Sunteți pe pagina 1din 94

Arch. Comput. Meth. Engng.

Vol. 11, 1, 3-96 (2004) Archives of Computational


Methods in Engineering
State of the art reviews

Modelling the Plastic Anisotropy of Metals


A.M. Habraken
Senior Research Associate FNRS
Département de Mécanique des Matériaux et Structures
Université de Liège
1, Chemin des Chevreuils, Bât B52/3, 4000 Liège, Belgique
e-mail: Anne.Habraken@ulg.ac.be

Summary
This work is an overview of available constitutive laws used in finite element codes to model elastoplastic
metal anisotropy behaviour at a macroscopic level. It focuses on models with strong links with the phenom-
ena occurring at microscopic level. Starting from macroscopic well-known models such as Hill or Barlat’s
laws, the limits of these macroscopic phenomenological yield loci are defined, which helps to understand
the current trends to develop micro-macro laws. The characteristics of micro-macro laws, where physical
behaviour at the level of grains and crystals are taken into account to provide an average macroscopic answer
are described. Some basic knowledge about crystal plasticity models is given for non-specialists, so every
one can understand the microscopic models used to reach macroscopic values. The assumptions defining
the transition between the microscopic and macroscopic scales are summarized: full constraint or relaxed
Taylor’s model, self-consistent approach, homogenisation technique. Then, the two generic families of micro-
macro models are presented: macroscopic laws without yield locus where computations on discrete set of
crystals provide the macroscopic material behaviour and macroscopic laws with macroscopic yield locus
defined by microscopic computations. The models proposed by Anand, Dawson, Miehe, Geers, Kalidindi
or Nakamachi belong to the first family when proposals by Montheillet, Lequeu, Darrieulat, Arminjon,
Van Houtte, Habraken enter the second family. The characteristics of all these models are presented and
commented. This paper enhances interests of each model and suggests possible future developments.

1 INTRODUCTION

Micro-macro modelling is a typical multi-disciplinarily field as it requires mechanical, met-


allurgical and computational knowledge. Here “micro” means at the crystal level, not at the
atomistic scale. This state-of-the-art review mainly comes out of “Aggregation in Higher
Education Thesis”, Habraken (2001). It aims to gather the necessary information to un-
derstand the various proposals found in the literature in the field of micro-macro models
dedicated to the anisotropic plastic behaviour of metals.
However the first question is probably why micro-macro models appear around 1985
(Lequeu et al. 1987a and b, Arminjon, 1988, Van Houtte 1988, Mathur & Dawson 1989)
and are currently more and more developed (Miehe et al. 1999, Geers et al. 2000, Feyel &
Chaboche 2000, Hoferlin 2001). What are the interests of industrial groups like ARCELOR,
CHRYSLER, ALCOA, ... in such models? In practice, different metal forming processes
such as deep drawing or folding processes are required to manufacture automotive parts,
beverage or food cans, steel sheet panels used in aeronautics or civil engineering applications.
Computer models try to replace the expensive and time-consuming trial-and-error methods
used in conventional design. The Finite Element Method (FEM) is quite successful to
simulate metal forming processes, but accuracy relies on the used constitutive law and
the material parameter identification. Following paragraphs emphasize the necessity of
efficient constitutive models to predict final shape, springback, rupture, fatigue behaviour
of the forming products.
For instance, final shape of the product is strongly linked to the plastic material flow.
Plastic anisotropy explains the ondulated rims called ears and present in a cup produced
by cylindrical tools applied on a circular blank. Figure 1 describes the typical case of a

2004
c by CIMNE, Barcelona (Spain). ISSN: 1134–3060 Received: December 2002
4 A.M. Habraken

mild steel supplied by SOLLAC and compares experimental and simulated earing profiles.
The classical isotropic elasto-plastic von Mises’law predicts no rim. In Figure 1, the simple
quadratic anisotropic elasto-plastic Hill’s law simulates a wrong earing profile when Ferron-
Tourki’s proposal (Tourki et al. 1996) is closer to experimental points. Such ears have to
be removed in order to produce a usable can. So material suppliers are required by can
manufactures to minimize the earing tendency of their material and model suppliers are
required to provide accurate earing predictions.

50.00

MILD STELL Ferron - Tourki


Hill 1948
48.00 Experimental points

46.00

Cup height (mm)


Cup 44.00
height
RD
42.00
angle

40.00

38.00
0.00 60.00 120.00 180.00 240.00 300.00 360.00
Angle from Rolling Direction (degrees)

Figure 1. Deep drawing of a cylindrical cup: comparison between experimental


and calculated earing profiles (from Tourki et al. 1996)

Springback (Figure 2a) is the shape distortion of sheet metal when removed from the
die after forming. It is a major problem in stamping high-strength steel and even more so
with aluminium sheet used to manufacture lightweight vehicles. Accurate analysis of this
phenomenon requires that at the end of the forming process before removal from the die
and trimming, the stress state and material state (isotropic and kinematic hardening) are
correctly computed. If the yield locus size and shape and the stress field are well known,
then the elastic return inducing springback during the die removal can be simulated. Fig-
ure 2c presents ADINA results for the two-dimensional draw bending benchmark problem
proposed by the organizers of NUMISHEET 93 Conference.
During the forming operation, rupture events can appear. Research has already brought
some solution. For instance, Forming Limit Diagram (FLD) helps to define a safe process. It
divides the plane of major principal strain versus minor principal strain into a safe zone and
a failure zone (see Figure 3 adapted from Barata da Rocha 1985). Again, this approach
requires accurate models to predict strain field during the forming operation. The FLD
prediction itself provides a very sensitive way to check behaviour model coupled or not
with FEM. The following references give only a small idea of the important effort dedicated
to theoretical FLD prediction: Narasimhan & Wagoner 1991, Boudeau et al. 1996, Hora et
al. 1996, Cayssials 1998 and 1999, Hoferlin et al. 1998, Vegter et al. 1999a. This proves the
interest in FLD from the automotive or can deep drawing industries. Note that the model
proposed by Cayssials 1999 is very effective for classical steels. Coupling plastic instability
theory with a damage approach, Cayssials’s model accurately predicts FLD of the metal
sheet with data measured by tensile tests.
When the piece manufacturing has been successful, the piece is used. Here, it is a key
factor to provide fatigue life prediction under periodical or random multi-axial stress state.
Modelling the Plastic Anisotropy of Metals 5

Figure 2. Draw bending test. (a) Shape of a metal sheet: initial, after stamping, after
spring back; (b) measurement technique; (c) comparison between experimental
data (circles and squares) and simulated results (dashed lines), (from Kawka
& Bathe 2001)

As checked by Weber et al. 1999, the forming process can strongly modify the material
state and by consequence, its fatigue characteristics. So life estimation techniques rely on
the final plastic state of the material (accurate size and shape of its yield locus).
Previous comments clearly show why accurate behaviour models are required in FEM
to simulate forming operations. A complete answer to the question why developing micro-
macro approaches, must however explain the interests of such laws by comparison with
classical phenomenological laws. Advantages and limitations of phenomenological constitu-
tive laws are detailed in Section 2. As such models roughly consist of the fitting of functions
on experimental results, it is not surprising that they provide only average behaviour when
too simple functions are chosen and that they are quite difficult to identify when complex
functions covering large fields of behaviour are applied. When models are based on the real
physical phenomena, their internal parameters are usually less numerous and their identi-
fication of reasonable difficulty. As crystal plasticity has been studied for more than 100
years (Ewing & Rosenhain 1900), material scientists know quite accurately the behaviour
of monocrystals. Average field theory and homogenisation theory, steps to polycrystalline
behaviour were already proposed by Sachs 1928, Taylor 1938 in metal. Such knowledge
as well as FEM and computer developments explain that about 20 years ago, micro-macro
6 A.M. Habraken

zo ne sstr
treetc
tchh
i
ined eedd z
ttrra zoo
rress ffailure zone
ailure zone nnee
εε11

0
0.5
=0
α == 0.5
state β =
plane α
ββ ==

strain state
αα

αα ==
==
--11

-- 00
ββ =

00
= 1
=-

plane
αα =
..55

strain
-11

ppuu aarr

uunn
1
= 1
ββ =
rree
aall
tteen

ia
sshh

iaxxi
rticric
nssio
ee

e t

a
iall
ionn mme ionon
yymm ennssi
SS
ltte
saf
e zone
safe zone iaal
iaiaxxi
bb

minor
minorprincipal
principalstrain
strain
εε2
2

Fig 3 Figure 3. Forming Limit Diagram for an isotropic material, whith α stress ratio
(α = σ2 /σ1 ) and β strain increment ratio (β = dε2 /dε1 ), adapted from
Barata da Rocha 1985
M
T

models have appeared. 1.0 T


M
This paper summarizes the required Hill notions
1948 about crystal plasticity in Section 3, then
σy / σF

the different possibilities to go from


Brass
the crystal to the polycrystal are presented in Section 4.
More details are provided on FEM M - Momicro-macro
S2 Hill 1979models without or with microscopic yield
locus in Sections 5 and 6 respectively.
T - Teflon Finally, Section 7 concludes with comments on the
0.5
interests of the different model families and the perspectives of the micro-macro models in
Hosford 1979
metal forming simulations.

2 ANISOTROPIC PHENOMENOLOGICAL YIELD LOCI


The yield locus is the boundary between the elastic and plastic domains. It is a continuous
0 0.5 1.0
surface in stress space, Fp (σij ) = 0, corresponding
σ x / σ F to all stress states σij that cause yield-
ing. During plastic deformation, the updated yield locus will expand or contract, translate
and
Fig 4distort. This phenomenon is described by a hardening rule. In this section, the ini-
tial shape of the yield locus associated with the first yielding is analysed. Experimental
evidence and theoretical considerations concerning the plastic behaviour of materials have
led to some restrictions on the mathematical representation of a yield surface. Following
Bridgman’s experimental observations (Bridgman 1923, 1952), hydrostatic pressure (25000
bars) does not induce plasticity in metals. More recently, Barlat et al. 1991 have measured
a relative density change due to plastic deformation of the order of 10−3 . Extrapolating
an approximate limit analysis of a porous medium, Gurson’s model 1977 proposes a plastic
behaviour law affected by pressure. His goal is fracture prediction but far from rupture, the
pressure effect is limited. In a hot sintering process, the initial porosity can be very high
and the pressure effect is important. However, in general cases of sound classical metals
characterized by low porosity, yield surfaces are taken to be pressure independent. Drucker
1951 showed that, based on a stability postulate, the yield surface must be convex. The
relationship between the components of the strain-rate tensor and the stress tensor is called
the flow rule.
Modelling the Plastic Anisotropy of Metals 7

Another consequence of Drucker’s 1951 postulate is that the flow rule for stable materials
is directly associated to the yield criterion: the strain rate vector is orthogonal to the yield
surface. The associated flow rule is:
∂Fp
ε̇pij = λ̇ (1)
∂σij

Hecker 1976 or Phillips 1979, who reviewed numerous critical experiments to assess the
yield surface shape, found that the normality rule was never violated. This is also confirmed
by the work of Hayakawa & Murakami 1998 in damage mechanics. Indeed, all the yield
surfaces listed hereafter respect these 3 characteristics. Table 1 summarizes the evolution
of the well-known Hill’s 1948 equation, then Table 2 introduces the models of Barlat and
Karafellis. Finally, Vegter’s models are shortly presented since they seem to be well vali-
dated and adapted to FEM computations. In addition to these tables, some comments help
to understand the origin of the presented models, their parameters identification method
as well as their experimental validation. Let us note that except for Karafillis’model and
Vegter’s model, all the others are restricted to anisotropic materials exhibiting orthotropic
symmetry, i.e. to materials which possess 3 mutually orthogonal planes of symmetry at
every point. Actually, most scientists agree that this restriction is not a very significant
constraint because most mechanically processed materials are orthotropic in their initial
state. Typical examples are rolled sheets or plates.
Reviews of phenomenological anisotropic models can be found in Vial et al. 1983,
Kobayashiet al. 1985, Barlat 1987, Arminjon et al. 1994, Mahmudi 1995, Kuwabara & Van
Bael 1999, Banabic 2000, Habraken 2001, Barlat et al. 2002. Note that sheet anisotropy
is usually characterized by the Lankford’s coefficient. This coefficient is determined by
uni-axial tensile tests on sheet specimens in the form of a strip. This anisotropy coefficient
is defined by:
ε2
r= (2)
ε3
where ε2 , ε3 are the strains in the width and thickness directions, respectively. This ratio
is strongly correlated with the deep drawability of the sheet.

2.1 Hill’s Approach


Table 1 is dedicated to Hill’s research. The classical quadratic yield criterion, Hill 1948 (see
Table 1) is well suited to specific metals and textures, but lacks flexibility. This criterion
gives a better correlation with metals having an average Lankford’s coefficient r̄ greater
than 1 (steel) but is less acceptable when r̄ is less than 1 (aluminium). This average value
is computed by:
1
(r0 + 2r45 + r90 )
r̄ = (3)
4
It can be seen that if r̄ < 1, the yield locus predicted by Hill 1948 criterion is located
inside the one given by von Mises; if r̄ > 1, the Hill 1948 yield locus is outside the von Mises
yield locus. The angle α defines a direction, the reference is always the Rolling Direction
(RD). For instance, the notation σα means the current yield stress in uni-axial tension along
a direction at an angle α with the Rolling Direction.
Only 3 tensile tests at 0◦ , 45◦ , 90◦ are required to determine the material parameters of
Hill 1948 in plane stress state but a biaxial test is necessary for the more recent versions.
Figure 4 compares experimental data and analytical yield loci for experiments performed
on brass. Cases of a steel sheet and an aluminium one are treated by Figure 5.
8

Analytical description Main features, drawbacks Advantages


General quadratic equation with 6 Basic assumptions easy to understand.
2 2 2 parameters, crude approximation Simple to implement in a FEM code.
F (σ y − σ z ) + G (σ z − σ x ) + H (σ x − σ y )
Hill 1948 2 2 2 2
of the yield surface shape Only 3 tensile tests at 0°, 45°, 90° required
+ 2 Lτ yz + 2 Mτ zx + 2 Nτ xy = 2σ F computed from polycrystalline to determine the material parameters in
models. plane stress state.
a a a Hill 48 with variable exponent
F σy −σz + Gσz −σx + H σx −σy (non integer) and no shear stress Able to describe material with r <1 and
a a term, orthotropic material axes and
σ b > σ F or the reciprocal (anomaly A1).
Hill 1979 + A 2σ x − σ y − σ z + B 2σ y − σ z − σ x principal stress axes must be
superimposed. Analytical expression of the associated flow
a a
+ C 2σ z − σ x − σ y = σF Biaxial test is required for rule.
identification.
a
a −1 Extension of Hill 79 that
a σ a 2 suppresses its limitation in loading
b 2 2
σ +σ + σ −σ + σ +σ
1 2 a 1 2 1 2 directions but is only defined for Able to describe anomaly A1.
Hill 1990 τ
plane stress case.
2 2 2 2 Biaxial test is required for
{− 2 A(σ 1 − σ 2 ) + B (σ − σ ) cos 2 β = ( 2σ b ) }
1 2
identification.

Table 1. Hill’s models


2 2
σx σ xσ y σy Only defined for plane stress case Able to describe anomaly A1.
−C + +
2 2 and loads applied along orthotropic Able to model different r0 and r90 values,
σ0 σ 0σ 90 σ 90
Hill 1993 axes (no shear component). when uni-axial stresses in rolling σ0 and
 ( Aσ x + Bσ y  σ xσ y Biaxial test is required for
=1
transversal σ90 direction are almost equal
( A + B ) −  identification.
 σb  σ 0σ 90 and reciprocal (anomaly A2).
A,B,C,F,G,H,L,M,N, a = material parameters,
σ x ,σ y ,σ z ,τ xy ,τ zx ,τ yz = stress components in the material orthotropic axes,
σ 1 ,σ 2 = principal stress components oriented by an anticlockwise angle β with the RD axis,
σ b = yield value under plane equibiaxial stress state,
τ = yield stress in pure shear test parallel to orthotropic axes (plane stress case),
σF= yield stress under uni-axial tension in a reference direction.
A.M. Habraken
Analytical description Main features, drawbacks Advantages
a a a a Good correlation with yield locus
A K1 + K 2 + A K1 − K 2 + ( 2 − A) 2 K 2 = 2σ F (a) Generalization of isotropic Hosford’s
1972 equation with a shear term. predicted by polycrystalline
Barlat 1989 K1 = σ x + hσ y / 2
( ) (b) Defined by 4 parameters, a, h, p, A. model, especially in the range of
Plane stress case only. equibiaxial tension.
K2 = [(σ x − hσ y )/ 2]2 + ( pσ xy ) 2 (c)
a a a a Generalization of isotropic Hosford’s Adapted for 3D stress state and
s1 − s2 + s2 − s3 + s3 − s1 = 2σ F
(a) 1972 equation with a shear term. orthotropic symmetric material.
with s = L σ (b) Defined by 6 anisotropy coefficients
c1 to c6 + exponent a Easy to implement in FEM
( c 2 + c 3 ) / 3 − c3 / 3 − c2 / 3 0 0 0  codes.
Modelling the Plastic Anisotropy of Metals

The 4th order tensor L can be


− c / 3 ( c 3 + c1 ) / 3 − c1 / 3 0 0 0

Barlat 1991  3  represented by a 6x6 matrix in the
axis of orthotropic symmetry (RD, Good predictions of the
− c 2 / 3 − c1 / 3 ( c1 + c 3 ) / 3 0 0 0 
L =   (c) TD, ND). distribution of uni-axial yield
 0 0 0 c4 0 0  stress and coefficient r in the
The methodology to reach the
 0  plane of the sheet.
0 0 0 c5 0 associative flow rules is rather
  difficult.
 0 0 0 0 0 c6 
2a 2a 2a 2a Good correlation with yield locus
Φ1 = s1 − s 2 + s 2 − s3 + s3 − s1 = 2σ F (a) Generalization of Barlat’s 91 work to
non orthotropic material. predicted by polycrystalline
2a
2 +2 model.
Φ 2 = s12 a + s22 a + s32 a = σ F2 a (b) In case of plane stress state and
Karafillis 32 a orthotropic material, it is based on 8 Good predictions of the
1993 32 a parameters, once more than Barlat distribution of uni-axial yield
Φ = (1 − C )Φ1 + C 2 a −1
Φ 2 = 2σ F2a

Table 2. Barlat’s and Karafillis’ models


(c) 1991. stress and coefficient r in the
2 +1
s = L σ with L tensor 4th order (d) plane of the sheet.
a a a a Extension of Barlat 1991 to model Good prediction of earing in
Φ = A s1 − s2 + B s2 − s3 + C s3 − s1 = 2σ F
Barlat 1997 high pure shear yield stress and to FEM simulations of deep-
with s and L defined as in Barlat 1991 better fit r0 and r90 (plane stress case). drawing of cylindrical cups.
s1, s2, s3 = eigenvalues of tensor s, σ x , σ y , σ xy = stress components on the orthotropic axes,
σ = stress tensor in orthotropic axes, σF = uni-axial plastic stress in a reference direction,
L = linear operator.
9
10 A.M. Habraken

M
T

1.0 T
M
Hill 1948

σy / σF Brass
M - Mo S 2 Hill 1979
T - Teflon
0.5
Hosford 1979

0 0.5 1.0
σx / σF

Figure 4. Brass yield loci normalized by uniaxial tension at strain of 0.1, biaxial
experimental points, horizontal and vertical tangent defined by plane
strain tests (from Vial et al. 1983)

Experiment Experiment

Hill 1993
160
Hill 1993
200 Hill 1948 140

120
von Mises
150 von Mises
100
σy σy
80 Hill 1948
100
60

40
50
20

0 0
0 50 100 150 200 0 40 80 120 160
σx σx

a) ST 1405 b) Al Mg Si 1

Figure 5. Biaxial tensile points compared with analytical yield loci for a steel and
an aluminium alloy (from Banabic et al. 1999)

2.2 Karafellis-Barlat-Banabic Family


The presentation of Karafillis et al. 1993 is reproduced as it provides a nice general frame
to explain his proposal as well as Barlat’s 1989, 1991, 2002. In fact, the first idea is to
find a general form of isotropic yield surfaces. Then, this form is extended to anisotropic
behaviour with the help of tensor s, a projection of the stress tensor σ .
Modelling the Plastic Anisotropy of Metals 11

First looking at isotropic yield surfaces, Karafillis recalls that Mendelson 1968 has de-
rived from symmetry and convexity considerations, bounds in an isotropic yield surface.
The lower bound coincides with the maximum shear stress yield surface as described by
Tresca’s 1864 criterion, whereas the upper bound was proposed by Hosford 1972. In the de-
viatoric plane section of the yield surface, von Mises’ yield circle lies between upper bound
and lower one (Figure 6). This deviatoric plane, also called Π plane, is perpendicular to
the line representing hydrostatic pressure state and contains the stress origin point. The
relations to compute the projection of a stress state onto the Π plane can be found in Van
Bael 1994 or Khan & Huang 1995.
σ
^
π - plane 2

von Mises

Tresca

upper bound

σ
^
3
σ
^
1

Figure 6. The upper bound, the lower bound and the von Mises’ yield surfaces
in the Π-plane (from Karafillis et al. 1993)

Isotropic yield surfaces lying between the bounds defined by von Mises’ yield surface
and Tresca’s yield surface can be mathematically described by Hosford’s 1972 criterion. His
proposal is a modification of the von Mises’ mathematical description of a yield surface,
where an exponent other than 2 is used:

(σ̂1 − σ̂2 )2a + (σ̂2 − σ̂3 )2a + (σ̂3 − σ̂1 )2a = 2σF2a (4)
with a integer > 1, σ̂1 σ̂2 σ̂3 the principal values of σ̂ the deviatoric stress tensor and σF
the yield stress under uni-axial tension. This relation is equivalent to von Mises’ equation
when a = 1 and to Tresca’s equation when a = ∞. Intermediate values of a describe all
the yield surfaces lying between the two proposals (Figure 7). Equation (4) is identical to
relation “Karafillis 1993 (a)” in Table 2 if σ̂ is substituted by s.
In fact, the isotropic surfaces lying between von Mises’ yield surface and the upper bound
(Figure 8) are described by relation “Karafillis 1993 (b)” in Table 2 if s is substituted by σ̂ .
When a = 1 , the yield surface corresponds to von Mises’ yield surface whereas when a = ∞,
the upper bound yield surface is recovered. So to describe a generic isotropic yield surface
lying between the lower bound and the upper bound, a generalized mathematical relation
mixing Karafillis 1993 relations (a) and (b) in Table 2 is proposed (with s substituted
by σ̂ ). It is the goal of relation (c) where the constant C (Figure 9) belongs to [0,1]. By
varying the value of the mixing factor C, a family of yield surfaces is created which spans
the space between the 2 bounds, set by the choice of a (Figure 9). In this approach, there
are two parameters C and a that may be used to “adjust” the shape of this isotropic yield
locus whereas the other criteria use only one parameter (exponent a for Hosford 1972) for
this purpose.
12 A.M. Habraken

1.50

1.25

1.00

0.75 a = 1 (von Mises)


σ2 / σF a=3
0.50 a=5
a=8
0.25 a = 16
a = +∝ (Tresca)
0.00
σ1 - σ2
- 0.25
section
- 0.50 Φ1
- 0.75
- 0.50 0.00 0.50 1.00 1.50
σ1 / σF

Figure 7. Influence of parameter a in φ1 function (from Karafillis et al. 1993)

1.50

1.25

1.00

0.75 a = 1 (von Mises)


a=3
σ2 / σF

0.50 a=5
a=8
0.25 a = 15
a = + ∝ (Upper
0.00 bound)
σ1 - σ2
- 0.25
section
- 0.50 Φ2
- 0.75
- 0.50 0.00 0.50 1.00 1.50
σ1 / σF

Figure 8. Influence of parameter a in φ2 function (from Karafillis et al. 1993)

The extension of the previous isotropic approach to anisotropic cases depends on the
choice of s tensor. Different proposals by Barlat, Karafellis, Banabic ... are known. A short
review is proposed hereafter.
Barlat et al. 1991 use a linear transformation of the 6 components of a stress state
(Barlat 1991 (b), Table 2). The obtained transformed stress state is then used in Hosford’s
1972 yield criterion (4). Different identification procedures are proposed in Banabic 2000.
Practically, a is larger or equal 6, depending on the anisotropy induced by texture and on
Modelling the Plastic Anisotropy of Metals 13

1.50

1.25

1.00

0.75
σ2 / σ F c=1
0.50 c = 0.99
c = 0.95
0.25 c = 0.75
c = 0.00
0.00

- 0.25
a = 15
- 0.50

- 0.75
- 0.50 0.00 0.50 1.00 1.50
σ 1 / σF

Figure 9. Influence of C parameter in global φ function (from Karafillis et al. 1993)

the crystal type. In fact, a can be tuned to optimise the prediction of the yield locus shape.
Barlat’s 1989 work already contains the same type of approach but is limited to plane
stress state (see Table 2). This formulation has the advantage of clearly showing the effect
of the shear component, which modifies the sections of the yield locus. Figure 10 presents
yield locus section of a material containing 50 % brass texture and 50 % of randomly
distributed grains. The 4 parameters A, h, p, a of Barlat’s 1989 model are determined as
explained by Barlat & Lian 1989 or Berg et al. 1998:
a = 6 for b.c.c. metals; a = 8 for f.c.c. metals (from polycrystalline model comparisons)
 
r0 r90 r0 r90
A =2−2 h= (5)
1 + r0 1 + r90 1 + r0 1 + r90
 1
σF 2 a
p= a
(6)
τ 2A + 2 (2 − A)
where σF is the yield stress in uni-axial tension, τ is the yield stress for pure shear. In
practice, p varies since the ratio σF /τ is not a constant function of the equivalent strain.
This affects the shape of the yield surface as demonstrated by Berg et al. 1998.
For anisotropic material, Karafillis et al. 1993 consider a tensor s , resulting of a linear
transformation L of the actual stress tensor σ . In this proposal, the transformed tensor s
is introduced in the general isotropic yield criterion defined by relation “Karafellis 1993 (c)”
in Table 2. The fourth order tensor L chosen by Karafillis is more flexible than the simple
choice made by Barlat and can be associated with material symmetries which range from
the lowest level of triclinic symmetry to the highest level of full symmetry. This provides
a potential for the representation of any anisotropic state of the material. More details on
symmetry properties and constraints applied to L can be found in Karafillis et al. 1993.
Finally, for orthotropic materials, 3 uni-axial tensile tests of specimens cut at 0◦ , 45◦ , 90◦
with respect to Rolling Direction allow to completely determine C and L. Remark that, in
this orthotropic case, Karafillis’ transformation tensor L reduces to Barlat’s 1991 (Table 2).
Figure 11 shows Karafillis’results compared with other models and experiments.
14 A.M. Habraken

1.50

1.00
Hill
1948
s = 0.0
0.50
s = 0.2
σy / σF
s = 0.3
s = 0.4
s = 0.432
0.00

a = 14
r 0 = 0.7
- 0.50
r 45 = 5.0

r 90 = 1.0

- 1.00
- 1.00 - 0.50 0.00 0.50 1.00 1.50
σ x / σF

Figure 10. Sections of Barlat & Lian 1989 yield surface, with s = σxy /σF (adapted from
Barlat & Lian 1989)

2.0
1.8 B=0
Experiment
1.6 Karafillis 1993
2008 - T4
Barlat 1991
1.4
TBH
1.2
1.0
r

0.8
0.6
0.4
0.2
0.0
0.0 15.0 30.0 45.0 60.0 75.0 90.0
ANGLE FROM ROLLINGS DIRECTION [DEG]

Figure 11. Curves of Lankford coefficient versus angle with the rolling direction for 2008-
T4 aluminum alloy , TBR =Taylor 1938-Bishop & Hill 1951 polycrystal model
(from Karafillis et al. 1993)

Hill 1948 (Table 1), Barlat 1989 and Karafillis 1993 (Table 2) were implemented in
the commercial code LS-DYNA3D. Andersson et al. 1999 use it to simulate deep drawing
tests of aluminum sheets and compare their FEM results to experiments. Their results show
that Karafillis’model yields the best agreement with the experimental results and that CPU
times are in the same range: Hill 1948 = 1, Barlat 1989 = 1.5 and Karafillis 1993 = 1.1.
Modelling the Plastic Anisotropy of Metals 15

Barlat’s 1997 proposal (Table 2) generalizes his 1991 model. For the plane stress case,
only 6 independent coefficients characterize anisotropy. They can be deduced from 4 tests:
uni-axial tension at 0◦ , 45◦ , 90◦ to the Rolling Direction and equal biaxial bulge test.
Let us note that Banabic et al. 2000 propose a new yield criterion BBC2000 for or-
thotropic sheet metals under plane stress conditions. It is derived from Barlat 1989 criterion.
Two additional coefficients have been introduced in order to allow a better representation:

[b(cΓ + dΨ)2a + b(cΓ − dΨ)2a + (1 − b)(2dΨ)2a ]1/2a = σF (7)


where a, b, c, d are material parameters, while Γ and Ψ are functions of the second and
third invariants of the fictitious stress tensor s defined by Karafellis (relation “Karafellis
1993 (d)” in Table 2).
Drucker’s 1949 isotropic criterion:

J23 − cJ32 = k 2 (8)


where J2 = 1/3trσ̂ 2 and J3 = 1/3trσ̂ 3 are the second and third invariants of the stress
deviator has been extended by Cacazu & Barlat 2001 to anisotropy. They replace the devi-
atoric stress invariants J2 and J3 by homogeneous functions of degree 2 and 3 respectively.
These functions are pressure insensitive, invariant to any transformation belonging to the
symmetry group of the material and reduce to J2 and J3 in isotropic conditions.
Barlat et al. 2002 propose to use two different linear transformations L and L of the
stress tensor σ to define 2 tensors s and s (see relation “Barlat 1991 (b)” Table 2). Then
they introduce them in isotropic functions:

Φ (s) = |s1 − s2 |a Φ (s) = |2s2 + s1 |a + |2s1 + s2 |a (9)


leading to the resulting anisotropic yield function:

Φ (s ) + Φ (s ) = 2σFa (10)


This approach is only developed in plane stress and uses 10 anisotropy coefficients, 2
are zero and the other 8 coefficients can be determined using stress and r values in tension
along three directions, the balanced biaxial flow stress σb and rb . The latter coefficient rb is
defined by analogy to Lankford’s coefficient but tensile test is replaced by a disk compression
test (see Barlat et al. 2002 for further details).
Banabic’s et al. 2000 (BBC2000), Cacazu & Barlat’s 2001 (CB2001) and Barlat et al.
2002 (Yld2000-2d) models have been identified for AA3103-0 aluminum alloy (Barlat et al.
2002). Figures 12 and 13 present comparison between experimental and predicted values
of the distributions of the uni-axial yield stress and Lankford’s coefficient with respect to
the angle with the Rolling Direction.

2.3 Vegter’s Approach


Vegter’s approach (Vegter et al. 1999a,b) proposes a yield criterion directly based on ex-
perimental data of the pure shear test, the uni-axial test, the plane strain test and the
biaxial test. Figure 14 clearly summarizes this criterion for a planar isotropic material
behavior. A complete yield surface is described by 12 Bezier interpolation functions. When
σ1 and σ1 are defined in such a way that σ1 ≥ σ2 , only the part of the surface beneath the
line σ1 = σ2 , is required. As the material is assumed to behave identically in tension and
compression, the part of the surface beneath the line σ1 = −σ2 can be derived from the
part above that line. As a result, the complete material yielding can be described using
16 A.M. Habraken

1.15 0.65
Yld2000-2d
CB2001
BBC2000
1.10

•Normalized stress
stress 0.6

r value
1.05


0.55
r value
1.00

3103-O

0.95 0.5
0 20 40 60 80
Angle from rolling

Figure 12. Curves of uniaxial yield stress and Lankford coefficient versus angle with the
rolling direction for AA31303-0 aluminum alloy (from Banabic et al. 2002)

1.50
3103-O
1.25

1.00
σ / σ

0.75
yy

0.50

Yld2000-2d
0.25
CB2001
BBC2000
0.00
0.00 0.25 0.50 0.75 1.00 1.25 1.50
σ / σ
xx

Figure 13. Sections of analytical yield surfaces and experimental points (circles)
for AA31303-0 aluminum alloy (from Banabic et al. 2002)

the measurements for a quarter of the yield surface. Vegter’s yield function is restricted to
plane stress state. It is defined as:

σveg (σx , σy , σxy ) − σyield = 0 (11)


where σveg is a kind of equivalent stress depending on the stress components in orthotropic
axes σx , σy , σxy and σyield is the yield stress.
In case of planar isotropic material behavior, σyield is the uni-axial yield stress σF . The
second order Bezier interpolations allow to describe the normalized yield surface. For a
Modelling the Plastic Anisotropy of Metals 17

σ2
Biaxial
H

Plane
strain
H
Uniaxial
σ1

Pure H
shear

Figure 14. Vegter’s yield locus in principal stress space, black circles= experimental
points, white circles = hinge points, adapted from Carleer et al. 1997

reference point 2 : p ref 2


σ2

hinge point : H

reference point 1 : p ref 1

σ1
Figure 15. Bezier curve in Vegter’s yield locus, adapted from Carleer et al. 1997

part going from reference point 1 to reference point 2 with hinge point H (Figure 15), one
has: σ1
= (1 − β)2 pref1
1 + 2β(1 − β)phinge
1 + β 2 pref2
1 (12)
σveg
σ2
= (1 − β)2 pref1
2 + 2β(1 − β)phinge
2 + β 2 pref2
2 (13)
σveg
where σ1 , σ2 are principal stresses, β is a curvilinear coordinate increasing from 0 to 1 and
representing the location on the curve between the reference points, prefj i and phinge
i are
respectively the components i of the reference point j and the hinge point. Both equations
(12) and (13) give an expression for σveg . Equating these expressions allows to find β
as a solution of a second order equation. This solution is chosen to satisfy the condition
0 ≤ β ≤ 1. The expression of σveg is then found by substituting β into (11).
For planar anisotropy, the 4 tests (pure shear, uni-axial, plane strain and biaxial tests)
are performed for directions of 0◦ , 45◦ , 90◦ from the Rolling Direction and an interpolation
function pref (α) is used to find interpolated reference points from the measured points prefα :

pref ref ref


0 + 2p45 + p90 pref − pref pref − 2pref ref
45 + p90
pref (α) = + 0 90
cos 2α + 0 cos 4α (14)
4 2 4
18 A.M. Habraken

Here α is the angle of the principal stresses to the Rolling Direction. The function pref (α)
can be regarded as a Fourier series with only 3 terms; it can be extended if orthotropic
properties are not respected. So, with the new interpolated reference points, Vegter’s yield
criterion can be constructed.
Figure 16 presents the benchmark of the Numisheet 96 conference. Vegter’s model
predicts minor strains along a symmetry axis of a Limiting Dome Height test that are very
closed from experiments for the studied draw quality mild IF steel. Figure 17 shows the
quality of FLD predictions with Vegter’s model.

0.02 experiment
Hill 1948
203.3 Vegter

minor strain
R 6.35 105.7

DIE

R 50.8
Punch - 0.08
0 60
132.6
original x-coordinate (mm)

Figure 16. Predicted and measured minor strains along a symmetry axis of a Lim-
iting Dome Height test, draw quality mild IF steel (Pijlman et al. 1998)

0.8

0.6

ε1
0.4
Vegter
Hill 1990
0.2 Hill 1993
Barlat 1991
Experiment
0
-0.4 -0.2 0 0.2 0.4 0.6
ε2
Figure 17. FLD predictions applied on bi-axial pre-strained specimens of an IF
steel (from Vegter et al. 1999a)

However the identification of Vegter’s model is quite expensive: pure shear, uni-axial,
plane strain and biaxial tests in one direction in case of planar isotropy or in 3 directions for
planar anisotropy. As pointed by Vegter et al. 1999b, it is important to reach high experi-
mental accuracy and to correct experimental results if non homogeneity or non isothermal
conditions happen (Piljman et al. 1999).
An important characteristic of Vegter’s model is its flexibility. For instance, Figure 18
presents a variant called Vegter-flat, in which the reference point corresponding to plane
strain is very close to the hinge point located between plane strain and biaxial states. The
Modelling the Plastic Anisotropy of Metals 19

A 1.1 B
x x xxx
1 x xx
xx
1 xx
x
x
σ2 / σF

xx

σ2 / σF
x
0.9 x
x
Vegter Vegter x
0.5 x
Hill 1990 Barlat 1991
0.8
Hill 1993 x Vegter - flat

x
0 0.7
0 0.5 1 0.7 0.8 0.9 1 1.1
σ 1 / σF σ1 / σF

Figure 18. Comparison of yield loci descriptions for Al-6000, A global yield locus, B zoom
around biaxial state (from Vegter et al. 1999a)

Figure 19. FLD predictions and measurements for Al-6000 (90◦ to RD) (from Veg-
ter et al. 1999a)

agreement with the measured Forming Limit Diagram shown on Figure 19 is significantly
improved. This last effect shows the extreme sensitivity of the FLD prediction to the shape
of the yield locus.

2.4 Summary of Described Phenomenological Models


To summarize this section devoted to phenomenological yield loci, scientists have found new
formulations of anisotropic yield loci that seem closer to experimental evidence. However,
the number of parameters used to determine these new criteria increases with their flexibil-
ity. Table 3 summarizes the experimental values necessary to identify the yield loci listed
above in plane stress state, in their initial shape, without hardening description. Usually,
the exponent a is fitted according to polycrystalline yield locus, common value are a = 6
for b.c.c. metals and a = 8 for f.c.c. metals. This table also tells if the model is limited to
plane strain or developed for three dimensional state and if it is able to simulate anomalies
A1 , A2 defined in Table 1.
Note that in some cases, the user can choose which data he uses to identify his model. For
instance, the parameters of Hill 1948 can be identified from the yield stresses or the Lank-
ford’s coefficients given by a uni-axial tensile tests. Choosing stress values or Lankford’s
20 A.M. Habraken

Name 3D σ0 σ45 σ90 σb τ r0 r45 r90 Addit. A1 A2


param.
Hill 1948 X X X X X
Hill 1979 X X X X X
Hill 1990 X X X X X X X
Hill 1993 X X X X X X
Barlat 1989 X X X X X
Barlat 1991 X X X X X X
Karafellis 1993 X X X X X X X X X
Barlat 1997 X X X X X X X X X
Banabic 2000 X X X X X X X X X
Cacazu 2001 X X X X X X X X σ30 σ75 X X
r30 r75
Barlat 2002 X X X X X X X rb X X
Vegter 1999 X X X X in X in plane strain X X
3 dir. 3 dir. in 3 dir.

Table 3. Identification of phenomenological models

coefficients is not equivalent and provides two sets of parameters. One set of parameters
recovers stress information and the other strain information. This apparent difficulty comes
from the approximations of this phenomenological model, unable to reproduce experiments
with perfect accuracy.
2.5 Discussion on Phenomenological Yield Criteria
In practice, the specific strain path or loading history undergone by the material induces
micro-structural and textural evolutions. The initial phenomenological yield criteria de-
scribed in the preceding sections must be associated with hardening rules to model plastic
deformation beyond plasticity initiation. Hardening models can be isotropic or anisotropic.
The former corresponds to an expansion of the yield surface without distortion. Any other
form of hardening, such as kinematic hardening, which is defined by the translation of the
yield surface, is anisotropic. This type of hardening leads to different properties in different
directions after deformation, even if the material is initially isotropic. The hardening mod-
els depend on internal variables and allow to compute the new size, shape and position of
the yield locus. The updated yield locus connect points representing stress states charac-
terized by identical values of internal variables. For simple macroscopic hardening models,
the plastic work is often taken as the only internal state variable, which validates the use of
the work contours to measure updated yield loci, even though this surface connects stress
points related to different material textures. However, for hardening approaches with strong
microscopic roots, the equality of internal hardening state variables effectively implies that
all the points connected by the yield locus are related to the same material state.
Note that for sheet forming simulations, Barlat et al. 2002 propose a kind of alternative
approach to accurate hardening models. They use stresses (called flow stresses) and r
values defined at a given amount of plastic work to characterize the average behaviour of
the material over a finite deformation range. In this case, it would be more appropriate
to talk about flow function and flow surface instead of yield function and yield surface
although, mathematically, yield or flow functions are identical.
Classical arguments for the use of phenomenological yield loci are:
1. Easy to understand for mechanical engineers;
Modelling the Plastic Anisotropy of Metals 21

2. Available in commercial codes (Hill 1948 is already in most of them);


3. Low CPU time and storage memory requirement compared to micro-macro approach;
4. Easy to implement in a Finite Element code;
5. Accurate enough;
6. Low number of parameters, easy to identify.
If the two first points are clearly true, the other ones are a source of controversy:
3. With hardware improvement, parallel computation and efficient formulation of micro-
macro models, the gap concerning CPU time decreases, even if it is still a strong
limitation for some of them.
4. As reported by Banabic 2000, all models are not straightforward to implement, some
have a quite unfriendly form of the yield function and require mathematical abilities.
5. The list of models should stop growing if scientists were happy with their accuracy.
According Barlat et al. 2002 because mechanical data are used as input, phenomeno-
logical models can be more accurate than polycrystalline models when the strain
amount is moderate, which is generally the case for sheet forming. However, for
larger strains or for abrupt strain path changes, microstructure evolution is an issue
and is the subject of much research at the present time (Teodosiu and Hu 1998, Lopes
et al. 2002). Because polycrystalline models can track the lattice rotation of each
individual grain, the material anisotropy is naturally evolutional, which makes this
approach very attractive.
6. To increase accuracy, authors add new parameters that require more and more me-
chanical tests for their identification. Such tests must be performed with care and
cost time and money. Is it really less expensive to do biaxial and bulk tests than to
measure a texture?
This discussion explains why micro-macro models were worth to be developed in addition
to phenomenological models. Another argument is that micro-macro models help to deeply
understand the material behaviour, they allow to numerically verify the microscopic models
and to select the driving mechanisms of plastic deformation in polycrystals.

3 BASIC KNOWLEDGE ABOUT MICROSCOPIC MODELS


3.1 Description of Crystal Events During Plastic Deformation
Ewing & Rosenhain 1900 have established that the plastic deformation of a crystal occurs by
slipping on some crystallographic planes (slip or glide planes) and in some crystallographic
directions (slip or glide direction) or by twinning, another crystallographic phenomenon.
However, the elementary theoretical estimation of the shear strength of a perfect crystal
given by Frenckel 1926 yields values of the order of 10−1 G (G = shear modulus of the
crystal). Knowing that the observed shear strength is of the order of 10−6 ≈ 10−4 G,
scientists have proposed the concept of crystal defects, called dislocations, that reduce the
strength of the crystal. Figures 20 and 21 present two basic types of dislocations, the edge
and screw dislocations, and show the dislocation line AB. The motion of dislocations results
in crystal shearing.
Burgers’ vector b is introduced to describe the slip direction of a dislocation. It is de-
termined by Burgers’circuit, a closed path involving two lattice directions and surrounding
the dislocation line (Figure 22).
22 A.M. Habraken

A
B

Figure 20. Edge dislocation (from Magnée, 1994)

Figure 21. Screw dislocation (from Khan & Huang, 1995)

F S

A F
b
S

(a) (b)

Figure 22. Burgers’ circuit for edge (a) and screw (b) dislocation defining Burgers’
vector (from Khan & Huang, 1995)
Modelling the Plastic Anisotropy of Metals 23

It should be pointed out that, for screw dislocation, Burgers’ vector is parallel to its
dislocation line, while for edge dislocation, Burgers’ vector is perpendicular to the dislo-
cation line. A slip or glide system is characterized by the combination of a slip plane and
a slip direction. Experimental observations show that, in most metals, the slip planes are
usually those planes with the closest atomic packing, while the slip directions are always
the closest packed directions on the slip planes. Once reference axes are defined, a slip
s
system is described by its geometrical matrix Kij also called Schmid’s tensor:
s
Kij = bsi nsj (15)
nsj unit vector normal to the slip plane for the slip system s, bsj unit vector in the slip
direction for the slip system s, Asij and Zij
s are respectively the symmetric and the anti-
s
symmetric parts of Kij .
Another approach to characterize one slip system is to use Miller’s indices to represent
the slip planes and directions. For instance, for b.c.c metals, 24 {110} < 111 > and
{112} < 111 > slip systems are available and, for f.c.c metals, 12 {111} < 110 > are
generally assumed (Van Houtte 1995).
The critical resolved shear stress CRSS τcs is the shear stress to apply in order to sus-
tain the long range propagation of a dislocation according to a slip system s, under the
superposed effects of all coexisting structural features which represent obstacles of differ-
ent strengths. In practice, the resolved shear stress τ s acting on a slip system s can be
derived by projecting the microscopic stress σ micro on the slip plane with the help of the
corresponding Asij matrix. This determination can be checked, if one computes the energy
dissipated by the strain rate tensor ε˙ associated to a single shear of one active slip system:

γ̇ s τ s = ε̇pij σ̇ij
micro
= Asij γ̇ s σij
micro

τ s = Asij σij
micro
(16)

The slip s occurs according to Schmid’s law if the shear stress τ s reaches the CRSS τcs .
This defines the yield locus of a single crystal:

−τcs ≤ τ s ≤ τcs (17)


Equals signs hold for plastic deformation, while inequalities apply to the elastic domain.
In this elastoplastic formulation, the detection of active slip systems can be done either by
Bishop-Hill’s approach or by Taylor’s approach. These methods are dual ones, shortly
summarized hereafter, more details can be find in Van Houtte 1988.

3.2 Taylor’s Single Crystal Plasticity Model


At the crystal level, the plastic microscopic velocity gradient generated by a particular slip
system s is given by:

Lsij = Kij
s s
γ̇ (18)
where Schmid’s tensor Kij s is defined by relation (15) and γ̇ s is the slip rate acting on

this slip system s. In practice, multiple slips occur together. Neglecting the elastic strains
as in Asaro & Needleman 1985, Becker 1990, Neale 1993 or Dawson & Kumar 1997, the
microscopic velocity gradient applied on a crystal Lmicro is given by:

Lmicro = Ls + Ω L
(19)
s
24 A.M. Habraken

(a) Ω p= ∑Z γ
s
s
&
s
= plastic spin or slip induced spin

(b) Ω L = rate of crystal lattice rotation as a rigid body, used to update texture

Figure 23. Representation of the 2 terms of the crystal microscopic spin

where Ω L is the rate of crystal lattice rotation. The establishment of this well-known
relation is recalled in Van Houtte 1995. Its link with the classical formalism of Continuum
Mechanics is summarized in Section 3.5. This microscopic velocity gradient can be split
into a microscopic plastic deviatoric strain rate ε˙ p micro and a microscopic spin Ω micro :

ε˙ p micro
= sym(Lmicro ) = As γ̇ s (20)
s

Ω micro
= skw(Lmicro ) = Zs γ̇ s + Ω L
(21)
s
micro
The 2 components of the microscopic
 spin Ω are symbolically represented on Fi-
gure 23. The first term of (21), Zs γ̇ s , is called plastic spin Ω p .
s
Several different combinations of slip rates may achieve the prescribed strain rate. Ac-
cording to Taylor 1938, the one which minimizes the power dissipation is chosen:

Ẇ p = τcs |γ̇ s | = min (22)
s

Taylor roughly assumes that all slip systems have a common value τc of their Critical
Resolved Shear Stress τcs (CRSS). This common value τc is only reasonable for annealed
condition and crystals like b.c.c. ones for instance, where no strong variations exists between
slip systems. Computing different CRSS is a classical improvement for scientists using this
approach.
The prescribed strain rate ε˙ p micro can be split into a scalar magnitude ε̇peq computed
by the classical von Mises’ formula and a strain rate mode Uε̇p :
ε˙ p micro
Uε˙ p = (23)
ε̇peqmicro
Modelling the Plastic Anisotropy of Metals 25

s
Introducing also the slip rate γ̇scaled per unit equivalent strain rate:

s γ̇ s
γ̇scaled = (24)
ε̇peqmicro
advantage can be taken of the assumed strain-rate insensitivity to simplify the formulation.
Indeed, dividing relation (22) by the strain rate magnitude and the reference CRSS τc leads
to:

Ẇ p 
s
= |γ̇scaled | = minimum (25)
ε̇peqmicro τc s
with, according to equation (20), the minimization constraint:

Uε̇p = As γ̇scaled
s
(26)
s
These two equations are called Taylor’s equations and may be efficiently solved by means
of linear programming (Simplex Method) as explained by Van Houtte 1988. The solution
s
γ̇scaled relies only on the prescribed strain mode Uε̇p , it depends neither on the magnitude
ε̇eq nor on the value of the reference CRSS τc .
In the Simplex Method, relation (25) is called the cost function, to be minimized under
the constraint (26). This primal problem in the space of slip rates is transformed into a
dual problem where the stresses are now the independent variables. In the stress space, the
value Ẇ p /ε̇peqmicro τc represents a maximum:

Ẇ p σ micro
= Uε̇p :
= maximum (27)
ε̇peqmicro τc τc
This retrieves the approach proposed by Bishop-Hill, which assumes that the admissible
stress states satisfy the yield locus constraints expressed by relation (17) and that the real
stress state maximizes the external plastic work. Taylor’s and Bishop-Hill’s methods are
strictly equivalent, that is why they are very often referred to as the Taylor-Bishop-Hill
(TBH) crystal model.
The Simplex Method produces both slip rates and microscopic stresses at the crystal
level. In practice, using the same CRSS τc for all slip systems leads to multiple solutions.
Different selections of activated slip systems achieve the prescribed strain rate (20) with
the same minimum power dissipation (22). Van Houtte 1988 proposes different methods to
choose one particular solution. From an energy point of view, all the solutions are equiv-
alent. However, they produce different slip rates, microscopic stresses and hence different
crystal lattice rotations. Once the set of active slip systems and the corresponding slip rates
are known by the resolution of equations (25), (26), the crystal rotation Ω L produced by
one imposed velocity gradient can be reached using relation (21). It is expressed by:

L
Ω =Ω micro
− Zs γ̇ s (28)
s
Let us define M (g, Uε̇p ) the Taylor’s factor, which is an important scalar in all the
microscopic models. This factor is associated with a crystal of orientation g for a given
strain mode Uε̇p . It is conventionally derived from the plastic power dissipation per unit
volume Ẇ p by the following relation:

Ẇ p  1 micro
s
M (g, Uε̇p ) = = |γ̇scaled |= σ (Uε̇p )ij (29)
ε̇peqmicro τc s τc ij
26 A.M. Habraken

Physically, it represents a certain amount of dislocation glide rate associated to the


crystal orientation and to the applied strain rate.
3.3 Strain Rate Sensitivity Approach for Single Crystal Plasticity Model
Another single crystal plasticity model is issued from the theory of thermally activated
dislocation glide. It consists in adopting a viscoplastic flow rule:
 s1
τ n
γ̇ = γ̇0  s  sign(τs )
s
(30)
τ c

where γ̇ s is the slip rate on the slip system s, γ̇0 is a reference slip rate defined so that
|γ̇ s | = γ̇0 when |τ s | = τcs and the parameter n characterizes the material rate-sensitivity.
Under isothermal conditions and in a narrow range of strain rates, the above simple power-
law equation has proven its validity. For cold deformation, the rate sensitivity is rather low
and the stress dependence on the slip rate can be reasonably approximated with n close to 0.
The slip system shear rate γ̇ s does not vanish as long as the resolved shear stress τ s on the
corresponding slip system s is not identically zero. It keeps however a low value if τ s is not
close to τcs . This equation applies a posteriori the activation condition defined by Schmid’s
law by filtering out the inactive systems. For low value of n, this approach provides an
interesting alternative to TBH model as it overcomes the problem of non uniqueness in
selecting a set of active slip systems by the TBH approach. It is a popular approach used
by many scientists (Asaro & Needleman 1985, Anand et al. 1997, Dawson et al. 1997)
its equivalence to TBH approach has been studied and verified by Neale 1993, Anand &
Kothari 1996.
3.4 Evolution Rule for CRSS Value
Whatever the solution is chosen to model single crystal behaviour (TBH model or vis-
coplastic flow rule), the CRSS τcs appears and evolves in a different way for each individual
slip system s. This implies the knowledge of the initial value of τcs as well as its evolution
equation. Slip on any slip system generally induces hardening for all slip systems. This is
taken into account by adopting an evolution equation of the CRSS of the form:

τ̇cs = hsu γ̇ u (31)
u
where hsu is the so-called hardening matrix. Diagonal components of this matrix corre-
spond to the self-hardening effect, while off-diagonal components describe cross-hardening
effects. Franciosi 1988 proposes an evolution rule for the components of this matrix. This
rule depends on the pair of slip systems, their shearing rates and their temperature. This
approach could seem sophisticated. However, it is still not accurate enough. The fact that
hardening defined by one fixed structure of dislocations is affected by structure modifica-
tion is not taken into account. The choice for hsu constitutes a distinction between the
microscopic models used in micro-macro approaches implemented in FEM codes. Khan &
Cheng 1996 and Teodosiu 1997 review some different proposals.
3.5 Mechanical Frame for Single Crystal Plasticity
In the frame of Continuum Mechanics, researchers as Asaro 1983, Anand & Kothari 1996 use
an isoclinic configuration, which is an intermediate conceptual local configuration defined
by Figure 24, to decompose the deformation gradient tensor:
∂x ∂x ∂x∗
F= = = F∗ Fp (32)
∂X ∂x∗ ∂X
Modelling the Plastic Anisotropy of Metals 27

( σ, T )

x2
x1
x1 = F * X 1

x 2 = F * -T X 2

Deformed configuration

p
F = F* F
F*

γs

X2 Fp X2
( Ο, T o ) ( Ο, T o )
X1
X1

Reference configuration ``Isoclinic´´ relaxed configuration


Fig.24
Figure 24. Reference, deformed and isoclinic configurations, multiplicative decom-
position of the deformation gradient tensor F = F ∗ F p

X coordinates in the initial configuration expressed in axes Xi , x coordinates in the de-


formed configuration expressed in axes xi , x∗ coordinates in the isoclinic configuration
expressed in axes Xi , F∗ sum of an overall “elastic” distortion of the lattice and the rigid
rotation of the lattice, Fp “plastic” simple shears that do not disturb the geometry of the
lattice.
The lattice in the isoclinic relaxed configuration has the same orientation as the lattice
in the reference configuration. The incremental deformation of a crystal is taken as the
result of the contributions from two independent atomic mechanisms:

• the sum of an overall “elastic” distortion of the lattice and a rigid rotation of the
lattice (F∗ );
• “plastic” simple shears that do not disturb the lattice geometry (Fp ).

The microscopic velocity gradient is linked to the deformation gradient tensor:


∂ ẋ ∂ ẋ ∂X
Lmicro = = = ḞF−1 (33)
∂x ∂X ∂x
−1 −1
Lmicro = Ḟ∗ F∗ + F∗ Ḟp Fp−1 F∗ (34)
28 A.M. Habraken

The plastic shear rate is expressed through a certain number of slip systems s (the active
ones in the TBH model and all slip systems in the rate sensitive approach):
−1 
Lmicro = Ḟ∗ F∗ + γ̇ s K∗s (35)
s

Schmid’s tensor K∗s is expressed in deformed configuration, but as the crystal lattice
is not affected by Fp , one has:
−1
K∗s = F∗ Ks F∗ (36)
So in the micro-mechanical frame, one gets (35) which should be equivalent to relation
(19) in Taylor’s model. Since in (19), the elastic part of F∗ has been neglected, the strain
rates deduced from (35) and (19) are different:

(35) → ε˙ = sym(Lmicro ) = ε˙ elastic + ε˙ p micro
= ε˙ elastic
+ A∗s γ̇ s (37)
s

(19) → ε˙ = sym(Lmicro ) = ε˙ p micro
= A∗s γ̇ s (38)
s

The skew-symmetric parts of both velocity gradients are identical and are called the
spin:

L p L
Ω micro
=Ω +Ω =Ω + Z∗s γ̇ s (39)
s

where Ω L is the crystal lattice rotation due to both the global rigid body rotation of the
macroscopic body and the particular crystal rotation due to texture updating.

4 MICRO-MACRO APPROACHES
On a macroscopic point of view, one works with polycrystals that contain a lot of crystallites.
In consequence, models defined for single crystals cannot directly be applied. A micro-macro
transition has to be developed. Before reviewing the different micro-macro proposals in
Section 4.2, let us look at the decomposition of a polycrystal into its crystallite orientations.
4.1 Crystallographic Texture
In a polycrystalline sample, each crystal is characterized by its volume fraction dV and by
its orientation, symbolically designated by g. More specifically, the set of Euler’s angles
(ϕ1 , φ, ϕ2 ), describing the orientation of the crystal reference system with respect to the
sample (external) reference system, is most often used (Bunge, 1982).
Figure 25 shows the physical meaning of ϕ1 , φ, ϕ2 . With dV (g)/V as the volume
fraction of crystals having their orientation within dg around g, the statistical crystallite
orientation distribution function (ODF) f (g) is then defined as:
dV (g) 1
= f (g) dg with dg = 2 sin φ dϕ1 dφ dϕ2 (40)
V 8π
It provides a quantitative description of the crystallographic texture of the polycrystal;
high values indicate preferred orientations, and f (g) ≡ 1 a completely random texture.
Because of the crystal symmetry, several symmetrically equivalent choices exist for a
crystal reference system. The classical ranges for the three Euler’s angles (0 ≤ ϕ1 ≤ 2π,
0 ≤ φ ≤ π and 0 ≤ ϕ2 ≤ 2π) may therefore be reduced. Additionally, symmetries in the
forming process may lead to an initial texture with similar statistical symmetries. This
Modelling the Plastic Anisotropy of Metals 29

z z

ϕ
1

z’ z’

y’
y’ ϕ
1 y
y

x’ ϕ
K B = Crystal 1
x’
K A= Specimen
x
x
(a) (b)

z z

ϕ ϕ
1 1

y’
z’ z’
y’
φ ϕ φ ϕ
2 2
φ φ
ϕ ϕ
1 y 1 y
ϕ
2
ϕ x’
ϕ 1
1
x’

x φ x φ

(c) (d)

Figure 25. Definition of Euler’s angles (from Van Houtte 1995)

so-called sample symmetry allows again for a reduction of the part of Euler’s space to
be considered. For example, in the case of cubic crystals without sample symmetry, the
ranges [0, 2π]x[0, π/2]x[0, π/2] are to be used; with additional orthorombic sample symme-
try, [0, π/2]x[0, π/2]x[0, π/2] will be sufficient (Van Bael 1994).
Using the harmonic method proposed by Bunge 1982, the ODF f (g) can be represented
by a series expansion:

max µmax
l (l) ν
max µν
f (g) ∼
= Clµν T̈˙ l (g) (41)
l=0 µ=1 ν=1

µν
with lmax the maximum degree of the series expansion, T̈˙ l (g) harmonic functions of Euler’s
angles and Clµν their Fourier’s coefficients describing the texture. For instance, truncating
at lmax = 22, an ODF for cubic crystals and orthorombic sample symmetry is represented
by a set of 185 Clµν - coefficients (355 in the absence of sample symmetry). These coefficients
can for example be obtained, after appropriate mathematical processing, from a set of X-
ray diffraction pole figures, measured on a sample with the help of a texture goniometer
(Bunge, 1982).
30 A.M. Habraken

4.2 Polycrystalline Plasticity Models


Starting from the state of individual crystallites, different assumptions have been proposed
to deduce the state of the corresponding polycrystal characterized by the macroscopic strain
rate tensor ε˙ macro and the stress tensor σ macro . The most logical approach is a volume
average weighted by the ODF function f (g) defined by relation (40):

ε˙ macro = ε˙ micro (g)f (g) dg (42)


σ macro
= σ micro
(g)f (g) dg (43)

Although the assumption of homogeneous microscopic stress tensor σ micro and strain
rate tensor ε˙ micro in each single crystal is already a simplification, most macroscopic models
are bound by stronger assumptions:

• Sachs 1928 applies the loading state (uni-axial tension) at each individual crystallite
as if it were a free-standing single crystal and he assumes that each crystal has only
one active slip system (Van Houtte 1995). This assumption leads to more or less
severe violations of geometric compatibility at the grain boundaries and, in general,
quite unsatisfactory results (Sevillano et al. 1980). In particular, the predictions of
texture evolutions due to deformation do generally not compare favourably with the
experimental results.

• Full Constraint Taylor 1938 (FC) assumes an homogeneous distribution of plastic


velocity gradient, that induces an homogeneous plastic strain rate distribution:

Lmacro = Lmicro =⇒ ε˙ p macro


= ε˙ p micro
(44)

This assumption leads to equilibrium violations at grain boundaries. For cubic metals,
as reported by Van Houtte et al. 2002, the predictions of this model are first-order
approximations of deformation textures for all possible strain histories and all possible
initial textures, though they do not perform well for low stacking-fault energy f.c.c.
(brass, silver, etc.). “First-order approximation” means that predictions are good
from a qualitative point of view but not from a quantitative point of view, since
the final locations of the final orientations may be up to 10◦ off, and intensities and
volume fractions may be up to a factor 2 wrong. However this model often computes,
for many practical applications, more acceptable results than the previous one. The
assumption (44) provides an upper bound solution when models derived from Sachs
approach computes lower bounds. It is generally the Full Constraint Taylor hypothesis
linked with Taylor-Bishop-Hill single crystal plasticity model that is applied when
literature provides “polycrystalline TBH yield locus” (Figure 26).

• The Full Constraint Taylor’s model enforces rigorous geometric compatibility at the
expense of stress equilibrium. The so-called Relaxed Constraint Taylor’s (RC) model
drops this strict assumption by relaxing the compatibility of well chosen components
of the velocity gradient tensor. In the lath model, component L(RD)(N D) is not pre-
scribed and the pancake model does not enforce the component L(T D)(N D) either,
where RD, TD and ND are the rolling, transverse and normal directions. The pan-
cake model is advocated for materials with flat, elongated grains produced by rolling.
Figure 27 shows for instance L(RD)(N D) relaxation for such a flat grain. L(T D)(N D)
Modelling the Plastic Anisotropy of Metals 31

4 σ22/τ0
1.5
Experiments
Barlat 1991 3
Barlat 1997
1.0
TBH
2
0.5
1
σy / σF

0.0 σ11/τ0
0
-4 -3 -2 -1 0 1 2 3 4
- 0.5
-1

- 1.0
-2
σy / σ x = - 1
- 1.5 -3 Crystal plasticity
- 1.5 - 1.0 - 0.5 0.0 0.5 1.0 1.5
6th-order plastic
σx / σ F potential
-4

(a) (b)
Figure 26. TBH yield loci compared with other predicted yield loci, a) Al-2.5%Mg
(from Barlat et al. 1997 b) IF steel (from Li et al. in press)

ND

TD

L (RD)(ND)

RD

Figure 27. Relaxation of L(RD)(N D) in a flat elongated grain of the pancake model
(from Van Houtte et al. 2002)

and L(RD)(N D) represent shears that would not cause any rotation of the top or bot-
tom surfaces of the flat elongated grains. It is believed by pancake model’s advocates
that the relaxation of such shears would cause fewer problems of strain misfits with
neighbouring grains that other kind of relaxations would. These components are cal-
culated by methods assuming that the associated shear stresses are set to zero. These
components take different values in each crystallite, depending on the lattice orien-
tation. Such approach does not take the interaction with other grains into account.
As shown by Van Houtte 1988, these kinds of relaxations can be implemented in an
elegant way using pseudo-slip system.
Compared to FC models, RC models sometimes produce better and more detailed
results, especially for rolling texture predictions, although their justification is still
under debate. The problem of accommodating the strain misfits inherent to RC
models leads scientists like Wagner, Lücke, Arminjon, Imbault, Van Houtte to propose
32 A.M. Habraken

several more advanced models which try to take these misfits into account. The
summary presented in Van Houtte 1996 confirms that such models lie between TBH
and self-consistent approaches and yield improved results compared to classical FC
or RC models.
In general, Taylor type models are quite successful for f.c.c and b.c.c metals where
a large number of slip systems ensures that individual crystals can accommodate an
arbitrary deformation. However, these types of models are not valid for materials
whose crystals have an insufficient number of deformation modes to sustain an arbi-
trary strain. Such kinematically rank deficient materials are not rare: semi-crystalline
polymers, minerals and other geological materials, superconducting ceramics, metals
of hexagonal close-packed crystal structure such as zinc, zirconium and titanium. Two
proposals adapted to such cases are described in Prantil, Dawson and Chastel 1995.

• Self-consistent models (Berveiller & Zaoui 1979, Canova & Lebensohn 1995, Molinari
1997, Masson & Zaoui 1999) consider a grain as a solid inhomogeneity embedded in
a homogeneous infinite matrix subjected to macroscopic loading. All the grains are
treated that way one after the other, the matrix behaviour results from the weighted
average of the individual contributions of all the grains. Both strain rate and stress
heterogeneities are allowed, but they are linked by an interaction formula based on
Eshelby’s 1957 work. For elastic cases, Kröner 1961 uses the interaction formula at
the grain level:
(σ micro − σ macro ) = −L∗ : (ε micro − ε macro ) (45)

where L∗ is a 4th rank tensor called the “interaction tensor”. The interested reader
can find the method to identify L∗ in Van Houtte 1995.
This approach is not straightforward to extend to nonlinear cases. For elastoplastic
cases, an incremental linearization proposed by Hill 1965 seems to be adapted:

σ̇ micro
− σ̇ macro
= −LH : (ε˙ micro − ε˙ macro ) (46)

where LH is called the Hill’s constraint tensor. It depends on the elastoplastic mod-
ulus, on the shape and orientation of the crystals. Masson & Zaoui 1999 summarizes
the scientific controversies on this topic and demonstrates that Hill’s conception could
be adopted even for elastoviscoplasticity.
Ponte Castaneda’s variational procedure (Ponte Castaneda 1991) developed for non
linear composites has recently been applied by Gilormini et al. 2002 to self-consistent
approach describing titanium polycrystals behaviour. The variational procedure in-
troduces a linear comparison polycrystal, (identified by an optimisation problem),
which is then used to give a self-consistent estimate of the response of the non linear
polycrystal.
Whatever self-consistent approach is used, a rather long iterative solution procedure
is required. This is mainly due to the fact that, at the end of the iterative process, the
macroscopic values must coincide with the average of the grain responses. However,
unlike the Taylor’s models, the self-consistent models allow to take into account both
effects of texture and grain morphology on the mechanical response of the material.
They provide an answer that respects in average both equilibrium and compatibility
between grains. Self-consistent models coupled with FEM are clearly one way to
increase accuracy of micro-macro models, but more efficient numerical approaches
are still required to solve non academic problems.
Modelling the Plastic Anisotropy of Metals 33

Figure 28. Illustration of the three types of relaxation considered by the Lamel
model. Type I corresponds to 12-shear; Type II to 23-shear; Type III
to 12-shear

• N-point models propose to look at n grains simultaneously to caputre some specific


effects due to the interaction of neighbouring grains. For instance, Van Houtte et
al. 1999 and 2002 propose different versions of the Lamel model, that studies two
grains at the same time. As the pancake model, it has been developped for rolling and
assumes that grains tend to become flattened and elongated. Each of the two grains
taken separately undergoes a homogeneous strain, which however does not need to be
equal to the macroscopic average strain. In its first version, the FC Taylor condition
was satified for the two grains taken together, when the second version of this model
relaxes this requirement. All the relaxations that maintain the geometry integrity of
the boundary between the two grains and keep the boundary plane parallel with the
rolling plane are possible. Figure 28 presents the three types of relaxation considered.
The local velocity of two interacting grains are computed by minimising the plastic
work under the following conditions:

– Plastic deformation proceeds by dislocation slip.


– The two grains coopeatively accommodate the macroscopic deformation (except
L12 in the second version of Lamel).
– No grain boundary sliding occurs along the planar boundary separating the grain.
– Minimisation of the rate of plastic work in each grains.

This model still closed to RC Taylor model allows however to take into account the
grain interaction.

• Homogenization technique, used by Smit et al. 1998, Miehe et al. 1999 or Geers et al.
2000, is directly based on a mathematical procedure already applied to find micro-
macro links in composite materials. Two levels of finite element models are used:
a mesh of the entire structure and a mesh of the Representative Volume Element
(RVE). At each interpolation point of the macroscopic mesh, the finite element model
of the RVE is called to provide the stress-strain behavior of the material.

Sections 5 and 6 present some finite element applications of the above micro-macro
hypotheses.
34 A.M. Habraken

4.3 Texture Updating


If a texture is described by a set of representative crystals orientations (Tóth & Van Houtte
1992), the rotation of each representative crystal leads to the up-dated texture. Taylor’s
assumption about an homogeneous velocity gradient:

Lmacro = Lmicro (47)

provides the rotation of each crystal, if Taylor’s single crystal plasticity model is applied
for each crystal. In a first step, relation (20) identifies the active slip systems and then
relation (21) defines the rate of crystal lattice rotation. This approach will be extensively
used in the micro-macro approaches presented in Sections 5 and 6 to update textures, it is
of common use in scientific works dedicated to micro-macro approaches (Van Houtte et al.
1989, Hirsch 1991, Winther et al. 1997, Aukrust et al. 1997).

4.4 Average Taylor’s Factor


The TBH theory described in Section 3.2 for one crystal can directly be applied to poly-
crystals with Taylor’s assumption (44). So, plastic strain rate tensors ε˙ p are supposed to
be homogeneous throughout the polycrystal. In addition, Taylor assumes that the common
reference CRSS τc (common reference value for all the slip systems in one crystal) is the
same for all crystallites in the polycrystal in one representative volume element. It is called
the average common reference CRSS τ̄c :

τc = τ̄c (48)

Again, this assumption seems acceptable for materials in their annealed state, though
it becomes questionable after accumulation of a certain deformation (Bunge et al. 1985).
With relations (29), the average plastic power dissipation in a polycrystal is easily computed
from its expression for a single crystal (22) and the ODF (40):

p

Ẇ (ε˙ p macro
)= Ẇ p (ε˙ p micro
, g)f (g) dg = ε̇peqmacro τ̄c M̄ (Uε̄˙ p ) = σeq
macro p macro
ε̇eq (49)

with the average Taylor’s factor:



M̄ (Uε̄˙ ) =p M (Uε̄˙ p , g)f (g) dg (50)

Hereafter any over-lined variable indicates a value which is assumed to be an average


for all crystallites belonging to the same elementary volume considered in the macroscopic
approach. Note that the average Taylor’s factor only depends on the texture of the poly-
crystalline material f (g) and on the given strain rate mode Uε̄˙ p but not on the strain rate
magnitude ε̇peqmacro .
Relation (49) yields the important micro-macro link:
macro
σeq = τ̄c M̄ (Uε̄˙ p ) (51)

Both geometric (textural) hardening related to M̄ and material (strain) hardening re-
lated to τ̄c (linked to dislocation density) coexist and appear in a micro-macro approach.
Experimental works also demonstrate that the anisotropic behaviour cannot be attributed
to only one of these factors. For instance, Jensen & Hansen 1987 measured the 0.2 %
tensile yield stress on specimens cut at different angles α to the Rolling Direction. Their
material was a sheet of commercially pure aluminum, which was previously cold-rolled at an
Modelling the Plastic Anisotropy of Metals 35

45

STRESS (0.2%)/M (MN/m2)


ε = 2.0

40
FLOW

30

ε = 0.2

25

20 40 60 80
ANGLE TO ROLLING DIRECTION

Figure 29. Results of tensile tests performed in different directions: yield stress divided
by M̄ in a pure aluminium submitted to cold rolling of 20 or 200%

equivalent true strain of up to 200 %. Textures of the rolled sheet were determined and the
average Taylor’s factors M̄ were computed for two different levels of deformation. Figure 29
macro /M̄ versus the angle α to the Rolling Direction, this
shows the evolution of the ratio σeq
ratio corresponds to τ̄c according to (51). At very large strains, τ̄c is almost constant, indi-
cating that the observed plastic anisotropy can mainly be attributed to the crystallographic
texture. On the contrary, after moderately large monotonic strains, τ̄c increases with α,
showing a strong influence of the intragranular microstructure on the plastic anisotropy.
According to Teodosiu’s 1997 review, this conclusion seems true for polycrystalline f.c.c:

• for moderately cold-rolled sheet, plastic anisotropy seems due to the orientation of
the dislocation structures (sheets of high dislocation densities more or less parallel to
the {111} slip planes);
• for heavily cold-rolled sheets, plastic anisotropy mainly seems due to the crystal-
lographic texture, because dislocations are arranged in thick-walled, equiaxed cells,
providing an almost isotropic hardening.

Let us note that the further work of Winther, Jensen and Hansen 1997 studies the com-
bined effect of texture and microstructure on the flow stress anisotropy of metals containing
dislocation boundaries with a macroscopic orientation with respect to the sample axes. As-
suming that dense dislocation walls and micro-bands resist like ordinary grain boundaries,
a value of the CRSS depending on a Petch-Hall equation is adopted. In their approach,
36 A.M. Habraken

Taylor’s or Sachs’ polycrystal assumption is used on a set of 1152 crystals, representative of


the texture. It has been proved that the presence of dislocation boundaries has a significant
effect on anisotropic yield.
4.5 Link Between the Evolution of the Reference CRSS and Macroscopic Strain
Hardening
Taylor 1938 proposes that the common reference CRSS τc evolves as a function of the total
slip Γ in each crystal:
 t
Γ(g) = Γ̇(g) dt (52)
0

where g defines the crystal orientation and the total slip rate Γ̇ is:

Γ̇(g, ε˙ p micro
)= |γ̇s | = ε̇peqmicro M (g, Uε̄˙ p ) (53)
s

where ε˙ p is a strain rate tensor represented by its mode Uε̄˙ p and magnitude ε̇peqmicro .
Using Taylor’s assumptions (ε¯˙ = ε˙ p micro = ε˙ p macro and τ̄c = τc ) the total polycrys-
¯ is defined by:
talline slip rate Γ̇

¯ p p
Γ̇(ε¯˙ ) = Γ̇(g, ε¯˙ )f (g) dg = ε¯˙ eq M̄ (g, Uε̄˙ p ) (54)

The total polycrystalline slip Γ̄ is then computed by integration of the slip rate and an
important micro-macro link appears:

dΓ̄ = dε̄peq M̄ (55)


If one works with a uni-axial test where εpeqmacro
= ε and macro
σeq
= σ, the macroscopic
work hardening can easily be deduced from relations (51) and (55):

dσ dτc dM̄
= M̄ 2 + τc (56)
dε dΓ dε
The first term (material hardening) on the right side of equation (56) indicates an
isotropic hardening at the polycrystal level, since an average is done whatever slip system
is activated and for all crystals. The second term (geometrical hardening) is due to the
evolution of texture resulting from plastic deformation. It is often neglected in macroscopic
models or even in simple micro-macro models (Schmitz 1995, Winters 1996). Relation (56)
is well known and comments can be found in Aernoudt et al. 1987.

5 FEM MICRO-MACRO MODELS WITHOUT MACROSCOPIC YIELD


LOCUS
5.1 Introduction
The Finite Element Method is currently used to simulate the behaviour of the materials
at the microscopic level (Teodosiu et al., Kalidindi et al. 1992, Anand & Kothari 1996,
Bertram et al. 1997, Barbe et al. 1999, Acharya & Beaudoin 2000, Van Houtte et al. 2002
...). Finite element simulations allow to treat a crystalline aggregate as a continuum, by
simply requiring the equilibrium of the stress tensors and the continuity of the displacements
across the grain boundaries. In such a model, each crystallite consists of at least one element
of the FE mesh. Figure 30 shows the mesh used by Kalidindi to model an IF-steel that has
been industrially cold-rolled to a reduction of 70 % (Van Houtte et al. 2002). In this case,
Modelling the Plastic Anisotropy of Metals 37

Figure 30. Initial (a) and final mesh (b) of an IF steel 70% cold-rolling simulation
(from Van Houtte et al. 2002)

each element represents one crystal, which is simply assumed cubic. In other cases, exact
boundaries of observed grains are represented as in Figure 31, where the eleven grains
detected on the front face of a pure copper tensile sample have been meshed (Teodosiu
et al. 1992). Another possibility is to use simple cubic elements to mesh grains. For
instance, Acharya & Beaudoin 2000 discretizes one grain by 12 to 96 elements according
to coarse or fine discretizations (Figure 32). In all these simulations, the lattice orientation
characterizes each grain. The constitutive equations of the microscopic model (crystal
plasticity model) are implemented as a material model. Such simulations aim to validate
the chosen microscopic model (Acharya & Beaudoin 2000), to study the effect of grains
interactions on texture prediction (Van Houtte et al. 2002) or to allow parametric studies,
difficult to carry out experimentally.
Regarding microscopic strain heterogeneity, microscopic FEM models are in principle
superior to methods such as the FC or RC Taylor’s models or Lamel model. However, the
required calculation time is at least one order of magnitude larger, which makes the method
unsuited for large scale FE simulations of a forming process of a fine-grained material.
This article is focused on another use of the FEM coupled with microscopic models.
It concerns accurate macroscopic simulations, where the material behaviour is taken into
account at a microscopic level. This time, each finite element has a macroscopic size, it
represents numerous crystallites. Clearly the FEM discretization is on a macroscopic scale.
In such a micro-macro approach, a lot of averaging methods to extract the macroscopic
behaviour from microscopic analyses are possible. Some examples are listed in Sections
5.2. to 5.3. They concern approaches where no macroscopic yield locus is computed, they
present one generic family of micro-macro models.
38 A.M. Habraken

8 11
2 3

6 4
1

7 5 9

10

(a)

(b)

Figure 31. (a) Observed 11 crystallites on the front face of a copper tensile sample; (b)
single layer of pentahedral finite elements (from Teodosiu et al. 1992)

Figure 32. Von Mises equivalent stress (Mpa) computed by a FEM mesh for a
tensile strain of 20% of nickel polycrystal aggregate (grain size 32µm),
from Acharya & Beaudoin 2000
Modelling the Plastic Anisotropy of Metals 39

The most simple micro-macro FEM model of this type is proposed by Nakamachi, it is a
direct extension of the FEM analysis applied at the microscopic level. At each macroscopic
integration point, the behaviour law of a single crystallite characterized by its orientation
and the hardening state of each of its slip system is adopted. So no average procedure of the
polycrystalline behaviour is applied. Nakamachi & Dong 1997 have applied their approach
to the Limiting Dome Height test (NUMISHEET’96 benchmark test, Figure 16). The
quarter of the 180 × 100 mm blank is simulated by 1125 eight-node SRI (Selected Reduced
Integration) solid elements. A dynamic explicit finite element code is used with an elas-
tic/crystalline viscoplastic constitutive law. The strain heterogeneity between integration
points directly follows from the nodal displacements. The equilibrium is assumed but not
checked since an explicit finite element scheme has been used. Such a simple micro-macro
approach can gives interesting results at least close to experimental ones (Nakamachi et al.
1997,1999a, 1999b).
Section 6 is dedicated to the macroscopic approximation of yield loci based on micro-
scopic assumptions (3G model, Aifantis’models) or microscopic computations (see described
models of Montheillet, Darrieulat, Arminjon, Bacroix, Imbault, Van Houtte). These models
belongs to a second generic family of micro-macro models.
5.2 Macroscopic FEM Simulations Relying on Discrete Set of Crystals
In this type of macroscopic FEM simulations, the response of each integration point de-
pends on the response of a multitude of single grains, representative of this material point.
Each crystallite is described by a microscopic model. This can be done only thanks to
massive parallel computations. Simulations like hydroforming process, performed by Daw-
son already 10 years ago (Dawson et al. 1992), show that, with powerful computers, this
approach can be applied to real problems of limited size. Clearly, in addition to the usual
choices in a FEM approach (lagrangian, eulerian formulation, explicit, implicit scheme, ...),
the scientist must determine further assumptions:
• Step 1: the micro-macro link (Full Constraint, Relaxed Constraint Taylor’s model
or one of their variants, self-consistent model, homogenisation technique). It allows
to go from a macroscopic velocity gradient to microscopic values and to provide,
after computations at microscopic level, the macroscopic stress as well as the stiffness
matrix for implicit scheme.
• Step 2: the set of representative crystals. The number of crystals must be deter-
mined as well as their orientation, shape, size, slip system and associated CRSS. The
initialisation of these crystal data is essential for the accuracy of the results.
• Step 3: the behaviour model of each crystal. In this type of models, elasticity can be
neglected or taken into account. The crystal plasticity model defines the link between
the resolved shear stress and the slip system rate. The CRSS hardening rule is also
a key function that induces differences between the models. As shown hereafter, two
different proposals exist: viscoplasticity or rate independent plasticity. Finally, the
microscopic stress related to the microscopic velocity gradient is computed.
Of course, all these options interact and have strong effects on the accuracy and com-
putation time. Well-known scientists have proposed various models in order to describe
polycrystalline models. Some of these models are presented hereafter, together with their
advantages and drawbacks.
5.2.1 TBH single crystal model + FC Taylor’s polycrystalline model
Such polycrystalline models are described for instance in Asro & Needleman 1985, Mathur
& Dawson 1989, Becker 1990, Neale 1993, Beaudoin et al. 1994, Kalidindi & Anand 1994,
40 A.M. Habraken

Anand et al. 1997. The three general choices: micro-macro link, set of representative
crystals and behaviour model of each crystal are successively identified hereafter and some
applications are presented to illustrate this type of models.
a) Micro-Macro link
Full Constraint Taylor’s polycrystalline model is used. This implies that the local de-
formation gradient in each grain is set homogeneous and identical with the macroscopic
deformation gradient.
The elasticity is generally neglected in the work performed by Dawson’s team (Mathur
& Dawson 1989, Beaudoin et al. 1994):

Fmacro = Fmicro = R∗ Fp (57)


R∗ being a rotation matrix. But elasticity is taken into consideration by scientists working
with Anand (Kalidindi & Anand 1994, Anand et al. 1997):

Fmacro = Fmicro = F∗ Fp (58)


with F∗ defined by Figure 24.
If the N representative grains or crystals have equal volume, a simple average is used
to link micro and macro stress tensors (Kalidindi & Anand 1994).
N
1 
σ macro
= σ micro(k)
(59)
N k=1
Otherwise, a weighted average based on the volume fraction Wk of each crystal orien-
tation is used (Beaudoin et al. 1994).
N

σ macro
= W kσ micro(k)
(60)
k=1

Clearly, the average used to reach the macroscopic stress tensor depends on the choice
of the representative set of crystals.
b) Representative set of crystals
Number of crystals: N varies depending on simulations and authors. For instance, 180 or
200 crystals per integration point are used for compression tests (Kalidindi & Anand 1994),
32 for cup-drawing (Anand et al. 1997), 256 for hydroforming process (Beaudoin et al.
1994), 200 for titanium rolling (Dawson & Kumar, 1997) ...
Orientation of crystals: experimental data from X-ray diffraction measurements of crystal-
lographic texture give discrete intensities of diffracted energy as a function of goniometric
angle position. This can be transformed into a “Crystal Orientation Distribution Function”
ODF (see Section 4.1) and can be used to generate a set of “weighted Euler angles”.
According to “popLA package” (Kallend et al. 1991), two approaches are proposed
to approximate a texture defined by a continuous density distribution through a set of
discrete orientations. In the first approach, all orientations have the same weight. Since the
crystals are assumed to have the same volume, they are located in the orientation space so
that higher density regions are more densely populated. The second approach consists in
randomly populating the orientation space with discrete orientations and assigning to each
of them an initial weight that minimizes the effect of the density fluctuations arising from the
Euler’s space distortion. The actual density in the associated volume of orientation space
then multiplies the weight of each discrete orientation. This provides a weighted orientation
Modelling the Plastic Anisotropy of Metals 41

approach generally used by Dawson and his co-workers (see relation (60)), while Anand’s
team generally uses the first approach with identical weights (see relation (59)).
This initial set of orientations is updated during the computations. Section 4.3 explains
how the computation of the crystal lattice rotation (Ω L for a single crystal is directly
applicable, thanks to Taylor’s assumption of identical deformations at macroscopic and
microscopic levels (see relations (19) - (21)).
Crystal slip systems: the slip systems are well known for f.c.c materials like aluminium,
copper or b.c.c materials like steel, tantalum, h.c.p. materials like zinc, zircalloy ...
CRSS: as explained in Section 3.4, the hardening of each CRSS associated to a slip system
should be taken into account as well as its initial value. Very often, the values of the
initial CRSS of all slip systems are assumed equal or close to each other (Van Bael 1994).
In an annealed state for f.c.c or b.c.c materials, this seems reasonable. From numerous
simulations of homogeneous deformations of f.c.c materials (Kalidindi et al. 1992), it has
been observed that, after large deformations, the values of the CRSS for the various slip
systems in an aggregate are quite close to each other. This common value is estimated from
a macroscopic simple compression test.
Crystal shape: this cannot be taken into account in a classical polycrystalline Taylor’s
model.

c) Behaviour model of each crystal


For instance, Mathur & Dawson 1989 and Beaudoin et al. 1994 neglect elasticity (57) and
use a rate dependent plasticity model described for each slip system by relation (30). They
also adopt a common average value for all CRSS in one crystal:

τcs = τc (61)

and choose a simple evolution law of Voce’s type:

τsat − τc
τ̇c = H0 and τsat = f (Γ̇) (62)
τsat − τc0

where Γ = total shear slip on all slip systems of the crystal; τsat = saturation value of the
common reference CRSS; τc0 = initial value of the common reference CRSS; H0 = material
parameter.
In Beaudoin et al. 1994, a hydroforming process (Figure 33) is chosen to validate the
numerical FEM model. This choice seems very well adapted to check the model of the
blank behaviour as i) applied pressure assures stability, ii) contact and friction models
do not introduce inaccuracy; punch and blank are in sticking conditions and blank-flange
contact is assumed frictionless.
Consequently the final deformed shape depends on the material anisotropy and is mea-
sured by the percentage of earring. As shown in Figure 34, numerical results are close to
experimental measurements. For this case, it was checked that the texture evolution during
the process and its effect on the shape of the yield locus are minimal.
Another example is provided by Anand & Kothari 1996 who compare two different mod-
els at the microscopic level: a viscoplastic flow rule (relation (30)) and a rate-independent
crystal plasticity model. A robust calculation scheme determines a unique set of active slip
systems and the corresponding shear increments in the rate-independent theory. In both
models, the evolution rule (31) of the CRSS associated to each slip system τcs is applied.
This relation comes from Asro & Needleman 1985 and the chosen hardening matrix hsu for
42 A.M. Habraken

OIL

RUBBER

CONTROLLED PRESSURE

STROKE

Figure 33. Axisymmetrical hydroforming process of an aluminium sheet (adapted


from Beaudoin et al. 1994)

5.0
5.5
Experiment
4.0
Simulation
3.5
3.0
% Earing

2.5
2.0
1.5
1.0
0.5
0
0 10 20 30 40 50 60 70 80 90
Radial distance from RD ( degrees )

Figure 34. Earing measurement and prediction at the final stage of a hydroforming
process (from Beaudoin et al. 1994)

the 12 slip systems of f.c.c crystals is:


 
 A qA qA qA   

 
 
1 1
 qA A qA qA  1 
hsu = hu with A= 1 1 1 (63)

 qA qA A qA 
 
 


 
 1 1 1
qA qA qA A
Modelling the Plastic Anisotropy of Metals 43

where systems 1, 2, 3 are coplanar, as systems 4, 5, 6 and 7, 8, 9 and 10, 11, 12 are. For
coplanar systems, the ratio of the latent hardening rate to the self hardening rate is equal
to unity. For non coplanar systems, it is evaluated by means of factor q = 1.4. The function
hu is defined by:
 a
τu
u
h = h0 1− c (64)
τsat

where h0 , a and τsat are slip system hardening parameters which are taken identical for all
slip systems. In fact, τsat should be an increasing function of strain rate but this can be
neglected at low temperature.

400
Strain Rate Jump Test 400 Constant True Strain Rate Test
300
( MPa )

300
( MPa )
200
200
σ

100
100

0 0
0.0 0.2 0.4 0.6 0.8 0.0 0.5 1.0 1.5
ε ε
a b

Figure 35. Experimental and numerical compression tests. (a) Strain rate jump
test; (b) Constant true strain rate test (from Kalidindi & Anand 1994)

These parameters are determined by curve-fitting between the results of simple com-
pression tests and numerical simulations using “Taylor simulation”. Figures 35 illustrate
the identification of copper parameters. By “Taylor simulation”, one means: the TBH
model applied at the crystal level with Taylor’s assumption of equality between micro-
scopic and macroscopic strain rates at the polycrystalline level. Here, this model uses (31)
and (64) to follow the evolution of the CRSS of each slip system in each crystal. The ini-
tial copper isotropic texture is represented by 200 crystals. One can check that the initial
value of CRSS τc0 influences the macroscopic initial yield, τsat the final saturated value of
macroscopic stress, h0 affects the initial hardening rate and a modifies the shape of the
polycrystalline stress-strain curve between the initial yield and saturation. If the TBH
model is replaced by a rate dependent visco-plastic law (30), two additional parameters:
γ̇0 , n appear. For copper, according to Kalidindi & Anand 1994, γ̇0 is assumed to have a
constant value of 0.012 and n is chosen equal to the corresponding macroscopic parame-
ter. This latter is identified thanks to a strain rate jump experiment on a polycrystalline
specimen in compression state at room temperature.
Once all the microscopic parameters τc0 , τsat , h0 , a and γ̇0 , n are identified, Anand
and co-workers (Kalidindi & Anand 1994, Anand & Kothari 1996, Anand et al. 1997)
can apply a macroscopic FEM model (ABAQUS) where each interpolation point averages
the microscopic behaviour of 180 crystallites. In particular, compression tests of cylindri-
cal specimens, plane strain compression experiments, forging experiments (Figure 36), cup
drawing experiments (Figure 37) are performed. Experimental and numerical loads, geo-
metric shapes and texture evolutions (Figure 38) are compared as well. Such a model has
been successfully applied to high deformation rate of tantalum (Anand et al. 1997).
44 A.M. Habraken

Figure 36. (a) Plane strain forging experiment performed on an isotropic copper;
(b) coarse FEM mesh with each integration point relying on 180 single
crystals (from Kalidindi & Anand 1994)

d) Conclusion
In short, the previous examples using Taylor’s model, coupled with finite elements, show
that such approach is feasible and has been validated for b.c.c and f.c.c materials. If a high
number of representative crystals at the microscopic level is used, this can provide accu-
rate predictions of texture, geometry and stress history during any cold forming process.
The texture evolution is directly implemented in this approach. This means that the “ge-
ometrical hardening”, responsible for the shape modification of the yield locus is directly
taken into account. This characteristic constitutes an advantage of this type of models
with respect to others, such as the phenomenological yield loci coupled with kinematic
and isotropic hardening. These ones cannot easily represent texture evolution effect. As
explained above, this micro-macro method does not use a global yield locus but an average
answer computed from the plastic behaviour of a set of representative crystals.
Taylor’s model coupled with finite elements requires a high CPU time and memory
storage, directly proportional to the number N of crystals associated to one integration
point. It is quite surprising that this number N has not been more investigated. Very
often, measured pole figures are qualitatively compared to the ones computed from the
discrete set of orientations (Anand & Kothary 1996). No further study from a mechanical
point of view, such as stress response computed for the same strain and different N values,
Modelling the Plastic Anisotropy of Metals 45

50

40

Cup Height ( mm ) 30

20 Experiment

Polycrystal model
10

0
0 90 180 270 360
Angle from Rolling Direction ( degrees )

Figure 37. Experimental and predicted earing profiles resulting of the deep draw-
ing of a cylindrical cup (from Anand et al. 1997)

is performed. For each of the N representative crystals associated with one integration
point, one must store:
• the crystal orientation, typically defined by 3 variables;
• hardening variable(s): either one reference CRSS τc , as in relation (61) if a common
reference value for all slip systems is adopted, or one CRSS τcs for each slip systems
s as in equation (31).
5.2.2 Self-consistent polycrystalline models
As explained in Section 4.2, the self-consistent approach respects, on the average, both
compatibility and equilibrium between grains. It is intensively used by scientists aiming to
understand and predict the macroscopic material behaviour thanks to micro-macro models
(Canova & Lebensohn 1995, Molinari 1997, Nikolov & Doghri 2000). However, this great
advantage is shaded by the increase of computation time. This fact explains why it is seldom
used for coupling with FEM models. It is not surprising that self-consistent approach
applications appear for hexagonal materials like Zircalloy (Chastel et al. 1998) because
the low number of deformation modes in such crystals yields inaccurate predictions with
Taylor’s model.
Chastel et al. 1998 present a viscoplastic self-consistent polycrystalline model (Leben-
sohn & Tomé 1993) coupled to a 3D eulerian finite element code LAM3 (Hacquin et al.
1995), applied to the hot extrusion of Zircalloy. In practice, the finite element calcula-
tion starts with an isotropic rheology, which provides a first deformation gradient at each
integration point. Then, the polycrystalline model is locally activated and provides an
anisotropic response of the material which induces subsequent calculation in LAM3. For
this case, the final flow patterns reached by the macroscopic approach and by the micro-
macro computation are very close to each other as they are mainly fixed by the kinematic
boundary computations. Discrepancies between predicted and measured texture evolutions
46 A.M. Habraken

Figure 38. Experiment and predicted {111} and {110} pole figures in the deformed specimen
at the point indicated in the deformed mesh (from Kalidindi & Anand 1994)

are attributed to the occurrence of recrystallization and/or recovery phenomena, which are
not taken into account in the model.
To summarize, such a method allows to consider directly “textural hardening” as Tay-
lor’s model coupled with finite elements does. It is more satisfactory from a scientific point
of view since, at the microscopic level, both compatibility and equilibrium are approached.
The grain shape can be taken into account at the microscopic level. The case of crystals
with a low number of deformation modes seems to require this type of model for accurate
predictions. The cost in memory is identical with Taylor’s approach. However, the CPU
time, which is already a problem with Taylor’s model, is even worse here, as an itera-
tive process is required to find the heterogeneous strain repartition between crystals. This
Modelling the Plastic Anisotropy of Metals 47

explains why so few macroscopic FEM models are linked to self-consistent microscopic mod-
els. Progresses in this direction are on their way. It must be pointed out that with “object
programming”, it is not a real problem to replace a Taylor’s model by a self-consistent one.
5.2.3 Homogenisation polycrystalline models
The three general choices: micro-macro link, set of representative crystals and behaviour
model of each crystal are successively identified hereafter and some applications are pre-
sented to illustrate this type of models.
a) Micro-macro link
Proposals by Smit et al. 1998, Miehe et al. 1999, Geers et al. 2000 or Feyel & Chaboche
2000 are directly based on mathematical procedures already applied in composite materials.
Two levels of finite element models are used: at every interpolation point of the macroscopic
FEM mesh, another microscopic FEM model, simulating a Representative Volume Element
(RVE), is called to provide the stress-strain behaviour of the material. In other words, the
constitutive law at a macroscopic point results from the global response given by the FEM
analysis of a set of representative crystals (RVE) described by a microscopic behaviour
model. These 2 levels of computation are represented in Figure 39.

Macro - Variables Micro - Variables

micro
macro X
X
macro
micro

Macro - Continuum Micro - Structure

Figure 39. Macroscopic and microscopic levels (adapted from Miehe et al. 1999)

The homogenisation approach provides the mathematical background to go from the


microscopic level to the macroscopic level. Figure 40 presents the mathematical description
of the same deformation at macroscopic and microscopic levels. In Figures 39, 40, 41
the superscript macro relates to the macroscopic mesh and the superscript micro identifies
variables attached to the RVE considered at the microscopic simulation level. The following
notations are introduced:
Γmacro or micro initial configuration,
γ macro or micro deformed configuration associated to the initial one,
Xmacro or micro coordinate tensor in the initial configuration,
xmacro or micro coordinate tensor in the deformed configuration,
∂xmacro
Fmacro = ∂X macro macroscopic deformation gradient tensor
micro ∂xmicro
F = ∂Xmicro microscopic deformation gradient tensor,
n Γ outward normal of the initial configuration of the RVE,
nγ outward normal of the deformed configuration of the RVE.
The following averaging relations define the macroscopic gradient tensor Fmacro and the
macroscopic first Piola Kirchhoff stress τ macro from their values at a microscopic scale:

1
Fmacro = Fmicro dV (65)
V Γmicro
48 A.M. Habraken

x macro

F macro

X macro x macro
macro macro

a. Deformation of Macro - Continuum

x micro = F macro X micro + w


~

X micro x micro
micro
micro
macro + ~
F micro = F F

b. Deformation of Micro - Structure

Figure 40. Definition of the deformation at macroscopic and microscopic level


(adapted from Miehe et al. 1999)

x micro macro micro + ~


X

~
F micro = F macro + F

Figure 41. (a) Details of the deformation at microscopic level; (b) (c) (d) initial
FEM meshes of the RVE where crystals are identified by different grey
colours (adapted from Miehe et al. 1999)
Modelling the Plastic Anisotropy of Metals 49


1
τ macro
= τ micro
dV (66)
V Γmicro
The application of Gauss’ theorem leads to:
 
1 1
Fmacro = Fmicro dV = xmicro ⊗ NΓ dA (67)
V Γmicro V ∂Γmicro
 
1 1
τ macro
= τ micro
dV = tmicro ⊗ Xmicro dA (68)
V Γmicro V ∂Γmicro

where the tension on the boundary ∂Γmicro is defined by:

tmicro = τ micro Γ
n (69)
Figure 41a. illustrates the above theoretical considerations. Figures 41b., c., d. present
different microscopic finite element meshes: one element per crystal grain (b), regular square
elements mesh (c), mesh of triangles applied on the crystals (d).
The deformation assumption in the RVE is related to macroscopic values by:

xmicro = FmacroXmacro + w̃ (70)

Fmicro = Fmacro + F̃ (71)


The assumption (71) coupled with the previous average equation (67) leads to:
  
1 1 − 1 +
F̃ dV = w̃ ⊗ nΓ dA + w̃ ⊗ nΓ dA = 0 (72)
V Γmicro V ∂Γmicro V ∂Γmicro
+ −
with NΓ = NΓ at 2 associated points of the contour (Figure 41).
This general mathematical frame shows that 3 alternative possibilities directly satisfy
relation (72):

• w̃ = 0 everywhere in Γmicro : Taylor’s assumption Fmicro = Fmacro is recovered, no


superimposed deformation field at the micro scale;
• w̃ = 0 on the contour ∂Γmicro : zero fluctuation of the superimposed deformation field
w̃ on the boundary but non zero fluctuation inside the RVE;
• w̃+ = w̃− on the contour ∂Γmicro : periodic fluctuation of the superimposed deforma-
tion field w̃ on the boundary.

b) Set of representative crystals


All the examples presented in Miehe et al. 1999 start from an isotropic texture easily
represented by random crystal orientations. According to the size of the macroscopic mesh,
the number of finite elements in the RVE varies.
For instance, in a validation test simulating a simple shear loading, one single macro-
scopic element is used, coupled with 100 finite elements in the RVE. Each of them has one
integration point associated with one crystal orientation. The texture predictions resulting
from the 3 alternative choices for the w̃ field are quite close to each other.
Another example of punch indentation was computed using 100 macroscopic finite ele-
ments, each of them linked with 400 crystal grains in the RVE.
50 A.M. Habraken

c) Behaviour model of each crystal


Miehe et al. 1999 provides one of the most complete micro-macro models: it considers
the anisotropic elasticity in each crystal coupled with a thermoplastic or thermoviscoplastic
anisotropic behaviour and temperature effect. The rate dependent plastic behaviour is quite
classical and the rate independent plastic approach proposes an extension of the work of
Kothari & Anand 1996.
d) Conclusion
The advantages of this approach are that texture and grain shape or size effects are directly
taken into consideration. Better than in the self-consistent approach, both equilibrium and
compatibility between crystals are accurately reached in the RVE. The effects of 3 different
assumptions at the microscopic scale (Taylor’s hypothesis, no fluctuation or periodic fluc-
tuation on the boundary) are easily considered. One significant drawback is the amount
of computations and memory requirements, which are even worse than with self-consistent
and Taylor’s models coupled with macroscopic FEM simulations. For instance, all the vari-
ables describing the microscopic FEM simulation (nodal positions, state variables at every
microscopic integration point: crystal orientation, reference CRSS or CRSS for each slip
systems ...) associated to every macroscopic integration point must be stored. Feyel &
Chaboche 2000 proposes however an industrial application.
5.3 FEM Analysis Applied on Both ODF Evolution and Mechanical Fields
The discrete set of representative crystals used in FEM simulations as described in Section
5.2.1 b) suffers from shortcomings because such a characterization comes with few analy-
tical tools. No direct means are available, for instance, to develop quantitative measures
to differentiate between textures associated with distinct discrete aggregates. As a result,
considerations on differences between textures are often qualitative, or obtained through
projections onto alternate representations. In dealing with spatially inhomogeneous tex-
tures, there is often a need to interpolate or to project across textures. This requirement
appears when initialising from experiment, or computing spatial gradients of texture as
measures of the inhomogeneity degree. Such measures are important when the considered
material is initially inhomogeneous and when substantial inhomogeneities develop over the
course of the process. In rolling, for instance, highly localized regions of inhomogeneity
develop through the thickness of the sheet due to roll induced shearing.
Kumar & Dawson 1995a, 1995b and 1996 and Dawson & Kumar 1997 propose a quite
complex approach that uses directly the ODF without relying on discrete sets of representa-
tive crystals. This approach offers the advantage of an easier and more accurate possibility
to compare and interpolate textures. 2 FEM simulations applied in different spaces are
connected to each other:

• in the crystal orientation space, a finite element mesh describes the ODF representing
the material; a “metallurgical-orientation” finite element analysis is used to solve the
ODF conservation equation (Clement 1982);
• in the classical geometrical space, a spatial steady state simulation using eulerian me-
chanical finite elements is applied to an industrial process such as rolling for instance.

It is fundamentally different from the approach proposed in Section 5.2.3 which consists of
2 coupled FEM simulations in the classical geometrical space but performed on different
scales: microscopic and macroscopic.
Of course, in Kumar and Dawson’s work, both FEM simulations are coupled:
• the microscopic FEM requires the strain evolution to compute texture evolution;
Modelling the Plastic Anisotropy of Metals 51

• the macroscopic FEM uses the up-dated texture as well as crystal plasticity with
Taylor’s or Sachs’ assumption to get the macroscopic constitutive behaviour.

5.3.1 Crystal orientation representation


Kumar and Dawson have not chosen Euler’s angles but Rodrigues’parameters to represent
the crystal orientations. The crystal orientation is defined as the rotation R required to
align the crystal lattice frame with a fixed sample reference frame. Three independent
parameters are sufficient to describe such a rotation. An alternative class of representation
uses an axis of rotation n and an angle of rotation φ. In this case:

R(n, φ) = n ⊗ n + (I − n ⊗ n) cos φ + I ∧ n sin φ (73)


where I is the second order unit tensor. A particular representation of the crystal orientation
is defined by Rodrigues by vector r:

r = n tg φ (74)

The parameters r 1 , r 2 , r 3 define the so-called Rodrigues’ space.


The advantages of this choice are the following ones:
• A simple fundamental region can be computed in the orientation parameters space.
Rodrigues’ proposal defines all the possible crystal orientations without redundancy
due to crystal symmetry. For instance, the case of f.c.c crystal leads to the funda-
mental region represented on Figure 42.
• The space distortion is limited and there is no singularity in this space. This can be
deduced from the form of the invariant volume element:
 
φ
dv = cos2 dr 1 dr 2 dr 3 (75)
4

Note that Euler’s space distortion and singularity can be identified from relation (40).

E C

D B
F

F
D
A
C A
E
A

B
B
F
E C B
r
3 D
E
D
A
r r
1 2 F C

Figure 42. F.c.c fundamental region for Rodrigues’space (from Kumar & Dawson 1995)
52 A.M. Habraken

With this parametrization, the Orientation Distribution Function (ODF) is defined by


the probability density f (r). The volume fraction V ∗ of crystals whose orientations belong
to a subset Ω∗ of the orientation space is given by:


V = f (r) dΩ (76)
Ω∗

The whole domain Ω of orientation space depends on symmetries exhibited by the


crystal. f (r) is scaled in such a way that:

l= f (r) dΩ (77)

5.3.2 Micro-macro link


The macroscopic behaviour is the average of the behaviour of the individual crystallites
of the aggregate. If xmicro is an arbitrary crystal quantity, its macroscopic value xmacro is
given by:

xmacro = xmicro (r)f (r) dΩ (78)

Here is a difference with respect to approaches using a discrete sample of orientations


and simply computing a weighted average of the crystal quantities (see relations (59) and
(60)). Dawson and Kumar apply a viscoplastic constitutive law at the crystal level as
proposed in Section 3.3, with identical rate sensitivity and CRSS for all slip systems and
for all the crystallites related to one interpolation point. Either Taylor’s assumption of
equality between micro and macro plastic strain rates or the assumption of micro and
macro stress equality (Prantil et al. 1995) is applied.
5.3.3 Evolution rule of the orientation distribution function
Restricting his attention to the lattice reorientation caused by crystallographic slips, Clement
1982 proposes to model texture evolution by integrating an equation for the rate of change
of the probability density f (r, t). This equation results from requiring that the material
derivative of relation (78) vanishes.
∂f
+ v grad f + f div v = 0 (79)
∂t
where v is the reorientation velocity. Note that Arminjon 1988 provides a demonstration of
this type of equation using physical hypotheses. The above formulation is an eulerian rep-
resentation in which f is associated to particular locations in the orientation space, rather
than to particular crystals whose orientations change. Alternatively, a reference texture f0
can be specified, for instance f at time t0 . Then, this reference position remains fixed for
all times and is used to define initial lagrangian coordinates for crystals. Relation (79) can
be written either with respect to current eulerian coordinates f (r, t) or with respect to the
lagrangian coordinates fˆ(r0 , t). Euler’s choice (79) leads to difficulties in the FEM formu-
lation because of the convective contribution associated with the term v grad f . This still
remains a problem, even if some solutions are proposed in computational fluid mechanics.
The alternative is to use the lagrangian representation. However, this adds the cost of an
explicit computation of crystal trajectories in the orientation space. Both approaches are
described in Kumar & Dawson 1996.
Another difficulty is the extreme behaviour of the ODF. The ODF can evolve expo-
nentially, sometimes tending asymptotically to Dirac’s function. Clearly, inaccuracies are
Modelling the Plastic Anisotropy of Metals 53

inevitable as the finite element size cannot tend to 0. An effective strategy applied by Ku-
mar and Dawson is to moderate the evolution of the ODF by the following transformation:

p = ln f (80)
Relation (79) to compute the new ODF uses the reorientation velocity field v, which is
linked to crystal plasticity and to the relationship between the crystal velocity gradient and
its macroscopic counterpart. Here again the relation (21) is applied and the relationship be-
tween the rate of crystal lattice rotation Ω L and v depends on the specific parametrization
employed for rotations. For Rodrigues’ parameters, Kumar & Dawson 1995a or Dawson &
Kumar 1997 have established the adapted formula.
5.3.4 Applications
a) Texture prediction under monotonic deformations
The following example shows a FEM analysis applied to solve the texture conservation
relation under the assumption of plane strain compression of f.c.c polycrystals. In Figure 43,
the 3D FEM mesh of 28672 - 4 nodes tetrahedral elements is showed.

r
3

r r
1 2

Figure 43. f.c.c fundamental region for Rodrigues’space (from Kumar & Dawson 1995)

As Taylor’s hypothesis is applied, the reorientation velocity, developed under a mono-


tonic deformation, is invariant with strain. So crystal computations are done only once.
The developed texture is adequately represented by the ODF on the boundaries of the
fundamental region; consequently only outside views of the ODF are represented. The
ideal components of f.c.c plane strain compression texture are compared with the com-
puted results on Figure 44. Texture development is dominated by two fibers: an α fiber
connecting the ideal Goss and Brass’ orientations and an β fiber connecting Brass and
Taylor’s orientations.
b) Application to aluminum rolling
The flat rolling of a 1100 aluminum (Figure 45) being a steady state process, a macroscopic
eulerian FEM approach is applied to model the macroscopic mechanical problem. In fact,
the FEM analysis of texture evolution is coupled with the macroscopic FEM computation.
Two FEM scales are present and the details on their parallel implementation can be found
54 A.M. Habraken

B G B

r
3 T

ε = 0.2
r
1
r
2 = 0.2

(a) (b)

Figure 44. Ideal (a) and computed (b) components of plane strain compression
texture of a f.c.c polycrystal (from Kumar & Dawson 1995b)

Free
surface Rigid
rolls
Traction Traction
free free
Workpiece
Sliding
friction
Constant angular
roll velocity

Figure 45. Schematic diagram of flat rolling (from Dawson & Kumar 1997)

in Kumar & Dawson 1995b. Figure 46 shows the equivalent plastic deformation rate, the
reference CRSS and the scalar measure of the spatial gradients of the ODF defined by:

∇A = |gradf (r, x)| dΩ (81)

where x identifies a material point of the workpiece.


As expected, the microstructure hardens primarily within the deformation zone under
the roll. Hardening and texture gradients appear through the workpiece thickness. It is
observed that the texture gradients are rather important.
5.3.5 Conclusion
Kumar and Dawson’s approach is interesting from a scientific point of view. It shows
another proposal for the micro-macro coupling, not limited to the mechanical point of
view. Here the FEM formulation is directly applied for texture prediction and, hence,
provides metallurgical information. However, the amount of computations seems to be
even greater than in Section 5.2 since the mesh discretization of the texture problem is
already of large size. Another problem, more “sociological” than scientific is the fact that
Modelling the Plastic Anisotropy of Metals 55

20.0 65.0 110.0


ε eq
p

28.0 41.5 55.0 MPa


τc

Z 0.9 2.8 8.5 x 10 6


Y Α

Figure 46. Equivalent plastic deformation rate εpeq , reference CRSS τc , scalar measure of the
spatial gradients of the ODF ∆A (from Kumar & Dawson 1995).

the ODF description is neither pole figures nor section in the Euler’s space well known by
people with a metallurgical background.
56 A.M. Habraken

5.4 Discussion about Micro-Macro Approaches without Yield Locus


Clearly, with parallel computers, the above selection of scientific works demonstrates that
micro-macro approaches attract a great interest and become closer to practical problems.
However, for each type of applications, it appears that more efforts should be devoted esti-
mating the size of the set of representative crystals or the RVE or the texture discretization.
This is directly linked to the accuracy of the results and to the CPU time.
From this review, it appears that a visco-plastic formulation is very often used at the
crystal level with the advantage of avoiding the choice between the multiple solutions of
Taylor’s model. However, at room temperature, the strain rate sensitivity coefficient is very
low, which leads to numerical difficulties and explains the further research dedicated to the
strain rate independent approach (Anand & Kothari 1996, Miehe et al. 1999). Finally both
models can be chosen according to their availability.
The use of a common CRSS evolution rule for all the slip systems in a crystal or even
for all crystallites in a polycrystal at an integration point, seems a logical simplification to
limit the number of state variables. However, a clear information on the accuracy benefit
of choosing a distinct CRSS for each slip system seems unavailable.
The effect of the micro-macro links which have been used (FC or RC or modified Taylor’s
model, self-consistent model, homogenisation method) has been the most studied problem.
The choice clearly depends on the number of available slip systems, the desired accuracy
and the CPU requirement. For b.c.c or f.c.c materials, Taylor’s model already seems to
give interesting results from a mechanical point of view (stress, strain) but it also gives
qualitative texture prediction with the lowest CPU time.

6 FEM MICRO-MACRO MODELS WITH MACROSCOPIC YIELD LOCUS


6.1 Introduction
In Section 5, the drawback of high CPU time to compute the state of representative crys-
tals, then to reach, by an averaging operation, the macroscopic behaviour has often been
mentioned. Scientists have investigated other micro-macro approaches that are less greedy
from a CPU time point of view.
A first option is to develop new macroscopic elasto-plastic or elasto-visco-plastic models
with general features imbued from plasticity models in single crystals. Such approaches are
described in Section 6.2. (3G model, Aifantis’model).
Another option is presented in Section 6.3. In this case, outside any FEM code, models
at crystal level and micro-macro links are applied to estimate an accurate expression of the
yield locus in polycrystal materials. Then, this accurate yield locus function is used during
macroscopic FEM computations. The evolution of size and position of this yield locus
during the process is defined by macroscopic isotropic and kinematic hardening rules. Such
hardening models can be macroscopic but with strong links to microscopic phenomena.
Even if their accuracy can be quite high (see Teodosiu & Hu 1998 and Miller & Mc Dowel
1996), such models generally neglect the “texture hardening”, i.e. the fact that due to
texture evolution, the yield locus shape should be updated. In some cases, this phenomenon
is really negligible and using an accurate description of the initial yield locus conjugated
with elaborate hardening models yields a very good accuracy at low CPU cost. The time
reduction of such approaches as compared to FEM codes directly coupled with microscopic
calculations is difficult to estimate. However we do not speak of a ratio of 2 or 3 but rather
10, 100 and even greater, if no parallel computation is applied.
About the memory requirement, such approaches spare all the variables required at
every integration point to store the orientation and the average CRSS (or even the CRSS
associated to each slip system) of each crystal of the representative set. This explains
Modelling the Plastic Anisotropy of Metals 57

why, even if some yield locus description and their hardening behaviour need 100 or more
constants and 50 internal variables, they are still much more economical than the micro-
macro approaches presented in Section 5.
Of course these FEM models predict only the mechanical behaviour. If the texture
evolution is an interesting result, the strain history during the process must be stored and
used in post-processor modules to predict the final texture. Such an approach is proposed
by Winters 1996, who simulates a cup drawing using the yield locus described in Section
6.3.4. in stress space. Then he uses the strain history to predict the texture evolution and
compares it to texture measurements. His simulation results are quite accurate. Another
example is the work by Schoenfeld & Asaro 1996. They study the texture gradients through
thickness in rolled polycrystalline alloys by means of FEM rolling simulations using a phe-
nomenological constitutive law. Once the displacement time history of the roll gap has
been calculated, Taylor’s model is applied at locations of interest through the thickness of
the workpiece to predict the final texture and material anisotropy.
6.2 Macroscopic Models Imbued from Single Crystal Plasticity
Such models are no micro-macro models but propose an interesting alternative to phe-
nomenological models described in Section 2. For instance, Khan and Cheng (1996, 1998)
propose to extend crystal plasticity model to polycrystalline model just by using a larger
number of slip systems. Two other examples are shortly summarized here: 3G and Aifantis’
models. Such proposals directly extend the knowledge of crystal plasticity to macroscopic
plasticity models. Each one makes some assumptions to provide “simple” useful macro-
scopic constitutive laws. Their applications verify that the actual behaviour under complex
loading paths can be predicted. It is however clear that each specific path (tension +
torsion, pre-straining ...) may need some adjustments of the models.
Such models are neither straightforward to implement in FEM code nor to identify. So
it is not surprising that they generally seem to be used only by the teams that have devel-
oped them. They provide more accurate results than simple phenomenological approaches.
Let us note that Aifantis’ model also gives a theoretical justification for more advanced
phenomenological models such as Karafillis & Boyce 1993.
6.2.1 3G model
Hage Chehade 1990, Monfort et al. 1991, Montfort & Defourny 1993, Montfort & Defourny
1994 have developed the so-called “3G model”, which is an anisotropic non associated visco-
plastic model. To understand this approach, first consider the plane stress state in a planar
isotropic material as described on Figure 47.
The strain tensor associated to this loading state is defined by (82) where G12 is associ-
ated to the shear stress τ12 applied on the planes oriented at 45◦ from principal directions
1 and 2.

   
G12 0 0 0 G12 0
p   p  
ε = 0 −G12 0 ε =  G12 0 0 (82)
0 0 0 ref axes 1,2,3 0 0 0 ref axes a,b,3

The extension of this simple state to general cases assumes that plastic strains happen
by plastic slips on planes oriented at 45◦ from principal axes and along directions where
shear stresses are maximal. As in b.c.c metals, 24 families of slip planes exist, one can
assume that there is always a crystallographic plane oriented nearly along a direction of
maximum shear stress. This generalisation is represented by Figure 48 and by relation (83)
where axes λ, µ, ν are not orthogonal axes.
58 A.M. Habraken

Figure 47. Pure shear state in axes a, b, rotated by 45◦ from the principal stress
axes 1,2 (adapted from Hage Chehade 1990)

ν 3 µ

2
1 λ

Figure 48. General view of assumed slip planes in general cases (adapted from
Hage Chehade 1990)

The plastic strain tensor is expressed by:


 
0 G12 G13
p  
ε =  G12 0 G32  (83)
G13 G32 0 ref axes λ,µ,ν
Modelling the Plastic Anisotropy of Metals 59

where the 3G shear strains, responsible for the model name, appear. For simplicity, relation
(83) is written in principal axes 1,2,3:

   
ε11 0 0 G12 − G13 0 0
p    
ε = 0 ε22 0 = 0 G23 − G12 0  (84)
0 0 ε33 0 0 G31 − G23 ref axes 1,2,3

Shear stresses acting on octahedral planes are associated to the 3G shear strains:

τ12 → G12, τ23 → G23 , τ13 → G13 (85)


With the previous concepts in mind, the isotropic planar version of the 3G model is
developed and applied to a thin sheet where the principal stress σ3 in the thickness direction
is low. It assumes that any plastic strain results from the superposition of plastic shears
oriented at 45◦ from principal stresses directions as represented on Figure 49.

σ3

σ1

σ2

σ2
σ1

σ3

Figure 49. Sheet strains decomposed into plastic slips in planes oriented at 45◦
from principal stresses (from Hage Cheade, 1990)

If the material presents a planar anisotropy, more crystals are oriented in specific di-
rections for which a macroscopic deformation is easier or more difficult. In other words,
the resistance to deformation varies with the direction in the sheet plane. Extending the
isotropic planar approach, macroscopic strain occurs in families of planes presenting the
most favourable ratio between the applied shear stress and the intrinsic resistance to defor-
mation. In general, such planes are deviated from the 45◦ directions to principal stresses.
The shift angle strongly depends on the Lankford coefficient. It can be mathematically
determined as well as its incidence on the total deformation.
As far as the hardening rule is concerned, the 3G model applies a generalization of
Bergström’s model based on dislocation balance (Bergström 1969). Using the 3G as strain
60 A.M. Habraken

measures allows to extend Bergström’s work on a more physical basis for any stress and
strain state. It is demonstrated (Monfort & Defourny, 1994) that the plastic shear stress
used in Tresca’s theory can be related to the 3G strain measures:
1
2τ = A + B[1 − exp(−ar45+β (|G12 | + r90+β |G23 | + rβ |G31 |)] 2 (86)

where A, B and a are material constants, rα is Lankford’s coefficient for a direction at an


angle α from the Rolling Direction, β is the angle between the first principal stress and the
Rolling Direction and τ is the stress limit in Tresca’s criterion.
As the 3G model is based on stresses acting on the maximum shear planes, a Tresca-like
criterion is used, which is consistent with the mechanism taken into account. However the
exact shape of the plasticity criterion is of secondary importance, since it only specifies
the limiting stress level of plasticity initiation. The subsequent plastic behaviour, in terms
of stresses and strains is obviously more important with this non associated visco-plastic
model. It is described by a flow rule developed according physical considerations. Return-
ing to the physics, the evolution of the shear strain dGij is due to the creation and the
propagation of mobile dislocations. It is proportional to:

• the time increment dt;

• the rate of mobile dislocation creation∗ itself proportional to the total amount of
immobile dislocations and to a thermal activation factor;

• the path free of mobile dislocations, which is assumed constant and equal to the
average dimension of the dislocation cell as in Bergström’s model;

• the inverse of Lankford’s coefficient;†

• the probability that a dislocation moves along a given shear plane, which is a function
of the shear stresses.

With its parameters related to fundamental physics, the 3G model potentially has the
ability to take into account the main macroscopic features of deformation under complex
strain paths. For instance, pre-strain effect, or change in dislocation cell shapes and sizes
can be handled with simple modifications. This advantage appears for any model based
on Bergström’s extension. The interesting feature of the 3G model is to take directly into
consideration changes in Lankford’s coefficients; however as their evolution is not provided
in the current version of the model, it neglects texture evolution effect.
The planar anisotropic behaviour of the 3G model is validated by theoretical and ex-
perimental considerations on the fracture of cylindrical cups (Montfort & Defourny 1994).
In conclusion, along simple strain paths, the 3G model gives slightly more accurate results
than the classical theory but its main advantage is to give a metallurgical background to the
mathematical formulation of the material behaviour. It has the basic ingredients to model
the events happening in complex loading paths but specific improvements are needed to
predict texture evolution or pre-strain effects. Its identification is not too heavy and its
FEM implementation is not straightforward but tractable.

The rate of mobile dislocation creation is related to a potential function of stresses (Mecking & Lucke
1970).

The texture effect responsible for Lankford’s coefficient value directly modifies the flow rule.
Modelling the Plastic Anisotropy of Metals 61

6.2.2 Aifantis’ models


Aifantis 1987 is a typical isotropic macroscopic model imbued from single crystal plasticity.
It is based on a scale invariance concept, which assumes that the structure of the resulting
equations is preserved during a transition from the micro-scale to the macro-scale.
The single crystal model proposed by Aifantis considers a single family of dislocations
moving along a slip. The vector n is the normal unit to the chosen slip plane and b a unit
vector in the slip direction. The microscopic state is represented by two sets of equations:
i) the mass and momentum balance of dislocations; ii) constitutive equations linking stress
state and dislocation state. Such a microscopic model helps to predict heterogeneity of
plastic flow such as shear bands or Portevin-Le Chatelier bands (Aifantis 1987).
The microscopic stress quantities σ micro , σ micro − α , α (dislocation stress) are as-
sumed to preserve their character and interrelationship during the transition to macro scale.
They are respectively identified with the macroscopic total stress σ macro , the macroscopic
effective stress σ effective and the macroscopic back stress α macro . As macroscopic plas-
ticity smoothes out plastic micro-heterogeneities, it is assumed that, at macroscopic level,
the divergence terms of the microscopic model have no influence and can be dropped. The
macroscopic plasticity theory neglecting volume changes, the climb process of dislocations
is also neglected and just the glide components are kept. The stress tn n⊗n, which accounts
for presence of dislocation dipole and decomposition, can be kept but is often neglected as
in Prager’s kinematic hardening rule. Consequently, the macroscopic relations chosen by
Aifantis are:
ρ̇ = ĉ(ρ, jb , τeffective ) (87)
a1b − a2b τeffective − a3b jb = 0 (88)

σ effective
=σ macro
−α macro
(89)
α macro
= tA A + tn n ⊗ n (90)
with ρ the dislocation density, jb the glide dislocation flux in the slip direction b , A is an
orientation tensor “equivalent” to the symmetric part of the Schmid’s tensor (see relation
(15)):
ε˙ p macro = γ̇ p A (91)
τeffective is the CRSS transported in a macroscopic scale: projection of σ effective on A
tensor, see relation (16). ĉ represents the generation, immobilisation or annihilation of
dislocations. a1b , a2b , a3b are assumed functions of ρ. The coefficient a1b measures the
lattice-dislocation interactions and is responsible for yielding. The coefficients a2b , measures
the effect of Peach-Koehler’s force, while a3b measures the drag associated with dislocation
motion and is responsible for internal damping and viscoplastic flow. The coefficient tA
measures the interaction forces between dislocations and is responsible for the development
of back stresses. The macroscopic yield condition is directly deduced from (88).
In Aifantis 1987, the macroscopic A tensor results from the macroscopic principle of
maximum dissipation for maximum entropy production. This is related to the power asso-
ciated with a dislocation motion along its slip plane:

τ effective jb > 0 ⇐⇒ (σ macro


−α macro
): ε˙ p macro
>0 (92)

This leads to a maximization problem:


effective
maximum of (σ ij Aij )
62 A.M. Habraken

with applied constraints:


1
trA = 0; trA2 = (93)
2
by analogy to the microscopic analysis where A is the symmetric part of Schmid’s tensor.
The solution found by Aifantis is:

σ̂ effective
A=  (94)
2 12 (σ̂ effective )2

where σ̂ effective is the deviatoric part of stress tensor σ effective . This solution allows to
retrieve Prandtl-Reuss’ flow rule and von Mises’ yield criterion:
γ̇ p  a1b
ε˙ p macro
= √ σ̂ effective
J2 = = σF (95)
2 J2 a2b
with J2 = (1/2)σ̂ effective
ij σ̂ effective
ij .
Using standard kinematic arguments in conjunction with relation (91), Aifantis obtains,
at the macroscopic level, the plastic spin formulation. His result is equivalent to the ex-
pression adopted in usual theories of crystal plasticity for the plastic spin if only one slip
system is considered (see Figure 23):
p
Ω = γ̇ p Z (96)
where Z is the non-symmetric part of Schmid’s tensor. By analogy to the microscopic state,
the practical macroscopic relation for the plastic spin is derived by eliminating tensor Z
from (96) thanks to the use of expressions (90) and (91):
p
Ω = −t−1
n (α
macro
ε˙ p − ε˙ p α macro
) (97)
The above relations define Aifantis’ macroscopic model, when a random texture results
in an isotropic macroscopic behaviour. Its interest is for instance to be able to predict
the development of axial strain due to torsion in free-end cylindrical specimens or the
development of axial stress due to torsion in fixed-end cylindrical specimens (Swift effect).
In cases where the effect of grain orientations cannot be neglected, Ning & Aifantis 1996
propose a micro-macro model relying on discrete set of crystals. This approach is closed
to the ones presented in Section 5.2., but has the advantage of providing final analytical
macroscopic relations. The simplified relations (87)-(90), (94), (95), (97) are now assumed
to be the constitutive relations for a single crystal. Aifantis’ anisotropic model applies
Taylor’s assumption of equality of the velocity gradient at microscopic and macroscopic
levels. Aifantis introduces an additional “texture spin” Ω t not mentioned in Section 3.2.
There the material rotation was only subdivided into a plastic spin Ω p and a crystal lattice
rotation Ω L . Ning & Aifantis’ texture spin Ω t results from grain boundary constraints
to maintain deformation compatibility. The crystal lattice rotation Ω L is derived as the
difference between the global spin (Ω micro = Ω macro ), related to the macroscopic velocity
gradient, and two types of plastic spins, the usual plastic spin Ω p due to crystal slip given
by relation (96) and the texture spin Ω t due to grain boundary effects:
L
Ω =Ω macro
−Ω p−Ω t
(98)
Ning & Aifantis define a crystal orientation by a unit vector a oriented in the crystal
direction, expressed as a function of the slip system identifiers n, b and of the angle between
Modelling the Plastic Anisotropy of Metals 63

a and b. Then, they use the classical concept of Orientation Distribution Function f , which
represents the probability for a grain to be oriented along a at time t. The evolution of
the ODF is prescribed by a conservation rule as already presented in Section 5.3.3, which
can be solved analytically in the case of an initial random texture and simple deformation
gradient tensor.
Ning & Aifantis’ micro-macro link differs from the classical volume average procedure
expressed in its integrated form (78), or in its discrete form (59) or (60). Such a classic
average procedure does not consider the effect of morphological texture. For instance, the
fact that the orientation of large grains may have a more pronounced effect than that of
small grains. The following improved average relation is proposed:

xmacro = K(a): xmicro f (a, t) da (99)

where xmicro is either the deviatoric Cauchy stress or the back stress at the crystal level and
xmacro is the associated macroscopic value. The fourth order tensor K, called “texture”
tensor, is a function of a. Due to the stress symmetry, it is a transversely isotropic tensor
finally defined by only three independent material parameters. The overall plastic flow
rule and yield condition are then obtained by combining the flow rule and yield condition
assumed at the crystal level (95) and the average stress relation (99). This results in the
macroscopic flow rule:
γ̇ p
ε˙ p = K−1 : σ̂ effective
(100)
2σF (γ p )
and the yield condition:

1 effectif
σ̂ K−1 K−1 σ̂ effectif = σF (γ p ) (101)
2
where the overall texture tensor K, is defined by the average formula:

K = Kf (a, t) da (102)

Relation (100) is similar to the previous phenomenological relations describing yield


behaviour of metallic materials by the introduction of “modified” stress tensors as, for
instance, in Karafillis & Boyce 1993, presented in Section 2.2.
The evolution of the macroscopic back stress and the plastic spin Ω p are deduced
from their microscopic value, Taylor’s assumption and the same averaging procedure. The
macroscopic counterpart of the additional texture spin Ω t describes the overall average
grain rotation due to grain boundary constraints. Thus, it cannot be expressed in terms
of the average slip processes alone. The average procedure (99) is slightly modified and
applied now to a vector associated with the skew-symmetric tensor Ω t . The fourth order
tensor K is reduced to a second order tensor, isotropic function of vector a, taking into
account 2 material parameters related to plastic strain history and temperature as well as
to grain size and shape.
Like Aifantis’isotropic macroscopic model, Aifantis’anisotropic macroscopic model has
been applied to the prediction of Swift’s effect. For this simple torsion loading state, it
is possible to find an analytical form of the ODF function. Analytical and experimental
results are very close as shown in Figure 50.
The yield behaviour in tension coupled to torsion has been further studied for a the-
oretical material. For instance, Figures 51 and 52 show the evolution of the initial yield
locus. They respectively correspond to a tension dominated deformation state and a shear
64 A.M. Habraken

-5

Axial stress ( MPa )


-10

-15

Experiment
-20 Theory

-25
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Shear strain

Figure 50. Predicted and experimental values for the non-monotonic evolution of axial stress
in a copper bar under fixed-end torsion (from Ning & Aifantis 1996)

4
0.0
3
0.1
Normalized Shear Stress

0.5
2

-1

-2
-4 -2 0 2 4
Normalized Axial Stress

Figure 51. Evolution of yield surface in tension-torsion for different tensile strains: 0, 0.1,
0.5, and a ratio shear/axial strain of 1 (from Ning & Aifantis 1996)

dominated deformation mode. The results show that the texture development causes both
rotation and distortion of the yield surface, the shear mode having a more pronounced ef-
fect than the tension mode. A comparison of model predictions with available experimental
data in small deformations (effect of plastic spin neglected) is presented in Figure 53 for
304 stainless steel in tension-torsion.
In short, these macroscopic models imbued from single crystal plasticity can predict
mechanical behaviour quite accurately. Ning & Aifantis 1996 analytically compute the ODF
evolution in some simple cases. However, the extension of this model to a general velocity
gradient seems heavy from both computational and theoretical points of view. The latter
model has the advantage of giving microscopic fundamental bases for phenomenological
models, which propose the same form for the yield locus function.
Modelling the Plastic Anisotropy of Metals 65

0.0
5
6

Normalized Shear Stress


0.1
4
0.50.0
5
3 0.1
Normalized Shear Stress

4
2 0.5
3
1
2 0

1 -1

0 -2
-5 -4 -3 -2 -1 0 1 2 3 4 5
-1 Normalized Axial Stress
-2
-5
Figure 52. Evolution -4 yield-3surface
of -2 in tension-torsion
-1 0 1for different
2 3tensile4 strains:
5 0; 0.1; 0.5
and a ratio shear/axial strain of 5 (from
Normalized AxialNing & Aifantis 1996)
Stress
Fig. 52
500
500 Experiment ( γ p = 0.0)
im ent(γ( γp ==0.
400 Experiment p 127
Exper 0) E-6)
im ent(γ( γp = =127
400 Experiment p 713 E-6)
Exper E-6)
300 Experim ent(γ p = 713 E-6)
Theory
Shear Stress (MPa)

300 Theory
Shear Stress(MPa)

200
200
100
100
0 0

-100
-100

-200
-200

-300
-300
-400
-400 -200
-200 00 200
200 400
400
AxialStress
Axial Stress(MPa)
(MPa)
Fig. 53
s 13 predictions and experimental data for
Figure 53. Comparison between theoretical
the yield surface of 304 stainless steel in tension-torsion, γ p means here
equivalent plastic strain (from Ning & Aifantis 1996)

6.3 Analytical Yield Loci Computed from Texture Data

Section 6.3.1 presents the classical way to obtain a macroscopic yield locus thanks to clas-
s11 - sAs
sical single crystal plasticity combined with texture description.
33
this approach does not
give an analytical formulation easy to implement in FEM codes, 2 different ways to derive an
analytic function for the flow surfaces have been proposed. Some are described in Sections
6.3.2 to 6.3.4.

Fig. 54
66 A.M. Habraken

6.3.1 Polycrystal yield locus obtained by Taylor’s approach


The average Taylor’s factor (see Section 4.4 relation (50)) appears as a material parameter
able to express the response of a textured material submitted to a given strain rate. At the
level of a polycrystal, the average plastic work dissipation per unit is (see relation (49)):

ε̇peqmacro τ̄c M̄ (Uε̄˙ p ) = σeq


macro p macro
ε̇eq =σ macro
: ε˙ p macro
(103)
One can see that, if the macroscopic strain rate ε˙ p macro is known, M̄ (Uε̄˙ p ) can be
computed and all stress tensors σ macro that satisfy (103) constitute hyperplanes in stress
vectors space. The strain rate vectors ε˙ p macro are perpendicular to these hyperplanes.
In practice, as already explained, only the deviatoric stress tensors are considered in the
plastic state, so the stress space is a 5 dimensional space. The deviatoric stress σ̂ macro
corresponding to ε˙ p macro is one point of the hyperplane. According to the normality rule,
the yield locus must be tangent to the hyperplane at the location of σ macro (see Figure 54).
Since the yield locus must be convex, it is the inner envelope of the hyperplanes associated
with all possible strain modes. The practical way to build 2-dimensional projections of the
yield surface is described in Canova et al. 1985, Lequeu et al. 1987a, Van Houtte 1992.
Some details on Van Houtte’s approach will be given in Section 6.3.4.

σ11 −σ33
Figure 54. Example of projection of the yield locus onto the subspace 2
, σ13
(from Canova et al 1985)

6.3.2 Proposals applying Montheillet’s concepts


Lequeu et al. 1987b follow the approach of Montheillet et al. 1985, where the principal axes
of anisotropy are chosen to coincide with the axes of the studied texture component, rather
than with the symmetry axes of the workpiece. In summary, this approach adjusts an
analytical function on Taylor-Bischop-Hill’s single crystal locus; then, by rotation, adapts
it according to macroscopic axes and uses a Sachs’ approach when the material has more
than one texture component.
Darrieulat & Piot 1996 apply the same type of approach to f.c.c materials characterised
by 12 slip systems. However, they consider a more accurate representation of the micro-
scopic behaviour and take into account the ODF function to represent the texture effect.
It is described hereafter.
Modelling the Plastic Anisotropy of Metals 67

Beginning with the plastic behaviour at the single crystal level, Darrieulat & Piot 1996
express Schmid’s law by an analytical form obtained through a classical mathematical
property of power averages already proposed by Arminjon 1988:
 12  1
 σ̂ : As n n
Fp (σ̂ ) =   =1 (104)
 τs 
s=1 c

where As is the symmetric part of Schmid’s tensor (15) associated with the slip system s.
Function (104) is differentiable, strictly convex and arbitrarily close to the inner envelope
of the hyperplanes of equations:
 
 σ̂ : As 
 
 τs  = 1 s = 1, . . . , 12 (105)
c
Applying the normality rule to the yield surface, the plastic strain rate is computed by:

  n−1
1
−1
12
 
ε˙ p = λ̇Fpn sgn(σ̂ : As )σ̂ : As  As (106)
s=1
It is interesting to note that this form is close to the macroscopic flow rule derived when
the rate sensitive approach is applied at the crystal level (see relation (30)). Figure 55
compares yield loci based on expression (104) for different values of n with results from
Lequeu et al. 1987a.

Figure 55. For cube orientation, yield stress versus the angle to the Rolling Direc-
tion, values predicted by Darrieulat & Piot and by Lequeu (Darrieulat
& Piot 1996)

When an orthotropic material is known through its N crystallographic components, re-


lation (104) provides a differentiable representation of the mechanical behaviour that can be
attributed to each component. The macroscopic behaviour is some average between them.
A simple physical assumption close to Khan & Cheng’s 1996 proposal is that polycrystals
behave like crystals possessing not 12 slip systems but 12 × N slip systems, the orientations
68 A.M. Habraken

of which are given by the texture data. Each of the N sets contributes proportionally to
its volume fraction.
Let the texture be defined by an ODF f (g) function of Euler’s angles ϕ1 , φ, ϕ2 . If Euler’s
space is covered by a step ∆, one value fi of the ODF represents the average intensity of the
orientations in the solid angle [ϕ1 −∆/2, ϕ1 +∆/2]x[φ−∆/2, φ+∆/2]x[ϕ3 −∆/2, ϕ3 +∆/2].
It is also the volume fraction of crystals belonging to this orientation space sector. The
number of different crystallographic orientations taken into account N directly depends on
the size of ∆.
Using the same power average as in relation (104), the plastic behaviour of the poly-
crystalline material is given by:
N 12  n  n1
  
Fp = fi σ̂ micro : As  = τ̄c (107)
 i i
i=1 s=1
where Taylor’s assumption of a common value of CRSS is adopted for all the slip systems
and all the crystals. The subscript i identifies the texture component. This tensorial
expression being the average of strictly convex and differentiable terms, it is also strictly
convex and differentiable. It can be easily computed no matter what the chosen axes are.
If n = 2, relation (107) is a quadratic criterion similar to Hill’s 1948 yield locus. Dar-
rieulat & Piot 1996 give the 6 coefficients of Hill’s 1948 criterion as functions of fi and As .
The macroscopic mechanical behaviour (σ̂ macro , ε˙ p macro ) of a polycrystalline materials
can be predicted from each of its component (σ̂ micro
i , ε˙ pi micro ) if:

• either a uniform stress (σ̂ micro


i = σ̂ macro ) or a uniform strain rate (ε˙ pi micro =
ε˙ macro
) assumption is accepted. In Darrieulat & Piot (1996), the stress uniformity is
chosen;
• a condition of homogenisation is applied:

N

fi σ̂ i : ε˙ i = σ̂ macro
: ε˙ macro (108)
i=1
The details are given in Darrieulat & Piot’s 1996 paper. Their conclusions are that pro-
posal (107) gives good results for low values of n (up to 10) but assigns a too large influence
to crystallographic texture when n increases. Their validations consist in predictions of
uniaxial yield stresses and Lankford’s coefficients as functions of the angle with the rolling
direction. This is done for a single component texture or real materials such as aluminium
3004 or 5182. Note that Darrieulat and Montheillet (2003) just proposed a new version of
their model restricted to Hill type quadratic assumption.
6.3.3 Arminjon, Bacroix, Imbault ... ’s potential formulation
Arminjon 1988, Arminjon & Bacroix 1991, Arminjon et al. 1994 propose an identification
of the plastic work rates to derive a yield criterion applied to polycrystalline materials.
When a quadratic form is assumed, the macroscopic anisotropy parameters become explicit
functions of the texture coefficients. Their approach is summarized hereafter.
In visco-plastic models, the existence of two dual convex potentials E and Ec is usually
assumed. They define the relationship giving the stress tensor as function of the plastic
strain rate tensor as well as the inverse relation:
∂E ∂Ec
σ = ε˙ p = (109a, b)
∂ ε˙ p ∂σ
Modelling the Plastic Anisotropy of Metals 69

where functions E and Ec exchange by Legendre’s transformation. In case of rate-independent


standard plasticity, function E is simply the rate of plastic work Ẇ p . The complementary
function Ec cannot be defined, since the strain rate ε˙ p is only determined by the stress
tensor σ up to the plastic multiplier λ̇(ε˙ p ). The yield locus function and the energy po-
tential are equal in associated plasticity model. The flow rule described by relation (1) is
used:

∂Fp
ε˙ p = λ̇(ε˙ p ) (110)
∂σ
In this section, the yield locus is expressed as:

Fp (σ ) = τ̄c (111)
Following Hill 1987, Fp can be formulated as an homogeneous function of order one with
respect to positive multipliers. This property implies :

∂Fp
Fp (σ ) = σ : (112)
∂σ
Relations (110) - (112) allow to give the following expression of the rate of plastic work:

Ẇ p = σ : ε˙ p = Fp (σ )λ̇(ε˙ p ) = τ̄c λ̇(ε˙ p ) (113)


Using (109a) and (113), Arminjon writes:

∂ Ẇ p ∂(τ̄c λ̇(ε˙ p )) ∂ λ̇(ε˙ p )


σ = p = p = Fp (σ ) (114)
∂ ε̇ ∂ ε˙ ∂ ε˙ p
Comparing the macroscopic expression of the plastic power dissipation (113) with the
micro-macro one (49), presented as an average value, one can find physical interpretations
for each term. τ̄c is the average CRSS in the polycrystal. λ̇ is naturally expressed from the
texture alone and is characterized by the average Taylor’s factor (relation (50)). The above
relations assume the equality between the macroscopic strain rate and the microscopic one.
So, all the further analytical formulation is based on the so-called Taylor’s hypothesis.
Arminjon & Bacroix 1991 define four requirements that functions λ̇(ε˙ p ) in strain rate
space, or Fp in stress space, have to fulfil:

• convexity,

• respect of orthotropic material symmetry,

• homogeneity,

• use of deviatoric tensors only.

If one chooses a polynomial function, the convexity must be checked. For instance,
Arminjon and Bacroix propose a 4th order homogeneous function F of plastic strain rate,
which respects the three other characteristics:


22
F = αk Xk
k=1
70 A.M. Habraken

X1 = (ε̇p11 )4 X2 = (ε̇p22 )4 X3 = (ε̇p23 )4 X4 = (ε̇p13 )4

X5 = (ε̇p12 )4 X6 = (ε̇p11 )3 ε̇p22 X7 = (ε̇p22 )3 ε̇p11 X8 = (ε̇p11 )2 (ε̇p22 )2

X9 = (ε̇p11 )2 (ε̇p23 )2 X10 = (ε̇p11 )2 (ε̇p13 )2 X11 = (ε̇p11 )2 (ε̇p12 )2 X12 = (ε̇p22 )2 (ε̇p23 )2

X13 = (ε̇p22 )2 (ε̇p13 )2 X14 = (ε̇p22 )2 (ε̇p12 )2 X15 = (ε̇p23 )2 (ε̇p13 )2 X16 = (ε̇p23 )2 (ε̇p12 )2

X17 = (ε̇p13 )2 (ε̇p12 )2 X18 = ε̇p11 ε̇p22 (ε̇p23 )2 X19 = ε̇p11 ε̇p22 (ε̇p13 )2 X20 = ε̇p11 ε̇p22 (ε̇p12 )2

X21 = ε̇p11 ε̇p23 ε̇p13 ε̇p12 X22 = ε̇p22 ε̇p23 ε̇p13 ε̇p12
(115)
where symmetry εij = εji is assumed and only i ≤ j terms are present. A scaled function
is used in practice as function λ̇:

Ẇ p 22
p p p 3
λ̇(ε˙ ) = = F (ε̇ )/(ε̇eq ) = αk Ψk (ε˙ p ) (116)
τ̄c k=1

with Ψk (ε˙ p ) = Xk (ε˙ p )/(ε̇peq )3 .


Note that symmetry σij = σji is not assumed in (114) but taken into account in (116).
So the stresses are computed by:

∂ λ̇
σij = τ̄c if i=j
∂ ε̇pij
1 ∂ λ̇
σij = τ̄c if i < j (117)
2 ∂ ε̇pij

Let us explain how the choice (116) leads to represent the coefficients αk by a linear
function of Clµν coefficients of the ODF. Here for simplicity, a single index notation Ci is
adopted to identify these Clµν coefficients (relation 41):

max µmax
l (l) ν
max µν  I
f (g) ∼
= Clµν T̈˙ l (g) =⇒ f (g) Ci Ti (g) (118)
l=0 µ=1 ν=1 i=1

where I is the maximum index of Fourier’s coefficients identified by a single index notation.
Looking at the above relations (116), (118), (49) and (50), one can express the function λ̇,
after algebraic manipulations, by a linear function of coefficients Ci :
I
p Ẇ p Ẇ p 
λ̇(ε˙ ) = = = Ci Mi∗ (ε˙ p ) (119)
τ̄c τ̄c i=1
where function Mi∗ is computed from Taylor’s factor and the harmonic function Ti . The
identification of the material parameters αk , introduced in relation (116), directly results
from the 2 expressions (119) and (116) of function λ̇(ε˙ p ):
I

αk = βki Ci k = 1, . . . , 22 (120)
i=1
Modelling the Plastic Anisotropy of Metals 71

Arminjon et al. 1994 have found that, for a given i, the best approximation of coefficients
βki is based on a set of 16200 values of ε̇ p and the comparison between:


22
Mi∗∗ (ε˙ p ) = βki Ψk (ε˙ p ) and Mi∗ (ε̇ p ) (121)
k=1

Such computation is done only once with a Taylor’s model and it has been checked
that for steel texture, I = 12 is enough. This fact means that the 22 coefficients αk are
not independent because relation (119) uses 13 coefficients to describe plastic dissipation
(12 coefficients Ci and the value of τ̄c ). The above process of parameters identification is
independent of the Ci values and hence independent of texture. This is why it can be done
only once.
The above explanations give the procedure to start from the choice of a scaled fourth
order polynomial function λ̇(ε˙ p ) and to finally provide an expression of this function pa-
rameters as linear relations of coefficients Ci . The same approach can be used if one chooses
λ̇ as a second order polynomial function or Hill’s yield locus. The application of this method
will give a direct computation of 2nd order series or Hill’s coefficient from Ci coefficients.
Figure 56 compare sections of the yield loci computed by different methods for 2 in-
dustrial steels I1 (Al killed) and I5 (IF, Ti). The following sections are represented:
σ12 /τ̄c = 0, 0.5, 1 and 1.5 and, at the center, the value of σ12 /τ̄c for σ11 = σ22 = 0 is
given. The above approach computes an analytical yield locus from the initial texture.
Figure 57 for steel I1 (Al killed) shows Lankford’s coefficients and yield loci deduced from
texture measurements in the initial state and after biaxial tension stopped at two different
levels of deformation. In this Figure, one observes that the effect of texture evolution can-
not always be neglected. Imbault & Arminjon 1993 propose a semi-analytical method to
take it into account in their analytical expression of the yield locus. In fact, the principle
is simple. They assume that a linear operator, which is adjusted by comparison with a
polycrystalline model, can express the evolution with strain of the Clµν coefficients. Of
course, this linear operator depends on strain rate, which means that numerous strain rate
tensors (1800) must be used to establish it. Once the Clµν coefficients are known, the above
identification process of the analytical expression for the yield locus can be activated.
6.3.4 Van Houtte’s potential formulation
As Arminjon, Van Houtte 1994 uses the method of the dual plastic potentials to derive
convenient formulae for calculating yield loci of rate insensitive anisotropic materials. In
practice, the implementation in FEM code of such a yield locus in strain space has been
performed by one of his Ph.D. student, Van Bael 1994, in collaboration with the university
of Birmingham (Wang et al. 1992). The yield locus in stress space has been implemented
in another FEM code by Winters 1996, another Ph.D. student in collaboration with M&S
team (Munhoven et al. 1995a and b). Further developments such as the coupling with
Teodosiu’s hardening model (Hiwatashi et al. 1997) and the formulation in strain rate
space in the LAGAMINE code (Hoferlin et al. 1998, Hoferlin et al. 1999b) is described in
details in Hoferlin’s Ph.D. thesis (2001). The hereafter description summarizes Van Houtte
and co-workers’ approach.
For rate-independent standard plasticity and formulation in strain rate space, the rela-
tion (109a) computes the stress tensor where the potential E is directly given by the rate
of plastic work Ẇ p macro . In practice, as already suggested by Lequeu et al. 1987a, Van
Houtte’s team works in a five dimensional space. The plastic strain rate is classically as-
sumed to be deviatoric and only the deviatoric stress tensor matters with regard to plastic
deformation (Van Houtte et al. 1989). So, the tensors have only 5 independent components
and are replaced by 5-dimensional vectors. This transformation, “tensor Vij - vector vp ”,
72 A.M. Habraken

22 c
- Taylor Serie 4
Hill 1948 - Bishop Arminjon
- Hill

11 c

22 c Serie 4
- Taylor
Hill 1948 - Bishop Arminjon
- Hill

11 c

Figure 56. Yield locus sections computed by classical Hill’s model, crystallographic
Taylor-Bishop-Hill’s model or Arminjon’s 4th order series, for 2 indus-
trial steels I1 (Al killed) and I5 (IF, Ti) (from Arminjon et al. 1994)

can be defined in different ways. The version chosen by Winters 1996 assumes, as Van
Houtte 1988:
 
1 3 3
v1 = √ (V11 − V22 ) v2 = (V11 + V22 ) = − V33
2 2 2
√ √ √
v3 = 2V23 v4 = 2V31 v5 = 2V12 (122)

This vector representation has the following property:

V: W = Vij Wij = vw = vp wp (123)


The vector forms of plastic strain rate and deviatoric stress tensors are respectively
noted ė and s and their components ėp and sp . Using the potential relation (109a) Taylor’s
hypothesis and micro-macro relations (49), (50), the deviatoric stress is computed by:

 
∂ ∂
σ̂ = Ẇ p micro
(ε˙ p micro
, g)f (g) dg = (ε̇p macro τ̄c M̄ (Uε̄˙ p )) (124)
∂ ε˙ p ∂ ε˙ p eq


σ̂ = τ̄c (ε̇p macro M̄ (Uε̄˙ p )) (125)
∂ ε˙ p eq
Modelling the Plastic Anisotropy of Metals 73

2.1 4

2.2 3

2 2
r Lankfor d coefficient

1.8 1

c
1.56
1.6 0.0

22
1.4 -1

1.2 initial -2
biaxial ε = 0.2
biaxial
1.0 biaxial ε = 0.52 -3
tension ε = 0.2
0.8 -4
0.0 1.5 3.0 4.5 6.0 7.5 9.0 -4 -3 -2 -1 0.0 1 2 3 4
Alpha angle
11 c
4 4

3 3

2 2

1 1
c
c

1.57 1.54
0.0 0.0
22
22

-1 -1

-2 -2

biaxial
-3 initial -3
tension ε = 0.52
-4 -4
-4 -3 -2 -1 0.0 1 2 3 4 -4 -3 -2 -1 0.0 1 2 3 4

11 c 11 c

Figure 57. Lankford’s coefficients and yield loci computed from texture measure-
ments for a steel in its initial state and after biaxial tests performed up
to 2 different levels (from Imbault & Arminjon 1993)

where Uε̄˙ p = ε˙ p macro /ε̇peqmacro is a strain mode as defined in relation (23). This is the strain
rate space formulation. The relation (125) assumes that the average CRSS is independent
of further applied strain rate modes. It helps to understand how Van Houtte and co-workers
completely dissociate the size (τ̄c ) and the shape of the yield locus ∂/∂ ε˙ p (ε̇peqmacro M̄ (Uε̄˙ p )).
Winters 1996 proposes to use an isotropic hardening model. He updates τ̄c via a simple
Swift’s law applied at the macroscopic stress strain level. The micro-macro link used to
identify τ̄c (Γ̄) is based on the first term of the relation (56) where Γ̄ is the total polycrystal
slip. The stress σ and the strain ε are those of a uniaxial test. Neglecting the second
term of (56) means that the texture evolution is dropped. This will be assumed in further
developments, except when clearly specified.
Hiwatashi et al. 1997 applies a kinematic hardening assumption and slightly modifies
74 A.M. Habraken

the relation (125) which becomes:



(ε̇p macro M̄ (Uε̄˙ p ))
σ̂ − α = τ̄c (126)
∂ ε˙ p eq
where α is the back-stress and defines the updated center of the yield locus. The evolutions
of α and τ̄c follow the model proposed by Teodosiu & Hu 1998.
The size and position of the yield locus being defined by the choice of the hardening
model, what about its shape? Relation (125) clearly shows the role of the texture via
the average Taylor’s factor M̄ (Uε̄˙ p ). u or up are the vector forms of the strain mode.
The average Taylor’s factor can be approximated by an analytical function Q(up ) of the
components up :

M̄ (Uε̄˙ p ) = M̄ (u) ≈ Q(u) = Fp1 p2 p3 ...pN up1 up2 up3 . . . upN


N = order of series expansion pi = 1, . . . , 5 i = 1, . . . , N (127)
For instance, if N is reduced to 2, there are 15 coefficients:

Q(u) = F11 u1 u1 + F12 u1 u2 + F13 u1 u3 + F14 u1 u4 + F15 u1 u5 + F22 u2 u2 + . . . F55 u5 u5 (128)


Van Bael 1994 has extensively described the symmetry properties of such an analytical
expression, the main drawback of which is its lack of convexity. He explains that an odd
order choice gives a non centro-symmetrical yield locus, which allows to model stress differ-
ential effects. It has been checked by Van Bael et al. 1996 and Munhoven et al. 1997 that
the 6th order is required to reproduce, with such an analytical description, the accuracy of
polycrystal approaches (Figure 58).

2.5 Taylor - Bishop - Hill 6


Lankford coefficient

M - series order

4
2.0

1.5
2

0 30 60 90
Angle α to RD (°)

Figure 58. Lankford’s coefficient of a classical interstitial steel computed by a poly-


crystal model or by means of 2nd, 4th, 6th order series in strain rate
space (from Munhoven et al. 1997)

This choice of N = 6 leads to 210 coefficients Fp1 ...p6 . To identify them, it is interesting to
note that any strain mode tensor can be represented by means of 4 independent parameters
and not 5 because Uε̄˙ p : Uε̄˙ p = 3/2. Van Houtte 1994 demonstrates:
 
U11 U12 U13
 
Uε̄˙ p =  U22 U23  Uε̄˙ p = RT UX
ε̄˙ R
p (129)
SYM U33
Modelling the Plastic Anisotropy of Metals 75

with:

R = R(β1 , β2 , β3 ) and (β1 , β2 , β3 ) ∈ (0, 2π) × (0, π/2) × (0, π) (130)

 X

U11
 X
U22  3
UX   tr (UX and UX X
ε̄˙ =  but ε̄˙ ) = 0 ε̄˙ : Uε̄˙ = (131)
p p p p
X
U33  2

   
X π X π X
U11 = cos β4 − , U22 = cos β4 + , U33 = − cos(β4 ) (132)
3 3
The 3 Euler’s angles β1 , β2 , β3 define the orientation of the strain rate mode principal
directions with respect to the sample reference system. The fourth angle gives the deviation
of the current strain rate mode with respect to an axisymmetric compression along the
third principal axis. Finally discrete variations of βi, for instance ∆β1 = ∆β2 = ∆β3 = 10◦
and ∆β4 = 7.5◦ define a discrete set of strain rate modes (around 70300). The average
Taylor factor is computed by the texture and crystalline approach (full Constrained Taylor-
Bischop-Hill’s model) for each of these modes. Then, the coefficients of the series expansion
are provided by a least square fit of Q function (127).
Relation (125) calculating the stress from the dissipation, is modified to take into ac-
count both the vector formulation and the analytical expression of the average Taylor’s
factor:
∂(ε̇peq Q(u))
sp = τ̄c (133)
∂ ėp
In FEM, one usually needs the yield locus point corresponding to a given stress direc-
tion s∗ . So, Legendre’s transformation must be applied. Let ss∗ describe one stress vector

belonging to the yield locus: the scalar factor is the vector norm sp sp of the stress point,
it is called stress radius, and s∗ is its unit vector direction. Assuming that u is the strain
rate mode associated with this stress direction s∗ and using the average Taylor’s factor
definition, one gets:
ss∗ 
M̄ (u) = min u u = minimum in strain rate space (134)
τ̄c
where u represents all the possible strain modes of the strain rate space. As u is actually
unknown, one has to minimize the ratio:

s M̄ (u )
= min u  ∗ (135)
τ̄c us
Or with the approximation (127):
s Q(u )
= min u  ∗ (136)
τ̄c us
The scaled factor τ̄c clearly shows that the shape does not depend on the size of the
yield locus. If a formulation in strain rate space is adopted, Q(u) is known and (136) is used
to perform the minimization. Figure 59 explains this procedure graphically. As reported
above, the field of strain rate modes can be covered by 4 independent parameters, which
limits the operation duration. Additionally, Hoferlin et al. 1999a propose a way to speed up
76 A.M. Habraken

σ 22/ τ c
u1

u2
M ( u1 )
M ( u2 )

s* σ11 / τ c

Figure 59. Graphical view of the minimization procedure to find a point on the
yield locus (adapted from Hoferlin et al. 1999a)

significantly the minimization: with a BFGS type method instead of a classical Newton’s
method and by splitting the minimization process.
As the convexity of the function Q(u) in strain rate space is not perfect, gathering all
stress points computed by (136) does not produce a convex yield locus in stress space. Some
fishtails appear as demonstrated by Figure 60. This drawing applies to an almost single
f.c.c. crystal texture generated around the Goss’ orientation (gaussian distribution with
11◦ spread around (011)[100]). Left Figure 60 shows the (π-section of one yield locus in
strain rate space. For a pure single crystal, linear segments compose this locus. The series
expansion reproduces linear segments by oscillating around them. The application of the
minimization (136) provides a yield locus in stress space with fishtails: see right Figure 60.
Real materials generally present a less sharp texture; so their yield loci are smoother
and the 4th or 6th order series expansions should be convex descriptions. However as non
convex loci fail to bring convergence in FEM simulations, some secure approach should be
implemented. Hoferlin’s proposal consists in repeating the minimization (136) with different
starting guesses for u and choosing the smallest result s, as the one giving the point on the
yield locus. This procedure cuts off the fishtails and prevents convergence problems due
to lack of convexity. Recent publication Van Bael & Van Houtte 2002 proposes another
analytical expression of the average Taylor coefficient (127) with the advantage that the
convexity can be strictly imposed.
Another choice can be to implement an analytical yield locus in stress space, computed
from texture and crystal plasticity approaches. The above presentation explains how to
find points belonging to the yield locus. For a set of directions s∗ in stress space, (135)
or (136) provide the stress radius s. These two ways to obtain sets of yield points are not
exactly equivalent:
Modelling the Plastic Anisotropy of Metals 77

ε
Pp  σ3ˆ
33c/ τ
3333

N=4
+

+
ε11pp
11 ε22Pp
22
σ1ˆ11/ cτ σˆ2 / cτ
22
N=6

Figure 60. π-plane sections of the yield loci expressed in strain rate space (left)
and in deviatoric stress space (right) for a f.c.c polycrystal with Goss’
texture component (from Hoferlin et al. 1999a)

p 3 σ 33 / τc 3
ε33
2 2

1 1

45 45
1 1 1 1
2 2 2 2
3 3 3 3
p
ε11 p
ε22 σ22 / τc σ 22 / τ c

Figure 61. π-plane sections of the yield loci expressed in strain rate space (left)
and in deviatoric stress space (right), relative to a polycrystal with cube
texture component and a Gaussian spreading of 16.5◦ (adapted from
Winters 1996)

• The first way uses relation (135), where M̄ (u ) has no analytical expression and is
computed for a certain number of strain rate modes. It is similar to the polycrystalline
method to get yield locus sections (see Section 6.3.1.). By analogy, this method is
called “geometrical approach”. These yield points generally describe a nearly convex
yield locus.
• The second way uses relation (136) and gives a set of points defining an approximate
yield locus, not necessarily convex as shown by Figure 61.

In both cases, one has not yet reached an analytical yield formulation of the stress yield
locus Fp . To reach this goal, one must fit an analytical function on these points. By analogy
78 A.M. Habraken

to the work in strain rate space, a series expansion is applied. Relations (137) and (138)
present two possible choices:

s s ∂(s/Q(s∗))
= Q(s∗ ) → Fp = ∗
− τc = 0 → u∗p = λ (137a, b, c)
τ̄c Q(s ) ∂s∗p

s 1 ∗ ∗ ∂(sQ(s∗ ))
= → F p = sQ(s ) − τ c = 0 → up = λ (138a, b, c)
τ̄c Q(s∗ ) ∂s∗p
where u∗ is chosen as a unit vector.
√ Its relation with the previous identification of the strain
modes in vector form is u∗ = u/ uu. It defines the direction of the strain rate mode u.
The third relations (137c), (138c) express a weak form of the normality rule in which
the vector norms are not defined since both s∗ and u∗ are unit vectors. The scalar λ is
computed to keep the norm of the strain rate mode tensor equal to unity.
In practice, Winters 1996 chooses (138) in order to use all the routines already developed
to compute the formulation (133) in strain rate space. His function Q(s∗ ) is a 6th order
series expansion, the coefficients of which are fitted to the inverse of the stress radii:
τ̄c
Q(s∗ ) = Gp1 p2 ...pN s∗p1 s∗p2 . . . s∗pN = (139)
s
This set of points τ̄c /s is provided by the first way presented above. So, the yield locus
will exhibit no fishtail. However as illustrated by Figure 61, it must be observed that this
yield locus in stress space is not completely convex. The locus on the left in strain rate
space results from the approximation (127) in strain rate space while the one in stress space
on the right is computed by (138b) and the first way to get the set of stress points.
Figure 62 summarizes all the required steps to reach the yield locus shape expressed in
stress or strain rate spaces. Figure 63 explains that if the FEM approach is coupled with,
for instance, a FC Taylor model that computes texture updating at each integration point
with the FEM computed velocity gradient, then the yield locus shape can be updated and
texture evolution can be taken into account during the FEM computation.
Finally, is it better to use a formulation in strain rate or in stress space? The answer
depends on your primary interest: accuracy, low CPU time, necessity of texture updating.
Further investigations have computed Lankford’s coefficient: from calculations in stress
space, from calculations in strain rate space and from Taylor-Bishop Hill’s model. Both
works from Winters 1996 and Van Bael et al. 1996 reach the same conclusion that the
yield locus in stress space is less accurate than the formulation in strain rate space. Table 4
summarizes the advantages and drawbacks of each approach.
One point is completely missing in this description: the numerical way to identify
the position of material axes. The solution proposed by Munhoven et al. 1995a can be
applied: each step is characterized by a constant local velocity gradient, which determines
the evolution of material axes. Another explanation of the same mathematical approach
is proposed by Hoferlin et al. 1999b, using Ponthot’s 1995 constant co-rotational strain
rate tensor. Another possibility is to determine the material axes position by the Mandel
spin (Peeters et al. 2001), which is the average of the spins of all the crystal lattices of the
polycrystal. An efficient way to compute Mandel spins for all possible strain modes has
been proposed by Van Houtte 2001.
6.3.5 Discussion about micro-macro approaches with yield locus
Sections 6.3.2 to 6.3.4 propose models able to describe the behaviour of real materials. They
can be implemented in FEM and each one presents some advantages and drawbacks. Dar-
rieulat’s micro-macro approach seems more accurate than Lequeu’s one. However, for real
Modelling the Plastic Anisotropy of Metals 79

Initial texture measurements

Identification of C µν coefficients of the Orientation Distribution


l

Computation of average Taylor’s factor M (U ε P


) for a set of strain rates modes

Minimization of Taylor’s factor to


find the stress radii s (s*) for a set of
stress directions s *
Fitting of a 6th order series on the
average Taylor’s factor

Fitting of a 6th order series on the


inverse of the stress radii Coefficients describing the shape of
the yield locus in strain rate space

Coefficients describing the shape of the


yield locus in stress space

Figure 62. Flowchart to reach the shape description of the yield locus in stress or
strain rate space

Stress space τ̄sc = Q(s1 ∗) Strain-rate space Ẇ p = τ̄c ε̇peqQ(u)


Speed of FEM Faster slower 4-dim. Minimization
Texture Slower Faster
evolution texture → M̄ → texture → M̄ → Q
4-dim. minimization → Q
Sharp textures one stress direction one stress direction
→ one stress point → fishtails (extra cost)
Accuracy Lower Higher
Table 4. Comparison between stress and strain rate formulations for analytical yield loci
implemented in FEM codes (adapted from Hoferlin et al. 1999a)

materials where important number of texture components are present, the FEM computa-
tions must be quite lengthy as their approach finally considers the yield locus associated to
each texture component. Uniform stress approaches are not straightforward to implement
in usual FEM integration scheme. The proposals by Arminjon, Van Houtte and co-authors
seem quite interesting. The identification of yield locus coefficients from texture coefficients
can be optimized outside the FEM code. Taylor’s assumption leads to a direct macroscopic
stress-strain formulation. Arminjon’s semi-analytical method to take texture updating into
80 A.M. Habraken

ODF of initial texture

Discretization technique

Discrete set of orientations


with associated volume fraction
Slip
systems
CRSS

updating Micro - Macro
transition

Texture prediction
For
each
Interpolation
Point
Yield locus
Macroscopie shape in stress
velocity gradient space

FEM
code

Geometry
Forces
Hardening

data

Figure 63. Complete flowchart for coupling yield locus and crystallographic texture

account is probably difficult to use for arbitrary velocity gradient. The approach proposed
by Hoferlin (yield locus in strain rate space) seems more adapted in a FEM context to
update the material state with the texture prediction during FEM simulations (Li et al.
2003). However, the CPU time is still important, as working in the strain rate space is
slower than in stress space.

7 A FEM MICRO-MACRO MODEL WITH LOCAL DESCRIPTION OF MA-


CROSCOPIC YIELD LOCUS
7.1 Model Description
This section summarizes the method developed by Habraken and co-workers. It aims to
take into account texture updating during the FEM computation in an efficient way. This
method relies on a set of representative crystal orientations defined at each integration
point of the FE mesh but most of the time no microscopic computations are performed,
an approximated local zone of the yield locus is used. So this approach states between the
models described in Sections 5.2 and 6.3.
This local description of the yield locus is assumed sufficient as the stress state generally
Modelling the Plastic Anisotropy of Metals 81

remains in a local zone of the yield locus for few increments of a FEM computation. In
the deviatoric stress space, five or six points in the interesting part of the yield locus are
computed by the full constraint Taylor’s model applied on a set of representative crystals.
This is done at each integration point of the FE mesh. No global yield locus is defined.
Just a direct interpolation between these points is achieved. These microscopic points must
be computed in two cases:
• when, at the current integration point, the plastic strains significantly deform the
material and induce changes in the crystallographic orientations (texture updating);
• when, during a loading increment, the new stress state at the current integration
point leaves the described part of the yield locus so that an updated local zone of
the yield locus is required. In most cases, the definition of this updated local zone
only requires the calculation of one new point in the vicinity of the points already
available.
Compared to micro-macro models without macroscopic yield locus, which systematically
call the microscopic model, the present approach proposes an important computation time
decrease as five or six calls to the microscopic model are only required from time to time.
Compared to micro-macro models with macroscopic yield locus, these five or six calls replace
the thousands calls required to update the full yield locus description. Hoferlin’s thesis
2001 and Van Houtte 2001 define the exact amount of runs of the Taylor’s program in
their approach. As verified by Imbault & Arminjon 1993, important effects on yield locus
anisotropy are induced by texture updating, so the interest of the proposed method is clear.
All the numerical details of this approach can be found in Duchêne 2000 or in Habraken
& Duchêne (in press). In fact two local descriptions have been tested. The most simple
one “Hyperplane approach” was a set of five pieces of hyperplane, each one containing one
central point and four from the five points describing the local zone of the yield locus. This
local description of the yield locus suffers of strong discontinuities as shown on Figure 64 and
introduces convergence problems in FEM simulations. The second approach “Interpolation
approach” extends the interpolation concept of isoparametric triangle finite element to the
five-dimensions deviatoric stress space and provides a more continuous and stable approach.
For a SPXI steel sheet, Figure 64 shows π-sections of the yield locus computed by five
different methods:

• 6th order = global yield locus in stress space described by a 6th order series from Van
Houtte (see Section 6.3.4).
• Hyp. 1◦ or 20◦ = Hyperplane approach with an angle of 1◦ or 20◦ between the strain
rate directions used as domain limit vectors.
• S-s I 1◦ or 20◦ = Stress-strain rate Interpolation approach with an angle of 1◦ or 20◦
between the strain rate directions used as domain limit vectors.

For the local approaches with large domains (Hyp. 20◦ and S-s I 20◦ ), the associated
yield locus normals, which represent the deviatoric plastic strain rates, are also plotted.
Normals from the hyperplane approach are clearly discontinuous. When small local domains
are used (1◦ ), their associated yield loci are superposed and very close to the yield locus
computed by the 6th order series. The latter locus is validated and close to Taylor-Bishop-
Hill locus. For larger domain size (20◦ ), the results of the hyperplane and interpolation
methods strongly differ: while the interpolation approach is continuous and close to the
yield loci computed by the 6th order series, the hyperplane result is discontinuous and
diverges from the 6th order series result.
82 A.M. Habraken

S1 S2

Norm. (s-s I 20°)


Norm. (hyp. 20°)
S-s I 1°
Hyp. 1°
S-s I 20°
Hyp. 20°
S3 6th order

Figure 64. π-sections of the yield locus and normals to the yield loci computed by
stress-strain rate interpolation and hyperplane method with a size of
20◦ (from Duchêne et al. 1999c)

7.2 Deep-Drawing Simulations


In order to show up the influence of the texture evolution during a forming process, the
deep drawing of a cylindrical cup with a hemispherical punch has been simulated. It
is one of the benchmarks proposed by NUMISHEET 1999. The geometry is defined in
Figure 65. The drawing ratio is 2.0; the blankholder force is 80 kN; the simulation is
achieved up to a drawing depth of 85 mm. The proposed material “DDQ” (mild steel) has
been chosen. A Coulomb law is used to model the friction with a coefficient of 0.15. All the
material parameters such has hardening behaviour and Lankford coefficients were measured
by simple tensile tests in different directions and available for benchmark’s participants (see
Numisheet 1999). As we focus on the texture, the shape of the yield locus is deduced from
its Orientation Distribution Function (ODF), which has been measured by X-ray diffraction
by Nakamachi, who provided us the C coefficients describing the ODF of this mild steel.
LAGAMINE FEM code developed by M & S department has been used with two layers of
solid 3D-mixed type finite elements BLZ3D (Zhu & Cescotto 1996) and contact elements
CFI3D (Habraken & Cescotto 1994).The simulation results computed with the stress-strain
interpolation method with and without texture updating are compared to a classical Hill
1948 constitutive law and to 2 experimental results (4 experimental results were available
but for Figure readability we kept only 2 defining the experimental range). Figure 66 defines
draw-in, curvilinear abscissa and sections where results are compared.
Modelling the Plastic Anisotropy of Metals 83

Figure 65. Geometry of the chosen benchmark of Numisheet 1999

Rolling Direction
RD

A
A: Draw-in at 90°

Cup

Draw-in: Blank
Distance between 45°
blank boundary line and B
part boundary line in a section 1)
C

B: Draw-in at 45° 45°

C: Draw-in at 0°

Figure 66. Definitions of sections A, B, C, curvilinear abscissa s and draw-in (from


NUMISHEET 1999)

Figure 67 shows the maximum principal strain distribution on a section along the trans-
verse direction for Hill and texture based laws and for experimental results. It should be
noticed that the strain is relatively small at the top of the cup (near the pole; for s smaller
than 50 mm) while large displacements take place. Indeed, in that part of the cup, the steel
sheet is applied against the punch and follows the punch travel without large deformation.
Then, at the flange of the cup, the maximum principal strain suddenly increases and is
maximum for s being equal to 90 mm. This maximum is located at the vertical part of the
flange where the cup is free of contact with the punch and the matrix. After the maximum,
the principal strain ε1 decreases as the contact with the matrix reduces the tension of the
sheet. The resulting maximum principal strain obtained with Hill law is a little bit too high
while texture based laws are in agreement with experimental results.
84 A.M. Habraken

ε1 Strain Distribution ε1 along Transverse Direction


0.8

0.7 Constant Texture


Evolving Texture
0.6 Hill
Experimental
0.5 Experimental

0.4

0.3

0.2

0.1

0
0 20 40 60 80 100 120 140
s: distance from pole (mm)

Figure 67. Maximum principal strain distribution ε1 along section A (transverse


section)

ε2 Strain Distribution ε2 along Transverse Direction


0.15

0.05

-0.05

-0.15

-0.25 Constant Texture


Evolving Texture
-0.35 Hill
Experimental
-0.45
Experimental
-0.55
0 20 40 60 80 100 120 140
s: distance from pole (mm)

Figure 68. Second principal strain distribution ε2 along section A (transverse section)

Figure 68 shows the second (smallest) principal strain along the same section. Here
again, the same evolution can be noticed: low constant value near the pole, followed by
a minimum (or a maximum in absolute value) in the flange of the cup and then lower
deformations under the blankholder. Looking ε1 and ε2 together, we can see that four
typical regions can be identified:

• Near the pole (s < 40 mm), an equi-biaxial tension state or stretched zone is present
(ε1 ≈ ε2 ),
• In the flange (40 < s < 90), increasing tensile and compression strains can be noticed;
a restrained zone is determined (ε1 > 0, ε2 < 0 and ε1  > ε2 ),
Modelling the Plastic Anisotropy of Metals 85

ε3 Strain Distribution ε3 along Transverse Direction


0.1
Constant Texture
0.05 Evolving Texture
0 Hill
Experimental
-0.05 Experimental
-0.1

-0.15

-0.2

-0.25

-0.3
0 20 40 60 80 100 120 140
s: distance from pole (mm)

Figure 69. Thickness strain distribution ε3 along section A (transverse section)

• Near the matrix curvature (90 < s < 110), the restrained zone presents a decrease in
the absolute value of the strains (ε1 > 0, ε2 < 0 and ε1  > ε2 ),

• Under the blankholder (110 < s < 135), the compression strain ε2 becomes larger
than the tensile strain ε1 (ε1 > 0, ε2 < 0 and ε1  < ε2 ). This strain state would
give rise to instability and wrinkling without the action of the blankholder.

Then Figure 69 shows the third principal strain distribution along section A (transverse
direction). It corresponds to the thickness strains. Near the pole, in the equi-biaxial tension
strain zone, ε3 is negative: corresponding to a thickness reduction. Then progressively, ε3
grows and becomes positive (the thickness is increasing during the process) in the fourth
region described here above.
The punch force as a function of the punch travel is presented on Figure 70. These
curves are not linked to the anisotropy of the steel sheet but to the global stiffness of the
material and then to the hardening behaviour. It can be noticed that the curve with the
evolving texture is very close to the experimental curve; the constant texture is a little bit
too low at the end of the process and Hill is too high.
Finally, Figure 71 is directly linked to the anisotropy of the steel sheet. For finite
element simulations, this anisotropy is introduced in the constitutive law (either the Hill
coefficients or the texture data). The Flange Draw-In is defined as the length of the flange
that is swallowed under the blankholder (see Figure 66). It is the opposite of the earing
profile. The experimental results exhibit a maximum draw-in along section B (45◦ ) and
lower values along sections A and C. The Hill based constitutive law shows a good draw-in
profile (maximum at section B) but the amplitude is too high (variation between the three
sections). The draw-in profiles obtained by the texture based laws are very similar to the
experimental one. The anisotropy is then well represented by this constitutive law but a
shift is observed along both directions: the draw-in is too low with these texture based
laws. The behaviour of the material is less ductile numerically than experimentally. The
amplitude of the draw-in profile is a little bit higher when the texture is updated during
the process.
86 A.M. Habraken

Punch force as a function of punch travel


160

140

120
Punch force (kN)

100

80
Constant Texture
60
Evolving Texture
40 Hill
Experimental
20
Experimental
0
0 10 20 30 40 50 60 70 80
Punch travel (mm)

Figure 70. Punch force as function of punch displacement

Flange Draw-In at sections A, B and C


40

35

30
Draw-In (mm)

25

20
Constant Texture
15
Evolving Texture
10 Hill
Experimental
5
Experimental
0
A : 90° B : 45° C : 0°
Section : angle from Rolling Direction

Figure 71. Flange draw-in at sections A,B,C

8 CONCLUSION

This overview work provides an idea of the main models implemented today in FEM codes
to represent macroscopic plastic anisotropic material behaviour. Concerning micro-macro
approach with yield loci, it is focused on the initial yield locus shape and on geometric or
textural hardening. The latter describes the yield locus updating in shape, position and size
due to the effect of crystallites rotation. Clearly no review of the material hardening, related
to dislocation density is provided. As dislocations constitute obstacles to the production
and motion of further dislocations, this phenomenon also induces shape, position and size of
the yield locus (Bouvier et al. 2002, Tedodosiu, Peeters 2002, Lopes et al. 2003). Complete
state of the art review of material hardening is another story, let us just underline that
Modelling the Plastic Anisotropy of Metals 87

proposals based on microscopic events like the ones from Bergström-van Liempt-Vegter
(Vegter’s et al. 1999a), Follansbee & Kocks 1988, Teodosiu & Hu 1998 are interesting
alternatives to classical isotropic and kinematic hardening models. Another key point in
FEM simulations not addressed here is the choice of orthotropic material axes evolution in
the FEM description. Peeters et al. 2001 summarizes clearly various choices.
Some care has been brought to provide not only the theoretical models, but their iden-
tification methods as well. Whenever possible, links between models are presented, such as,
for instance, Aifantis’s proposal that gives physical basis to von Mises and Karafillis’ laws.
In the future, hardware development and parallel computation will reduce problems
of CPU time. However, each time that this happens, engineers increase the sizes of the
problems that they want to solve by finite elements ... So, simple phenomenological laws
described in Section 2 that allow escaping to microscopic computations in the macroscopic
FEM simulations, retain their interest. This also forces researchers to identify the important
features necessary to capture material behaviour. For instance, it is clear that texture
evolution effects on yield locus are not necessary in all deep-drawing simulations.
The actual question is: what is really useful to take into account? The answer is not the
same according to the goal of the simulations: shape prediction after spring back, texture
prediction, residual stress field, wrinkling and necking prediction. It is clear that criteria
exist to predict necking and wrinkling, but they rely on accurate stress and strain field
computations. The spring back prediction is quite hard if your model neglects elasticity.
The final shape and size of the yield locus after forming processes are important if your goal
is to apply accurate fatigue models to predict the life of the pieces. The model descriptions
provided in this review should help to choose the adapted model to fit one’s requirements.
One direction not investigated in this overview is the formulation of the Finite Element
itself. Going from a simple displacement formulation to a mixed or hybrid formulation can
already provide a better convergence and a smoother stress answer, even if a low number of
crystals is used per integration point (Beaudoin et al. 1995). Another possibility is to apply
a simple macroscopic analysis coupled with a micro-macro analysis only where some event,
such as strain localization, appears and requires a finer scale. For instance, Garikipati &
Hughes (2000) propose such a so-called variational multiscale approach.

ACKNOWLEDGEMENT

A. M. Habraken is mandated by the National Fund for Scientific Research (Belgium). She
also thanks the Belgian Federal Science Policy Office (Contract P5/08) for its financial
support.

REFERENCES
1 A. Acharya and A. J. Beaudoin (2000). Grain-size effect in viscoplastic polycrystals at moderate
strains. J. Mech. Phys. Solids, 48(10), 2213–2229.
2 E. Aernoudt, J. Gil-Sevillano and P. Van Houtte (1987). Constitutive Relations and Their
Physical Basis, S.I. Andersen et al. (Eds.), RisøNational Laboratory, Roskilde, Denmark, 1–38.
3 E. C. Aifantis (1987). The physics of Plastic Deformations. Int. J. Plasticity, 3, 211–247.
4 R. J. Asaro (1983). Micromechanics of crystals and polycrystals, Advances in Applied Mechan-
ics, 23, 1–115.
5 R. J. Asaro and A. Needleman (1985). Texture development and strain hardening in rate
dependent polycrystals. Acta Metallurgica, 33, 923–953.
6 L. Anand and M. Kothari (1996). A Computational Procedure for Rate-Independent crystal
plasticity. J. Mech. Phys. Solids, 44(4), 525–558.
88 A.M. Habraken

7 L. Anand, S. Balasubramanian and M. Kothari (1997). Constitutive Modeling of Polycrystalline


Metals at Large Strains: Application to Deformation Processing, Large plastic deformation of
crystalline aggregates. International Centre for Mechanical Sciences, Courses and Lectures n◦
376, Springer Verlag, 109–172.
8 A. Andersson, C. A. Ohlsson, K. Mattiasson and B. Persson (1999). Implementation and Eval-
uation of the Karafillis-Boyce Material Model for Anisotropic Metal Sheets. Numisheet’99,
13-17 September 1999, 1, Besançon, France, 13-17 September 1999, JC Gélin, P. Picart (Eds.),
Université de Franche-Comté, 115–121.
9 M. Arminjon (1988). Lois de comportement homogénéisées pour la plasticité des polycristaux.
Mém. d’habilitation, Univ. Paris-Nord, Villetaneuse.
10 M. Arminjon and B. Bacroix (1991). On plastic potentials for anisotropic metals and their
derivation from the texture function. Acta Mechanica, 88, 219–243.
11 M. Arminjon, B. Bacroix, D. Imbault and J. K. Raphanel (1994). A fourth-order plastic poten-
tial for anisotropic metals and its analytical calculation from texture function. Acta Mechanica,
107, (33).
12 I. Aukrust, S. Tjotta, H. E. Vatne and P. Van Houtte (1997). Coupled FEM and texture
modelling of plane strain extrusion of an Aluminium alloy. Int. J. of Plasticity, 13(1,2).
13 D. Banabic (2000). In: Formability of Metallic Materials. D. Banabic (Ed.), Springer Verlag,
Berlin, p. 119–172.
14 D. Banabic, D. S. Comsa and T. Balan (2000). Proc. 7th Conf. TPR 2000, Cluj Napoca,
215–224.
15 A. Barata da Rocha (1985). Mise en forme des tôles minces, instabilité plastique, anisotropie
et endommagement. These de Doctorat, Institut National Polytechnique de Grenoble.
16 F. Barbe, G. Cailletaud, S. Forest and S. Quilici (1999). Large Scale Parallel Computation
applied to Polycrystalline Aggregates, Book of Abstracts 5th U.S. National Congress on Com-
putational Mechanics, edited by A. Carosio, P. Smolarkiewicz, K. Willam and J. Yang; in
University of Colorado Printing Services
17 F. Barlat (1987). Crystallographic texture, anisotropic yield surfaces and forming limits of
sheet metal. Materials Science and Engineering, 91(55).
18 F. Barlat and J. Lian (1989). Plastic behaviour and stretchability of sheet metals. Part 1: A
yield function for orthotropic sheets under plane stress conditions. Int. J. of Plasticity, 5, 51.
19 F. Barlat, D. J. Lege and J. C. Brem (1991). A six-component yield function for anisotropic
materials. Int. J. of Plasticity, 7, 693.
20 F. Barlat, R. C. Becker, J. C. Brem, D. J. Lege, D. J. Murtha, Y. Hayashida, Y. Maeda,
M. Yanagawa, K. Chung, K. Matsui and S. Hattori (1997). Yielding description for solution
strengthened aluminum alloys, Int. J. of Plasticity, 13, 385–401.
21 F. Barlat, D. Banabic and O. Cazacu (2002). Anisotropy in sheet metals Numisheet 2002,
Numerical Simulation of 3D Sheet Forming Processes. Dong-Yol Yang, Soo Ik Oh, Hoon Huh,
Yong Hwan Kim (Eds.), Design innovation Through Virtual Manufacturing, 515–524
22 A. J. Beaudoin, P. R. Dawson, K. K. Mathur, U. F. Kocks and D. A. Korzekwa (1994).
Application of polycrystal plasticity to sheet forming. Comp. Methods Appl. Mech. Eng., 117,
49–70.
23 A. J. Beaudoin, P. R. Dawson, K. K. Mathur, U. F. Kocks, (1995). A hybrid finite element
formulation for polycrystal plasticity with consideration of macrostructural and microstructural
link. Int. J. of Plasticity, 11(5), 501–521.
24 Becker (1990). An analysis of shear localization during bending of a polycrystalline sheet.
Microstructural Evolution in Metal Processing, 46.
Modelling the Plastic Anisotropy of Metals 89

25 H. Berg, P. Hora and J. Reissner (1998). Simulation of sheet metal forming processes using
different anisotropic constitutive models, Simulation of materials processing: theory, methods
and applications, Huetink and Baaijens (Eds.), Balkema.
26 Y. Bergström (1969). A dislocation model for the stress strain behaviour of polycrystalline α-Fe
with special emphasis on the variation of the densities of mobile and immobile dislocations.
Mat. Sci. Eng., 5, 179–192.
27 M. Berveiller and A. Zaoui (1979). An extension of the self-consistent scheme to plastically-
flowing polycrystals. J. Mech. Phys. Solids, 26, 325–344.
28 A. Bertram, T. Böhlke and M. Kraska (1997). Numerical simulation of texture development
of polycrystals undergoing large plastic deformations. Computational Plasticity. Fundamentals
and Applications, D. R. J. Owen, E. Oñate and E. Hinton (Eds.).
29 N. Boudeau and J. C. Gelin (1996). Post-processing of finite element results and prediction
of the localized necking in sheet metal forming. J. of Materials Processing Technology, 60,
325–330.
30 S. Bouvier, C. Teodosiu, H. Haddadi and V. Tabacaru (2002). Anisotropic Work-Hardening
Behaviour of Structural Steels and Aluminium Alloys at Large Strains. Proceedings of the 6th
European Mechanics of Materials Conference, Non Linear Mechanics of Anisotropic Materials,
Liege, Cescotto (Ed.).
31 P. W. Bridgman (1923). The compressibility of thirty metal as a function of pressure and
temperature. Proc. Am. Acad. Arts Sci, 58, 165.
32 P. W. Bridgman (1952). Studies in large plastic flow and fracture. Metallurgy and Metallurgical
Engineering Series, New-York, McGraw-Hill.
33 H. J. Bunge (1982). Texture Analysis in Materials Science, Butterworths Publishers, London.
34 G. R. Canova, U. F. Kocks and C. N. Tomé (1985). The yield surface of textured polycrystals.
Mech. Phys. Solids, 33, 4, 371–397.
35 G. R. Canova and R. Lebensohn (1995). Micro-macro modelling, Computer Simulation in
Materials Science, NATO ASI, Ile d’Oleron, France, June 6–16.
36 O. Cazacu and F. Barlat (2001). Generalization of Drucker’s yield criterion to orthotropy
Mathematics and Mechanics of Solids, 6, 613–630.
37 F. Cayssials (1998). A new method for predicting FLC, IDDRG. Conference Geneval, Brussel
6/98.
38 F. Cayssials (1999). The new version of the Sollac model, Working Group of the IDDRG 99,
Birmingham.
39 Y. Chastel, R. Loge, M. Perrin, V. Lamy and S. Zaefferer (1998). Microscopic and macroscopic
length scales in hot extrusion of Zircaloy 4. First ESAFORM Conference on Material Forming,
Sophia-Antipolis, France.
40 A. Clément (1982). Prediction of Deformation Texture Using a Physical Principle of Conser-
vation. Mater. Sci. Eng., 55, 203–210.
41 M. Darrieulat and D. Piot (1996). A method of generating analytical yield surfaces of crystalline
materials. Int. J. Plasticity, 12(10), 1221–1240.
42 M. Darrieulat and F. Montheillet (2003). A texture based continuum approach for predicting
the plastic behaviour of rolled sheet. Int. Jour. of Plasticity, 19(4), 517–546.
43 P. R. Dawson, A. J. Beaudoin and K. K. Mathur (1992). Simulating deformation-induced
texture in metal forming. Num. Meth. in Ind. Form. Proc..
44 P. R. Dawson and A. Kumar (1997). Deformation Process Simulations Using Polycrystal Plas-
ticity. Large plastic deformation of crystalline aggregates, International Centre for Mechanical
Sciences, Courses and Lectures n◦ 376, Springer Verlag, 247.
90 A.M. Habraken

45 D. C. Drucker (1949). J. Appl Mech, 16, 349–357.


46 D. C. Drucker (1951). A more fundamental approach to plastic stress-strain relations. Proc.
First US Nat. Congr. Applied Mechanics, ASME, New-York, 487–491.
47 L. Duchêne (2000). Implementation of a Yield Locus Interpolation Method in the Finite Ele-
ment Code Lagamine, DEA Graduation Work, Université de Liege M & S Department, Bel-
gium.
48 J. D. Eshelby (1957). The determination of the elastic field of an ellipsoidal inclusion and
related problems. Proc. Roy. Soc. London, A241, 376–396.
49 Ewing, J.A. and W. Rosenhain (1900). The crystalline structure of metals; Phil. Trans. R.
Soc., London, 193, 353.
50 F. Feyel and J. L. Chaboche (2000). Mutiscale non linear FE analysis of composite structures:
damage and fiber size effets. Euromech 417, 2-4 October 2000, University of Technology of
Troyes, France.
51 P. S. Follansbee and U. F. Kocks (1988). A constitutive description of the deformation of copper
based on the use of the mechanical threshold stress as an internal state variable. Acta Metall.,
36(1), 81–93.
52 P. Franciosi (1988). On flow and a work hardening expression correlation in metallic single
crystal plasticity. Revue Phys. Appl., 23, 383–394.
53 J. Frenckel (1926). Zur Theorie der Plastizitätsgrenze und der Gestigheit kristallinischer
Körper. Z. Phys., 37, 572–609.
54 K. Garikipat and J. R. T. Hughes (2000). A Variational Multiscale Approach to Strain Local-
ization, Formulation for Multidimensional Problems. Computer Methods in Applied Mechanics
and Engineering, 188, (2000), 39–60
55 M. G. D. Geers, V. Kouznetsova and W. A. M. Brekelmans (2000). Constitutive approaches
for the multi-level analysis of the mechanics of microstructures. 5th National Congress on
Theoretical and Applied Mechanics, Louvain-La-Neuve, May 23–24, 2000.
56 P. Gilormini, Y. Liy and P. Ponte Castaneda (2002). Application of the variational self-
consistent model to the deformation textures of titanium. Int. Journal of Forming Processes,
5(2,3,4), 327–336.
57 A. L. Gurson (1977). Continuum theory of ductile rupture by void nucleation and growth. J.
Engng. Materials Technology, 99, 2–15.
58 A. M. Habraken and S. Cescotto (1994). Contact between deformable solids, the fully coupled
approach. Mathematical and Computer Modelling, 28(4-8), 153–169.
59 A. M. Habraken (2001). Contributions to Constitutive laws of metals: micro-macro and damage
models, These d’Agrégé de l’Enseignement Supérieur, Université de Liege, Département M &
S.
60 A. M. Habraken and L. Duchêne (in press). Micro-Macro Simulations of Polycrystalline Metals.
Int. Jour. of Plasticity, (electronic version available since March 2004).
61 A. Hacquin, P. Montmitonnet and J. P. Guillerault (1995). Coupling of roll and strip deforma-
tion in three-dimensional simulation of hot rolling. Simulation of Materials Processing: Theory,
Methods and Applications, 921.
62 I. Hage Chehade (1990). Simulation de l’emboutissage des tôles anisotropes par éléments finis
avec prédiction des risques de striction, These de doctorat, Institut National des Sciences
Appliquées de Lyon.
63 K. Hayakawa and S. Murakami (1998). Space of damage conjugate force and damage potential
of elastic-plastic-damage materials, Damage Mechanics in Engineering Materials, Voyiadjis,
G.Z., Ju, J.W., Chaboche J.L. Eds, 27–44.
Modelling the Plastic Anisotropy of Metals 91

64 S. S. Hecker (1976). Experimental studies of yield phenomena in biaxially loaded metals, J. A.


Strick-Lin, K. J. Saczalski (Eds.), Constitutive equations in viscoplasticity: computational and
engineering aspects, ASME, New-York, 1–33.
65 R. Hill (1948). A theory of the yielding and plastic flow of anisotropic materials. Proc. Royal
Soc. London, A193, 281–297.
66 R. Hill (1965). Continuum micro-mechanics of elastoplastic polycrystals. J. Mech. Phys. Solids,
13, 89–101.
67 R. Hill (1979). theoretical plasticity of textured aggregates. Math. Proc. Cambridge Philosoph-
ical Soc., 85, 179–191.
68 R. Hill (1987). Constitutive dual potentials in classical plasticity. J. Mech. Phys. Solids, 35,
23–33.
69 R. Hill (1990). Constitutive Modelling of Orthotropic Plasticity in Sheet Metals. J. Mech. Phys.
Solids, 38(3), 405–417.
70 R. Hill (1993). A user-friendly theory of orthotropic plasticity in sheet metals. Int. J. Mech.
Sci., 35(1), 19–25.
71 J. R. Hirsch (1990). Correlation of deformation texture and microstructure. Materials Science
and Technology, 6, 1048.
72 S. Hiwatashi, A. Van Bael, P. Van Houtte and C. Teodosiu (1997). Modelling of plastic
anisotropy based on texture and dislocation structure. Computational materials science, 9,
274–284.
73 E. Hoferlin, A. Van Bael, S. Hiwatashi and P. Van Houtte (1998). Influence of texture and
microstructure on the prediction of forming limit diagram. 19th RISO Symposium on Materials
Science, 7-11 Sept. 1998.
74 E. Hoferlin, A. Van Bael and P. Van Houtte (1999a). Comparison between stress-based and
strain-rate based elasto-plastic finite element models for anisotropic metals. Plasticity’99, Con-
stitutive and damage modelling of inelastic deformation and phase transformation, A. S. Khan
(Ed.), Neat Press Fulton, Maryland.
75 E. Hoferlin, A. Van Bael and P. Van Houtte, C. Teodosiu (1999b). An accurate model of texture
and strain-path induced anisotropy. Numisheet’99, Numerical Simulation of 3D Sheet Forming
Processes, J. C. Gelin, P. Picart (Eds.), Université de Franche Comté, Besançon, France.
76 E. Hoferlin (2001). Incorporation of an accurate model of texture and strain-path induced
anisotropy in simulations of sheet metal forming. Ph. D. thesis Katholieke Universiteit Leuven.
77 P. Hora, L. Tong and J. Reissner (1996). A prediction method for ductile sheet metal failure in
f.e. simulation. Proceedings of the 3rd Int. Conf. Numisheet’ 96 Numerical Simulation of 3-D
Sheet Metal Forming Processes - Verification of Simulations with Experiments, Lee, Kinzel,
Wagoner (Eds.), Ohio State University.
78 W. F. Hosford (1972). A generalized isotropic yield criterion. J. Appl. Mech. Trans. ASME,
39, 607–609.
79 D. Imbault and M. Arminjon (1993). Theoretical and numerical study of the initial and induced
plastic anisotropy of steel sheets, using a texture-based methodology, Final report of contract
Univ. J. Fourier n◦ 17191401, Laboratoire 3S, Université Joseph Fourier, France.
80 J. Jensen and D. N. Hansen (1987). Relations Between Texture and Flow Stress in Commercially
Pure Aluminium, Constitutive Relations and their Physical Basis. 8th Riso Int. Symp. on
Metallurgy and Mat. Sci., S. I. Andersen et al. (Eds.), Riso Nat. Lab., Roskilde, 353–360.
81 S. R. Kalidindi, C. A. Bronkhorst and L. Anand (1992). Crystallographic texture evolution
during bulk deformation processing of FCC metals. J. Mech. Phys. Solids, 40, 537–579.
82 S. R. Kalidindi and L. Anand (1994). Macroscopic shape change and evolution of crystallo-
graphic texture in pre-textured FCC metals. J. Mech. Phys. Solids, 42(3), 459–490.
92 A.M. Habraken

83 J. S. Kallend, U. F. Kocks, A. D. Rollett and H. R. Wenk (1991). popLA - an integrated


software system for texture analysis. Text. microstruct., 14-18, 1203–1208.
84 A. P. Karafillis and M. C. Boyce (1993). A general anisotropic yield criterion using bounds and
a transformation weighting tensor. J. Mech. Phys. Solids, 41(12), 1859–1886.
85 A. S. Khan and S.Huang (1995). Continuum theory of plasticity, Wiley & Sons.
86 A. S. Khan and P. Cheng (1996). An anisotropic elastic-plastic constitutive model for single
and polycrystalline metals. I - theoretical developments. Int. J. Plasticity, 12(2), 147–162.
87 A. S. Khan and P. Cheng (1998). An anisotropic elastic-plastic constitutive model for single
and polycrystalline metals. II - experiments and predictions concerning thin-walled tubular
IFHC copper. Int. J. Plasticity, 1(3), 209.
88 S. Kobayashi, R. M. Caddell and W. F. Hosford (1985). Examination of Hill’s latest yield
criterion using experimental data for various anisotropic sheet metals. Int. J. of Mech. Sci.,
27, 509.
89 E. Kröner (1961). Zur plastischen Verformung des Vielkristalls. Acta Metall., 9, 155–161.
90 A. Kumar and P. R. Dawson (1995a). The simulation of texture evolution during bulk deforma-
tion processes using finite elements over orientation space. Simulation of Materials Processing:
Theory, Methods and Applications, Shen & Dawson, Balkema.
91 A. Kumar and P. R. Dawson (1995b). Polycrystal plasticity modeling of bulk forming with
finite elements over orientation space. Comp. Mech., 17, 10–25.
92 A. Kumar and P. R. Dawson (1996). The simulation of texture evolution with finite elements
over orientation space. I. Development, II. Application to planar crystals. Comp. Methods Appl.
Mech. Eng., 130, 227–261.
93 T. Kuwabara and A. Van Bael (1999). Measurement and Analysis of Yield Locus of Sheet
al.uminium Alloy 6XXX. Numisheet’99, 13-17 September 1999, 1, Besançon, France, 13-17
September 1999, J.C. Gélin and P. Picart (Eds.), Université de Franche-Comté, 85–90.
94 R. A. Lebensohn and C. N. Tome (1993). A self-consistant anisotropic approach for the simula-
tion of plastic deformation and texture development of polycrystals : application to Zirconium
alloys. Acta Metall. Mater., 41, 2611–2624.
95 Ph. Lequeu, P. Gilormini, F. Montheillet, B. Bacroix and J. J. Jonas (1987a). Yield surfaces
for textured polycrystals. I. Crystallographic Approach. Acta Metall., 35(2), 439–451.
96 Ph. Lequeu, P. Gilormini, F. Montheillet, B. Bacroix and J. J. Jonas (1987b). Yield surfaces
for textured polycrystals. II. Analytical Approach. Acta Metall., 35(5), 1159–1174.
97 S. Li, E. Hoferlin, A. Van Bael, P. Van Houtte and C. Teodosiu (2003). Finite element modeling
of plastic anisotropy induced by texture and strain-path change. Int. J. of Plasticity. 19 (5),
647–674.
98 A. B. Lopes et al. (2003). Effect of texture and microstructure on strain hardening anisotropy
for aluminum deformed in uniaxial tension and simple shear. Int. J. Plasticity, 19 (1), 1–22.
99 A. Magnée (1994). Physique du solide, notes de cours de la Faculté des Sciences Appliquées,
Université de Liege.
100 Mahmudi (1995) Yield loci of anisotropic aluminium sheets. Int. J. of Mech. Sci., 37, 919.
101 R. Masson and A. Zaoui (1999). Self-consistent estimates for the rate-dependent elastoplastic
behaviour of polycrystalline materials. J. of Mechanics and Physics of Solids, 47, 1543–1568.
102 K. K. Mathur and P. R. Dawson (1989). On modeling the development crystallographic texture
in bulk forming processes. Int. J. Plasticity, 5, 67–94.
103 A. Mendelson (1968). Plasticity: Theory and Application, MacMillan, New-York, 87.
Modelling the Plastic Anisotropy of Metals 93

104 C. Miehe, J. Schröder and J. Schotte (1999). Computational homogenization analysis in finite
plasticity, simulation of texture development in polycrystalline materials. Computer methods
in applied mechanics and engineering, 171, 387–418.
105 M. P. Miller and D. L. Mc Dowell (1996). Modeling large strain multiaxial effects in FCC
polycrystals. Int. J. of Plasticity, 12(7), 875–902.
106 A. Molinari (1997). Deformation Process Simulations Using Polycrystal Plasticity. Large plastic
deformation of crystalline aggregates, International Centre for Mechanical Sciences, Courses
and Lectures n◦ 376, Springer Verlag, 173–246.
107 G. Monfort, J. P. Adriaens, J. Defourny, P. Jodogne, M. Brunet and J. M. Detraux (1991).
FEM simulation of non-axisymmetric press formed parts using anisotropic constitutive laws
for steel. IDDRG, Pisa.
108 G. Monfort and J. Defourny (1993). A new orthotropic plasticity model for complex sheet
forming. Centre for Theoretical Physics - External activities. (ICTP-OEA), Conference on
interface between physics and mathematics. Hangzhou, China.
109 G. Monfort and J. Defourny (1994). The 3G plasticity model, Metallurgical bases - Mechanical
evaluation - Application to finite element simulation of steel forming, 14 septembre, Centre de
recherches métallurgiques, Liege.
110 F. Montheillet, P. Gilormini and J. J. Jonas (1985). Relation between axial stresses and texture
development during torsion testing: a simplified theory. Acta Metall., 33(4), 705–717.
111 S. Munhoven, A. M. Habraken, J. Winters, R. Schouwenaars and P. Van Houtte (1995a).
Application of an anisotropic yield locus based on texture to a deep drawing simulation, NU-
MIFORM 95, Simulation of Materials Processing: Theory, Methods and Applications, Shen &
Dawson, Balkema, 767–772.
112 S. Munhoven, J. Winters and A. M. Habraken (1995b). Finitie element applications of an
anisotropic yield locus based on crystallographic texture. Computer simulations in materials
science Nano/Meso/Macroscopic Space and teine scales. Nato Advenced study Institute, Ile
d’Oléron France 6-16 juin, Université de Liege, Département MSM.
113 S. Munhoven, A. M. Habraken and J. P. Radu (1997). Anisotropic plasticity based on crystal-
lographic texture. 4eme Congres de Mécanique théorique et appliquée, Leuven.
114 E. Nakamachi and X. H. Dong (1997). Study of Texture Effect on Sheet Failure in a Limit
Dome Height Test by Using Elastic/Crystalline Viscoplastic Finite Element Analysis. J. Appl.
Mech. Trans. ASME(E), 64, 519–524.
115 E. Nakamachi, E. Oñate, P. Bergan, M. H. Boduroglu and C. R. Kaykayoglu (1999a). The study
of crystalline morphology effects on sheet metal forming. IACM Expressions, 7, Spring-Summer
1999.
116 E. Nakamachi, C. L. Xie, K. Hiraiwa and M. Harimoto (1999b). Development of elas-
tic/crystalline viscoplastic finite element analysis code based on the meso-phenomenological
material modeling. Numisheet’99, 1, Besançon, France, J.C. Gélin and P. Picart (Eds.), Uni-
versité de Franche-Comté 79–84.
117 K. Narasimhan and R. H. Wagoner (1991). Finite element modeling simulation of in-plane
forming limit diagrams of sheets containing finite defects. Metallurgical Transaction A, 22A,
1991, 2655.
118 K. W. Neale (1993). Use of Crystal Plasticity in Metal Forming Simulations, Int. J. Mech. Sci.,
35(12), 1053–1063.
119 S. Nikolov and I. Doghri (2000). A micro-macro constitutive model for the small deformation
behavior of polyethylene. Polymer, 41, 1883–1891.
120 J. Ning and E. C. Aifantis (1996). Anisotropic yield and plastic flow of polycrystalline solids.
Int. J. of Plasticity, 12(10), 1221–1240.
94 A.M. Habraken

121 Numisheet (1993). Proceedings of the 2nd Int. Conf. and Workshop on Numerical Simulation
of 3D Sheet Forming Processes, Isehara, Japan, August 31-September 2.
122 Numisheet (1996). Proceedings of the 3nd Int. Conf. and Workshop on Numerical Simulation
of 3D Sheet Forming Processes, G. L. Kinzel and R. H. Wagoner (Eds.), Ohio State University,
Columbus.
123 Numisheet (1999). Proceedings of the 4th Int. Conf. and Workshop on Numerical Simulation of
3D Sheet Forming Processes, Benchmarks, Besançon, France, 13-17 September 1999, 2, J. C.
Gélin and P. Picart (Eds.), Université de Franche-Comté.
124 B. Peeters, E. Hoferlin, P. Van Houtte and E. Aernoudt (2001). Assessment of crystal plasticity
based calculation of the lattice spin of polycrystalline metals for FE implementation. Int. J.
Plasticity, 17, 819–836.
125 B. Peeters (2002). Multiscale Modelling of the induced plastic anisotropy in IF steel during
sheet forming. Ph.D. thesis, Katholieke Universiteit Leuven, MTM.
126 A. Phillips (1986). A review of quasi-static experimental plasticity, and viscoplasticity. Int J;
of Plasticity, 2, 315.
127 A. Phillips and C. W. Lee (1979). Yield surfaces and loading surfaces. Experiments and rec-
ommendations. Int. J. Solids & Structures, 15, 715–729.
128 H. H. Pijlman, J. Brinkman, J. Huetink and H. Vegter (1999). The Vegter Yield Criterion
Based on Multi-Axial Measurements. Numisheet’99, 1, Besançon, France, J. C. Gélin and P.
Picart (Eds.), Université de Franche-Comté, 109–114.
129 P. Ponte Castaneda (1991). The effective mechanical properties of nonlinear isotropic compos-
ites. Journal of the Mechanics and Physics of Solids, 39, 45–71.
130 J. Ph. Ponthot (1995). Traitement unifié de la mécanique des milieux continus solides en grandes
déformations par la méthode des éléments finis. Ph.D. thesis, LTAS, Université de Liege.
131 V. C. Prantil, P. R. Dawson and Y. B. Chastel (1995). Comparison of equilibrium-based plastic-
ity models and a Taylor-like hybrid formulation for deformations of constrained crystal systems.
Modelling Simul. Mater. Sci. Eng., 3, 215–234.
132 G. Sachs (1928). Zur Ableitung einer Fliessbedingung. Z. Verein Deutscher Ing., 72, 734–736.
133 A. Schmitz (1995). Development and experimental validation of a coupled thermal, mechani-
cal and textural model for ferritic hot-rolling of steel. Faculty of Engineering of the Catholic
University of Leuven.
134 S. E. Schoenfeld and R. J. Asaro (1996). Through thickness texture gradients in rolled poly-
crystalline alloys. Pergamon, Int. J. Mech. Sci., 38(6), 661–683.
135 G. Sevillano, P. Van Houtte and E. Aernoudt (1980). Large strain work hardening and textures.
Progress in Materials Science, 25, 111.
136 R. J. M. Smit, W. A. M. Brekelmans and H. E. H. Meijer (1998). Prediction of the mechanical
behavior of nonlinear heterogeneous systems by multi-level finite element modeling. Comp.
Meth. Appl. Mech. Eng., 155, 181–192.
137 G. I. Taylor (1938). Plastic strains in metals. J. Inst. Metals, 62, 307–324.
138 C. Teodosiu, J. L. Raphanel and L. Tabourot (1992). Finite Element Simulation of the Large
Elastoplastic Deformation of Multicrystals, Large Plastic Deformations. Fundamentals and Ap-
plications to Metal Forming. Proc. MECAMAT’91, C. Teodosiu, F. Sidoroff and J. L. Raphanel
(Eds.), Balkema, Rotterdam, 153–168.
139 C. Teodosiu (1997). Dislocation modelling of crystal plasticity, Large plastic deformation of
crystalline aggregates. International Centre for Mechanical Sciences, Courses and Lectures n◦
376, Springer Verlag, 21–80.
Modelling the Plastic Anisotropy of Metals 95

140 C. Teodosiu and Z. Hu (1998). Microstructure in the continuum modelling of plastic anisotropy.
Proceedings of the 19th Riso Int. Symp. on Materials Science: Modelling of Structure and
Mechanics of Materials from Microscale to Products, J. V. Carstensen and T. Leffers (Eds.).
141 L. S. Tóth and P. Van Houtte (1992). Discretization techniques for orientation distribution
functions. Textures and Microstructures, 19, 229–244.
142 Z. Tourki, A. Zeghloul and G. Ferron (1996). Sheet metal forming simulations using a new
model for orthotropic plasticity. Computational Materials Science, 5, 255–262.
143 H. Tresca (1864). On the yield of solids at high temperature (in French). Comptes Rendus
Academie des Sciences, 59, Paris, 754.
144 A. Van Bael (1994). Anisotropic yield loci derived from crystallographic data and their applica-
tion in finite element simulations of plastic forming processes, proefschrift voorgedragen tot het
behalen van het doctoraat in de toegepaste wetenschappen, Katholieke Universiteit Leuven.
145 A. Van Bael, J. Winters and P. Van Houtte (1996). A semi-analytical approach for incorporating
crystallographic data into elasto-plastic finite element formulations, Textures of Materials.
Proceedings of the 11th Int. Conf. on Textures of Materials, 1, ICOTOM-11, Sept. 16-20, Z.
Liang, L. Zuo and Y. Chu (Eds.).
146 A. Van Bael and P. Van Houtte (2002). Convex Fourth and Sixth-Order Plastic Potentials
derived from Crystallographic Texture, Proc. of the 6th European Mechanics of Materials
Conference (EMMC6), Liege Cescotto editor.
147 P. Van Houtte (1988). A comprehensive mathematical formulation of an extended Taylor-
Bishop-Hill model featuring relaxed constraints, the Renouard-Wintenberger theory and a
strain rate sensitivity model. Textures and Microstructures, 8-9, 313–350.
148 P. Van Houtte, K. Mols, A. Van Bael and E. Aernoudt (1989). Application of yield loci calcu-
lated from texture data. Textures and microstructures, 11, 23–39.
149 P. Van Houtte (1992). Anisotropic Plasticity, Numerical Modelling of Material Deformation
Processes. Research, Development and Applicaitons, P. Hartley snd I. Pillinger (Eds.), Springer
Verglag.
150 P. Van Houtte (1994). Application of plastic potentials to strain rate sensitive and insensitive
anisotropic materials. Int. J. Plasticity, 10, 719–748.
151 P. Van Houtte (1995). Micromechanics of polycrystalline materials. Chaire Francqui, Université
de Liege.
152 P. Van Houtte (1996). Microscopic strain heterogeneity and deformation texture prediction,
Textures of Materials. Proceedings of the 11th Int. Conf. on Textures of Materials, 1, ICOTOM-
11, Z. Liang, L. Zuo and Y. Chu (Eds.).
153 P. Van Houtte, L. Delannay and I. Samajdar (1999). Quantitative prediction of cold rolling
textures in low-carbon steel by means of the LAMEL model. Texture and Microstructure, 31,
109–149.
154 P. Van Houtte (2001). Fast calculation of average Taylor factors and Mandel spins for all
possible strain modes. Int. J. Plasticity, 17, 807–818.
155 P. Van Houtte, L. Delannay and S. R. Kalidindi (2002). Comparison of two grain interaction
models for polycrystal plasticity and deformation texture prediction. Int. Jour of Plasticity,
18, 359–377.
156 H. Vegter, Y. An, H. H. Pijlman and J. Huetink (1999a). Different approaches to describe
the plastic material behaviour of steel and aluminium-alloys in sheet forming. 2nd ESAFORM
Conference on Material Forming, Guimaraes, Portugal, J. A. Covas (Ed.).
157 H. Vegter, Y. An, H. H. Pijlman and J. Huetink (1999b). Advanced Mechanical Testing on
Aluminium Alloys and Low Carbon Steels for Sheet Forming. Numisheet’99, 13-17 September
1999, 1, Besançon, France, J. C. Gélin and P. Picart (Eds.), Université de Franche-Comté 3–8.
96 A.M. Habraken

158 C. Vial, R. M. Caddell and W, F. Hosford (1983). Yield loci of anisotropic sheet metals. Int.
J. of Mech. Sci., 25, 899.
159 N. Wang, F. R. Hall, I. Pillinger, P. Hartley and C. E. N. Sturgess (1992). Finite-element
prediction of texture evolution in material forming. Numerical Methods in Industrial Forming
Processes, Chenot, Wood and Zienkiewicz, (Eds.), 193.
160 B. Weber, A. Carmet, M. Duchêt, B. Bomprezzi, J. Bourgasser, J. L. Robert (1999). Estimation
de durée de vie sous chargement quelconque: application a un composant automobile. Actes
de la conférence de la Société Française de Métallurgie et de Matériaux “Dimensionnement en
fatigue des structures, Démarches et Outils”, Paris 2-3 juin.
161 J. Winters (1996). Implementation of a texture-based yield locus into an elastoplastic finite el-
ement code. Application to sheet forming. Katholieke Universiteit Leuven, proefschrift voorge-
dragen tot het behalen van het doctoraat in de toegepaste wetenschappen, Katholieke Univer-
siteit Leuven.
162 G. Winther, D. J. Jensen and N. Hansen (1997). Modelling flow stress anisotropy caused by
deformation induced dislocation boundaries. Acta Mater., 45(6), 2455–2465.
163 Y. Y. Zhu and S. Cescotto (1996). Unified and mixed formulation of the 8-node hexahedral
elements by assumed strain method. Comput. Methods Appl. Mech. Engrg., 129, 177–209.

Please address your comments or questions on this paper to:


International Center for Numerical Methods in Engineering
Edificio C-1, Campus Norte UPC
Grand Capitán s/n
08034 Barcelona, Spain
Phone: 34-93-4016035; Fax: 34-93-4016517
E-mail: onate@cimne.upc.es

S-ar putea să vă placă și