F ðr Fp Þ becomes "p! for infinitesimal deformation (Hwang et al.,
2003b). Accordingly, the effective plastic strain gradient p can be defined by the
quadratic invariant of "p! as
p
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
¼ ð"p rÞ : ð"p rÞ ¼ ð"p rÞ : ð"p rÞ: ð4:8Þ
Since the present strain gradient plasticity theory does not involve the higher-
order stress, equilibrium equations and traction boundary conditions remain the
same as the conventional continuum theories, i.e.,
ji;j þ fi ¼ 0; ð4:9Þ
ji nj ¼ ti ; ð4:10Þ
where fi is the body force, and nj and ti are the unit normal and stress traction on the
boundary.
5. Analyses
We present a few examples to illustrate the CMSG plasticity theory without the
higher-order stress.
0 0
The non-vanishing deviatoric stresses and effective stress are 11 ¼ 222 ¼
0 2
233 ¼ 3 11 and e= 11. The non-vanishing strain rates are obtained from (4.4) as
: :
: 11 : 11 m : : 11 1 : 11 m
" 11 ¼ þ" ; "22 ¼ " 33 ¼ " ; ð5:2Þ
E flow E 2 flow
where E and are the Young’s modulus and Poisson’s ratio, respectively, and
: : : : : :
" ¼ 23 ð"11 "22 Þ. Elimination of "22 gives the following relation between " 11 and 11 ,
2 3
11 m
: 6 2 flow 7 : 11
" 11 ¼6
4 1 þ ð1 þ Þ 7 : ð5:3Þ
3 11 m 5 E
1
flow
The flow stress Ðflow in (5.2) and (5.3) is given in (3.7) in terms of the equivalent
: :
plastic strain "p ¼ "p dt and effective plastic strain gradient p. Here "p is obtained
from (4.3) by
11 m
:
: 2 flow 11
" p ¼ ð1 þ Þ m ; ð5:4Þ
3 11 E
1
flow
Fig. 3. A schematic diagram of a bar subjected to a constant body force g and uniform stress at the free
end along the direction of the bar.
766 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
and the initial value of "p (i.e., prior to imposed loading and g) is zero. The
effective plastic strain gradient p in (4.6) and (4.8) becomes
rffiffiffi rffiffiffi p
p 5 d"p 5 d"
¼ ¼ ; ð5:5Þ
8dx1 8 dx1
and
pffiffiffi p pffiffiffi p
p 3 d" 3 d"
¼ ¼ ; ð5:6Þ
2 dx1 2 dx1
respectively, and their difference is less than 10%. For 0 < x1 < L, a central difference
scheme is used to calculate
and
d"p "p ðL 2Dx1 Þ 4"p ðL Dx1 Þ þ 3"p ðLÞ
¼
dx1 x ¼L 2Dx1
1
The plastic strain distribution in the bar ("p versus x1 =L) is shown in Fig. 4 for
several ratios of intrinsic material length l to the bar length L. The effective plastic
strain gradient p in (5.5) is used. [Eq. (5.6) gives essentially the same plastic strain
distribution as (5.5).] The limit l=L ¼ 0 corresponds to conventional plasticity, while
l=L ¼ 1 and 10 represent the bar length on the order of microns and sub-microns,
respectively. The material parameters are plastic work hardening exponent N=0.2,
Poisson’s ration =0.3, ratio of yield stress to Young’s Modulus Y =E ¼ 0:2%, and
rate sensitivity exponent m=20. These material parameters are used in all examples
in the paper unless otherwise specified. The applied stress at the free end of the bar
(x1=0) and the body force g are imposed proportionally, with the maximum values
at ¼ Y and g ¼ Y =L. All curves in Fig. 4 give a vanishing plastic strain at x1=0,
but strong size effect is clearly observed as x1 increases, since the curve for l=L ¼ 10
is significantly lower than that for l=L ¼ 1 and for conventional plasticity ðl=L ¼ 0Þ.
Shi et al. (2001) studied the same problem using the MSG plasticity theory that is
also based on the Taylor dislocation model but involves the higher-order stress. For
the special case of plastic work hardening exponent N=0.5 and negligible elastic
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 767
Fig. 4. The plastic strain distribution in the bar; l is the intrinsic material length in strain gradient plasti-
city and L the bar length. The limit l=L ¼ 0 corresponds to conventional plasticity. The material para-
meters are the same as those in Fig. 1.
Fig. 5. The distribution of strain gradient d"11 =dx1 in the bar predicted by the present (solid curve) and
MSG plasticity (dashed curve) theories for bar length L=0.1l, where l is the intrinsic material length in
strain gradient plasticity. The material parameters are N=0.5, ¼ 1=2, Y =E ¼ 0:2%, and m=20.
x 2
p 1
initial ¼ 0 1 ; ð5:7Þ
D
which has a maximum value 0 at x1=0, and vanishes on the boundaries x1= D.
Niordson and Hutchinson (2003) found that the plastic strain distribution predicted
by Bassani’s (2001) theory exhibits a vertex at the symmetry plane x1=0, which
contradicts to the symmetry condition d p =dx1 ¼ 0 at x1=0. We use the present
strain gradient plasticity theory to study the same problem and examine the possi-
bility of vertex in the plastic strain distribution.
The non-vanishing stresses are 12= 21=, which are uniform p inffiffiffi the layer as
required by the equilibrium Eq. (4.9). The effective stress is e ¼ 3. The non-
vanishing strain rates and effective strain rate are
: : 1: : 1 : 1 :
" 12 ¼ "21 ¼ ; " ¼ pffiffiffi ¼ pffiffiffi ; ð5:8Þ
2 3 3
:
where is the engineering shear strain rate and it is non-uniform in the layer due to
the residual strain given in (5.7). For these stresses and strain rates, (4.4) gives the
: :
following relation between and
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 769
Fig. 6. A schematic diagram of an infinite layer with thickness 2D under uniform shear stress .
:
: 1
¼ p ffiffi
ffi m : ð5:9Þ
3
1
flow
The flow Ðstress flow in (5.9) is given in (3.7) in terms of the equivalent plastic
: :
strain "p ¼ "p dt and effective plastic strain gradient p. Here " p is then obtained
from (4.3) by
pffiffiffi m
3
:
: 1 flow
" p ¼ pffiffiffi p ffiffi
ffi m ; ð5:10Þ
3 3
1
flow
and the initial value of "p (i.e., prior to imposed shear stress ) is obtained
pffiffiffi from the
p p
initial residual plastic shear strain initial in (5.7) by "p ¼ initial = 3. The effective
plastic strain gradient p in (4.6) and (4.8) becomes identical, and is given by
pffiffiffi p
p 3 d"
¼ : ð5:11Þ
2 dx1
Only 04x14D is studied due to symmetry. Similar to Section 5.1, the central and
skew finite differences are used for 0 < x1 < D, and x1=0,D, respectively.
770 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
p
Fig. 7 shows the distribution of normalized engineering plastic shear strain
pffiffiffi p =0
p
versus x1 =D for several levels of applied shear stress =0 , where ¼ 3" and
0 ¼ 0 . The ratio of intrinsic material length to half layer width l=D ¼ 3 was
suggested by Niordson
pffiffiffi and Hutchinson (2003), so was the maximum residual plastic
shear strain 0 ¼ 3 0:2% ¼ 0:0035. Here p =0 starts from the residual plastic
p 2
shear strain initial =0 ¼ 1 xD1 in (5.7) at =0, which has a maximum plastic
strain gradient on the boundary (x1=D) and vanishes at the center (x1=0). The
initial flow stress obtained from (3.7) is also nonuniform in the layer. As the applied
shear stress increases, p near the boundary x1=D increases faster than p near the
center x1=0 such that the plastic strain distribution becomes more uniform. For
=0 5 1:16, the plastic strain becomes uniform in the layer, and the effect of non-
uniform residual plastic shear strain completely disappears. Therefore, the present
strain gradient plasticity theory does not show any ‘‘vertex’’ in the plastic strain
distribution.
We have also examined even thinner layers (e.g., l=D around 10) and have not
observed any vertex in the plastic strain distribution. We must point out that the
existence of possible vertex in the plastic strain distribution cannot be ruled out for
l=D >> 10. However, as discussed at the end of Section 3, l=D >> 10 corresponds
to a layer thinner than sub-microns, which falls outside the intended application
range of any continuum (including strain gradient) plasticity theories.
Fig. 7. The distribution of normalized engineering plastic shear strain p =0 in layer for l=D ¼ 3, and
several levels of applied shear stress =0 , where l is the intrinsic material length in strain gradient plasticity,
D the layer thickness, and 0= 0. The material parameters are the same as those in Fig. 1.
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 771
The thin-wire torsion experiments of Fleck et al. (1994) clearly showed that plastic
work hardening increases significantly as the wire diameter decreases from 170 to 12
mm. We use the present strain gradient plasticity theory to study a thin wire under
torsion.
The wire has a radius a and is subject to a twist per unit length . The non-van-
ishing strain rates in the cylindrical coordinates (r, , z) and effective strain rate are
: : 1: : 1 :
" z ¼ "z ¼ r; " ¼ pffiffiffi r; ð5:12Þ
2 3
where z denotes the central axis of the pffiffiwire.
ffi The non-vanishing stresses are z= z,
which give the effective stress e ¼ 3z . The substitution of above stresses and
: :
strain rates into (4.5) yields the relation between z and ,
" pffiffiffi m #
: : 3z
z ¼ r 1 : ð5:13Þ
flow
TheÐ flow stress flow in (5.13) is given in (3.7) in terms of equivalent plastic strain
: :
"p ¼ "p dt and effective plastic strain gradient p. Here "p is obtained from (4.3) by
pffiffiffi m
:p 1 : 3z
" ¼ pffiffiffi r ; ð5:14Þ
3 flow
and the initial value of "p (i.e., prior to imposed twist ) is zero. The effective plastic
strain gradient p in (4.6) and (4.8) becomes
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p 2 p 2
p 3 d" " 2 d"p "p
¼ pffiffiffi þ þ ; ð5:15Þ
2 2 dr r 3 dr r
and
rffiffiffisffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi
p 3 d"p 2 "p 2 d"p "p
¼ þ þ ; ð5:16Þ
2 dr r dr r
respectively.
The torque T in the wire cross section is obtained from z by
ða
T ¼ z 2r rdr ð5:17Þ
0
The normalized torque, T= Y a3 , versus the normalized twist per unit length, a,
is shown in Fig. 8 for the effective plastic strain gradient in (5.15) and several ratios
of intrinsic material length l to wire radius a. [The torque based on p in (5.16) is
only slightly larger and is not presented here.] The limit l=a ¼ 0 corresponds to
conventional plasticity, while l=a ¼ 1 and 10 correspond to micron and sub-micron
772 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
Fig. 8. The normalized torque, T= Y a3 , versus the normalized twist per unit length, a, predicted by the
present (solid curves) and MSG plasticity (dashed curves) theories for several ratios of intrinsic material
length l to wire radius a, where Y is the initial yield stress. The limit l=a ¼ 0 corresponds to conventional
plasticity. The material parameters are the same as those in Fig. 1.
Stolken and Evans (1998) observed strong size effect on plastic work hardening in
bending of thin nickel beams with beam thickness ranging from 12.5 to 100 mm. We
use the present strain gradient plasticity theory to study bending of a thin beam. For
simplicity, we assume the material is incompressible, and the beam is subject to
plane-strain bending within the (x1, x2) plane, with x1-axis being the neutral axis of
the beam.
The beam has a thickness h, and is subject to a bending curvature . The non-
vanishing strain rates and effective strain rate are
: : : : 2 :
" 11 ¼ "22 ¼ x2 ; " ¼ pffiffiffi jx2 j: ð5:18Þ
3
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 773
The flow stress Ðflow in (5.19) is once again given in (3.7) in terms of equivalent
: :
plastic strain "p ¼ "p dt and effective plastic strain gradient p. Here "p is obtained
from (4.3) by
pffiffiffi m
: 2 : 3j11 j
" p ¼ pffiffiffi jx2 j ; ð5:20Þ
3 2flow
and the initial value of "p (i.e., prior to imposed curvature ) is zero. The effective
strain gradient p in (4.6) and (4.8) becomes the same, and is given by
pffiffiffi p
p 3 d"
¼ : ð5:21Þ
2 dx2
The bending moment in the cross section is obtained from the axial stress 11 by
ðh
2
M¼ 11 x2 dx2 : ð5:22Þ
h2
Fig. 9 shows the normalized bending moment, M= Y h2 , versus the normalized
curvature, h, predicted by the present and MSG (Huang et al., 2000a) plasticity
theories for several ratios of intrinsic material length l to beam thickness h. Once
again the difference between two theories is very small. The size effect in beam
bending is clearly observed since the normalized bending moment M= Y h2 for a
micron thick beam ðl=h¼ 1Þ is nearly 20% larger than that for thick beams
ðl=h ¼ 0Þ, while M= Y h2 for a sub-micron thick beam ðl=a ¼ 10Þ is about 2.5 times
of that for l=h ¼ 0. Therefore, the present strain gradient plasticity theory can
capture the size effect at the micron and sub-micron scales.
Based on the Taylor dislocation model, Liu et al. (2003) extended the Rice and
Tracey (1969) model of void growth to account for the void size effect. They estab-
lished that, at the same remote stress level, small voids tend to grow much slower
than large voids. We use the present strain gradient plasticity theory to study the
growth of spherical and cylindrical microvoids.
774 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
Fig. 9. The normalized bending moment, M= Y h2 , versus the normalized curvature, h, predicted by
the present (solid curves) and MSG plasticity (dashed curves) theories for several ratios of intrinsic
material length l to beam thickness h. The limit l=h ¼ 0 corresponds to conventional plasticity. The
material parameters are the same as those in Fig. 1, except ¼ 1=2.
: R2 : e m : R2 : e m
0RR ¼ 4 03 u0 1 ; e ¼ 6 03 u0 1 : ð5:25Þ
R flow R flow
The flow Ðstress flow in (5.25) is given in (3.7) in terms of the equivalent plastic
: :
strain "p ¼ "p dt and the effective plastic strain gradient p. Here "p is obtained
from (4.3) by
: 2R 2 : e m
" p ¼ 30 u0 ; ð5:26Þ
R flow
and the initial value of "p (i.e., prior to imposed deformation) is zero. The effective
plastic strain gradient p in (4.6) and (4.8) becomes
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p 2
p 1 5 d"p 2 d"p "p "
¼ pffiffiffi þ3 þ9 ; ð5:27Þ
2 4 dR dR R R
and
p 1 d"p "p
¼ pffiffiffi þ 3 ; ð5:28Þ
2 dR R
respectively.
From equilibrium Eq. (4.9) and the traction-free boundary condition on the void
surface, we obtain the remotely applied stress 1 as
ð1
0 ð1
1 RR e
¼ 3 dR ¼ 2 dR: ð5:29Þ
R0 R R0 R
Fig. 10 shows the normalized remotely applied stress, 1 =Y , versus the normalized
displacement on the void surface, u0 =R0 , for the effective plastic strain gradient in
(5.27) and several ratios of intrinsic material length l to void radius R0. [The results
based on p in (5.28) are essentially the same as those in Fig. 10, and are not pre-
sented here.] The limit l=R0 ¼ 0 corresponds to conventional plasticity, while l=R0 ¼
1 and l=R0 ¼ 10 correspond to micron and sub-micron size microvoids, respectively.
The curve for l=R0 ¼ 1 is very close to that for conventional plasticity ðl=R0 ¼ 0Þ.
Therefore, the micron-size voids ðl=R0 ¼ 1Þ have little size effect on the void growth
rate, which is consistent with Liu et al.’s (2003) studies. However, the sub-micron
size voids ðl=R0 ¼ 10Þ clearly display some size effect since the corresponding curve
is above that for conventional plasticity ðl=R0 ¼ 0Þ.
776 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
Fig. 10. The normalized remotely applied stress, 1 =Y , versus the normalized displacement on the spherical
void surface, u0 =R0 , for several ratios of intrinsic material length l to void radius R0. The limit l=R0 ¼ 0 cor-
responds to conventional plasticity. The material parameters are the same as those in Fig. 1, except ¼ 1=2.
where u0, r0 and r are the radial displacement on the void surface, void radius, and
radial coordinate, respectively. The non-vanishing strain rates in the cylindrical
coordinates (r, , z) and effective strain rate are
: : r0 : : 2 r0 :
" rr ¼ " ¼ 2 u0 ; " ¼ pffiffiffi 2 u0 : ð5:31Þ
r 3r
rr þ
The non-vanishing stresses rr ; ; zz ¼ 2 can be decomposed to a
hydrostatic stress, rr þ
2 p , ffiffiand 0 0 0 0
ffi 0 deviatoric stresses rr ; ¼ rr ; zz ¼ 0 . The
effective stress is e ¼ 3rr . The substitution
: of above stresses and strain rates
: : :
into (4.5) gives the relation between 0 rr and u0 , or equivalently, between e and u0 ,
:0 r0 : e m : pffiffiffi r0 : e m
rr ¼ 2 2 u0 1 ; e ¼ 2 3 2 u 0 1 : ð5:32Þ
r flow r flow
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 777
The flow Ðstress flow in (5.32) is given in (3.7) in terms of the equivalent plastic
: :
strain "p ¼ "p dt and the effective plastic strain gradient p. Here "p is obtained
from (4.3) by
: 2 r0 : e m
" p ¼ pffiffiffi 2 u0 ; ð5:33Þ
3r flow
and the initial value of "p (i.e., prior to imposed deformation) is zero. The effective
plastic strain gradient p in (4.6) and (4.8) becomes
pffiffiffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
p 2 p p
p 2
ffi
p 3 d" d" " "
¼ þ2 þ4 ; ð5:34Þ
2 dr dr r r
and
pffiffiffi p
p 3 d" "p
¼ þ 2 ; ð5:35Þ
2 dr r
respectively.
Fig. 11. The normalized remotely applied stress, 1 =Y , versus the normalized displacement on the
cylindrical void surface, u0 =r0 , for several ratios of intrinsic material length l to void radius r0. The limit
l=r0 ¼ 0 corresponds to conventional plasticity. The material parameters are the same as those in Fig. 1,
except ¼ 1=2.
778 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
From equilibrium equation (4.9) and the traction-free boundary condition on the
void surface, we obtain the remotely applied stress 1 as
ð1 ð
1 2 0 1 12
¼ rr dr ¼ pffiffiffi e dr: ð5:36Þ
r0 r 3 r0 r
The normalized remotely applied stress, 1 =Y , versus the normalized displace-
ment on the void surface, u0 =r0 is shown in Fig. 11 for the effective plastic strain
gradient in (5.34) and several ratios of intrinsic material length l to void radius r0.
[The results based on p in (5.35) are essentially the same as those in Fig. 11, and are
not presented here.] The results are very similar to those in Fig. 10 for spherical
voids; the submicron-size voids display some void size effect, while the micron-size
voids show very little dependence on the void radius r0.
6. Concluding remarks
Acknowledgements
Y.H. and S.Q. are grateful for many insightful discussion with Professor A.J.
Beaudoin. Y.H. acknowledges the support from NSF (grant CMS-0084980). The
support from NSFC is also acknowledged.
References
Acharya, A., Bassani, J.L., 2000. Lattice incompatibility and a gradient theory of crystal plasticity.
Journal of the Mechanics and Physics of Solids 48 (8), 1565–1595.
Acharya, A., Beaudoin, A.J., 2000. Grain-size effect in viscoplastic polycrystals at moderate strains.
Journal of the Mechanics and Physics of Solids 48 (10), 2213–2230.
Aifantis, E.C., 1984. On the microstructural origin of certain inelastic models. Journal of Engineering
Materials and Technology—Transactions of the ASME 106 (4), 326–330.
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 779
Aifantis, E.C., 1992. On the role of gradients in the localization of deformation and fracture.
International Journal of Engineering and Science 30 (10), 1279–1299.
Arsenlis, A., Parks, D.M., 1999. Crystallographic aspects of geometrically-necessary and statistically-
stored dislocation density. Acta Materialia 47 (5), 1597–1611.
Asaro, R.J., Needleman, A., 1985. Texture development and strain hardening in rate dependent
polycrystals. Acta Metallurgica et Materialia 33, 923–953.
Ashby, M.F., 1970. The deformation of plastically non-homogeneous alloys. Philosophical Magazine 21,
399–424.
Atkinson, M., 1995. Further analysis of the size effect in indentation hardness tests of some metals.
Journal of Materials Research 10, 2908–2915.
Bailey, J.E., Hirsch, P.B., 1960. The dislocation distribution, flow stress, and stored energy in cold-worked
polycrystalline silver. Philosophical Magazine 5, 485–497.
Bassani, J.L., 2001. Incompatibility and a simple gradient theory of plasticity. Journal of the Mechanics
and Physics of Solids 49 (9), 1983–1996.
Beaudoin, A.J., Acharya, A., 2001. A model for rate-dependent flow of metal polycrystals based on the
slip plane lattice incompatibility. Materials Science and Engineering A 309, 411–415.
Bilby, B.A., Bullough, R., Smith, E., 1955. Continuous distributions of dislocation: a new application of
the methods of non-Riemannian geometry. Proceedings of the Royal Society of London A 231, 263–273.
Bishop, J.F.W., Hill, R., 1951a. A theory of plastic distortion of a polycrystalline aggregate under
combined stresses. Philosophical Magazine 42, 414–427.
Bishop, J.F.W., Hill, R., 1951b. A theoretical derivation of the plastic properties of a polycrystalline face-
centered metal. Philosophical Magazine 42, 1298–1307.
Canova, G.R., Kocks, U.F., 1984. The development of deformation textures and resulting properties of
fcc metals. In: Brakman, C.M., Jongenburger, P., Mittemeijer, E.J. (Eds.), Seventh Int. Conf. on
Textures of Materials. Soc. for Materials Science, Netherlands, pp. 573–579.
Cermelli, P., Gurtin, M.E., 2001. On the characterization of geometrically necessary dislocations in finite
plasticity. Journal of the Mechanics and Physics of Solids 49 (7), 1539–1568.
Chen, S.H., Wang, T.C., 2000. A new hardening law for strain gradient plasticity. Acta Materialia 48 (16),
3997–4005.
Chen, S.H., Wang, T.C., 2002. A new deformation theory with strain gradient effects. International
Journal of Plasticity 18 (8), 971–995.
Cleveringa, H.H.M., VanderGiessen, E., Needleman, A., 1997. Comparison of discrete dislocation and
continuum plasticity predictions for a composite material. Acta Materialia 45 (8), 3163–3179.
Cleveringa, H.H.M., VanderGiessen, E., Needleman, A., 1998. Discrete dislocation simulations and size
dependent hardening in single slip. Journal de Physique IV 8 (P4), 83–92.
Cleveringa, H.H.M., VanderGiessen, E., Needleman, A., 1999a. A discrete dislocation analysis of
bending. International Journal of Plasticity 15 (8), 837–868.
Cleveringa, H.H.M., VanderGiessen, E., Needleman, A., 1999b. A discrete dislocation analysis of residual
stresses in a composite material. Philosophical Magazine A 79 (4), 893–920.
Cottrell, A.H., 1964. The mechanical Properties of Materials. J. Wiley, New York. p. 277.
Dai, H., Parks, D.M., 2003. Geometrically-necessary Dislocation Density in Continuum Crystal Plasticity
Theory and FEM Implementation (unpublished manuscript).
de Borst, R., Mühlaus, H.B., 1992. Gradient-dependent plasticity: formulation and algorithmic aspects.
International Journal for Numerical Methods in Engineering 35 (3), 521–539.
De Guzman, M.S., Neubauer, G., Flinn, P., Nix, W.D., 1993. The role of indentation depth on the
measured hardness of materials. Materials Research Symposium Proceedings 308, 613–618.
Douglass, M.R., 1998. Lifetime estimates and unique failure mechanisms of the digital micromirror device
(DMD). In: Annual Proceedings–Reliability Physics Symposium (Sponsored by IEEE), 31 March–2
April 1998, pp. 9–16. Available from <http://www.dlp.com/dlp/resources/whitepapers/pdf/
ieeeir.pdf >.
Elssner, G., Korn, D., Ruhle, M., 1994. The influence of interface impurities on fracture energy of UHV
diffusion-bonded metal-ceramic bicrystals. Scripta Metallurgica et Materialia 31 (8), 1037–1042.
780 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
Eshelby, J.D., 1956. The continuum theory of lattice defects. In: Seitz, F.D., Turnbull, D. (Eds.), Solid
State Physics, Vol. 3. Academic Press, New York, pp. 79–144.
Evers, L.P., Parks, D.M., Brekelmans, W.A.M., Geers, M.G.D., 2002. Crystal plasticity model with
enhanced hardening by geometrically necessary dislocation accumulation. Journal of the Mechanics
and Physics of Solids 50, 2403–2424.
Fleck, N.A., Hutchinson, J.W., 1993. A phenomenological theory for strain gradient effects in plasticity.
Journal of the Mechanics and Physics of Solids 41 (12), 1825–1857.
Fleck, N.A., Hutchinson, J.W., 1997. Strain gradient plasticity. In: Hutchinson, J.W., Wu, T.Y. (Eds.),
Advances in Applied Mechanics, Vol. 33. Academic Press, New York, pp. 295–361.
Fleck, N.A., Hutchinson, J.W., 2001. A reformulation of strain gradient plasticity. Journal of the
Mechanics and Physics of Solids 49 (10), 2245–2271.
Fleck, N.A., Muller, G.M., Ashby, M.F., Hutchinson, J.W., 1994. Strain gradient plasticity: theory and
experiments. Acta Metallurgica et Materialia 42 (2), 475–487.
Fox, N., 1966. A continuum theory of dislocations for single crystals. Journal of Institute for
Mathematics and its Applications 2, 285–298.
Gao, H., Huang, Y., Nix, W.D., 1999a. Modeling plasticity at the micrometer scale. Naturwissenschaften
86, 507–515.
Gao, H., Huang, Y., Nix, W.D., Hutchinson, J.W., 1999b. Mechanism-based strain gradient plasticity- I.
Theory. Journal of the Mechanics and Physics of Solids 47 (6), 1239–1263.
Gurtin, M.E., 2000. On the plasticity of single crystals: free energy, microforces, plastic-strain gradient.
Journal of the Mechanics and Physics of Solids 48 (5), 989–1036.
Gurtin, M.E., 2002. A gradient theory of single-crystal viscoplasticity that accounts for geometrically
necessary dislocations. Journal of the Mechanics and Physics of Solids 50 (1), 5–32.
Haque, M.A., Saif, M.T.A. Strain gradient effect in nanoscale thin films. Acta Materialia, S1 (11), 3053–
3061.
Huang, Y., Gao, H., Nix, W.D., Hutchinson, J.W., 2000a. Mechanism-based strain gradient plasticity—
II. Analysis. Journal of the Mechanics and Physics of Solids 48 (1), 99–128.
Huang, Y., Xue, Z., Gao, H., Nix, W.D., Xia, Z.C., 2000b. A study of microindentation hardness tests by
mechanism-based strain gradient plasticity. Journal of Materials Research 15 (8), 1786–1796.
Hutchinson, J.W., 1976. Bounds and self-consistent estimates for creep of polycrystalline materials.
Proceedings of the Royal Society of London A 348, 101–127.
Hwang, K.C., Jiang, H., Huang, Y., Gao, H., Hu, N., 2002. A finite deformation theory of strain gradient
plasticity. Journal of the Mechanics and Physics of Solids 50 (1), 81–99.
Hwang, K.C., Jiang, H., Huang, Y., Gao, H., 2003a. Finite deformation analysis of mechanism-based
strain gradient plasticity: torsion and crack tip field. International Journal of Plasticity 19 (2), 235–251.
Hwang, K.C., Yun, G., Huang, Y., 2003b. A reformulation of mechanism-based strain gradient (MSG)
plasticity (submitted for publication).
Jiang, H., Huang, Y., Zhuang, Z., Hwang, K.C., 2001. Fracture in mechanism-based strain gradient
plasticity. Journal of the Mechanics and Physics of Solids 49 (5), 979–993.
Kocks, U.F., 1970. The relation between polycrystal deformation and single crystal deformation.
Metallurgical and Materials Transactions 1, 1121–1144.
Kocks, U.F., 1985. Dislocation interactions: flow stress and strain hardening. In: Loretto, M.H. (Ed.),
Dislocations and Properties of Real Materials. The Institute of Metals, London, pp. 125–143.
Kok, S., Beaudoin, A.J., Tortorelli, D.A., 2002a. A polycrystal plasticity model based on the mechanical
threshold. International Journal of Plasticity 18 (5-6), 715–741.
Kok, S., Beaudoin, A.J., Tortorelli, D.A., 2002b. On the development of stage IV hardening using a
model based on the mechanical threshold. Acta Materialia 50 (7), 1653–1667. APR.
Kok, S., Beaudoin, A.J., Tortorelli, D.A., Lebyodkin, M., 2002c. A finite element model for the Portevin-
Le Chatelier effect based on polycrystal plasticity. Modelling and Simulation in Materials Science and
Engineering 10 (6), 745–763.
Kondo, K., 1952. On the geometrical and physical foundations of the theory of yielding. Proceedings
Japan National Congress of Applied Mechanics 2, 41–47.
Kondo, K., 1955. Non-Riemannian geometry of imperfect crystals from a macroscopic viewpoint. In:
Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782 781
Kondo, K. (Ed.), RAAG Memoirs of the Unifying Study of Basic Problems in Engineering and
Physical Science by Means of Geometry, Vol. 1. Gakuyusty Bunken Fukin-Kay, Tokyo.
Kröner, E., 1960. Allgemeine kontinuumstheorie der Versetzungen und Eigenspannungen. Archive for
Rational Mechanics and Analysis 4, 273–334.
Lasry, D., Belytschko, T., 1988. Localization limiters in transient problems. International Journal of
Solids and Structures 24 (6), 581–597.
Lee, E.H., 1969. Elastic-plastic deformation at finite strains. Journal of Applied Mechanics 36, 1–6.
Liu, B, Qiu, X., Huang, Y., Hwang, K.C., Li, M., Liu, C., 2003. The size effect of void growth in ductile
materials. Journal of the Mechanics and Physics of Solids (in press).
Lloyd, D.J., 1994. Particle reinforced aluminum and magnesium matrix composites. International
Materials Reviews 39, 1–23.
Lücke, K., Mecking, H., 1973. Dynamic recovery. In: Reed-Hill, R.E. (Ed.), The Inhomogeneity of Plastic
Deformation. American Society for Metals, p. 223.
Ma, Q., Clarke, D.R., 1995. Size dependent hardness of silver single crystals. Journal of Materials
Research 10, 853–863.
Mariano, P.M., 2002. A note on Ceradini-Capurso-Maier’s theorem in plasticity. International Journal of
Plasticity 18 (12), 1749–1773.
McElhaney, K.W., Vlasssak, J.J., Nix, W.D., 1998. Determination of indenter tip geometry and indenta-
tion contact area for depth-sensing indentation experiments. Journal of Materials Research 13, 1300–
1306.
Mindlin, R.D., 1964. Micro-structure in linear elasticity. Archive Rational Mechanics Analysis 16, 51–78.
Mindlin, R.D., 1965. Second gradient of strain and surface tension in linear elasticity. International
Journal of Solids and Structures 1, 417–438.
Mühlaus, H.B., Aifantis, E.C., 1991. A variational principle for gradient plasticity. International Journal
of Solids and Structures 28 (7), 845–857.
Nabarro, F.R.N., Basinski, Z.S., Holt, D.B., 1964. The plasticity of pure single crystals. Advances in
Physics 13, 193–323.
Niordson, C.F., Hutchinson, J.W., 2003. On lower order strain gradient plasticity theories. European
Journal of Mechanics. A/Solids (in press).
Nix, W.D., 1989. Mechanical properties of thin films. Materials Transactions 20A, 2217–2245.
Nix, W.D., 1997. Elastic and plastic properties of thin films on substrates: nanoindentation techniques.
Materials Science and Engineering A 234, 37–44.
Nix, W.D., Gao, H., 1998. Indentation size effects in crystalline materials: A law for strain gradient
plasticity. Journal of the Mechanics and Physics of Solids 46 (3), 411–425.
Nye, J., 1953. F, Some geometrical relations in dislocated crystals. Acta Metallurgica et Materialia 1, 153–
162.
Pan, J., Rice, J.R., 1983. Rate sensitivity of plastic flow and implications for yield-surface vertices.
International Journal of Solids and Structures 19, 973–987.
Peirce, D., Asaro, R.J., Needleman, A., 1983. Material rate sensitivity and localized deformation in
crystalline solids. Acta Metallurgica et Materialia 31, 1951–1976.
Poole, W.J., Ashby, M.F., Fleck, N.A., 1996. Micro-hardness of annealed and work-hardened copper
polycrystals. Scripta Materialia 34, 559–564.
Press, W.H., Flannery, B.P., Teukolsky, S.A., Vetterling, W.T., 1986. Numerical Recipes. Cambridge
University Press, Cambridge.
Qiu, X., Huang, Y., Nix, W.D., Hwang, K.C., Gao, H., 2001. Effect of intrinsic lattice resistance in strain
gradient plasticity. Acta Materialia 49 (19), 3949–3958.
Qiu, X., Huang, Y., Wei, Y., Gao, H., Hwang, K.C., 2003. The flow theory of mechanism-based strain
gradient plasticity. Mechanics of Materials 35 (3-6), 245–258.
Rice, J.R., Tracey, D.M., 1969. On the ductile enlargement of voids in triaxial stress fields. Journal of the
Mechanics and Physics of Solids 17, 201–217.
Saha, R., Xue, Z., Huang, Y., Nix, W.D., 2001. Indentation of a soft metal film on a hard substrate:
strain gradient hardening effects. Journal of the Mechanics and Physics of Solids 49, 1997–2014.
Seeger, A., 1958. In: Flügge, S. (Ed.), Handbuch der Physi, Vol. VII/2. Springer, p. 1.
782 Y. Huang et al. / International Journal of Plasticity 20 (2004) 753–782
Shi, M., Huang, Y., Jiang, H., Hwang, K.C., Li, M., 2001. The boundary-layer effect on the crack tip field
in mechanism-based strain gradient plasticity. International Journal of Fracture 112, 23–41.
Shizawa, K., Zbib, H.M., 1999. A thermodynamical theory of gradient elastoplasticity with dislocation
density tensor. I: Fundamentals. International Journal of Plasticity 15 (9), 899–938.
Shu, J.Y., Fleck, N.A., 1999. Strain gradient crystal plasticity: size-dependent deformation of bicrystals.
Journal of the Mechanics and Physics of Solids 47 (2), 297–324.
Sluys, L.J., de Borst, R., Mühlaus, H.B., 1993. Wave-propagation, localization and dispersion in a
gradient-dependent medium. International Journal of Solids and Structures 30 (9), 1153–1171.
Steinmann, P., 1996. Views on multiplicative elastoplasticity and the continuum theory of dislocations.
International Journal of Engineering Science 34, 1717–1735.
Stelmashenko, N.A., Walls, A.G., Brown, L.M., Milman, Y.V., 1993. Microindentation on W and Mo
oriented single crystals: an STM study. Acta Metallurgica et Materialia 41, 2855–2865.
Stolken, J.S., Evans, A.G., 1998. A microbend test method for measuring the plasticity length scale. Acta
Materialia 46 (14), 5109–5115.
Suresh, S., Nieh, T.G., Choi, B.W., 1999. Nano-indentation of copper thin films on silicon substrates.
Scripta Materialia 41, 951–957.
Swadener, J.G., George, E.P., Pharr, G.M., 2002. The correlation of the indentation size effect measured
with indenters of various shapes. Journal of the Mechanics and Physics of Solids 50 (4), 681–694.
Taylor, G.I., 1934. The mechanism of plastic deformation of crystals. Part I.—theoretical. Proceedings of
the Royal Society of London A 145, 362–387.
Taylor, G.I., 1938. Plastic strain in metals. Journal of the Institute of Metals 62, 307–324.
Tymiak, N.I., Kramer, D.E., Bahr, D.F., Wyrobek, T.J., Gerberich, W.W., 2001. Plastic strain and strain
gradients at very small indentation depths. Acta Materialia 49 (6), 1021–1034.
Wang, W., Huang, Y., Hsia, K.J., Hu, K.X., Chandra, A., 2003. A study of microbend test by strain
gradient plasticity. International Journal of Plasticity 19 (3), 365–382.
Xue, Z., Huang, Y., Li, M., 2002a. Particle size effect in metallic materials: a study by the theory of
mechanism-based strain gradient plasticity. Acta Materialia 50 (1), 149–160.
Xue, Z., Saif, M.T.A., Huang, Y., 2002b. The strain gradient effect in microelectromechanical systems
(MEMS). Journal of Microelectromechanical Systems 11 (1), 27–35.
Zbib, H.M., Aifantis, E.C., 1988. On the localization and postlocalization behavior of plastic deforma-
tion. Part I. On the initiation of shear bands; Part II. On the evolution and thickness of shear bands;
Part III. On the structure and velocity of Portevin-Le Chatelier bands. Res Mechanica 23 (2–3), 261–
277, 279–292, 293–305.
Zhu, H.T., Zbib, H.M., 1995. Flow strength and size effect of an Al-Si-Mg composite model system under
multiaxial loading. Scripta Metallurgica et Materialia 32, 1895–1902.
Zhu, H.T., Zbib, H.M., 1997. Strain gradients and continuum modeling of size effect in metal matrix
composites. Acta Mechanica 121, 165–176.